You are on page 1of 187

The Pennsylvania State University

The Graduate School

College of Agricultural Sciences


IRON AND ALUMINUM HYDROXIDE NANOPARTICLES IN THE
ENVIRONMENT: FROM NANO-SCALE TO THE FIELD PROCESSES

A Dissertation in

Soil Science and Biogeochemistry

by

Ekaterina Bazilevskaya
2009 Ekaterina Bazilevskaya
Submitted in Partial Fulfillment
of the Requirements
for the Degree of
Doctor of Philosophy

December 2009

The dissertation of Ekaterina Bazilevskaya was reviewed and approved* by the


following:

Carmen Enid Martnez


Associate Professor of Environmental and Soil Chemistry
Dissertation Advisor
Chair of Committee

Douglas Archibald
Research Associate in Agricultural Analytical Chemistry

Edward Ciolkosz
Professor Emeritus of Soil Genesis and Morphology

James Kubicki
Professor of Geosciences

Kwadwo Osseo-Asare
Distinguished Professor of Metallurgy and Energy and Geo-environmental
Engineering

David Sylvia
Professor of Soil Microbiology
Head of the Department of Crop and Soil Sciences

*Signatures are on file in the Graduate School

ii

ABSTRACT
The objective of this doctoral research was to increase scientific understanding of
the behavior of Fe and Al hydroxide nanoparticles in soils. These particles are of great
environmental importance due to their ability to retain and transport nutrients and
contaminants. Three studies were undertaken at different scales, which are documented in
three manuscripts included in this dissertation
The first study examined the rate constants for goethite (-FeOOH) crystallization
from nano-particulate Fe hydroxide suspensions in the absence (0% Al) and presence
(2% Al) of aluminum. One of the merits of this study was the application of a
multivariate curve resolution analysis (MCR) of infrared spectra to environmentally
important mixed Fe-Al hydroxide colloids in order to quantify goethite content in poorlycrystalline mixtures. Obtained rate constants were found to be equal to (7.640.67)10-7
s-1 for 0% Al and (4.50.21) 10-7 s-1 for 2% Al hydroxides. Dissolution-precipitation
mechanism was dominant in the process of goethite transformation to ferrihydrite.
Further growth of goethite crystals took place either by aggregation mechanism to form
polycrystalline agglomerates or alternatively by Oswald ripening to form large single
crystals. The presence of aqueous Al species poisoned goethites surface by disrupting
the formation of hydrogen bonds thus increasing the number of non-stoichiometric
hydroxyls.
The second study addressed changes of mineral composition in mixed Fe-Al
hydroxide nanoparticles as a function of Al-substitution and reaction time. It was found
that low Al concentrations (2-8 mol. %) lead to formation of moderately crystalline Aliii

goethite upon ageing, while at medium Al concentrations (10-20%) colloidal suspensions


remained stable for the duration of the whole experiment (54 days), goethite formation
was completely retarded, and less crystalline intermediate structure were formed. At 25%
Al substitution, gibbsite Al(OH)3 microcrystalline structures appeared within the first
days of experiment. In addition, to understand the mechanism of Al substitution in
goethite, we explored the most energetically favorable arrangement of Al atoms within
goethite by ab initio periodic density functional theory (DFT) calculations. These
calculations showed that Al may form Al-O-Al clusters as opposed to evenly distributed
isolated Al atoms (Al-O-Fe) in goethite structure.
The third study was conducted to investigate the relative importance of Fe and Al
hydroxide nanoparticles in migration and accumulation of these elements in soils. I
approached this goal by studying soil water and coatings on mineral grains from
Spodosol soil, which is characterized by intensive leaching of Fe, Al, and organic matter
(OM) and their accumulation in Spodosol profile. While Fe, Al and Si were mostly
transported as inorganic colloids, Al also showed close association with organic matter.
Fe, Al, and Si have the highest mobility in organic-rich A and Bh horizons of the
spodosol profile, which suggests that the presence of organic matter facilitates transport
of these elements by stabilizing the inorganic colloids. The two major mechanisms of
immobilization of Fe, Al, Si and OM are (1) polymerization of metal-OM complexes and
(2) surface charge neutralization of OM-inorganic colloidal aggregates. Both of these
processes presumably occur when the (Fe+Al) to C ratios in the colloidal fraction
increase in Bh horizon.
iv

The presented research findings contribute to our understanding of the


fundamental properties of mixed Fe- and Al-hydroxide nanoparticles and also help to
predict the interactions and environmental fate of natural nanocolloids in soil profile.

TABLE OF CONTENTS
LIST OF FIGURES .................................................................................................................ix
LIST OF TABLES...................................................................................................................xiii
ACKNOWLEDGMENTS .......................................................................................................xiv
Chapter 1 INTRODUCTION...................................................................................................1
1.1. Background ...............................................................................................................1
1.2. Dissertation structure ................................................................................................2
References........................................................................................................................3
Chapter 2 NANO-GOETHITE CRYSTALLIZATION IN THE PRESENCE OF LOW
ALUMINUM CONCENTRATIONS UNDER ENVIRONMENTALLY
RELEVANT CONDITIONS ...........................................................................................5
Abstract ............................................................................................................................5
2.1. Introduction...............................................................................................................6
2.2. Experimental Section ................................................................................................8
2.2.1. Synthesis of Fe-hydroxide and Al-doped Fe-hydroxide nanoparticles ..........8
2.2.2. Synchrotron-based X-ray diffraction..............................................................10
2.2.3. ATR-FTIR spectroscopy .................................................................................10
2.2.3.1. ATR-FTIR data collection and Gaussian band analyses .....................10
2.2.3.2. Multivariate curve resolution (MCR) analysis of ATR-FTIR
spectra.......................................................................................................12
2.2.4. Dynamic light scattering (DLS) .....................................................................13
2.2.5. Transmission electron microscopy (TEM) .....................................................13
2.3. Results.......................................................................................................................14
2.3.1. XRD data........................................................................................................14
2.3.2. Infrared spectra..............................................................................................15
2.3.2.1. OH-bending vibrations ........................................................................15
2.3.2.2. OH-stretching vibrations: Gaussian deconvolution.............................15
2.3.2.3. MCR analysis of IR O-H stretching region.........................................18
2.4. Discussion .................................................................................................................20
2.4.1. Mechanisms for nano-goethite crystallization ...............................................20
2.4.2. Rate constants for ferrihydrite-goethite transformation ................................23
2.5. Conclusions...............................................................................................................25
Acknowledgments............................................................................................................26
References........................................................................................................................26
Figures..............................................................................................................................32
Chapter 3 ATR-FTIR, XRD AND PERIODIC-DENSITY-FUNCTIONAL
THEORY STUDIES OF MINERAL PHASES IN CO-PRECIPITATED FE
AND AL HYDROXIDE ..........................................................................................46
Abstract ............................................................................................................................46
3.1. Introduction...............................................................................................................47
3.2. Methods.....................................................................................................................50
3.2.1. Experimental ..................................................................................................50
vi

3.2.1.1. Synthesis and aging of Fe-hydroxide and Al-doped Fe-hydroxide


nanoparticles.............................................................................................50
3.2.1.2. Characterization of Fe-hydroxide and Al-doped Fe-hydroxide
nanoparticles.............................................................................................51
3.2.2. Periodic density functional theory calculations of bulk Al-goethite model
structures ..........................................................................................................52
3.3 Results and discussion ...............................................................................................54
3.3.1. Phase transformation in Fe-Al hydroxide nano-particle suspensions ...........54
3.3.2. Arrangement and concentration of Al in Al-goethites: energies and unit
cell parameters from DFT calculations ...........................................................57
3.3.3. Goethite versus Al-goethites: stability considerations ...................................60
3.3.4. Relative abundance of gibbsite and diaspore in soils ....................................62
3.4. Conclusions...............................................................................................................64
Acknowledgements ..........................................................................................................65
References........................................................................................................................65
Tables ...............................................................................................................................69
Figures..............................................................................................................................72
Chapter 4 SPECTROSCOPIC AND MICROSCOPIC INVESTIGATIONS OF THE
COMPOSITION OF SOIL WATER AND ACCUMULATIONS FORMED IN
SITU IN A SPODOSOL ..................................................................................................81
Abstract ............................................................................................................................81
4.1. Introduction...............................................................................................................82
4.2. Methods.....................................................................................................................84
4.2.1. Field site description and samples .................................................................84
4.2. 2. Analyses of soil solids ...................................................................................85
4.2.2.1. Soil pH and selective extractions ........................................................85
4.2.2.2. Total Organic Carbon (TOC) and Total Nitrogen ...............................86
4.2.2.3. X-ray diffraction (XRD) of clay size fraction (< 2 m) from
Spodosol horizons.....................................................................................87
4.2.3. Analyses of soil waters ...................................................................................87
4.2.3.1. Soil water extraction............................................................................87
4.2.3.2. Dissolved organic carbon (DOC) and Colloidal organic carbon
(COC) .......................................................................................................88
4.2.3.3. Attenuated Total Reflectance Fourier Transform Infrared ATRFTIR .........................................................................................................89
4.2.3.4. Scanning electron microscopy - Energy dispersive spectroscopy
(SEM-EDS) ..............................................................................................89
4.2.3.5. Transmission electron microscopy (TEM) ..........................................90
4.2.4. In situ coating formation within the Spodosol profile ....................................90
4.2.4.1. Experimental setup ..............................................................................90
4.2.4.2. Scanning Electron Microscopy - Energy Dispersive Spectroscopy
(SEM-EDS) ..............................................................................................90
4.2.4.3. Fe-GIXAS (Gracing Incidence X-ray Absorption Spectroscopy).......91
4.3. Results.......................................................................................................................92
4.3.1. Characteristics of Soil Solids .........................................................................92
4.3.2. Soil water: dissolved and colloidal fraction...................................................93
4.3.2.1. Dissolved concentration of Fe, Al, Si and C .......................................94
vii

4.3.2.2. Concentrations of Fe, Al, Si and C in colloidal fraction of soil


waters........................................................................................................94
4.3.2.3. ATR-FTIR analysis of soil water ........................................................95
4.3.2.4. SEM-EDS analyses of the colloidal fraction (0.45-1.2 m particles
size) of soil waters ....................................................................................98
4.3.2.5. TEM-EDS of the colloidal fraction of soil waters...............................99
4.3.3. Characteristics of in situ coatings formed in the Spodosol profile ................100
4.3.3.1. SEM-EDS of coatings of wafers .........................................................100
4.3.3.2. Fe-EXAFS analysis of quartz wafers ..................................................102
4.4. Discussion .................................................................................................................103
4.4.1. Transport and accumulation forms of Fe, Al, and Si in Spodosols................103
4.4.1.1. Aluminum............................................................................................103
4.4.1.2. Iron ......................................................................................................104
4.4.1.3. Silica....................................................................................................105
4.4.2. Role of organic matter in podzolization process............................................106
4.4.2.1. Transport forms ...................................................................................106
4.4.2.2. Immobilization ....................................................................................107
4.4.5. Environmental significance of this study........................................................108
4.5. Conclusions...............................................................................................................108
Acknowledgements ..........................................................................................................109
References........................................................................................................................109
Tables ...............................................................................................................................115
Figures..............................................................................................................................117
Chapter 5 CONCLUSIONS AND FUTURE WORK..............................................................129
APPENDIX A. Composition and morphology of soil colloids (0.45m to 1.2 m) by
SEM-EDS.........................................................................................................................131
APPENDIX B. SEM-EDS analyses of quartz wafers implanted in the Spodosol soil ............143
APPENDIX C. X-ray diffraction (XRD) of Spodosol clay fraction (< 2 m) ........................164

viii

LIST OF FIGURES
Figure 2-1. X-ray diffractograms for FeIII-based nanocolloids synthesized with 0 or 2 mol%
incorporated-Al at initial conditions (0 days) and after aging in dilute suspensions for 2, 9,
23, and 54 days. Go stands for goethite, numbers in parenthesis indicate crystal face; Fh
denotes ferrihydrite. ............................................................................................................... 32
Figure 2-2. Crystallite sizes for goethite as a function of time. Crystallite sizes were estimated
from the XRD (110) peak using the Scherrer equation (0%Al, diamonds; 2% Al, squares). 33
Figure 2-3. (A) Crystal structure of goethite. Fe is in octahedral coordination surrounded by O2and OH- ligands (O, red; H, white). (B) Closer look showing hydrogen bonds, O-HO
(dashed lines). ........................................................................................................................ 34
Figure 2-4. ATR-FTIR spectra as a function of time for Fe-hydroxide suspensions with 0% Al
(upper panel) and 2% Al (lower panel). Multiple lines at a given time reflect both batch and
instrument replicate variability. ............................................................................................. 35
Figure 2-5. (A) Example of Gaussian band analysis for the OH-stretching region (0%Al
suspensions: upper plot, suspension aged for 54 days; bottom plot, initial suspension at 0
days). Solid lines are experimental spectra, dotted lines are the Gaussian components, and
broken lines are the sum of the Gaussian components. Stoichiometric OH peak position (B)
and full width at half maximum (FWHM) (C) as a function of time for 0%Al and 2% Al
suspensions. ........................................................................................................................... 36
Figure 2-6. MCR analysis of OH-stretching region. (A) Components extracted from the 0% Al
experimental dataset; these two components describe all spectral variability. (B) Scores
(relative ratios) for Component 1 (ferrihydrite-like) and Component 2 (goethite-like) as
a function of time. (C) Gaussian band analysis indicates each component (for both 0% and
2% Al suspensions) is consistent with combination of stoichiometric hydroxyls, nonstoichiomeric hydroxyls, and adsorbed water........................................................................ 37
Figure 2-7. (A) Particle sizes as determined by dynamic light scattering for 0%Al (upper panel)
and 2%Al (bottom panel) suspensions. Dark grey symbols show the results for a suspension
measured after aging for 0, 2, 9, 23 and 54 days. Light symbols show data obtained by insitu heating of an initial suspension (0 days) at 50 C every 2 hours for about 2 days in
automatic mode. (B) Volume-based particle size distributions for 0%Al (upper panel) and
2%Al (bottom panel) suspensions aged for 0, 2, 9, 23, and 54 days (indicated by numbers
with arrows). .......................................................................................................................... 38
Figure 2-8. Representative TEM images of iron hydroxide nanoparticles. Images A and B show
non-aged particles (0 days) from 0%Al and 2%Al suspensions, respectively. TEM revealed
small (<~30 nm) goethite needles in Al-free suspensions at 0 days (insert in A), while no
goethite needles were observed in 2%Al suspensions (B). (C) Al-free suspensions aged for
54 days exhibit three populations of nanoparticles: round particles less than 10 nm in
diameter (ferrihydrite), thin needles of goethite about 50 nm in length and several nm wide,
and large crystals (prisms) of goethite 50-100 nm in length and 10-20 nm in diameter. (D)
Ferrihydrite particles are more abundant in 2%Al suspensions aged for 54 days although a
few goethite needles and prisms are observed....................................................................... 39

ix

Figure 2-9. TEM images of diluted samples. Images A and B show non-aged particles (0 days)
from 0%Al and 2%Al suspensions, respectively; C and D - Al-free and 2%Al suspensions
aged for 54 days, respectively................................................................................................ 40
Figure 2-10. High-resolution TEM images of 0% Al suspensions aged for 0 days (A) and 54 days
(B and C). (A) A needle formed from aggregated particles (0 days). (B) Large goethite
crystal shows single-crystal lattice fringes in agreement with a spot diffraction pattern
(insert). (C) Goethite needle consisted of smaller grains with different lattice orientation.
Diffraction pattern produced diffraction rings characteristic of polycrystalline materials. ... 41
Figure 2-11. Schematic diagram illustrating the transformation of ferrihydrite to goethite. ........ 42
Figure 2-12. Plots of ATR-FTIR- and XRD- derived parameters for 0%Al and 2%Al suspensions
as a function of time. These parameters reflect the presence of the crystallization product
(goethite). Solid circles for 0% Al and open circles for 2% Al suspensions. Solid (0% Al)
and dotted (2% Al) lines are data fits using the first order equation [Goethite] = [Goethite]0 *
(1 exp-kt). .............................................................................................................................. 43
Figure 2-13. Rate constants for goethite appearance obtained from ATR-FTIR- and XRDderived parameters for 0%Al and 2%Al suspensions............................................................ 44
Figure 2-14. Comparison of first-order rate constants for goethite crystallization obtained in this
work to available literature data. Authors are indicated by numbers: 1. Shaw et al. (2005); 2.
Nagano et al. (1994); 3. Fischer (1971); 4. Yee et al.(2006); 5. Ford et al. (1999, 2005).
Methods used in calculating the rate constants: - synchrotron XRD; - oxalate extraction;
x - colorimetry. The aging temperature (oC) at which experiments were conducted is also
indicated................................................................................................................................. 45
Figure 3-1. Titration curves for Fe-Al mixed solutions adjusted to pH 5 by addition of 0.1 M
KOH. Numbers on the plots show initial mol% Al of the 10 mmol [Al + Fe] in solution.
Titration curves of solutions with 30%, 50% and 75% Al have 2 inflection points which
indicate the formation of two separate phases (Fe-hydroxides and Al-hydroxides).............. 72
Figure. 3-2. Millimoles added hydroxide at half-conversion pH points (see Fig. 3-1) versus mol%
Al in initial solutions.............................................................................................................. 73
Figure 3-3. Aluminum in initial preparation solutions versus that observed in dialyzed nanoparticle suspensions. The solid line indicates a 1:1 ratio. ...................................................... 73
Figure 3-4. ATR-FTIR spectra of particles from Fe-Al-hydroxide suspensions (A) at initial
conditions (0 days) and (B) after aging at 50 C for 54 days. Numbers in the center of each
spectrum indicate the initial mol% Al as a fraction of [Fe + Al] in suspension.
Abbreviations: Go, goethite; Gib, gibbsite . .......................................................................... 74
Figure 3-5. XRD patterns for particles from Fe-Al-hydroxide suspensions: (A) at initial
conditions (0 days), and (B) after aging at 50 C for 54 days. Numbers between panels
indicate mol% Al in suspensions as a fraction of [Fe + Al]. Abbreviations: Fh, ferrihydrite;
Go, goethite; Hm, hematite; Gib, gibbsite. ............................................................................ 75
Figure 3-6. Theoretical solubility curves for (A) iron hydroxides and (B) aluminum hydroxides in
equilibrium with total aqueous Fe(III) and Al(III) species, respectively. Solubility curves
were constructed using Ksp values for Fe-hydroxides from Stumm and Morgan (1981), and
for gibbsite and diaspore from Peryea and Kittrick (1988). Shaded areas in panel B show Al
concentrations in our synthetic mixed oxide nanoparticles, with numerical ranges indicating
the corresponding mol%Al as a fraction of [Fe + Al]. .......................................................... 76
x

Figure 3-7. Optimized structural models for goethite with (A) 2 clustered Al atoms, and (B) 2
isolated Al atoms. The distance between two Al atoms is 3.15 and 6.73 in clustered and
isolated structures, respectively ............................................................................................. 77
Figure 3-8. Optimized structural models for goethite with (A) 6 clustered Al atoms, and (B) 6
isolated Al atoms. The distances among Al atoms are 3-3.5 and 4.5-5.5 in clustered and
isolated structures, respectively. ............................................................................................ 78
Figure 3-9. Experimental and calculated unit cell volume (A) and cell parameters a (B), b (C) and
c (D) for goethite and Al-substituted goethites as a function of mol% Al substitution and
type of substitution (clustered vs. isolated). Model calculations for clustered substitution
(upward pointing triangles) yield parameters that are closer to measured values.
Experimental unit cell parameters were obtained from Szytula et al. (1968); Schulze (1984);
Kosmas et al. (1986); Fazey and O'Connor (1991); Piszora and Wolska (1998); Gualtieri and
Venturelli (1999); Scheinost et al. (2001); Ruan et al. (2002); Majzlan and Navrotsky
(2003); Wells et al. (2006); Alvarez et al. (2007); Blanch et al. (2008). Thin solid lines
indicate theoretical dependence of goethite unit cell parameters on mol% of Al substitution
in goethite (Vegards law) ..................................................................................................... 79
Figure 3-10. VASP calculated free energies for iso-structural goethite and diaspore and their Alcontaining solid solutions ...................................................................................................... 80
Figure 4-1. Placement of quartz wafers into soil profile............................................................. 117
Figure 4-2. Fe, Al and Si forms in soil obtained from selective extractions............................... 118
Figure 4-3. Dissolved (<0.45 m soil water filtrates) concentrations of (A) Fe, Al, Si, and (B)
carbon. Also shown are the results from thermodynamic calculations of (C) Al and (D) Fe
speciation. Fe DOM1 and Al DOM1 denote Fe and Al complexes with dissolved organic
matter, respectively.............................................................................................................. 119
Figure 4-4. Concentrations of (A) Fe and Al, (B) Si, and (C) carbon and nitrogen in the colloidal
fraction (particles within the 0.45 m to 1.2 m size range) of soil water filtrates. (D)
Atomic ratios of (Fe+Al) to carbon present in the colloidal fraction of soil waters ........... 120
Figure 4-5. ATR-FTIR spectra of mineral and organic standards .............................................. 121
Figure 4-6. ATR-FTIR spectra of soil water extracted from A, E, Bh, Bhs and C horizons. A <0.45 m size fraction, and B - <1.2 m size fraction......................................................... 122
Figure 4-7. Typical SEM images of soil colloids (0.45-1.2 m) on 0.45 m pore size silver filter.
A. Blank filter; B. Colloids extracted from A-horizon; C. Colloids from Bh horizon; insert
are magnified particles from larger image; D. Colloids from Bhs horizon with Fe-enriched
particle. ................................................................................................................................ 123
Figure 4-8. Chemical composition of soil colloids (0.45 m< size <1.2 m) obtained from SEMEDS measurements for A, Bh, and Bhs horizons. Shown are the distribution of
concentrations (atomic %) of major elements and (Fe+Al)/C ratios in Spodosol horizons. Yaxis values indicate frequencies of occurrence versus concentration range ........................ 124
Figure 4-9. TEM-EDS data of colloids (0.45 to 1.2 m) from Bh horizon. A. Amorphous silica
particles ~100 nm size; B. Carbon-rich amorphous round structures with some traces of Fe
and Al; C. Organic matter with some crystalline inorganic particles containing Si, Fe, Ca and
Al as suggested by diffraction pattern and EDS spectra...................................................... 125
xi

Figure 4-10. TEM-EDS data of colloids (0.45 to 1.2 m) from Bhs horizon containing
amorphous organo-mineral colloids containing C, Al (major elements) and traces of Si, Ca,
Fe, P, and S .......................................................................................................................... 126
Figure 4-11. Fourier transforms of Fe-EXAFS spectra of Fe-organic standards (Fe-citrate, FeEDTA and Fe-catechol), Fe-mineral standards (goethite and hematite) and field samples.
These plots show inter atomic distances from Fe atom to its nearest neighbors (not corrected
for phase shifts).................................................................................................................... 127
Figure 4-12. Schematic diagram depicting suggested composition of metal-organo-complexes in
soil water. Two-end members are shown: mineral particles with organic groups attached to
their surface (left) and organic particles bridged by Fe and Al ions (right). The relative
proportion of these two end-members and composition of the intergrades between them
depends on the supply of metals and organics to the soil water. ......................................... 128
Figure A-1. SEM images of colloids (0.45-1.2 m) from A horizon ......................................... 133
Figure A-2. SEM images of colloids (0.45-1.2 m) from Bh horizon........................................ 135
Figure A-3. SEM images of colloids (0.45-1.2 m) from Bhs horizon ...................................... 138
Figure B-1-a. SEM image and EDS spectra of A-horizon, area1. .............................................. 145
Figure B-1-b. SEM image, EDS spectra and elemental map of A-horizon, area 2. ................... 146
Figure B-1-c. SEM image and elemental map of A-horizon, area 3........................................... 147
Figure B-1-d. SEM image, EDS spectra and elemental map of A-horizon, area 4. ................... 148
Figure B-1-e. SEM image, EDS spectra of A-horizon, area 5. ................................................... 149
Figure B-2-a. SEM image and EDS spectra of Bh-horizon, area 1. ........................................... 150
Figure B-2-b. SEM image, EDS spectra of Bh-horizon, area 2.................................................. 150
Figure B-2-c. SEM image and EDS spectra of Bh-horizon, area 3 ............................................ 151
Figure B-2-d. SEM image and EDS spectra of Bh-horizon, area 4. ........................................... 151
Figure B-2-e. SEM image and EDS spectra of Bh-horizon, area 5 ............................................ 152
Figure B-2-f. SEM image and EDS spectra of Bh-horizon, area 6............................................. 153
Figure B-2-f. SEM image and EDS spectra of Bh-horizon, area 6 (magnified). ........................ 154
Figure B-2-g. SEM image and EDS spectra of Bh-horizon, area 7 ............................................ 155
Figure B-2-h. SEM image and EDS spectra of Bh-horizon, area 8 ............................................ 156
Figure B-3-a SEM image and EDS spectra of Bhs-horizon, area 1............................................ 157
Figure B-3-b SEM image and EDS spectra of Bhs-horizon, area 2 ........................................... 158
Figure B-3-c SEM image and EDS spectra of Bhs-horizon, area 3............................................ 159
Figure B-3-d SEM image and EDS spectra of Bhs-horizon, area 4 ........................................... 160
Figure B-3-e SEM image and EDS spectra of Bhs-horizon, area 5............................................ 161
Figure B-3-f SEM image and elemental maps of Bhs-horizon, area 6 ....................................... 162
Figure B-3-g SEM image and EDS spectra of Bhs-horizon, area 7............................................ 163
xii

Figure B-3-h. SEM image and EDS spectra of Bhs-horizon, area 8 .......................................... 163
Figure C-1. XRD spectra of different treatments for the clay fraction (< 2 m) from A-horizon
............................................................................................................................................. 168
Figure C-2. XRD spectra of different treatments for the clay fraction (< 2 m) from E-horizon
............................................................................................................................................. 169
Figure C-3. XRD spectra of different treatments for the clay fraction (< 2 m) from Bh-horizon
............................................................................................................................................. 170
Figure C-4. XRD spectra of different treatments for the clay fraction (< 2 m) from Bhs-horizon
............................................................................................................................................. 171
Figure C-5. XRD spectra of different treatments for the clay fraction (< 2 m) from C-horizon
............................................................................................................................................. 172

LIST OF TABLES
Table 3-1. ATR-FTIR- and XRD-determined mineral phases in mixed Fe and Al
hydroxide nano-particle suspensions as a function of aging time and initial Al moles
as a fraction of total Al and Fe. Abbreviations: Fh, ferrihydrite, Go, goethite, Hm,
hematite, Gib, gibbsite. Numbers represent estimates of relative peak intensity
(higher numbers for higher intensities). .................................................................... 69
Table 3-2. Calculated and experimental crystallographic parameters and calculated free
energy for goethite, Al-goethites, diaspore and gibbsite........................................... 70
Table 3-3. Relative stability of Al-substituted goethites and diaspore with respect to
goethite (Hreaction) .................................................................................................... 71
Table 3-4. Relative stability of gibbsite and diaspore with respect to corundum (Hreaction)
................................................................................................................................... 71
Table 4-1. Description of soil horizons of Black Moshannon Spodosol site................. 115
Table 4-2. Infrared band assignments in the 4000-600 cm-1 region ............................... 116
Table A-1. Elemental composition of colloids (0.45m <size<1.2 m) on silver filters as
derived from SEM-EDS analysis. Elements are expressed in atomic percentages. All
areas analyzed were morphologically similar (see Figures A4-1-1, A4-1-2, and A41-3). ......................................................................................................................... 131
Table B-1. Elemental composition of accumulations on quartz wafers. Only the presence
and relative intensity (tr trace, x - medium, X - high) is presented ...................... 143
Figure B-1-a. SEM image and EDS spectra of A-horizon, area1.................................. 145
xiii

ACKNOWLEDGMENTS

First of all I would like to express my greatest appreciation to my academic


adviser, Dr. Carmen Enid Martinez for her guidance and constant support. It was a
pleasure to work with her. I am also deeply indebted to my PhD Committee members
Douglas Archibald, Edward Ciolkosz, James Kubicki, and Kwadwo Osseo-Asare for the
invaluable advice throughout my research and the comments to my manuscripts. They
helped me immensely to believe in myself as a scientist. I also want to thank Ephraim
Govere for his advice and discussions on different analytical approaches. I was really
lucky to be a Fellow of the Center for Environmental Kinetics Analysis (CEKA) at Penn
State, which gave me an opportunity to communicate with many bright scientists at and
outside Penn State and to present my work in the interdisciplinary environment.
On a more personal note, I would like to thank my children Nikita, Natasha, and
Kristina for their unselfish understanding and for keeping me sane during the years of my
graduate studies. Finally, I thank my husband Mark Fedkin for his constant
encouragement, support, and sacrifice.
Funding for this work was provided by the National Science Foundation, Grant #
CHE-0431328, and The Penn State Biogeochemical Research Initiative for Education
(BRIE) under National Science Foundation Grant # DGE-9972759.

xiv

Chapter 1
INTRODUCTION
1.1. Background
Migration of mineral nanoparticles through the soil profile and their interaction with
dissolved elements, organic matter and soil minerals are important processes in soil development
(Hochella et al., 2008, Banfield and Zhang, 2001). Due to a high specific surface area,
nanocolloids are known to have a greatly increased reactivity towards many geochemically
important solute species and hence are largely responsible for the retention and transport of metal
ions, contaminants, and nutrients in the environment (Zhang, 2003; Madden et al., 2006;
Novikov et al., 2006; Wigginton et al., 2007). Some of their distinct properties are believed to
result from variations (or distortions) in their atomic structure at the nano-scale (Waychunas et
al., 2005). It is also known that thermodynamic properties such as surface enthalpies, enthalpies
of mineral phase transition, and aqueous solubility may be functions of particle size, favoring the
stability of particular minerals at different size ranges (Navrotsky, 2008).
Among the mineral nanoparticles that exist in soils, iron and aluminum hydroxides
(Fe(OH)3, Al(OH)3,) are the most abundant and environmentally relevant. These low-crystalline
materials of colloidal size (<0.1 m) are formed as a result of soil weathering processes and
precipitate as metastable Fe-Al hydroxide mixtures. The stability of such hydroxide
nanoparticles, both with respect to aggregation and phase transformation, can be quite different
compared to their micron-size analogues. It should be recognized that there is a substantial
amount of research aimed specifically at studying the effects of Al substitution on the
crystallization of Fe-hydroxides (Fey and Dixon, 1981; Colombo and Violante, 1996; Blanch et
al., 2008), however, most of these studies were conducted at experimental conditions far from
environmental.
1

Some of the questions that need to be addressed are: (1) what is the mechanism of
goethite crystallization from amorphous Fe-hydroxide (ferrihydrite) precipitated in the presence
of Al, and how Al influences the rate of ferrihydrite-goethite transformation? (2) how does Al
incorporate in the structure of goethite during crystallization? (3) are co-precipitated hydroxides
more stable compared to pure Fe- and Al-hydroxide nanoparticles? (4) how do Fe- and Alhydroxide nanoparticles form, migrate and interact in complex soil environments?
To address these questions, in my dissertation I use a combination of laboratory
experiments, field work, and molecular modeling to explore the issues of Fe and Al hydroxide
nano-particle crystallization, transformation and interaction in soil environments.

1.2. Dissertation structure


This dissertation consists of three main chapters that will be submitted for publication in
peer reviewed journals. The chapter titled Nano-goethite (-FeOOH) crystallization in the
presence of low aluminum concentrations under environmentally relevant conditions (Chapter
2) focuses on quantification of the rate of ferrihydrite-goethite transformation in nano-colloidal
suspensions in the presence of low concentrations of Al, and on understanding the mechanisms
of goethite crystallization. My hypothesis was the following: even low Al concentrations in
nano-colloidal suspensions may significantly decrease the rate of goethite crystallization from
ferrihydrite suspensions. To quantify the rate of this transformation, we applied a novel
analytical approach to estimate goethite contents in nano-colloidal suspensions using infrared
data.
Chapter 3, Co-precipitation of Fe and Al hydroxide nanoparticles: ATR-FTIR, XRD and
Periodic Density Functional Theory Calculations tests the following hypotheses: (1) the kinetics
of crystallization of minerals (goethite and/or gibbsite) from mixed Fe-Al hydroxide
2

nanoparticles is much slower than that of pure phases due to the formation of intermediate
phases that are indefinitely metastable in low-temperature soil environments; (2) Al incorporates
into goethite by forming Al-clusters, which is energetically more favorable compared to isolated
substitution of Al for Fe; however, even in clustered arrangement, Al creates strain on the
goethite structure which delays its further crystallization. To test these hypotheses, we used a
combination of experimental data and ab initio molecular modeling to follow phase
transformations in co-precipitated Fe-Al hydroxides with a wide range of Fe/Al ratios.
Finally, Chapter 4 (Mobility and accumulation of Fe, Al and Si organic and inorganic
forms in soil and soil solutions) is an attempt to go across scales and extend knowledge on the
behavior of Fe- and Al-hydroxides nanoparticles to soil environments. For this purpose, we
studied a spodosol characterized by intensive transport and accumulation of Fe and Al within the
soil profile. We hypothesized that: in this soil, Fe and Al migrate from surface horizons to the
subsoil in the form of inorganic colloids (Fe and Al hydroxides, short-range aluminosilicates)
that accumulate in lower profiles due to interaction with soil organic matter. To test the
hypothesis, we used various spectroscopic and microscopic techniques to explore the
composition of soil water and soil coatings.

References
1. Banfield, J.F. and Zhang, H. (2001) Nanoparticles in the environment. In: Nanoparticles
and the environment. Eds: Banfield, J.F. and Navrotsky, A. Reviews in mineralogy and
geochemistry, V. 44, pp. 1-58
2. Blanch, A.J, Quinton, J.S., Lenehan, C.E. , and Pring, A.(2008) The crystal chemistry of
Al-bearing goethites: an infrared spectroscopic study. Miner. Magazine, 72, 1043-1056
3. Colombo, C. and Violante, A. (1996) Effect of time and temperature on the chemical
composition and crystallization of mixed iron and aluminum species. Clays and Clay
Minerals, 44, 113-120.

4. Fey, M.V., and Dixon J.B. (1981) Synthesis and properties of poorly crystalline hydrated
aluminous goethites. Clay and Clay Miner. 29, 91-100
5. Hochella, M.F., Lower, S.K., Maurice, P.A., Penn, R.L., Sahai, N., Sparks. D.L. and
Twining, B.S. (2008) Nanominerals, mineral nanoparticles and earth systems. Science,
319, 1631-1635
6. Madden, A.S., Hochella, M.F, and Luxton, T.D. (2006) Insights for size dependent
reactivity of hematite nanomineral surfaces through Cu2+ sorption, Geochim et
Cosmochim Acta, 70, 4095-4014
7. Navrotsky, A., Mazeina, L., and Majzlan, J. (2008) Size-driven structural and
thermodynamic complexity in iron oxides. Science, 319, 1635-1638
8. Novikov, A.P., Kalmykov, S.N., Utsunomiua, S., Ewing, R.C., Horreard, F., Merkulov,
A, Clark, S.B., Tkachev, V.V., and Myasoedov, B.F. (2006) Colloidal transport of
plutonium in the far-field of the Mayak Production association, Russia. Science, 314,
638-641
9. Waychunas, G. A., Kim, C.S, and Banfield, J.F. (2005) Nanoparticulate iron oxide
minerals in soils and sediments: unique properties and contaminant scavenging
mechanisms. J. of Nanoparticle Res. 7, 325-571
10. Wigginton, N.S., Haus, K.L., and Hochella, M.F. (2007) Aquatic environmental
nanoparticles. J. Env. Monitoring, 9, 1306-1316
11. Zhang, W.X. The nanoscale iron particles for environmental remediation: An overview.
Journal of nanoparticle research, 5, 323-332

Chapter 2
NANO-GOETHITE CRYSTALLIZATION IN THE PRESENCE OF LOW ALUMINUM
CONCENTRATIONS UNDER ENVIRONMENTALLY RELEVANT CONDITIONS

Abstract
We quantified rate constants for goethite (-FeOOH) crystallization from dilute nanoparticulate Fe hydroxide suspensions (Fe(III) = 10 mM) aged for up to 54 days at 50 oC in the
absence (0% Al) and presence (2% Al) of aluminum using attenuated total reflectance Fourier
transform infrared (ATR-FTIR) spectroscopy and synchrotron-based X-ray diffraction (s-XRD).
We used multivariate curve resolution (MCR) analyses of the OH-stretching region of the FTIR
spectra to derive ferrihydrite-like and goethite-like components. Each of these components
consisted of (1) non-stoichiometric OH stretching vibrations (reflect the combination of surface
hydroxyls and disordering of the Fe-hydroxide structure) and (2) stoichiometric OH-stretching
vibrations (reflect characteristic hydrogen bond (O-H...H) stretching vibrations in the goethite
structure) as assumed from band component analysis (Gaussian functions). Rate constants were
obtained by fitting the contribution of goethite-like MCR component to the overall spectra, and
were equal to (7.4 1.1)*10-7 s-1 for 0% Al and (4.2 0.4)*10-7 s-1 for 2% Al-doped iron
hydroxides. Rate constants derived from intensities of the 790 cm-1 and 890 cm-1 OH-bending
infrared vibration showed similar values (within error) for both 0% Al and 2% Al. The presence
of 2% Al decreased the rate constants determined from analyses of infrared OH-stretching and
OH-bending vibrations by 43-57%, but did not change the rate constant determined from
synchrotron-XRD. The presence of hydrogen bond types (O-HH) characteristic of goethite in
all nano-particle suspensions suggests nano-goethite- and amorphous Fe-hydroxide-like
structures co-existed and formed during the synthesis of Fe-nanoparticles. A mechanism for

nano-goethite crystallization under experimental conditions close to environmental is also


discussed.

2.1. Introduction
In terrestrial and aquatic environments, nano-sized Fe (oxy)hydroxides exhibit a variety
of amorphous and crystalline structures, among which ferrihydrite (Fe(OH)3H2O) and goethite
(-FeOOH) are the most common phases (Waychunas et al., 2005; Wigginton et al., 2007).
Goethite co-exist with ferrihydrite in spodosol, volcanic ash, bog iron, and hydromorphic soils
formed in moist and cool temperate climate (Schwertmann and Taylor, 1989; Bigham et al.,
2002), and in acid mine drainage waters at pH<5 (Hochella et al., 2005a, Filip et al., 2007). In
natural environments goethite may be formed by oxidation of Fe(II) in solution followed by
direct Fe(III) precipitation from solution (van der Zee et all., 2003; Banfield et al., 2000) or via
transformation from ferrihydrite (Cornell and Schwertmann, 2003). Conversion from less
ordered ferrihydrite to more crystalline goethite can significantly modify mineral surface
reactivity and efficiency in the retention of nutrients and/or contaminants.

For example,

transformation of ferrihydrite to the more crystalline iron oxides substantially reduced the
oxides ability to scavenge trace metals from solution (Ford et al., 1997; Buekers et al., 2008).
Chemical analyses have shown the presence of impurities (i.e., elements such as Al3+,
Cd2+, Cu2+, Cr3+, Ni2+, Pb2+, Zn2+) in natural Fe (oxy)hydroxides; these elements form a
coprecipitate with the primary solid and/or chemisorb onto its surface (Martinez and McBride,
2001). Besides influencing the size, surface area, crystallinity, and reactivity of the primary
precipitate, the presence of impurities within a mineral structure (coprecipitation) might affect
the rate, and perhaps the mechanism, of transformation of the primary precipitate. Aluminum is a
common constituent of natural Fe (oxy) hydroxides (up to 30%), and the laboratory experiments
6

have been shown Al retards ferrihydrite crystallization (Colombo and Violante, 1996; Cornell
and Schwertmann, 2003). To our knowledge, no attempt has been made to quantify the rate of
nano-goethite crystallization in the presence of Al, although numerous experimental studies have
shown that Al retards the formation and crystallization of Fe-hydroxides by forming Fe-Al
coprecipitates and/or by sorption to the surface of already formed Fe-hydroxides (Anand and
Gilkes, 1987; Fontes et al., 1992).
When studying structural transformations of nanoparticulate and poorly-crystalline
substances, it is always a challenge to properly identify and quantify mineral phases. Rate
constants for the transformation of ferrihydrite to goethite have been determined from
operationally defined chemical extractions using ammonium oxalate (Schwertmann and Murad,
1983; Cornell et al., 1989) or 0.4 M HCl (Ford et al., 1999), which are presumed to selectively
dissolve ferrihydrite. However, McCarty et al. (1998) showed that ammonium oxalate dissolved
considerable amounts of nano-crystalline goethite in Fe-rich precipitates from an acid mine zone,
which indicates that selective extraction methodologies might not be appropriate in the study of
nano-particle transformations. This concern is also supported by the earlier observation by
Schwertmann (1973) that oxalate extraction could dissolve nano-sized goethite crystallites.
Colorimetric techniques (Nagano et al., 1994; Hamzaoui et al., 2002) and synchrotron-based Xray diffraction (XRD) (Shaw et al., 2005; Yee et al., 2006) were also used to determine the rate
constants for the ferrihydrite to goethite transformation by detection of the fraction of goethite in
samples aged for up to 15 hours. However, most of these laboratory studies were conducted
under alkaline pH conditions (pH > 8), which are rare in natural environments. Rate constants of
goethite crystallization in alkaline solutions were found to be at least one order of magnitude
higher than rate constants obtained for more typical environmental pH (4 < pH < 8) conditions
(Fischer and Schwertmann, 1975). In addition, Hockridge et al. (2009) found that diffraction
7

patterns of goethite are not detectable in the mixtures of nanocyrstalline goethite and ferrihydrite
until the amount of goethite present exceeds about 10%. On the other hand, infrared
spectroscopy provides significant advantages over XRD for studying the transformation of
ferrihydrite to nano-goethite in nano-colloidal suspensions. It is more sensitive for the
determination of small amounts of goethite in poorly-crystalline ferrihydrite suspensions because
it probes short-order molecular vibrations.
In this work we synthesize iron hydroxide nanoparticulate suspension under
environmentally relevant conditions (pH=5, T=50 C, slow titration rate, ionic strength of
solution < 5x10-3). We use synchrotron-XRD and Attenuated total reflectance Fourier transform
infrared (ATR-FTIR) spectroscopy to (1) follow structural transformations of ferrihydrite to
goethite in nanoparticle suspensions, (2) quantify the relative amounts of ferrihydrite and
goethite, (3) calculate the rate constants of nano-goethite crystallization in the absence and in the
presence of low concentrations of aluminum, and (4) to obtain some insight on the mechanism of
ferrihydrite-to-goethite transformation.

2.2. Experimental Section


2.2.1. Synthesis of Fe-hydroxide and Al-doped Fe-hydroxide nanoparticles
To mimic realistic environmental conditions, pure Fe- and Al-doped-Fe-hydroxide
nanoparticles were synthesized at 25 oC using a relatively low Fe (III) concentration (0.01 M)
and a slow rate (1 mL/min-1 to 1 L) of base addition to a pH of 5. In preliminary experiments,
series of Fe-Al mixed hydroxide were prepared from homogeneous solutions with Al
concentrations 2, 4, 8, 12, 16, and 20. It was found that for 4 and 8% Al hematite appeared along
with goethite as a second crystallize phase, and at higher Al concentrations (12, 16 and 20)
goethite was not detected for the duration of experiment. Therefore, we used suspensions with 2
8

mol% Al to evaluate the effect of low Al concentrations on goethite crystallization. For each 0%
and 2%Al composition, two replicate suspensions were synthesized. The following procedure,
modified from Bakoyannakis et al. (2003), was specifically designed to produce nano-sized Fe
and Fe-Al hydroxide particles. Specific amounts of Fe(III)-nitrate and Al-nitrate salts were
dissolved in 100 mL of de-ionized (DI) water to obtain the desired 0% and 2% Al in solutions.
Solutions were slowly titrated to pH 5 with 0.1 N KOH (1 mL min-1) using a pH-controlled
Masterflex C/L TM Pump system for ~3-6 hours. The final volume of all samples was adjusted to
1 L, and a final total metal ([Fe+Al]) concentration of 10-2 M was obtained. The suspensions
were then dialyzed at 25 C against Milli-Q water using 1 nm pore size Spectropor/7 dialysis
tubes (molecular weight cutoff 1000) to remove excess salts and excess Al, so all the Fe and Al
contained in the dialysis tubes were in nano-particulate form. The water was replaced several
times over the course of several days, until the conductivity was reduced to 0.3 S/cm (ionic
strength value < 0.005 M). Aliquots of each suspension (~100 ml) were taken immediately after
dialysis (time = 0 days) and after the suspensions were aged for 2, 9, 23, and 54 days at 50C. No
precipitates were formed during the course of the experiment, and all suspensions remained
visually clear without any precipitates. A portion of the aliquots (~40 ml) collected at each time
interval was freeze-dried for s-XRD analysis, while another portion was stored at 4 C in the
dark in polystyrene bottles until used for ATR-FTIR, DLS, and TEM analyses as described
below. All nano-colloidal suspensions were yellow-brown in color and remained optically
transparent for the duration of the experiment and measurements.
The elemental composition of dialyzed suspensions was determined by dissolving 1 ml of
suspension in 2 ml of concentrated HNO3 and heating at ~80 C for 1 hour. The digest was
diluted 100 times with DI water, and total dissolved Fe and Al concentrations determined by
atomic absorption spectroscopy. As expected, for the samples with 2% Al in initial solution,
9

chemical analysis revealed 1.95% Al (n=2) , which indicate that almost all of the Al introduced
to the system either formed a coprecipitate with the iron hydroxides or retained at its surface
(Lovgren et al., 1990; Fey and Dixon, 1981).
2.2.2. Synchrotron-based X-ray diffraction
High-resolution synchrotron-based XRD patterns were collected at beamline X16C,
National Synchrotron Light Source, Brookhaven National Laboratory. An X-ray wavelength of
0.70051(7) was selected with a Si(111) double crystal monochromator. Freeze-dried powder
samples were loaded into a glass capillary to consistent density (0.7 mm internal diameter) and
rotated about the longitudinal axis during data collection. XRD patterns were collected in the 2range from 2 to 40 using a 0.05 step size and a counting time of 5 seconds per step. Jade+
software (version 8.2, Materials Data, Inc., Livermore, CA) was used for the identification of
crystalline phases and measurement of peak areas. Goethite crystallite sizes were estimated using
the Scherrer equation for the most intense and non-overlapping (110) peak. The full-width at
half-maximum (FWHM) values were corrected for instrument broadening using silicon powder
standard.
2.2.3. ATR-FTIR spectroscopy
2.2.3.1. ATR-FTIR data collection and Gaussian band analyses
ATR-FTIR measurements were performed using a Bruker Tensor 27 FTIR instrument
with a La DTGS detector and equipped with single-reflection diamond/ZnSe ATR accessory. A
0.5 L aliquot of the hydroxide suspension was placed on the sensor and air-dried for 5 min to
form a thin coating before data collection. The sensor background spectrum was measured before
each sample. Spectra of 8 cm1 resolution were acquired by co-addition of 256 scans within the
4000-600 cm-1 region. Three or more replicates were measured for each sample. Spectral
manipulation, including subtraction of background water and CO2 vapor, baseline correction, and
10

spectra normalization were performed using in-house developed codes under the MATLAB 7.4
computing environment. Baseline correction was performed by subtracting a 2-point linear
baseline in the background region and applying standard normal variate correction across the
selected range. The 4000-2500 cm-1 spectral regions were normalized to a unit area (i.e. rescaled
in order to set the area under the peak equal to 1). Due to intense sloping background in the
1000-650 cm-1 region, for each spectrum the absorption at each wavelength was divided by the
area of corresponding peak in the 4000-2500 cm-1 region.
Gaussian band analysis (deconvolution) of the iron (oxy)hydroxide OH-stretching (40002500 cm-1) region of each ATR-FTIR spectrum was performed using the nonlinear least-squares
fitting routine SOLVER in MS Excel (de Levie, 2001). The choice of number of bands and their
positions were made based on the available literature data (Farmer, 1974; Morterra et al., 1984;
Weckler and Lutz,, 1998; Ruan et al., 2002). Spectral de-convolution of measured FTIR peaks of
FeOOH is typically performed to extract information about the contributions of different OHgroups (Hug, 1997; Ruan et al., 2001; Boily et al., 2006). The choice of number of bands used in
fitting a band envelope and the band positions largely depends on a researchers model, and
sometimes mathematically the best fit may not be the best fit to explain the chemistry of the
system (Meier, 2005). This univariate approach, when each spectral group is assigned to a
specific component, can be challenging in the case of highly overlapped varying intervals bands
in the OH-stretching region. In addition to analyses of overlapping OH-stretching vibrations,
peak heights for non-overlapping OH bending vibrations were determined from Gaussian curves
fitted to experimental spectra.

11

2.2.3.2. Multivariate curve resolution (MCR) analysis of ATR-FTIR spectra


To complement our Gaussian band deconvolution approach we brought a new technique,
multivariate curve resolution-alternating least squares (MCR-ALS). MCR permits extraction of
chemically reasonable spectral components to describe the whole system (Garrido et al., 2008)
and assumes that each spectrum in a dataset in composed of pure component spectra. These
pure components contribute to the spectral intensity in various proportions (scores) and can
completely explain the observed spectral variance in the dataset, which in our case spans an
aging (crystallization) process. MCR analysis is able to reduce and analyze matrices of ATRFTIR spectra by extracting major spectral features (MCR components) and their correspondent
weights (concentrations) to describe the whole original experimental dataset (Tauler, 2003). This
approach is especially useful in its application to nano-particle crystallization studies where
infrared spectra are collected as a function of a perturbation to the chemical system (i.e., aging)
and when no end-member standards are available for quantification (Schoonover et al., 2003).
MCR analysis consists of 3 major steps: (1) estimation of the number of pure components; (2)
eigenvalue decomposition of the full dataset matrix to produce abstract factors and scores,
arranged with contributions of each factor to the measured spectra in decreasing order; and (3)
refining abstract factors into real spectra of pure components by fitting the sum of factors, each
with a particular score, to the experimental spectra by an iterative ALS process so that the
percent of the spectral variance not explained by MCR model was minimal. The non-negativity
constraint was applied where only positive values were allowed for pure components and scores.
The choice of the number of components was based on the physical meaning of extracted
components, on the percent of the variance explained by the fit, and on a plot of eigenvalues
versus number of factors that allows the differentiation of real signal from noise (Schoonover et
al., 2004). MCR analyses were conducted using MATLAB (The Mathworks) with the PLS
12

toolbox Version (Eigenvector Inc.) and in-house developed functions. Normalized spectra were
used in all analyses.
2.2.4. Dynamic light scattering (DLS)
Particle size distribution for the Fe- and Fe-Al- hydroxide nanoparticle suspensions was
determined by dynamic light scattering using a ZetaSizer Nano S (Malvern Instruments, UK).
Particle size distributions were obtained by fitting the correlation functions to multiple
exponentials using non-negatively constrained least squares (NNLS) analysis. The diffusion
coefficient (D) was calculated by fitting the correlation curve to an exponential function, with D
being proportional to the lifetime of the exponential decay. The hydrodynamic radius (RH) was
then calculated from the diffusion coefficient using the Stokes-Einstein equation, D =

kT
6R H

where k is the Boltzmann constant, T is the temperature, and is the medium viscosity. Since
particle size distributions were polydispersed for most of the samples, the intensity distributions
were converted to volume distributions according to the Mie theory (Dahneke, 1983).
Suspensions of iron (oxy)hydroxide with 0% and with 2% aluminum-doping aged for 0, 2, 9, 23
and 54 days were placed in disposable square polystyrene cuvettes and light scattering measured
with a laser beam at an angle of 173 between the incident and diffracted beams. Additionally, in
situ heating experiments at 50 C were performed for both Fe- and Fe-Al suspensions by
measuring the light scattered from an aliquot of the initial suspensions (time = 0 days) every 2
hours for about 2 days in automatic mode.

2.2.5. Transmission electron microscopy (TEM)


Samples for TEM were prepared by placing a single drop of the Fe and Fe-Al
suspensions onto a 200 mesh holey-carbon-coated copper grid and allowed to air-dry. TEM
images and selected area electron diffraction (SAED) patterns were collected using a JEOL EM-

13

2010 with a LaB6 filament operated at 200 kV. TEM images were obtained for non-diluted and
diluted suspensions. High-resolution TEM (HRTEM) images were obtained for diluted
suspensions using a JEOL EM-2010F (Field Emission TEM/STEM) operated at 300 kV.

2.3. Results
2.3.1. XRD data
The characteristic broad bands at d-spacings of 0.15 and 0.26 nm indicate that poorly
ordered 2-line ferrihydrite (Cornell and Schwertmann, 2003) was the dominant phase in undoped
and 2%-Al-doped ferric (oxy)hydroxide nano-colloid suspension at initial conditions and in 2day aged 2%Al suspension (Fig. 2-1). The intensity of the main XRD peak (110) (0.42 nm in dspace), which is characteristic of goethite, increased with time for the 0% and 2% Al
suspensions. X-ray diffraction patterns of Al-free suspensions, however, revealed fairly
crystalline goethite was present after 54 days of aging at 50C, while in suspensions with 2% Al
the intensities of the goethite peaks were lower in intensity with little goethite formed until 9
days (Fig. 2-1). No other crystalline phases except goethite were determined in suspensions.
Since the (110) peak reflaction is distinct from ferrihydrite and does not overlap with other
peaks, its area was used to determine the rate constant for the crystallization of goethite from
ferrihydrite (section 4.2). Goethite (110) reflection was also used to determine crystallite sizes
using the Scherrer equation (Patterson, 1939).

Analyses (Fig. 2-2) showed crystallite size

increased with time from 2.5 to 5.5 nm in 0% Al suspensions. In the presence of 2% Al,
crystallite sizes were in the range from 3.5 to 4.8 nm with no clear trend with time (Fig.2-2).
These results suggest the presence of Al decreased the degree of structural order in nanoparticle
suspensions, as is also evidenced from XRD patterns (Fig.2-1).

14

2.3.2. Infrared spectra


2.3.2.1. OH-bending vibrations
Goethites structure consists of double sheets of edge-sharing FeO3(OH)3 octahedra
connected to each other by corner-sharing (O bridges) that form 2x1 octahedra tunnels
(Schwertmann and Cornell, 1991). Hydrogen bonds (O-HO) that cross the tunnels and link the
chains of octahedra are structurally equivalent, but point in different directions (Fig. 2-3). The
presence of these hydrogen bonds in a Fe hydroxide mixture is an indication of the presence of
goethite and can be used to follow the ferrihydrite to goethite crystallization process. The
presence of these hydrogen bonds is verified by characteristic infrared stretching (3120-3200 cm1

) and bending (895 and 795 cm-1) vibrations in bulk goethite (Morterra et al., 1984, Cornell and

Schwertman, 2003). Hydroxyl bending vibrations at 895 cm-1 and 795 cm-1 are diagnostic for
goethite and reflect OH bending deformations in the a-b plane and out of the a-b plane
deformations, respectively (Schwertmann et al., 2003). Their intensities increased upon aging
due to the increase in the number of goethite hydrogen bonds in suspension (Fig. 2-4). It is
important to note that in Al-free suspensions the OH-bending vibrations, although broad and of
low intensity, started to appear even at 0 days. In contrast, OH-bending vibrations were not
observed at 0 days in suspensions with 2% Al, and lower intensity peaks were developed at
corresponding times compared to Al-free suspensions (Fig.2-4). The intensities (peak height) of
the OH-bending vibrations were used to determine rate constants of ferrihydrite to goethite
transformation (section 2.4.2).
2.3.2.2. OH-stretching vibrations: Gaussian deconvolution
Unlike OH-bending vibrations, IR bands in the OH-stretching region overlap strongly
making it difficult to directly estimate intensity changes due to system perturbation (Fig. 2-3).
The types of spectral envelopes of the OH-stretching region were similar for Al-free and 2% Al15

containing Fe-hydroxides suspensions. Replicate variability was higher in 2% Al suspensions


(Fig.2-4) possibly due to heterogeneity in particle shapes and sizes. As shown by our results, the
infrared OH-stretching region of iron hydroxides is characterized by relatively broad bands
which contain multiple OH contributions under broad peak areas (Kubicki et al., 2008). For our
analysis, the OH-stretching envelopes were deconvoluted using Gaussian functions to model
three OH contributions: stoichiometric OH vibrations, non-stoichiometric OH vibrations, and
adsorbed water (Fig. 2-5A). Stoichiometric, or bulk, OH groups are those not reactive in
adsorption/desorption processes and are difficult to exchange with deuterium (D2O) vapor
(Morterra et al., 1984). They usually yield infrared bands in the 3050 to 3240 cm-1 range
depending on crystallinity and crystal size of goethite and on the infrared technique used. These
OH groups form hydrogen bonds (O-HO) between double chains of octahedra (Weckler and
Lutz, 1998), and are characteristic of the goethite structure. Bands in the 3400-3480 cm-1 range
have been attributed to weakly H-bonded surface hydroxyls (Rochester and Topham, 1978), bulk
OH groups in the ferrihydrite structure (Russel, 1979), and to non-stoichiometric hydroxyl units
incorporated into the goethite structure during the process of synthesis, when, for example, one
iron ion is replaced by 3 protons (Ruan et al., 2001; Boily et al., 2006). Most of these hydroxyls
can be readily exchanged with deuterium, which suggest high reactivity and mobility, and by
definition, indicate greater degree of disorder in the goethite structure. Furthermore, Boily et al.
(2006), who studied goethite suspensions evaporated on an ATR cell under dry N2 gas
atmosphere, assigned OH-stretching vibrations in the 2600-3800 cm-1 region to contributions
from (1) structural, or stoichiometric, hydroxyls in bulk goethite at ~3027 cm-1, (2) nonstoichiometric hydroxyls in bulk goethite at ~3384 cm-1, and (3) OH stretching of isolated
surface hydroxyls at ~3640-3660 cm-1. In agreement with their data, our ATR-FTIR spectra also
had two broad bands that can be attributed to stoichiometric and non-stoichiometric hydroxyls,
16

but lack surface hydroxyl bands at high frequencies (Fig.2-5A). A possible explanation for the
absence of these bands is that our measurements were done under ambient atmosphere, and the
sorption of water vapor onto the Fe hydroxide film could have suppressed surface hydroxyl
features. Adsorbed water may also cause a significant asymmetric broadening at lower
frequencies (~2900-3000 cm-1; Figs. 2-3 and 2-4A) (Morterra et al., 1984). As depicted in
Figure 2-4A, band component analysis was used to differentiate among the contribution of
specific OH vibrations to the OH-stretching envelope, thus isolating stoichiometric OH
vibrations characteristic of the goethite structure. The best fit to the experimental spectra was
obtained by using three Gaussian functions, assigned as non-stoichiometric (~3460 cm-1),
stoichiometric (~3240 cm-1), and adsorbed water (2960 cm-1) (Fig. 2-5A). These band
component assignments are in good agreement with the data of Ruan et al. (2002) who obtained
bands for stoichiometric and non-stoichiometric hydroxyls at 32333206 and 34763450 cm1,
respectively, and two other bands (~2804, 2930 cm-1) at lower frequencies.

One of the

approaches we undertook to follow goethites crystallization process was to use changes in


frequency and in full width at half maximum (FWHM) values of stoichiometric hydroxyls peaks
with time, which were obtained by Gaussian deconvolution of the OH-stretching region. In
general, the width of the Gaussians that describe the OH stoichiometric components of aged (i.e.,
54 days) suspensions decreased and their intensity increased compared to initial (i.e., 0 days)
suspensions, thus suggesting a continuous increase in structural order of nano-goethite with
aging time. This is evidenced by a decrease in the peak position (wavenumbers at center of
stoichiometric OH vibrations) at which the 0%Al and 2% Al suspensions appeared with time
(Fig. 2-5B). In addition, FWHM values for the stoichiometric OH vibrations decreased steeply
within the first 10 days of aging (Fig. 2-4C). Peak positions and FWHM values are slightly
higher for 2% Al suspensions. The shape and position of adsorbed water remained constant for
17

all suspensions. It was suggested by Waychunas et al (2005) that during crystallization double
chains of edge-sharing FeO3(OH)3 octahedra, the building block of the goethite structure, begin
to stack together by Fe-O-Fe corner sharing and by the formation of hydrogen bonds between
these double chains (Fig. 2-3). We hypothesize that as the number of free hydroxyls decrease
(Fe-O-H bonds present in the octahedra) and the number of H-bonded hydroxyls increase (Fe-OHO-Fe that connect double chains, the energy of the hydroxyl bond decreases and their
frequencies shift to lower wavenumbers, as observed in our results (Fig. 2-5B). On the other
hand, the relatively large initial FWHM value indicates there is variability in the energy of Hbonded hydroxyls that contribute to stoichiometric OH vibrations (Fig. 2-5C). A substantial
decrease (Fig. 2-4C) in the FWHM value for stoichiometric OH vibrations with time
(approximately 40%) indicates a decrease in the variability of H-bonded hydroxyls and therefore
an increase in structural order. In the presence of 2% Al, the formation of hydrogen bonds
between double chains of octahedra is hindered compared to Al-free suspensions as evidenced
from slightly higher peak positions and FWHM values with time. A lower degree of structural
order in Al-doped suspensions may result from disturbance to the network of hydrogen bonds
that form across double chains of Fe octahedra by substitution with smaller Al octahedra. The
presence of Al might then increase the number of non-stoichiometric hydroxyls and/or increase
variability within stoichiometric hydroxyls in Al-doped nano-goethite.
2.3.2.3. MCR analysis of IR O-H stretching region
Multivariate curve resolution (MCR) analysis was used to extract pure components and to
describe their contribution to spectral variations that occurred during the crystallization process
(Tauler, 2003). MCR analysis is appropriate for systems in which there is a significant change in
spectral features as the system evolves with time, i.e. amorphous to crystalline transformations as
in the case of this study. A 2-component system, in which the components were inversely
18

related, resulted from MCR analysis of the OH-stretching region (Fig. 2-6). A 2-component
system is in agreement with our Gaussian deconvolution results and with principal component
analysis (PCA), both of which revealed the contribution of two primary components
(stoichiometric and non-stoichiometric OH vibrations in Gaussian deconvolution) in the dataset.
MCR is more advantageous than PCA because it is possible to add non-negativity constraints in
the MCR algorithm.
MCR-extracted components (spectra) and scores, along with experimental spectra for the
0%Al dataset are shown in Fig. 2-6. Component 1 characterized the system initially, while
Component 2 resembled the most aged samples. It is clear that none of these components can be
attributed to a single type of OH-stretching vibration, but rather consists of overlapping bands.
To bring chemical relevance to this statistical method and aid in its interpretation, each of the
main components was deconvolved into single bands in the same manner we did for the
experimental spectra (Gaussian deconvolution). Similar to the experimental spectra discussed
above (section 2.3.2.2), each component exhibited non-stoichiometric hydroxyls, stoichiometric
hydroxyls and adsorbed water (Fig. 2-6B). Conceptually, Component 1 can be presented as
disordered Fe-hydroxide or ferrihydrite-like phase, in which non-stoichiometric hydroxyls are
dominant (>30% of the total volume; Zhao et al., 1994) due to structural disorder and large
surface area characteristic for ferrihydrite, but minor concentrations of goethite nanoparticles are
already present. In contrast, Component 2 can be considered a goethite-like phase since it
exhibits well-expressed stoichiometric OH-stretching vibrations at ~3100 cm-1 with some
structural disorder/surface water at ~3300 cm-1 (Fig. 2-6B). These two components described
99.2% and 99.6% of the variance in the 0%Al and 2%Al datasets, respectively, and may be
considered the two end-members of the goethite crystallization reaction. Their contributions to
spectral changes as a function of time are reflected in the MCR-derived scores (Fig. 2-6B),
19

which clearly indicate a gradual decrease of the ferrihydrite-like component and a concomitant
increase of the goethite-like component (Fig. 2-6B). Scores for the goethite-like component
can be directly related to the number of goethite stoichiometric OH-groups in the suspension and
can therefore describe the kinetics of goethite crystallization.

2.4. Discussion
2.4.1. Mechanisms for nano-goethite crystallization
The rate and mechanism of ferrihydrite-goethite transformation in laboratory experiments
depend on many factors among which pH, Fe concentration, particle size, and the presence of
trace metal impurities are the most important ones. For example, the effect of particle size on
phase transformation kinetics becomes clear when several authors suggested that in addition to
conventional Ostwald ripening (defined by a first-order equation) the kinetics of phase
transformation in nanoparticulate systems can be described by oriented aggregation, a
mechanism specific for nanoparticles (Banfield et al., 2000; Waychunas, 2001; Zhong et al.,
2006; Jia et al., 2006; Penn et al., 2007). The kinetics of oriented aggregation was quantified
from particle sizes obtained by transmission electron microscopy (TEM) and dynamic light
scattering (DLS), and the rate law was found to be second order in the concentration of primary
nanoparticles (Penn, 2004). These two mechanisms are not exclusive: aggregation can
predominate at first, followed by Ostwald ripening (Waychunas et al., 2005), or they may occur
simultaneously. Thus, it is important to understand the mechanism of ferrihydrite-goethite
transformation in particular systems to correctly model kinetic data and derive rate constants.
Goethite crystallization kinetics is usually described by a dissolution-precipitation mechanism
(i.e. Ostwald-ripening) wherein that small particles with high surface potential dissolve due to
their higher solubility compared to larger particles which grow by incorporating Fe mono- and
20

poly-nuclear species from solution into the crystal structure by diffusion (Jolivet, 2004). This
mechanism obeys a first-order type reaction and considers only the volume of goethite
transformed with time; Ostwald ripening assumes intermediate Fe-hydroxide species are not
involved. In our suspensions, dynamic light scattering (DLS) revealed no measurable increase in
average particle size at 2 days (average particle size = 5.04 3.7). An increase in particle size
(to 40-55 nm) was observed for 0% Al suspensions after 9 days (Fig. 2-7). However, the 2% Al
containing suspensions showed more variability in particle sizes during the first 2 days of aging,
with values ranging from 2 to 30 nm and no trend with time (Fig. 2-7). A possible explanation
of such differences in particle size between pure and Al-doped suspensions is that the presence
of Al could increase the rapidity of Fe hydrolysis (Singh and Kodama, 1994) thus favoring the
formation of ferrihydrite as a more metastable phase. Alternation explanation is that Al
suppressed ferrihydrite dissolution and transformation, however it is not in agreement with
Jentzsch and Penn (2006) who found that doping of small amounts of Al (<2 mol %) increased
ferrihydrite reactivity.
Direct comparison of XRD-calculated crystallite size and DLS-determined particle size is
challenging since light scattering does not distinguish between ferrihydrite and goethite particles.
Goethite was not detected by XRD before 2 days (for 0% Al) and 9 days (for 2% Al) of aging
(Fig. 2-1) which suggests that small particles (under 10 nm) determined by DLS within the first 2
days represent ferrihydrite. However, FTIR (Fig. 2-3) indicated small amounts of nano-goethite
crystals were present in suspension at 0 days (0% Al) and 2 days (2% Al). Indeed, goethite peaks
in a XRD pattern result from detection of a diffracted x-ray beam and require the mineral to have
a minimum degree of structural order and size to yield a detectable signal. FTIR, on the other
hand, probes molecular vibrations and does not require long-range order to yield a signal. For
suspensions aged for more than 9 days, the average DLS particle size was ~10 times larger than
21

the corresponding nano-goethite crystallite size (Fig. 2-7A), which suggests aggregation of nanogoethite crystallites into larger polycrystalline goethite particles. This interpretation is in
agreement with particle size distribution (PSD) analysis of the DLS data (Fig. 2-7B). For time 0
and 2 days, PSD exhibited narrow size distributions within 10 nm; however, as the suspensions
aged, PSD broadened and presented a tail (right-skewed) of large particles with sizes up to 100
nm. If there was a constant particle growth rate (dissolution-precipitation mechanism), then PSD
would only shift laterally on the size axis (Li et al., 2005), however, PSD broadening signified
particle aggregation.
TEM and DLS results are in good agreement since TEM images showed small round
particles (< 5 nm) for both 0% and 2% Al suspensions at time 0 days (Figs. 2-8A, B and 2-9A,
B). Three populations of nanoparticles of various shapes and sizes (5-100 nm range) were
present in Al-free suspensions aged for 54 days at 50 C (Figs. 2-8C and 2-9C): (1) round
particles less than 10 nm in diameter; (2) thin goethite needles about 50 nm in length and several
nm wide; (3) large goethite crystals (prisms) 50-100 nm in length and 10-20 nm wide. In the
presence of 2% Al (Figures 2-8D and 2-9D), ferrihydrite (round) particles were the most
abundant, yet goethite needles and crystals were also present. It is important to note that TEM
images showed small needle-like (<~30 nm) structures in Al-free samples at 0 days (insert in
Fig. 2-8A, Fig. 2-9A). These needle-like structures were not detected by TEM in initial
suspensions containing Al.
HRTEM images of Al-free suspensions at initial (0 days) conditions showed needle-like
structures (Fig. 2-10A) contained primary round particles with lattice fringes (nano-goethite) and
primary round particles without lattice fringes (ferrihydrite) (Janney et al., 2000). This
observation suggests phase transformation from ferrihydrite particles to nano-goethite crystals
occurred. In Al-free suspensions aged for 54 days, HRTEM images indicated large goethite
22

crystals (Fig. 2-10B) contained single-crystal lattice fringes and showed a pattern of spots in the
SAED image; thus suggesting it was formed by monomeric addition (Ostwald ripening). In
contrast, goethite needles consisted of smaller grains with different lattice orientation (Fig. 210C) and a diffraction pattern showing characteristic polycrystalline rings, which suggests
goethite needles were formed by aggregation of round nano-goethite particles. Suspensions with
2% Al showed similar features.
Based on our (HR)TEM and DLS results, we hypothesize that at first, some of the
primary (round) particles of ferrihydrite were transformed to nano-goethite round particles by a
dissolution-precipitation mechanism that probably occurred at the dialysis stage (for non-doped
Fe-hydroxide suspensions) or during the first 2 days of aging. Then, nano-goethite particles
started to grow either by aggregation to form goethite needles or by monomeric addition to form
single-crystal goethite (Fig. 2-11). Since nano-goethite aggregation took place after ferrihydrite
to nano-goethite transformation of small round particles, we concluded Ostwald ripening was the
dominant mechanism controlling the rate of this reaction. Gradual increase with time of MCRderived goethite-like component supports Ostwald ripening mechanism. The presence of Al
does not change the mechanism of goethite crystallization, but retards formation of goethite
needles and prisms as evidenced from TEM data

2.4.2. Rate constants for ferrihydrite-goethite transformation


Ostwald ripening (dissolution-precipitation) is usually described by a pseudo-first order
equation when only the formation of a crystalline end-product (i.e., goethite) or the decrease in
the amorphous end-member (i.e., ferrihydrite) is considered; the concentration of additional
chemical species taking part in the reaction are not considered (Lasaga, 1998). This reaction can
be written as Ferrihydrite Goethite, or in exponential form:
[Goethite] = [Goethite]0 * (1 exp kt )

(eq.1)
23

where the concentration of goethite, [Goethite], is expressed in terms of XRD (110) peak
areas, intensities of OH-bending (790 cm-1 and 890 cm-1) vibrations, and MCR scores as
indicators of the presence of goethite in Fe-hydroxide and Al-doped Fe-hydroxide mixtures at a
given time t (sec); [Goethite]0, is the initial concentration of goethite in the mineral mixture, and
k (sec-1) is the first-order rate constant. Both [Goethite]0 and rate constant were determined by
fitting experimental data plotted versus reaction time to equation 1.
XRD peak areas, the intensity of OH-bending vibrations, and MCR scores increased with
time for Al-free and Al-doped Fe-hydroxide suspensions (Fig. 2-12). Corresponding values for
2% Al were lower compared to 0% Al, with the largest difference resulting from the intensities
of OH bending vibrations. The rate constants obtained from different methods were consistent
within experimental error for both 0% and 2% Al suspensions (Fig. 2-13). The presence of 2%
Al decreased the rate constants determined from OH-stretching and OH-bending vibrations by
~47-50%. For example, MCR derived rate constant decreased from (7.4 1.1)*10-7 (0%Al) to
(4.2 0.4)*10-7 s-1 (2%Al). However, the values of the rate constants determined from XRD
peak areas were the same for both 0% and 2% Al suspensions. This discrepancy can be
explained by a large experimental error due to lower sensitivity of XRD to structural changes in
poorly-crystalline mixtures (Burleson and Penn, 2006).
The fact that both the rate of goethite crystallization and the stability of colloidal
suspension are influenced by factors such as pH, temperature, [Fe], ionic strength of the solution
and the presence of foreign ions, makes it difficult to obtain a consistent picture from already
available experimental work (Fig. 2-14). Many studies on ferrihydrite to goethite transformation
were conducted under very high alkaline (pH 10-13) conditions (Nagano et al., 1992; Nagano et
al., 1994; Shaw et al., 2005), where the rate of goethite transformation is relatively fast compared
to environmental pH values of 4-9 (Fig. 2-14). Even studies conducted at circumneutral pH (Fig.
24

2-14) show that the rate constants highly depend on experimental parameters. For example, the
presence of Fe2+ in solution greatly accelerates the rate of goethite crystallization (Yee et al.,
2006). Our values of rate constants are similar to the values of Ford et al. (2002) and Fischer and
Schwertmann (1975), both of which derived their rate constants from ammonium oxalate
extractions. What we can deduce from published rate constants is that both increases in pH and
temperature result in a faster rate of transformation (Fig. 2-14). Differences in the numerical
value of the rate constants obtained by various methods (XRD, infrared, colorimerty, oxalate
extraction) may not be as significant as differences attributed to synthesis procedures or
environmental conditions (i.e., pH, temperature, [Fe], etc.).

2.5. Conclusions
We succesfully applied MCR analysis of overlapped IR OH-stretching bands to study
ferrihydrite-goethite transformations and to quantify rate constants in poorly-crystalline nanocolloidal suspensions. To our knowledge, this is the first time the rate constant for ferrihydrite
crystallization in the presence of low concentrations of Al has been quantified. The presence of
2% Al decreased the rate constants determined from analyses of infrared OH-stretching and OHbending vibrations by ~45%, while TEM data showed that in the presence of Al the growth of
larger goethite particles was hindered. Our results also indicated dissolution-precipitation was the
dominant mechanism in the reaction of ferrihydrite transformation to nano-goethite. Further
growth of nano-goethite crystals took place by aggregation to form polycrystalline goethite
crystals or by Ostwald ripening to form large single crystals.

25

Acknowledgments
This research was funded by the National Science Foundation under Grant No. CHE0431328. (HR)TEM work was performed in the Electron Microscopy Facility of the Materials
Characterization Laboratory at The Pennsylvania State University. DLS measurements were
made in the Particulate Materials Center (PMC) at The Pennsylvania State University.
Synchrotron-XRD data was collected in beamline X-16C at the National Synchrotron Light
Source (NSLS), Brookhaven National Laboratory. The NSLS is supported by the U.S.
Department of Energy (DE-AC02-98CH10886). We thank Dr. Trevor Clark (HR-TEM), Dr.
Rafaat Malek (DLS), and Dr. Peter Stephens (synchrotron-based XRD) for assistance in data
collection.

References
1. Anand, R.R. and Gilkes, R.J. (1987) Variations in the properties of iron oxides within
individual specimens of lateritic duricrust. Australian J. Soil Res., 25, 287-302
2. Bakoyannakis, D. N., Deliyanni, E. A., Zouboulis, A. I., Matis, K. A., Nalbandia, L.,
and Kehagias, T. (2003) Akaganeite and goethite-type nanocrystals: synthesis and
characterization. Microporous and Mesoporous Materials 59 35-42
3. Banfield, J.F., Welch, S.A., Zhang, H., Ebert, T.T., and Penn, R.L. (2000) Aggregationbased crystal growth and microstructure development in natural iron (oxy)hydroxide
biomineralization products. Science, 289, 751754
4. Buekers, J., Amery, F., Maes, A., and Smolders, E (2008) Long-term reactions of Ni,
Zn and Cd with iron (oxy)hydroxides depend on crystallinity and structure and on metal
concentrations. European J. Soil Sci. 59, 706-715
5. Bigham, J.M., Fitzpatrick, R.W., and Schulze, D.G. (2002). Iron Oxides. In: Soil
Mineralogy with Environmental Applications. Eds: Dixon, J.B., and Schulze, D.G. Soil
Science Society of America, Madison, WI, 2002, pp. 866
6. Boily, J.F., Szanyi, J., and Felmy, A.R (2006) A combined FTIR and TPD study on the
bulk and surface dehydroxylation and decarbonation of synthetic goethite. Geochim.
Cosmochim Acta 70 36133624
7. Burleson, D.J. and Penn, R.L. (2006) Two-step growth of goethite from ferrihydrite
Langmuir, 22, 402-409
8. Colombo, C., and Violante, A. (1996) Effect of time and temperature on the chemical
composition and crystallization of mixed iron and aluminum species. Clays Clay Miner.
45 113120

26

9. Cornell, R.M. and Giovanoli, R. (1985) Effect of solution conditions on the proportion
and morphology of goethite formed from ferrihydrite. Clays and Clay Minerals, 33,
424432
10. Cornell, R.M., Giovanoli, R., and Schneider, W. (1989) Review of the hydrolysis of
iron(III) and the crystallization of amorphous iron(III) hydroxide hydrate. J. Chem.
Tech. and Biotech., 46, 115-134
11. Cornell, R.M. and Schwertmann, U. (2003) The Iron oxides: Structure, Properties,
Reactions, occurrences and uses. Wiley-VCH, 2003
12. Cudennec, Y., Lecerf, A. (2006) The transformation of ferrihydrite into goethite or
hematite, revisited. J. Solid State Chem. 179, 716722
13. Cwiertny D. M., Handler, R.M., Schaefer, M.V., Grassian, V.H., and Scherer, M.M.
(2008) Interpreting nanoscale size-effects in aggregated Fe -oxide suspensions: reaction
of Fe(II) with goethite. Geochim. Cosmochim Acta 72 13651380
14. Dahneke, B.E. (1983) Measurement of Suspended Particles by Quasielastic Light
Scattering, Wiley, 1983.
15. Farmer V.C. (ed.) (1974) The Infrared Spectra of Minerals. Mineral.Soc., London
16. Fey, M.V., and Dixon J.B. (1981) Synthesis and properties of poorly crystalline
hydrated aluminous goethites. Clay and Clay Miner. 29, 91-100
17. Filip, J., Zboril, R., Schneeweiss, O., Zeman, J., Cernik, M., Kvapil P., and Otyepka M.
(2007) Environmental applications of chemically pure natural ferrihydrite. Environ. Sci.
Technol., 41, 4367-4374
18. Fischer, W.R. and Schwertmann, U. (1975) Formation of hematite from amorphous
iron(III) hydroxide. Clays and Clay Miner., 23, 33-38
19. Fontes, M.P.F. (1992) Iron oxide clay mineral association in Brazilian oxisols A
magnetic separation study. Clays and Clay Min., 40, 175-179
20. Ford, R.G., Bertsch, P.M. and Farley, K.J. (1997). Changes in transition and heavy
metal partitioning during hydrous iron oxide aging. Env. Sci Tech, 31, 20282033.
21. Ford, R.G., Kemner, K.M., and Bertsch, P.M. (1999) Influence of sorbate-sorbent
interactions on the crystallization kinetics of nickel- and lead-ferrihydrite
coprecipitates. Geochim. Cosmochim Acta 63, 39-48
22. Ford, R.G. (2002) Rates of hydrous ferric oxide crystallization and the influence on
coprecipitated arsenate. Environ. Sci. Technol., 36 (11), 2459 -2463
23. Garrido, M., Rius, F.X., and Larrechi, M.S. (2008) Multivariate curve resolution
alternating least squares (MCR-ALS) applied to spectroscopic data from monitoring
chemical reactions processes. Anal. Bioanal. Chem. 390, 20592066
27

24. Gilbert, B., Zhang, H., Huang, F., Finnegan, M.P, Waychunas, G.A., and Banfield, L.F.
(2003) Special phase transformation and crystal growth pathways observed in
nanoparticles. Geochem. Trans., 4, 2027
25. Gilbert, B., Lu, G., and Kim, C. S (2007) Stable cluster formation in aqueous
suspensions of iron (oxy)hydroxide nanoparticles. J. Colloid Interface Sci, 313, 152159
26. Hamzaoui, A., Mgaidi, A., Megriche, A., and Maaoui M.L. (2002) Kinetics study of
goethite formation from ferrihydrite in alkaline medium. Ind. Eng. Chem. Res. 41,
5226-5231
27. Hochella M.F., Jr., Kasama T., Putnis A., Putnis C., and Moore J.N. (2005)
Environmentally important, poorly crystalline Fe/Mn hydrous oxides: Ferrihydrite and
a vernadite-like mineral from a massive acid mine drainage system. American
Mineralogist, 90, 718-724.
28. Hochella M.F., Jr., Moore J.N., Putnis C., Putnis A., Kasama T., and Eberl D.D. (2005)
Direct observation of toxic metal-mineral association from a massive acid mine
drainage system: Implications for metal transport and bioavailability. Geochimica et
Cosmochimica Acta, 69, 1651-1663
29. Hockridge, J.G., Jones, F., Loan, M, and Richmond, W.R. (2009). An electron
microscopy study of the crystal growth of schwertmannite needles through oriented
aggregation of goethite nanocrystals. J. Crystal Growth, 311, 3876-3882
30. Hug, S.J. (1997) In Situ Fourier transform infrared measurements of sulfate adsorption
on hematite in aqueous solutions. J. Colloid Interface Sci. 188 415422
31. Jackson, K.A. (2004) Kinetic Processes: Crystal Growth, Diffusion, and Phase
Transitions in Materials. Wiley-VCH, 2000
32. Janney, D.E., M. Cowley, J.M. and Buseck, P.R. (2000) Structure of synthetic 2-line
ferrihydrite by electron nanodiffraction, Am. Miner. 85, 11801187
33. Jentzsch, T.L. and Penn, R.L. (2006) Influence of Aluminum Doping on Ferrihydrite
Nanoparticle Reactivity J. Phys. Chem. B,, 110 , 1174611750
34. Jia, C., Cheng, Y., Bao, F., Chen, D. and Wang, Y (2006) pH value-dependant growth
of alpha-Fe2O3 hierarchical nanostructures. J. Crystal Growth, 294, 353-357
35. Jolivet, J.-P., Chanac, C., and Tronc, E. (2004) Iron oxide chemistry. From molecular
clusters to extended solid networks Chem. Commun, 481 - 487
36. Kubicki, J.D., Paul, K.W., and Sparks, D.L (2008) Periodic density functional theory
calculations of bulk and the (010) surface of goethite. Geochem Trans. 9, 4-12
37. De Levie, R. (2001) How to use excel in analytical chemistry and in general scientific
data analysis. University press Cambridge UK

28

38. Li, H.X., Addai-Mensah, J., Thomas, J.C. and Gerson, A.R. (2005) The crystallization
mechanism of Al(OH)(3) from sodium aluminate solutions J. Crystal
Growth, 279, 508-520
39. Loan, M., Newman, O.G.M., Farrow, J.B., and Parkinson, G.M. (2008) Effect of rate of
crystallization on the continuous reactive crystallization of nanoscale 6-line
ferrihydrite. Crystal Growth & Design, 8, 1384-1389
40. Lovgren, L., Sjciberg, S, and Schindler, P. (1990) Acid/base reactions and Al(III)
complexation at the surface of goethite. Geochim. Cosmochim. Acta, 54, 1301-1306
41. Martnez, C.E. and McBride, M.B. (2001) Cd, Cu, Pb, and Zn coprecipitates in Fe
oxide formed at different pH: aging effects on metal solubility and extractability by
citrate. Env. Toxic. Chem. 20, 122126.
42. McCarty, D.K., Moore, J.N., and Marcus, W.A. (1998) Mineralogy and trace element
association in an acid mine drainage iron oxide precipitate; comparison of selective
extractions. Applied Geochem. 13, 165-176
43. Meier, R.J. (2005) On art and science in curvefitting vibrational spectra. Vibr.
Spectroscopy 39 266-269
44. Morterra, C., Chiorino, A., and Borello, E. (1984) An IR spectroscopic characterization
of -FeOOH (goethite). Mat Chem. Phys., 10, 119-138
45. Nagano, T., Nakashima, S., Nakayama, S., Osada., K., and Senoo, M. (1992) Color
variations associated with rapid formation of goethite from proto-ferrihydrite at pH 13
and 40 C. Clays and Clay Miner., 40, 6013-607
46. Nagano, T., Nakashima, S., Nakayama, S., and Senoo, M. (1994) The use of color to
quantify the effects of pH and temperature on the crystallization kinetics of goethite
under highly alkaline conditions. Clays and Clay Miner., 42, 226-234
47. Penn, R. L. (2004) Kinetics of Oriented Aggregation. J. Phys. Chem., 108, 1270712712.
48. Penn, R.L., Tanaka, K., and Erbs, J.J. (2007) Size Dependent Kinetics of Oriented
Aggregation. J. Crystal Growth, 309 97-102.
49. Patterson, A.L. (1939)The Scherrer Formula for X-Ray Particle Size Determination".
Phys. Rev. 56, 978982
50. Rochester, C.H., and Topham, S.A. (1978) Infrared study of surface hydroxyl groups
on goethite. J. Chem. Soc. Faraday Trans. I 75, 591602.
51. Ruan, H.D. Frost., R.L., and Kloprogge, J.T. (2001) the behavior of hydroxyl units of
synthetic goethite and its dehydroxylated product hematite. Specrtochimica Acta, Part
A Molecular and Biomolecular Spectroscopy, 57, 2575-2586

29

52. Ruan, H.D. Frost., R.L., Kloprogge, J.T., and Duong, L. (2002) Infrared spectroscopy
of goethite dehydroxylation. II. Effect of aluminum substitution on the behavior of
hydroxyl units. Spectrochim. Acta, 58 479491
53. Russel, G.D. (1979) Infrared spectroscopy of ferrihydrite: evidence for the presence of
structural hydroxyl groups. Clay Miner., 14, 109.
54. Schoonover, J.R., Marx, R., and Zhang, S.L. (2003) Multivariate curve resolution in the
analysis of vibrational spectroscopy data files. Applied Spectroscopy 57, 154A-170A
55. Schoonover, J.R., Marx, R., and Nichols, W.R (2004) Application of multivariate curve
resolution analysis to FTIR kinetics data. Anal Chim. Acta 500 195210
56. Schwertmann, U. (1973) Use of oxalate extraction from soils. Canadian J. Soil Sci., 53,
244246.
57. Schwertmann, U. and Murad, E. (1983) Effect of pH on the formation of goethite and
hematite from ferrihydrite. Clays and Clay Miner., 31, 277284.
58. Schwertmann, U., Friedl, J, and Stanjek, H. (1999). From Fe(III) ions to ferrihydrite
and then to hematite. J.Colloid Interface Sci., 209, 215-223
59. Schwertmann, U. and Taylor, R.M. (1989) Iron oxides. In: Dixon, J.B.Weed, S.R.,
editors. Minerals in Soils Environments. 2nd ed. Madison, Wisconsin: Soil Sci Soc Am,
379-439.
60. Schwertmann, U., and R.M. Cornell. 1991. Iron oxides in the laboratory. VCH Publ.,
Weinheim, Germany
61. Shaw, S., Pepper, S.E., Bryan, N.D., and Livens, F.R. (2005) the kinetics and
mechanisms of goethite and hematite crystallization under alkaline conditions, and in
the presence of phosphate. Amer. Mineral., 90, 18521860
62. Singh, S.S. and Kodama, H. (1994) Effect of the presence of aluminum ions in iron
solutions on the formation of iron oxy)hydroxides (FeOOH) at room temperatures
under acidic environment. Clays and Clay Miner., 42, 606-613
63. De Juan, A., and Tauler, R. (2007) Multivariate Curve Resolution (MCR) from 2000:
Progress in Concepts and Applications, Critical Rev. Anal. Chem, 36, 163 - 176
64. Tauler, A.J.R. (2003) Chemometrics applied to unravel multicomponent processes and
mixtures. Revisiting latest trends in multivariate resolution Anal Chim. Acta 500 195
210
65. van der Zee C., Roberts D.R., Rancourt D.G., and Slomp C.P. (2003) Nanogoethite is
the dominant reactive (oxy)hydroxide phase in lake and marine sediments. Geology 31,
993-996

30

66. Waychunas, G.A., Rea B.A., Fuller, C.C., and Davis, J.A. (1993) Surface chemistry of
ferrihydrite: Part 1. EXAFS studies of the geometry of coprecipitated and adsorbed
arsenate Geochimica et Cosmochimica Acta, 57, 2251-2269
67. Waychunas, G.A. (2001) Structure, aggregation and characterization of nanoparticles
Nanoparticles and Env. 44, 105-166
68. Waychunas, G. A., Kim, C.S, and Banfield, J.F. (2005) Nanoparticulate iron oxide
minerals in soils and sediments: unique properties and contaminant scavenging
mechanisms. J. of Nanoparticle Res. 7, 325-571
69. Weckler, B. and Lutz, H.D. (1998) Lattice vibration spectra. Part XCV. Infrared
spectroscopic studies on the iron oxide hydroxides goethite (), akagankite (),
lepidocrocite () and feroxyhyte (). Eur. J. Solid State Inorg. Chem. 35, 531-544
70. Wigginton, N.S., Haus,K.L., and Hochella, M.F. (2007) Aquatic environmental
nanoparticles. J. Env. Monitoring, 9, 1306-1316
71. Yee, N., Shaw, S., Benning, L.G. and Nguyen, T.H. (2006). The rate of ferrihydrite
transformation to goethite via the Fe(II) pathway. Amer. Mineralogist., 91 9296
72. Zhang, Z., Wu, S., Ren, M., and Xiao C. (2004) Model of cold crystallization of
uniaxially oriented poly(ethylene terephthalate) fibers Polymer, 45, 4361-4365
73. Zhao, J., Huggins, F.E., Feng, Z., and Huffman, G.P. (1994) Ferrihydrite surface
structure and its effects on phase transformation. Clays and Clay Minerals, 42, 737-746
74. Zhong, L.S, Hu, J.S; Liang, H.P et al., (2006) Self-assembled 3D flowerlike iron oxide
nanostructures and their application in water treatment. Adv. Mat., 2426

31

Figures
6000

6000

2% Al
Go(111)

Go(130)

5000

Go(110)

0% Al

5000

Go

Go

Go(221)
Go(022)

Go (101)
(110)

Relative intensity

Relative intensity

4000

4000

(110)
Go (101)

54 days

3000

3000

54 days

2000

23 days

23 days

2000
9 days

9 days

Fh

1000

1000

2 days

Fh

2 days
0 days

Fh
0
0.5

0.4

0 days

Fh

0.3
0.2
d-space, nm

0.1

0
0.5

0.4

0.3
d-space, nm

0.2

0.1

Figure 2-1. X-ray diffractograms for FeIII-based nanocolloids synthesized with 0 or 2 mol%
incorporated-Al at initial conditions (0 days) and after aging in dilute suspensions for 2, 9, 23,
and 54 days. Go stands for goethite, numbers in parenthesis indicate crystal face; Fh denotes
ferrihydrite.

32

6
Crystallite size, nm

0% Al
2% Al

2
1

10

100

Time, days (log scale)

Figure 2-2. Crystallite sizes for goethite as a function of time. Crystallite sizes were estimated
from the XRD (110) peak using the Scherrer equation (0%Al, diamonds; 2% Al, squares).

33

A.

.
B.

Figure 2-3. (A) Crystal structure of goethite. Fe is in octahedral coordination surrounded by O2and OH- ligands (O, red; H, white). (B) Closer look showing hydrogen bonds, O-HO (dashed
lines).

34

OH-stretching vibrations

890

-3

0%aAl x 10

790

3200

Absorbance (arbitrary units)

OH-bending vibrations

Time
days6

54

23

0
38003600 3400 32003000 2800 2600

3800

3400

3000

900
800
700
2600 950 900 850 800 750 700
wavenumber, cm-1

-3

x 10

3200

2%b Al
5

Time
days
-3

x 10

54

234

22
0

0
3800

3400

3000

2600

3800

3400

3000

2600

900

800

700

950 900 850 800 750 700

Wavenumber / cm-1
Figure 2-4. ATR-FTIR spectra as a function of time for Fe-hydroxide suspensions with 0% Al
(upper panel) and 2% Al (lower panel). Multiple lines at a given time reflect both batch and
instrument replicate variability.

35

stoichiometric OH

Stoichiometric OH
Non-stoichiometric OH

0Al
2Al

adsorbed water

peak position

3250

Time=54 days

3220
3190
3160
3130
-10

4000

3800

3600

3400

3200

3000

2800

10

2600

20

30

40

50

60

time, days

stoichiometric OH
0Al
2Al

FWHM

320
270
220

Time=0 days

170
4000

3800

3600

3400

3200

3000

Wavenumber, cm-1

2800

2600

-10

10

20

30

40

50

60

time, days

Figure 2-5. (A) Example of Gaussian band analysis for the OH-stretching region (0%Al
suspensions: upper plot, suspension aged for 54 days; bottom plot, initial suspension at 0 days).
Solid lines are experimental spectra, dotted lines are the Gaussian components, and broken lines
are the sum of the Gaussian components. Stoichiometric OH peak position (B) and full width at
half maximum (FWHM) (C) as a function of time for 0%Al and 2% Al suspensions.

36

Experimental data

Component 1

Component 2

B
b

Component 2
1

Component 2

Component 1
0.8

scores

0.8

0.6
0.6
0.4
0.4
0.2

0.2
0

10

20

30

40

50

60

10

20

Time, days

3800

60

0% Al
3600

3400

3200

3000

2800

2600

2% Al
4000

50

Component 2

0% Al
3800

40

Time, days

Component 1

4000

30

4000

3800

3600

3400

3200

3000

2800

2600

3400

3200

3000

2800

2600

2% Al
3600

3400

3200

3000

wavenumber

2800

2600

4000

3800

3600

wavenumber

Figure 2-6. MCR analysis of OH-stretching region. (A) Components extracted from the 0% Al
experimental dataset; these two components describe all spectral variability. (B) Scores (relative
ratios) for Component 1 (ferrihydrite-like) and Component 2 (goethite-like) as a function of
time. (C) Gaussian band analysis indicates each component (for both 0% and 2% Al
suspensions) is consistent with combination of stoichiometric hydroxyls, non-stoichiomeric
hydroxyls, and adsorbed water.

37

60
50
40
30
20
10
0

Particle diameter, nm

B.
A

0% Al

60
50
40
30
20
10
0

10

100

2% Al

10

100

time, days (log scale)


A.
B

30

Volume (%)

Volume (%)

54

20

0% Al

15

10
5
0
0

B.

Volume (%)

23

25

30

20

25

40

54

100

2% Al

15

80

23

20

60

10
5
0
0

20

40

60

80

100

size, diameter in nm

Figure 2-7. (A) Particle sizes as determined by dynamic light scattering for 0%Al (upper panel)
and 2%Al (bottom panel) suspensions. Dark grey symbols show the results for a suspension
measured after aging for 0, 2, 9, 23 and 54 days. Light symbols show data obtained by in-situ
heating of an initial suspension (0 days) at 50 C every 2 hours for about 2 days in automatic
mode. (B) Volume-based particle size distributions for 0%Al (upper panel) and 2%Al (bottom
panel) suspensions aged for 0, 2, 9, 23, and 54 days (indicated by numbers with arrows).

38

Figure 2-8. Representative TEM images of iron hydroxide nanoparticles. Images A and B show
non-aged particles (0 days) from 0%Al and 2%Al suspensions, respectively. TEM revealed small
(<~30 nm) goethite needles in Al-free suspensions at 0 days (insert in A), while no goethite
needles were observed in 2%Al suspensions (B). (C) Al-free suspensions aged for 54 days
exhibit three populations of nanoparticles: round particles less than 10 nm in diameter
(ferrihydrite), thin needles of goethite about 50 nm in length and several nm wide, and large
crystals (prisms) of goethite 50-100 nm in length and 10-20 nm in diameter. (D) Ferrihydrite
particles are more abundant in 2%Al suspensions aged for 54 days although a few goethite
needles and prisms are observed.

39

Figure 2-9. TEM images of diluted samples. Images A and B show non-aged particles (0 days)
from 0%Al and 2%Al suspensions, respectively; C and D - Al-free and 2%Al suspensions aged
for 54 days, respectively.

40

Figure 2-10. High-resolution TEM images of 0% Al suspensions aged for 0 days (A) and 54
days (B and C). (A) A needle formed from aggregated particles (0 days). (B) Large goethite
crystal shows single-crystal lattice fringes in agreement with a spot diffraction pattern (insert).
(C) Goethite needle consisted of smaller grains with different lattice orientation. Diffraction
pattern produced diffraction rings characteristic of polycrystalline materials.

41

Aggregation and growth of polycrystalline goethite particles

Goethite coprecipitates with


ferrihydrite

Goethite

Fe(III) aq

Fe(III) aq

Ferrihydrite

Single
crystal
goethite

Dissolution-precipitation by monomer addition


-

Time
Figure 2-11. Schematic diagram illustrating the transformation of ferrihydrite to goethite.

42

(110)
peak
XRD XRD
peak area
80000

0.04
0.8

0% Al

0% Al

60000

2% Al

0.4
0.02

Peak area

0.03
0.6

Scores

Fraction of component
in mixture

MCR Stoichiometric
component
region
Goethite-like
MCR- OH-stretching
component

0.01
0.2

2% Al

40000
20000

0.00
0.0

B
0

20

40

60

20

Time, days

-1 OH-bending
Gauss
height - OH- bending
region
890900
cmpeak
vibrations

0.005

0.005

0% Al

0.004
0.003

2% Al

0.002

0% Al

0.004

Height

Height

60

Time, days

Gauss
800-1peak
height - OH- bending
region
790 cm
OH-bending
vibrations

0.001

0.003
0.002

2% Al

0.001

0.000

40

0.000
0

10

20

30

Time, days

40

50

60

10

20

30

40

50

60

Time, days

Figure 2-12. Plots of ATR-FTIR- and XRD- derived parameters for 0%Al and 2%Al
suspensions as a function of time. These parameters reflect the presence of the crystallization
product (goethite). Solid circles for 0% Al and open circles for 2% Al suspensions. Solid (0%
Al) and dotted (2% Al) lines are data fits using the first order equation [Goethite] = [Goethite]0 *
(1 exp-kt).

43

0% Al
XRD: (110) peak area

2% Al
0% Al

FTIR: Goethite-like MCR component

2% Al
0% Al

FTIR: 790

cm-1 OH-bending

2% Al
0% Al

FTIR: 890 cm-1 OH-bending


2% Al
0

8
-7

rate constant 10 s

10

12

-1

Figure 2-13. Rate constants for goethite appearance obtained from ATR-FTIR- and XRDderived parameters for 0%Al and 2%Al suspensions.

44

1E-03
-3

Rate constant, s

-1

10

1
33-50C

1E-04
10-4

60C

50C

50C

80C
60C

40C

40C

70C
1E-05
10-5

5
25C

1E-06
10-6

0% Al this study, 50 C
2% Al
10-7

1E-07

10

11

12

13

14

pH

Figure 2-14. Comparison of first-order rate constants for goethite crystallization obtained in this
work to available literature data. Authors are indicated by numbers: 1. Shaw et al. (2005); 2.
Nagano et al. (1994); 3. Fischer (1971); 4. Yee et al.(2006); 5. Ford et al. (1999, 2005). Methods
used in calculating the rate constants: - synchrotron XRD; - oxalate extraction; x colorimetry. The aging temperature (oC) at which experiments were conducted is also indicated.

45

Chapter 3
ATR-FTIR, XRD AND PERIODIC-DENSITY-FUNCTIONAL THEORY STUDIES OF
MINERAL PHASES IN CO-PRECIPITATED FE AND AL HYDROXIDE

Abstract
In soils, iron and aluminum can co-precipitate forming metastable nanoparticles, which
eventually crystallize into more ordered Fe and Al hydroxides. Model Fe-Al mixed colloids with
a broad range of chemical composition (Fe/Al molar ratios) were synthesized under experimental
conditions close to environmental (pH 5, [Fe+Al] = 10 mM, low ionic strength). Changes in
mineral composition as a function of Al-substitution and suspension aging time were followed
using ATR-FTIR and synchrotron-based XRD. It was found that low amounts of Al substitution
(2-8 mol %) lead to the formation of moderately crystalline Al-goethite upon aging, while at
intermediate Al substitution (10-20%) we found long-lived relatively non-crystalline
nanocolloids. Gibbsite (Al(OH)3) microcrystalline structures and ferrihydrite appeared in initial
nano-colloidal suspensions containing 25% Al substitution. Aging of the nano-colloidal
suspensions containing 25% Al resulted in the formation of crystalline Fe oxides (goethite and
hematite) and gibbsite, thus suggesting nanocolloids formed two different phases. In addition,
periodic density functional theory (DFT) calculations were performed using bulk goethite to
investigate the structural arrangement of Al atoms that will lead to the most stable Al-goethite
structure. We tested two levels of Al substitution (8% and 25%) and two arrangements of Al
atoms (isolated and clustered) within the goethite structure. Al in diaspore-like clusters was
found to be more energetically favored than isolated Al substitution. Optimized unit cell
parameters for Al-goethites with clustered Al substitution showed good agreement with available
experimental literature data, while isolated Al substitution did not. These results suggest that
46

when Al co-precipitates with Fe-hydroxides from solution, it prevents goethite crystallization by


forming clusters within the goethite structure with futher formation of separate Al-hydroxide
phase (gibbsite).

3.1. Introduction
Iron and aluminum hydroxides (Fe(OH)3, Al(OH)3) are of great importance in
environmental, soil and material sciences, particularly as related to the sorption and transport of
nutrients and contaminants and in soil remediation applications (Kasprzyk-Holdern, 2004; Davis
et al., 2002; Farmer, 1974; Lundstrom and Bain, 2000). As a result of disequilibrium conditions
leading to dissolution of rock-forming minerals, Fe and Al can co-precipitate forming nanoparticulate mixtures or coatings on soil substrates (Coston et al., 1995; Penn et al., 2001). These
nano-phases are typically of low-crystallinity and are represented by particles less than 0.1 m in
size. With a high surface area and variable surface charge, these nano-colloidal materials can
extensively retain nutrients (Zn2+, PO43-) and contaminants (Pb2+, Cd2+, AsO43-) thus controlling
their biological availability (Coston et al., 1995; Davis et al., 2002). However, the reactivity of
mixed Fe-Al hydroxides may vary upon aging and might be the result of different chemical
composition (Fe/Al ratio), which can define the crystallization rate and pathway. To predict their
environmental behavior and long-term reactivity in soils and natural waters, the structural
transformations of Fe-Al hydroxide nanocolloids need to be understood as a function of time and
chemical composition.
Previous studies have shown that initial metastable mixed Fe-Al co-precipitates are
converted to more stable crystalline Fe and/or Al oxides (Fey and Dixon, 1981; Colombo and
Violante, 1996; Bazilevskaya et al., 2009), and that the Fe:Al ratio in mixed Fe-Al hydroxides is
likely to influence the reactivity of these hydroxides towards solutes (Coston et al., 1995;
47

Jentzsch and Penn, 2006) and may also dictate the rate of crystallization (Bazilevskaya et al.,
2009) and the identity of the crystalline products (Violante et al., 1998). Transformation products
may include goethite (-FeOOH), Al-substituted goethite ((Fe,Al)OOH), gibbsite (Al(OH)3) and
their poorly crystalline precursors. In particular, goethite, an ubiquitous and environmentally
important mineral in soils, rarely exists as a pure phase, but often co-precipitates and crystallizes
in the presence of Al. Substitution of Al for Fe in the crystal structure of goethite changes its
solubility, crystallization rate and surface properties (Eggleton, 1987; Goh et al., 1987;
Schwertmann and Taylor, 1989; Violante et al., 1998; Bazilevskaya et al., 2009). Also, low Al
substitution of goethite has been shown to decrease at least one of its unit cell parameters thus
leading to strain within the goethite structure (Schulze, 1982; Marjzlan and Navrotsky, 2003;
Alvarez et al., 2007; Blanch et al., 2008). The degree of such strain may depend on the structural
arrangement of Al atoms after incorporation into the goethite structure. Structural arrangements
of Al atoms include their distribution at isolated positions within the goethite structure, the
formation of Al clusters within the goethite structure, and/or the formation of an epitaxial phase
or a separate Al-phase.
Although a considerable amount of research on Fe-Al co-precipitates exist, especially on
Al-substituted goethites, most of the experiments were conducted using high concentrations of
metals, high temperatures, and rapid neutralization with concentrated base at high pH to produce
well-crystalline micron-size precipitates, i.e. conditions which are not environmentally relevant
(Schulze, 1984; Cornell and Schwertmann, 2003; Wells et al., 2006; Alvarez et al., 2007). In
soils, however, mixed Fe-Al hydroxides are typically of low-crystallinity and represented by
colloids less than 0.1 m in size. These Fe-Al co-precipitates may remain suspended in soil
water depending on their Fe:Al ratio, soil pH, soil redox status, and the presence of other cations
and organic matter.
48

Substitution of Al has two important effects on the goethite structure. First, substitution
of Al3+ for Fe3+ induces strain because Al-O bonds are shorter than Fe-O bonds. Second, it was
found that Al3+ is less suitable for incorporation into goethite due to its higher charge compared
to Al(OH)2+, and this is supported by the observation that Al-substitution is not favored at very
low pH (Wolska, 1997). Therefore, Al-substitution is often accompanied by replacement of 3O2by 3OH- in a vacant site, creating excess structural OH and disruption of hydrogen bonds in the
structure which also induces strain in the structure (Stanjek and Schwertmann, 1992; Ruan et al.,
2002). Unfortunately, there is not enough information on how Al is distributed within the
goethite structure to estimate its effect on the stability of goethite. Blanch et al. (2008) calculated
the probability distribution of next nearest Al neighbors assuming a true random solid solution
with no site preference for Al substitution onto Fe sites. They found that at 6 mol % Al
substitution, 39% of Al sites would have one or more Al in adjacent Al sites, while at ~33% Al
content, there will be two or three Al ions in the second coordination sphere which are linked by
edge-sharing to form Al-rich clusters. To investigate the most favorable or most likely structural
arrangement of Al incorporation in goethite (isolated versus clustered atoms), we performed ab
initio calculations on Al-substituted goethite structures with two levels of substitution (8 and 25
mole percent).
We use laboratory-synthesized Fe-Al coprecipitates to determine the mineral composition
of nanoparticles as a function of Al mole fraction and aging time in dilute suspension at pH 5,
conditions chosen to be close to a realistic soil water environment. We also investigate the most
favorable arrangement of Al atoms (isolated versus clustered) within the goethite structure by ab
initio periodic density functional theory (DFT) calculations. In this respect, our research
hypotheses were that (1) increasing concentrations of Al in mixed Fe-Al hydroxide nanoparticles
suspensions lead to the formation of intermediate phases that are metastable in long-term in low49

temperature soil environments, and (2) at low Al concentrations, Al incorporates into the
goethite nano-particle structure by forming interstitial or surface clusters of Al in the same
structural arrangement, which create less strain on the goethite structure compared to isolated
interstitial substitution of Al for Fe.

3.2. Methods
3.2.1. Experimental
3.2.1.1. Synthesis and aging of Fe-hydroxide and Al-doped Fe-hydroxide nanoparticles
The following procedure, modified from Bakoyannakis et al. (2003) was specifically
designed to produce nano-sized iron hydroxide particles. Precise masses of Fe(III)-nitrate and
Al-nitrate salts were dissolved in 100 mL of Milli-Q water to yield 0, 2, 4, 8, 10, 12, 16, 20, 25,
30, 50, 75, and 100 mole % Al in solutions with a 10-1 M total metal, M, (=Fe+Al) concentration.
Solutions were stirred at 25 C and titrated to pH 5.0 with 0.100 N KOH additions at the rate of
equal to or less than 1 mL min-1 using a pH-controlled Masterflex C/LTM pump system. Typically
after about 1 h the pH had stabilized; the final volume of was then adjusted to 1 L, yielding a
final total (Fe+Al) concentration of 10-2 M. The suspensions were then dialyzed at 25 C against
Milli-Q water using 1 nm pore size Spectropor/7 dialysis tubes (molecular weight cutoff 1000) to
remove excess salts and Al so presumably all Fe and Al contained in the dialysis tubes were in
nano-particulate form. The water was replaced several times over about XX h until conductivity
was reduced to 0.3 S/cm (ionic strength value < 5x10-3). The average particle size of these
yellow-brown, optically transparent dialyzed suspensions was in the range 100-300 nm, as
determined by dynamic light scattering with Malvern Zetasizer Nano Series instrument. Aliquots
for the aging experiment at 50 C in glass vials were taken immediately after dialysis (time=0),
and the suspensions were characterized at 2, 5, 9, 15, 23, and 54 days. Stock starting suspensions
50

were also stored at 4 C in the dark in polypropylene bottles. During the course of experiment
both the stock and incubated suspensions stayed visually clear without any precipitates.
3.2.1.2. Characterization of Fe-hydroxide and Al-doped Fe-hydroxide nanoparticles
The metal composition of aged mixed Fe-Al hydroxide nanoparticles was determined by
digesting 1 ml of suspension with 2 ml of concentrated HNO3 at 80 C for 1 hour. The resulting
solution was diluted by a factor of 100 with Milli-Q water, and the Fe and Al were determined
by inductively coupled plasma atomic emission spectroscopy (ICP-AES).
ATR-FTIR measurements were performed using a Bruker Tensor 27 Fourier transform
infrared spectrometer equipped with a diamond-ZnSe attenuated total reflection sensor. Thin
films of the nanoparticles were prepared directly on the ATR sensor by air-drying 0.5 L
aliquots for approximately 5 min. Spectra at 8 cm1 resolution were acquired by co-addition of
256 scans within the 4000-600 cm-1 region. A sensor background was taken before each sample,
and measurements were replicated at least three times. Spectral manipulation including
subtraction of background water and CO2 vapor, baseline correction, and spectra normalization
were performed using in-house developed codes under MATLAB 7.4 computing environment.
High-resolution synchrotron-based XRD patterns were collected in beamline X16C at the
National Synchrotron Light Source, Brookhaven National Laboratory (Upton, New York) using
a X-ray of wavelength of 0.700517 selected with a Si(111) double crystal monochromator.
Freeze-dried powder samples were loaded into a glass capillary (0.7 mm diameter) and rotated
about the longitudinal axis during data collection. XRD patterns were collected in the 2 range
from 2 to 40 in steps of 0.05. Jade+ software (version 8.2, Materials Data, Inc., Livermore,
CA) was used for the determination of crystalline phases.

51

3.2.2. Periodic density functional theory calculations of bulk Al-goethite model structures
Bulk -FeOOH and Al-substituted -FeOOH model structures were built using the
Crystal Builder in Cerius2 software (Accelrys Inc., San Diego, CA). Initial unit cell parameters
and atomic positions for -FeOOH were taken from experimental neutron X-ray diffraction
measurements of goethite (Szytula et al., 1968). The unit cell of goethite contains 4 Fe atoms, but
calculations were performed using a 1x3x2 supercell model (24 Fe atoms, a = b = c = ~9 ) to
increase the possible number of Al-substitutions and make the energy calculations more
isotropic. Four Al-goethite structures were created: (1) two Al atoms substituting for 2 Fe atoms
(8% Al substitution) in a cluster (Fig. 3-7A); (2) two Al atoms substituting for 2 Fe atoms (8%
Al substitution) that are isolated from each other (Fig. 3-7B); (3) six Al atoms substituting for 6
Fe atoms (25% Al substitution) in a cluster (Fig. 3-8A); and (4) six Al atoms substituting for 6
Fe atoms (25% Al substitution) all isolated from each other (Fig. 3-8B). Our distinction between
isolated and clustered arrangements of Al atoms is based on the degree of Al-Al separation. The
diaspore crystal was also modeled, using experimental data of Hill (1979).
DFT calculations were performed using the Vienna Ab Initio Simulation Package (VASP;
Kresse and Furthmller, 1996). All atomic positions and lattice parameters were allowed to relax
during geometry optimization. The Fe_pv, Al_h, O_h and H_h pseudopotentials were used for
Fe, Al, O and H atoms, respectively. An energy difference between iterations of 0.0004 eV/atom
was taken as the convergence criterion. L(S)DA+U parameters (on-site columbic interaction)
taking into account strong intra-atomic interaction were used to better describe atoms with
localized (strongly correlated) d and f electrons, such as Fe atoms within the goethite structure
(Kubicki et al., 2008).
In choosing positions for the Al atoms within the goethite structure, the magnetic
properties were considered. Fe is in a high-spin ground-state in goethite (Kubicki et al., 2008).
52

Pure bulk crystalline goethite orders antiferromagnetically, that is, the structure has equal
number of positive and negative spins with neighboring spins pointing in opposite directions so
the net magnetic moment is equal to 0. The spins are oriented along the b-axis with up and down
spins in alternate chains of octahedra (Cornel and Schwertmann, 2003).
Pollard et al. (1991) found that Al substitution induces ferrimagnetism in Al-substituted
goethite, i.e. the alignment of spins is unparallel as for antiferromagnetic materials, but the
different spins have unequal moments so that the structure has a net magnetic moment.
According to Murad and Bowen (1987), Al can lower the Nel temperature which would result
in a paramagnetic structure (spins become parallel under an external magnetic field and exhibit
magnetism). The magnetic hyperfine field (Bh) determined from Mssbauer spectroscopy has
been shown to decrease as structural Al increases (Cornell and Schwertmann, 2003). Aluminum
acts here in two ways: (1) directly, when the diamagnetic Al does not contribute to the magnetic
field, and (2) indirectly, when Al reduces crystal size which also decreases Bhf. Aluminum
substitution results in the asymmetric broadening of six narrow lines in the goethite Mssbauer
spectra.
In our calculations the number of unpaired electrons for the Fe atoms were set to five
(i.e., d5 electronic configuration), and the Fe atoms were assigned positive and negative spins in
alternating layers to keep the overall magnetization close to zero using the MAGMOM
parameter. For Al-substituted structures, we chose the positions of the Al atoms so that the
number of Al in positive and negative layers and number of positive and negative Fe spins
were the same; this resulted in an antiferromagnetic structure, as for pure goethite.
Diaspore and gibbsite model structures were created using experimental data of Hill
(1979) and Saalfeld and Wedde (1974), respectively. Both models had 24 aluminum atoms and
the same input parameters as for Al-goethite calculations, except for MAGMOM.
53

3.3 Results and discussion


3.3.1. Phase transformation in Fe-Al hydroxide nano-particle suspensions
Fe-Al hydroxide nanoparticles were synthesized by slow base titration of initial
Fe(III) and Al(III)-nitrate containing solutions to pH 5. Titration curves of these solutions
showed that the extent of the buffer region depended on the Fe:Al ratio in initial solutions (Fig.
3-1). In titrations such as these, the pH is buffered during neutralization of iron [Fe(H2O)6]3+ ions
to produce a complex of lower charge, according to the reaction:
[Fe(H2O)6]3+(aq) + OH- [ Fe(H2O)5(OH)]2+ (aq) + H2O
This reaction continues until a neutral complex is produced, which rapidly polymerizes to
form an iron hydroxide:
[Fe(H2O)4(OH)2] +(aq) + OH- [Fe(H2O)3(OH)3] (s) + H2O
The buffer region was the largest for solutions with initial 0, 2 and 4% Al (Fig. 3-1),
which can be explained by the fact that Fe3+ has a higher hydrolysis constant than Al3+ (i.e., pKh1
= 2.19 for Fe3+ and pKh1 = 5.00 for Al3+). The presence of 8-25% Al, however, facilitated rapid
hydrolysis of Fe, which was indicated by a shorter buffer region, in agreement with observations
by Sigh and Kodama (1994). Values of OH concentration at the inflection point versus mole%
Al in initial solutions are plotted in Fig. 3-2. Inflection points were taken as the point of
maximum slope where pH changed most rapidly with base addition (Bertsch et al., 1989). Such
rapid pH change means that half the metal aqua/hydroxide complex at that pH that could react
with hydroxide has done so. The steepest inflections indicate reactions that are favored in the
forward direction - therefore with limited pH buffering. Two distinct inflection points for 30, 50
and 75% Al suggested the formation of two separate hydroxide phases: first Fe-hydroxide and
then Al-hydroxide (Figs. 3-1 and 3-2). The amount of Al in nanoparticles was the same as in
54

initial solutions (Fig. 3-3), except for 50% and 75% Al suspensions where significant amounts of
Al did not polymerize at pH 5 and were leached out in ionic form during dialysis.
ATR-FTIR spectra of nanoparticles revealed gibbsite formation at initial conditions (0
days, 25 C) in suspensions with 25% Al and higher (Fig. 3-4A), while little or no crystalline
phases formed in suspensions with lower Al concentrations, as indicated by broad peaks in the
OH-stretching region and few peaks in the fingerprint range. Upon aging for 54 days at 50 C,
goethite was present in suspensions with 0%, 2% and 4% Al (Fig. 3-4B); however, mostly
amorphous materials were present in suspensions with 8 to 20% Al. The strength of gibbsite
peaks in aged suspensions grew with increasing Al mole fraction (25 to 100% Al) in
nanoparticles. It can be seen that the presence of Al (4% and higher) (panel B) significantly
suppresses formation of goethite.
Synchrotron-XRD patterns were in good agreement with infrared analysis of initial
suspensions (Fig. 3-5A) showing a broad 2-line pattern characteristic of poorly-crystalline
material assigned as ferrihydrite (Schwertmann and Cornell, 1991) for suspensions with 0 to
20% Al, and a gibbsite pattern for suspensions with 25-100% Al. However, in aged suspensions,
hematite was observed in nanoparticles with low (4% and 8%) and high (25-75%) Al
concentrations (Fig. 3-5B). These results were in agreement with earlier observations that the
presence of Al favors formation of hematite (Ruan et al., 2002; Cornell and Schwertmann, 2003).
At intermediate (12 to 20%) Al concentrations, poorly crystalline Fe and Al hydroxides persisted
for the duration of the experiment.
ATR-FTIR- and XRD-determined mineral phases present in mixed Fe-Al hydroxide
suspensions as a function of Al mole fraction and aging time are summarized in Table 3-1.
Gibbsite formed at time = 0 days in suspensions with 25% Al and higher and persisted through
all aging times. For those suspensions with >25% Al, goethite (and hematite) were formed upon
55

aging. No diaspore was detected. The more interesting range is 20 mol% Al and below, where it
appears the Fe-based nanoparticles may be incorporating Al. Goethite and ferrihydrite are the
most important nano-particulate phases in this compositional range.
Theoretical solubility curves based on published Fe and Al hydroxide solubilities (Fig. 36A), and other studies (Casey et al., 2009) support our interpretation that Fe is hydrolyzed first in
our experiments. In Al-free and low Al (0-4%) nanoparticles, goethite (and sometimes hematite)
formed upon aging of the initially formed ferrihydrite; however, the intensity of goethite peaks
(both infrared and XRD) decreased dramatically as Al mole fraction was increased just slightly.
At low Al initial concentrations (< 12%), the solutions would be undersaturated with respect to
Al-hydroxide (Fig. 3-6B). Aqueous Al species could be scavenged from solution during coprecipitation with Fe-hydroxides and create disorder in Fe-hydroxides thus preventing their
transformation to goethite, but be expelled with greater aging time (Wolska, 1997). For the
intermediate Al-concentrations (up to 20%) goethite formation was completely prevented for a
long period of time due to formation of more stable, but poorly crystalline mixed Fe-Al
hydroxides. According to Fey and Dixon (1981), such mixed phases consist of nearly spherical
particles about 3-6 nm size. Aluminum probably suppresses the growth of iron hydroxide
particles by adsorption onto their surfaces, and the excess Al could remain as an amorphous
phase. At 25% Al and higher, Al solution concentrations are high enough to reach oversaturation
with respect to the amorphous pure aluminum hydroxide phase which can then transform to even
more insoluble gibbsite (Fig. 3-6). Since most of the Al was consumed to form quite insoluble
gibbsite, we suggest that over time this scavenged aluminum ions that otherwise slow the growth
of goethite. The beginning of this process that results in generation of two different nanoparticle
phases (gibbsite and goethite) is supported by the presence of two inflection points in our initial
titration data (Fig. 3-2).
56

3.3.2. Arrangement and concentration of Al in Al-goethites: energies and unit cell parameters
from DFT calculations
Optimized structural models for goethite with 2 and 6 Al atoms are shown in Figures 3-7
and 3-8, respectively. In the goethite model with 2 Al (8% Al substitution), the distance between
Al atoms (Al-Al) in clustered arrangement was 3.15 compared to Al-Al distance of 6.73 in
the model with isolated arrangement of Al atoms. In the structure with 6 Al atoms (25% Al
substitution) the average distance among Al atoms in clustered arrangement was 3.28 0.22 ,
while in isolated arrangement Al-Al distance was 4.98 0.57 .
To test the results from our calculations, it was imperative to present available literature
data on Al-substituted goethites in order to compare the observed and calculated crystal
structures. The cell volumes and three unit cell dimensions (a, b, and c) of synthetic goethites
from various sources were plotted as a function of Al substitution (Fig. 3-9). All dimensions
showed the general trend of smaller values with increasing Al-substitution; however, the cdimension (Fig. 3-9D) varied considerably for samples with the same amount of Al substitution
and showed the largest deviation from a Vegard line, which represents linear relationships
between the size of unit cells of goethite and diaspore (Schulze and Schwertmann, 1984). The
large amount of scatter in the c dimension might be caused by structural defects because
hydrogen bonds between double chains along the c-axis can be easily disrupted during crystal
growth whereas stronger bonds (51% ionic character) are present within double chains.
Therefore, the integrity of the double chains is preserved in the direction of the a and b axes
(Schulze, 1984).
In addition, it was found that both the extent of Al substitution into the goethite structure
and the cell parameters were sensitive to the mode of synthesis. Samples prepared using Fe(II)
and Fe(III) both followed Vergard law, when Al substitution was 20 mol% and lower. However,
57

samples prepared from Fe(II) reach much higher degrees of substitution than goethites prepared
from Fe(III) solutions and deviate from Vergard law at Al substitution >20 mol% (Fig. 3-9A).
Synthesis in Fe(II) systems involves oxidative hydrolysis of Fe(II) solutions and slower goethite
crystallization rates compared to Fe(III) systems, which leads to the larger number of structural
defects and lower crystallinity and make easier to incorporate defects such as Al in more
disordered structure.
Comparison of unit cell parameters for calculated goethite models with experimental data
showed that the models with clustered arrangements of Al within the goethite structure were in
good agreement with available experimental data (Fig. 3-9). However, isolated arrangement of
Al atoms in the goethite structure led to large deformations of the crystal structure resulting in a
substantial increase and decrease of the cell volume for 2Al-isolated and 6Al-isolated model
structures, respectively (Fig. 3-9A). Such discrepancies between experimental and calculated
structures suggest that the incorporation of Al as isolated atoms within the goethite structure is
highly unlikely. Goethite and both clustered models (with 2 and 6 Al) were in good agreement
with experimental data. This finding suggests that Al in clustered arrangement may be more
favorable compared to isolated Al (This is also supported by energy values, Table 3-2).
Optimized unit cell parameters and minimum energies are shown in Table 3-2 and in
Figures 3-9 and 3-10. One cannot compare directly the energies of the structures with different
numbers of Fe and Al atoms because the self-energy, i.e. the energy between electron and
nucleus in an atom, is different for Fe and Al. For each configuration within the same
stoichiometry, i.e., 2 and 6 Al-goethite models respectively, the clustered arrangement of Al
atoms was lower in energy than the isolated arrangement: -604.479 eV for 2Al-cluster versus 603.567 eV for 2Al-isolated; and -607.937 eV for 6Al-cluster versus -602.180 for 6Al-isolated
(Fig. 3-10). These results suggest that clustered Al atoms create less local strain on the goethite
58

structure than isolated Al atoms. Indeed, since there is a substantial difference in size between
the larger Fe- and the smaller Al-occupied octahedra in the solid solution (difference in Pauling
crystal radii between Fe3+ and Al3+ is ~17%) (Wolska (1997)), there will be local elastic strain
within the structure due to size mismatch. Isolated Al would cause deformation in a higher
number of Fe-octahedra, while Al in clustered arrangement would crystallize to form diasporelike clusters within the goethite structure.
We can also compare the results from our DFT calculations with the experimental results
of Majzlan and Navrotsky (2003). Following their approach, the reaction for the formation of
2Al-substituted goethite (8% Al substitution) from pure end-members, i.e., goethite and diaspore,
can be written as follows:
0.92*GT=0(Go) + 0.08* GT=0 (Dia) GT=0(2AlGo)
where GT=0(Go), GT=0 (Dia), and GT=0(2AlGo are corresponding VASP free energies (eV)
for goethite, diaspore, and goethite with 2 Al atoms in cluster arrangement, respectively (Table
3-2). The resulting GVASPrnx of this reaction was negative (-0.474 eV or -45 kJ/mol), which
suggests that the formation of goethite with 8% Al substitution is favorable. However, GVASPrnix
for goethite with 25% Al substitution was positive (1.13 eV or 109 kJ/mol), which means that its
formation is not favorable.
Comparing values of experimental enthalpy of formation Hf0 (5.81 kJ/mol) for
8% substituted goethite, obtained from high-temperature calorimerty measurements by Marzlan
and Navrotsky (2003) to GVASPrnix results we observed a ~9 times difference between these
values. It should be noted that comparison of VASP energies to thermodynamic data is not
entirely appropriate, because VASP energies are calculated at 0 K and do not include atomic
vibrations. In a strict sense, we would need to use a correction for the changes in formation
enthalpies (H0-298) over the temperature range from 0 K to 298.15 K:
59

Hrnx = GT=0(2AlGo) + H0-298(2AlGo) {0.92*[GT=0(Go) + H0-298(Go)] +


0.08*[GT=0(Dia) + H0-298(Dia)]}
While the values for H0-298(Go) and H0-298(Dia) are available from the literature (10.74
kJ/mol for goethite, from Majzlan et al., 2003; and 6.85 kJ/mol for diaspore, from Perkins et al.,
1979), no H0-298 values are available for Al-goethites, i.e. no dependence of heat capacities
versus T (0298.15) have been studied. When this information becomes available, the
comparison between VASP energies and thermodynamic data could be made.

3.3.3. Goethite versus Al-goethites: stability considerations


Ruan et al. (2001) reported that infrared vibrations of metal-OH bonds shifted to higher
energy going from goethite to diaspore end-members. This suggests Al can retain OH more
effectively than Fe in the structure of goethite because of the higher ionic potential of Al3+ due to
its smaller ionic radius (Ruan et al., 1995; Blanch et al., 2008). Also, Al-substitution was found
not only to favor incorporation of excess OH into goethites structure, but to retain hydroxyl
units within the goethite structure at higher temperatures, which suggests Al-goethite is
thermally more stable than pure goethite (Ruan et al., 2002). Based on these considerations, it
can be proposed that Al-goethites should be more stable than pure goethites. However, using
high temperature calorimerty Majzlan and Navrotsky (2003) found that Al-goethite was
metastable with respect to goethite and diaspore, and that the enthalpy of mixing for goethitediaspore solid solution can be described by the equation: H = W*Xgoethite*(1-Xgoethite), where
W=79.14 kJ/mol and X is the mole fraction of goethite. Using their values for the enthalpy of
mixing and other thermodynamic data (Stumm and Morgan, 1981), we calculated relative
stabilities for diaspore and Al-goethites (8 and 25 mol %) with respect to goethite from
differences between formation enthalpies of reaction products and reactants (Table 3-3). These
calculations showed the formation of Al-substituted goethites from pure goethite requires very
60

large excess of enthalpy, and therefore, they are thermodynamically less stable compared to endmember goethite and diaspore. It should be noted that the enthalpy value for the reaction
between goethite and 25% Al-goethite may not be valid, since at high Al concentration the
formation of goethite-diaspore solid solution will be highly unfavorable. The results from these
calculations should be viewed with great caution, however, because Al substitution within a Fe
hydroxide that result in the formation of bulk Al-goethites occurs while the primary precipitate is
forming, and not from Al diffusion into a crystalline goethite structure.
Furthermore, not all the Al in Al-goethites may be in solid solution, but possibly create
gibbsite like inclusions or coatings at Al concentrations > 10 mol % (Cervini-Silva and Sposito,
2002). Such inclusions, or clusters, can increase the stability of Al-goethites, which is supported
by our ab initio calculations that show goethite structures with Al clusters are more energetically
favorable. However, one needs to keep in mind that our DFT calculations were made using bulk
goethite models with Al directly substituting for Fe in the crystal structure of goethite, i.e. the socalled substitutional solid solution (Wolska, 1997). However, Al can occupy sites that are
normally empty in the crystal structure (interstitial solid solution). This is common for Algoethites formed from Fe(II) solutions, when simultaneous Fe oxidation and slower coprecipitation rates allow more Al to be incorporated into the Fe hydroxide structure nonstoichiometrically (Cornell and Schwertamnn, 2003). This can explain the deviation of
experimental cell parameters observed for Fe(II)-goethites from the Vegard line at high Al
concentrations (Fig. 3-9). In addition, Al can adsorb onto the surface of Fe-hydroxide clusters,
thus preventing these colloids from growing and making them stable in soil water. Such stability
was demonstrated in our experimental colloidal suspensions with intermediate Al concentrations
(4-20 mol %), which were stable for the duration of the experiment with respect to both
aggregation and phase transformation. For future work it will be useful to explore additional
61

arrangements of Al atoms including positions at the surface of goethite nanoparticles by creating


(010) -FeOOH surface models following the procedure of Kubicki et al. (2008).

3.3.4. Relative abundance of gibbsite and diaspore in soils


Our calculations suggested that goethite with diaspore-like clusters is energetically
favorable and this suggestion fits well our experimental data (Fig. 3-9A). However, diaspore was
not detected in any of our experimental suspensions, although some initial solutions were
oversaturated with respect to both diaspore and gibbsite (Fig. 3-6). This apparent discrepancy
may be explained by: (1) the concentration of diaspore was below the detection limit of ATRFTIR and synchrotron-based XRD; or (2) the formation of gibbsite was thermodynamically
favored. VASP free energies for optimized diaspore and gibbsite structures were -631.402 and 986.953 eV, respectively (Table. 3-2). It should be noted that we cannot compare directly the
VASP free energies of these two minerals because they contain different amounts of water in
their structures. On the other hand, we can estimate relative stabilities of gibbsite and diaspore
with respect to corundum (as the most stable Al-oxide) using thermodynamic data. Formation of
gibbsite from corundum can be described by the reaction:
Al2O3 + 3H2O 2Al(OH)3
therefore, the enthalpy of reaction, Hrxn0, can be calculated as the difference between
reaction enthalpies of product minus reactants:

Hrxn0 = 2Hgibbsite0 [Hdiaspore0 + 3Hwater0]


Our calculations showed that gibbsite is thermodynamically more stable than diaspore
with respect to corundum (Table 3-4). These calculations confirm natural observations that in
soil environments gibbsite is more stable than diaspore and boehmite (-AlOOH) in the presence
of water (Chesworth, 1975).

62

Direct formation of gibbsite can occur under conditions of high rainfall, tropical
temperatures, and in the presence of basic old rocks of advanced weathering (Hsu, 1989).
However, the presence of certain organic acids can promote the formation of boehmite, which
can persist in soils if dry conditions occur before its conversion to gibbsite. Diaspore will form
over gibbsite only if water activity is less than 1, for example, upon evaporation in clays when
water activity progressively diminishes with decreases in pore size (Tardy and Nahon, 1985).
However, the formation of diaspore directly from Al saturated aqueous solutions, where water
activity is equal to 1, is highly unlikely. Also, it has been reported that surface Al-OH is more
stable than Al-O in aqueous systems (Hsu and Bates, 1964), therefore diaspore and boehmite can
be unstable when immersed in water. Thus, occurrences of diaspore and boehmite in soils are
rare. Even in extremely advanced weathering environments such as those encountered in bauxite
deposits, gibbsite is the dominant phase with much less abundance of diaspore or boehmite (Hsu,
1989).
When soil water is undersaturated with respect to Al solid phases, and aqueous Al species
co-precipitate with Fe-hydroxide, Al-substituted goethite will eventually form. Even as Al in Algoethites will most probably form clusters, as suggested by our calculations; there is no
unequivocal evidence that Al3+ ions will directly substitute for Fe3+ ions in the structure. In the
typical soil pH range (3 to 8), hydrolyzed Al species will dominate in solution (for example
Al(OH)2+ at pH 5) (Stumm and Morgan, 1981). In this case, interstitial solid solutions with
excess of non-stoichiometric OH may form within the goethite structure (Wolska, 1997; Ruan et
al., 2002). Nonetheless, since Al hydroxides were not detected in nano-particle suspensions with
low concentrations of Al (neither gibbsite nor diaspore) and gibbsite was only detected when Al
concentrations were 25% or higher, it is still possible that diaspore-like structures, which are isostructural with goethite, were present in our Fe-Al mixed suspensions with low % Al. As shown
63

by our experimental results, even if diaspore clusters formed, their concentration was too low for
detection.

3.4. Conclusions
Dilute solutions Fe+3 and Al+3 slowly hydrolyzed with base to pH 5 and
containing low initial Al concentrations (2-8 mol %) led to the formation of moderately
crystalline Al-goethite nanoparticulate suspensions upon aging. Initial hydrolysis curves suggest
the formation of a single initial disordered ferrihydrite-like phase. Infrared, but not XRD,
indicates the formation of a strained goethite structure, that could possibly be a combination of
goethite nanoparticles with a coating of isostructural diaspore or Al within the goethite structures
up to a level of about 2-4 mole % substitution. The synthetically produced nano-goethites are
small enough that a monolayer or bilayer of Al hydroxide could constitute a significant mole
fraction of total Fe and Al. Higher levels of added Al disrupted formation of goethite-like phases
which resulted in the formation of amorphous Fe-l hydroxides co-precipitates. As shown by our
DFT calculations, Al in low (8%) and high (25%) substitution Al-goethites preferred a clustered
arrangement, i.e. forming diaspore-like clusters, which was more energetically favored than Al
substitution at isolated positions. Optimized unit cell parameters for low (8%) and high (25%)
substitution Al-goethite with Al in clustered positions were in good agreement with available
experimental data. Future modeling work should address clusters that are large enough to form
coatings on gibbsite nanoparticles. Nano-colloidal suspensions with intermediate Al
concentrations (12-20 mol %) remained stable for the duration of the experiment, and formation
of crystalline phases was hindered. In soil environments such colloids could be formed by
weathering of acidic parent rocks with considerable concentrations of Al and they could persist
for a long time in soil water, thus becoming important for the retention and transport of trace
64

elements. In nano-colloidal suspensions with high Al concentrations (>25 mol %), gibbsite was
formed initially, and co-existed with goethite in aged suspensions. Diaspore was not detected.
Thermodynamic calculations using literature data showed diaspore formation is less favorable
than gibbsite formation, which explains the rare abundance of diaspore in natural environments.

Acknowledgements
This research was funded by the National Science Foundation under Grant No. CHE0431328. Synchrotron-XRD measurements were performed in beamline X-16C at the National
Synchrotron Light Source (NSLS), Brookhaven National Laboratory. The NSLS is supported by
the U.S. Department of Energy (DE-AC02-98CH10886). We thank Dr. Peter Stephens for
assistance in synchrotron-based XRD data collection. James Kubicki and Masoud Aryanpour
helped with DFT calculations.

References
1. Alvarez, M., Rueda, E.H., and Sileo, E.E. (2007) Simultaneous incorporation of Mn and
Al in the goethite structure. Geoderma, 140, 8-16
2. Bakoyannakis, D. N., Deliyanni, E. A., Zouboulis, A. I., Matis, K. A., Nalbandia, L.,
Kehagias, T. (2003) Akaganeite and goethite-type nanocrystals: synthesis and
characterization. Microporous and Mesoporous Materials 59 35-42
3. Bazilevskaya, E.A., Archibald, D.D, and Martinez, C.E. (2009) Nano-goethite (FeOOH) crystallization in the presence of low aluminum concentrations under
environmentally relevant conditions. In preparation
4. Blanch, A.J, Quinton, J.S., Lenehan, C.E. , and Pring, A.(2008) The crystal chemistry of
Al-bearing goethites: an infrared spectroscopic study. Miner. Magazine, 72, 1043-1056
5. Bertsch, P.M., Miller, W.P., Anderson, M.A., and Zelazny, L.W. (1989) Co-precipitation
of iron and aluminum during titration of mixed Al3+, Fe3+, and Fe2+ solutions. Clays and
Clay Minerals, 37, 12-18.
6. Busing W.R, Levy, H.A. (1958) A single crystal neutron diffraction study of diaspore,
AlO(OH). Acta Cryst., 11, 798-803
7. Casey, W.H., Rustad, J.R. and Spiccia, L. (2009) Minerals as Molecules-Use of Aqueous
Oxide and Hydroxide Clusters to Understand Geochemical Reactions. Chem. A Eur. J.,
15, 4496-4515
8. Cervini-Silva, J., and Sposito, G. (2002) Steady-state dissolution kinetics of aluminumgoethite in the presence of desferrioxamine-B and oxalate ligands. Env. Sci Tech., 36,
337-342
65

9. Chesworth, W. (1975) System SiO2-AlOOH-Fe2O3-H2O and kaolinitic stage of goethite


facies. Clays and Clay Minerals, 23, 389-392
10. Cornell, R.M. and Schwertmann, U.(2003) The Iron oxides: Structure, Properties,
Reactions, occurrences and uses. Wiley-VCH, 2003
11. Colombo, C. and Violante, A. (1996) Effect of time and temperature on the chemical
composition and crystallization of mixed iron and aluminum species. Clays and Clay
Minerals, 44, 113-120.
12. Coston, J.A., Fuller, C.C. and Davis, J.A. (1995) Pb2+ and Zn2+ adsorption by natural
aluminum and iron-bearing surface coatings on an aquifier sand. Geochim. Cosmochim.
Acta, 59, 3535-3547.

13. Davis, C., Knocke, W.R. and Edwards, M. (2002) Implications of aqueous silica sorption
to iron hydroxide: Mobilization of iron colloids and interference with sorption of arsenate
and humic substances. Environ. Sci. Technol., 35 (15), 3158 -3162
14. Eggleton, R. A. (1987) Non-crystalline Fe-Si-Al-oxyhydroxides. Clays and Clay
Minerals, 35, 29-37.
15. Fazey, P.G., and O'Connor, B.H. (1991) X-ray powder diffraction rietveld
characterization of synthetic aluminum substituted goethite. Clays and Clay Minerals, 39,
248-253
16. Farmer V.C. (ed.) (1974) The Infrared Spectra of Minerals. Mineral.Soc., London
17. Fey, M.V., and Dixon J.B. (1981) Synthesis and properties of poorly crystalline hydrated
aluminous goethites. Clay and Clay Miner. 29, 91-100
18. Goh, T.B., Huang, P.M., Dudas, M.G. and Pawluk, S. (1987) Effect of iron on the nature
of precipitation products of aluminum. Can. J.Soil Sci. 67, 135-145.
19. Gualtieri A., and Venturelli P. (1999) The goethite-hematite phase transformation. Amer.
Mineral., 84, 895-904
20. Hill, R.J. (1979). Crystal structure refinement and electron density distribution in
diaspore Phys. Chem. Minerals, 5, 179-200
21. Hsu, P.H. (1989) Aluminum hydroxides and oxyhydroxydes. Ch. 7., 331-378. In: J.B.
Dixon and S.B. Weed (eds.), Minerals in Soil Environment, 2nd Edition.
22. Hsu, P.H. and Bates, T.F. (1964) Formation of X-ray amorphous and crystalline Alhydroxides. Mineral. Mag., 33, 749-768
23. Jentzsch, T.L. and Penn R.L. (2006) Influence of aluminum doping on ferrihydrite
nanoparticle reactivity. J.Phys.Chem.B 110, 11746-11750
24. Kasprzyk-Holdern, B. (2004) Chemistry of alumina, reactions in aqueous solutions and
its application in water treatment. Adv. Colloid Interface Sci. 10, 19-48
25. Kosmas, C.S., Franzmeier D.P. and Schulze, D.G. (1986) Relationship among derivative
spectroscopy, color, crystallite dimensions, and Al substitution of synthetic goethites and
hematites. Clay and Clay Miner. 34, 625-634
26. Kubicki, J.D., Paul, K.W., and Sparks, D.L (2008) Periodic density functional theory
calculations of bulk and the (010) surface of goethite. Geochem Trans. 9, 4-12
66

27. Lundstrom, U.S., van Breeman, N., and Bain, D. (2000) The podzolization process. A
review. Geoderma, 94, 91-107
28. Majzlan, J., Lang, B.E., Stevens, R., Navrotsky, A., Woodfield, B.F., and Boerio-Goates,
J. (2003) Thermodynamics of Fe oxides: Part I. Entropy at standard temperature and
pressure and heat capacity of goethite (alpha-FeOOH), lepidocrocite (gamma-FeOOH),
and maghemite (gamma-Fe2O3) Amer. Min., 88, 846-854
29. Majzlan, J., and Navrotsky (2003) Thermodynamics of the goethite-diaspore solid
solution. Eur. J. Min, 15, 495-501, 2003
30. Murad, E., and Bowen, L.H. (1987) Magnetic-ordering in Al-rich goethites influence of
crystallinity. Amer. Mineral., 72, 194-200
31. Navrotsky, A., Mazeina, L., and Majzlan, J. (2008) Size-driven structural and
thermodynamic complexity in iron oxides. Science, 319, 1635-1638
32. Penn, R.L., Zhu, C., and Xu, H. (2001) Iron oxide coatings on sand grains from the
Atlantic coastal plain: High-resolution transmission electron microscopy characterization
Geology, 29, 843-846
33. Peryea, F.J., and Kittrick, J.A. (1988) Relative solubility of corundum, gibbsite, boemite,
and diaspore at standard state conditions. Clays Clay minerals, 36, 39134. Piszora, P. and Wolska, E. (1998) X-ray powder diffraction study on the solubility limits
in the goethite-diaspore solid solutions. Mater. Sci. Forum, 278-2, 584-588
35. Pollard, R.J., Pankhurst, Q.A., and Zientek, P. (1991) Magnetism in aluminous goethite.
Phys. Chem. Min., 18, 259-264
36. Ruan, H.D., and Gilkes, R.J. (1995) Dehydroxylation of aluminous goethite unit-cell
dimentions, crystal size and surface area. Clays Clay mineral , 43, 196-212
37. Ruan, H.D. Frost., R.L., and Kloprogge, J.T. (2001) The behavior of hydroxyl units of
synthetic goethite and its dehydroxylated product hematite. Specrtochimica Acta, Part A
Molecular and Biomolecular Spectroscopy, 57, 2575-2586
38. Ruan, H.D. Frost., R.L., Kloprogge, J.T., and Duong, L. (2002) Infrared spectroscopy of
goethite dehydroxylation. II. Effect of aluminum substitution on the behavior of hydroxyl
units. Spectrochim. Acta, 58 479491
39. Saalfeld, H. and Wedde, M. (1974). Z. Kristallogr., 139, 129-135
40. Scheinost, A.C., Stanjek, H., Schulze, D.G. (2001) Structural environment and oxidation
state of Mn in goethite-groutite solid-solutions. Amer. Mineral., 86, 136-146
41. Schulze, D.G. (1984) The influence of aluminum on iron-oxides. 8. Unit cell dimensions
of Al-substituted goethites and estimation of Al from them.
42. Schulze, D.G., and Schwertmann, U. (1984) The influence of aluminium on iron oxides:
X. Properties of A1-substituted goethites. Clay Miner. 19, 521 539.
43. Schwertmann, U. and Taylor, R.M. (1989) Iron oxides. In:Dixon, J.B.Weed, S.R.,
editors. Minerals in Soils Environments. 2nd ed. Madison, Wisconsin:Soil Sci Soc Am,
379-439.
67

44. Schwertmann, U., and R.M. Cornell. 1991. Iron oxides in the laboratory. VCH Publ.,
Weinheim, Germany
45. Sigh, S.S., and Kodama,H. (1994) Effect on the presence of aluminum ions in iron
solutions on the formation of iron oxy)hydroxides (FeOOH) at room temperature under
acidic environment. Clays and Clay Miner. 42, 606-613
46. Stanjek, H. & Schwertmann, U. (1992) Influence of aluminum on iron oxides. Part XVI:
Hydroxyl and aluminum substitution in synthetic hematites. Clays & Clay Minerals, 40,
347-354. Stumm and Morgan, 1981
47. Stumm, W., and Morgan, J.J. Aquatic Chemistry: An Introduction Emphasizing Chemical
Equilibria in Natural Waters, 2nd ed. New York: John Wiley & Sons, 1981
48. Szytula., A., Burewicz A., and Dmitrijevic, Z. (1968) Neutron Diffraction studies of FeOOH. Physica Status Solidi, 26, 429-434
49. Tardy, Y. and and Nahon, D. (1985) Geochemistry of laterites, stability of Al-goethite,
Al-hematite, and Fe2+ -kaolinie in bauxites and ferricretes: an approach to the mechanism
of concretion formation. Amer. J. Sci. 285, 865-903
50. Violante, A., Colombo, C., Cinquegrani, G., Adamo, P. and Violante, P. (1998) Nature of
mixed iron and aluminum gels as affected by Fe/Al molar ratio, pH and citrate. Clay
Minerals, 33, 511-519.
51. Violante, A., Ricciardella M. and Pigma, M. (2003) Adsorption of heavy metals on
mixed Fe-Al oxides in the absence or presence of organic ligands. Water, Air, and Soil
Pollution, 145, 289-306.
52. Wells, M.A., Fizpatrick, R.W., and Gilkes, R.J. (2006) Thermal and mineral properties of
Al-, Cr-, Mn-, Ni-, and Ti-substituted goethite. Clays and Clay Miner. 54, 176-194
53. Perkins, D., Essene, E.J., Westrum, F.F., and Wall, V.G. (1979) New thermodynamic
data for diaspore and their application to the system AL2O3-SIO2-H2O. Amer. Miner.,
64, 1080-1090
54. Wolska, E. (1997) The substitution of Fe, Al, and OH in synthetic and natural iron oxides
and hydroxides. Advances of Soil Ecology, 30, 271-282

68

Tables
Table 3-1. ATR-FTIR- and XRD-determined mineral phases in mixed Fe and Al hydroxide nano-particle suspensions as a function of
aging time and initial Al moles as a fraction of total Al and Fe. Abbreviations: Fh, ferrihydrite, Go, goethite, Hm, hematite, Gib,
gibbsite. Numbers represent estimates of relative peak intensity (higher numbers for higher intensities).

Time, days

mol% Al in
initial suspension

23

54

0
2
4
8
12
16
20
25
30
50
75
100

Fh
Fh
Fh
Fh
Fh
Fh
Fh
Gib1-Fh
Gib-Fh
Gib-Fh
Gib
Gib2

Fh
Fh
Fh
Fh
Fh
Fh
Fh
Gib2 -Fh
Gib-Fh
Gib-Fh
Gib
Gib2

Go2
Go1
Go1
Fh
Fh
Fh
Fh
Gib3-Go1
Gib-Fh
Gib-Fh
Gib
Gib3

Go3
Go2
Go2
Go1-Hm1
Fh-Go1
Fh
Fh
Gib2-Go1
Gib-Go
Gib-Go
Gib
Gib4

Go4
Go3
Go3-Hm1
Go2-Hm2
Go2
Fh-Go1
Fh-Go
Gib2-Go1
Gib-Go-Hm
Gib-Go-Hm
Gib-Go-Hm
Gib4

69

Table 3-2. Calculated and experimental crystallographic parameters and calculated free energy for goethite, Al-goethites, diaspore and
gibbsite.

, degree , degree

Unit cell
volume,
3

Free energy,
eV

Unit cell parameters


a,

b,

c,

, degree

Goethite experimental*
Goethite calculated
Goethite 2Al-cluster
Goethite 2Al-isolated
Goethite 6Al-cluster
Goethite 6Al-isolated

9.95
9.947
9.950
10.357
9.948
9.583

3.01
2.999
3.015
3.100
3.017
2.885

4.62
4.595
4.577
4.714
4.591
4.447

90.0000
90.0067
89.9840
89.6800
89.9940
89.6550

90.0000
90.0075
90.0190
90.7380
89.6020
89.5960

90.0000
90.0673
90.0320
90.3800
90.0690
90.1710

138.367
137.085
137.303
151.380
137.754
122.936

n/a
-601.623
-604.479
-603.567
-607.937
-602.180

Diaspore experimental*
Diaspore calculated

9.425
9.426

2.845
2.846

4.401
4.386

90.0000
90.0000

90.0000
90.0000

90.0000
90.0000

118.013
117.552

n/a
-631.402

Gibbsite experimental***
Gibbsite calculated

8.684
8.586

5.078
5.053

9.736
9.498

90.0000
90.0000

94.5400
92.8951

90.0000
90.0000

429.332
412.021

n/a
-986.953

* Szytula et al. (1968); ** Hill, R.J. (1979); ***Saalfeld and Wedde (1974)

70

Table 3-3. Relative stability of Al-substituted goethites and diaspore with respect to goethite (Hreaction)
Species
Goethite, -FeOOH
8 mol% Al-goethite

Standard
enthalpy of
formation
Hf0 (kJ/mol)*
-559.3

Reaction

Enthalpy of reaction
(kJ/mol)

n/a

5.81**

-FeOOH + 0.08Al+3 = -(Fe0.92 Al0.08)OOH +0.08Fe3+

603.7144

14.81**

-FeOOH + 0.25Al+3 = -(Fe0.75 Al0.25)OOH +0.25Fe3+

694.7375

Diaspore, -AlOOH

-1002.7

-FeOOH + Al+3 = -AlOOH +Fe3+

Al3+ (aq)

-531.0

n/a

n/a

Fe3+ (aq)

-48.5

n/a

n/a

-(Fe0.92 Al0.08)OOH
25 mol% Al-goethite
-(Fe0.75 Al0.25)OOH

41.8

* Source: Stumm and Morgan (1981) unless otherwise noted; **The values for mixing enthalpies of Al-goethites were calculated using
the equation H = W*Xgoethite*(1-Xgoethite), where W=79.14 kJ/mol and X is the mole fraction of goethite (Majzlan and Navrotsky, 2003)
Table 3-4. Relative stability of gibbsite and diaspore with respect to corundum (Hreaction)

Corrundum, Al2O3

Standard enthalpy of
formation
Hf0 (kJ/mol)*
-1675.7

Gibbsite, Al(OH)3

-1293.13

Al2O3+3H2O =2Al(OH)3

-53.16

Diaspore, AlOOH

-1002.7

Al2O3+H2O = 2AlOOH

-43.9

H2O

-285.8

n/a

Species

Reaction

Enthalpy of
reaction (kJ/mol)

n/a

n/a

Source: *Stumm and Morgan (1981)

71

Figures

5.5
5.0
4.5

pH

4.0

0-4

100

3.5

8-25
8-20

50

75

30

3.0
2.5
2.0
1.5
0

10

20

30

40

50

60

- added, mM
OH
KOH,
mM

Figure 3-1. Titration curves for Fe-Al mixed solutions adjusted to pH 5 by addition of
0.1 M KOH. Numbers on the plots show initial mol% Al of the 10 mmol [Al + Fe] in
solution. Titration curves of solutions with 30%, 50% and 75% Al have 2 inflection
points which indicate the formation of two separate phases (Fe-hydroxides and Alhydroxides).

72

OH added, mM

70
60
50
40
30
20
10
0
0

10

20

30

40

50

60

70

80

90 100

% Al in initial solutions

% Al in nanoparticles after
dialysis

Figure. 3-2. Millimoles added hydroxide at half-conversion pH points (see Fig. 3-1)
versus mol% Al in initial solutions.
100
90
80
70
60
50
40
30
20
10
0

:1

10

20

30

40

50

60

70

80

90

100

% Al in initial solutions

Figure 3-3. Aluminum in initial preparation solutions versus that observed in dialyzed
nano-particle suspensions. The solid line indicates a 1:1 ratio.

73

Figure 3-4. ATR-FTIR spectra of particles from Fe-Al-hydroxide suspensions (A) at


initial conditions (0 days) and (B) after aging at 50 C for 54 days. Numbers in the center
of each spectrum indicate the initial mol% Al as a fraction of [Fe + Al] in suspension.
Abbreviations: Go, goethite; Gib, gibbsite .

74

Figure 3-5. XRD patterns for particles from Fe-Al-hydroxide suspensions: (A) at initial
conditions (0 days), and (B) after aging at 50 C for 54 days. Numbers between panels
indicate mol% Al in suspensions as a fraction of [Fe + Al]. Abbreviations: Fh,
ferrihydrite; Go, goethite; Hm, hematite; Gib, gibbsite.

75

0
-1
-2

range of experimental [Fe]

log [Fe] total

-3
-4
-5
-6
-7

Fe(OH)3(am)

-8
-9

Goethite

-10
-11
1

A.

3
pH

0
-1
-2

25-100
12-20
0-8

log [Al] total

-3
-4

Al(OH)3(am)

-5
-6

gibbsite

-7

diaspore

-8
-9
-10
-11
1

B.

3
pH

Figure 3-6. Theoretical solubility curves for (A) iron hydroxides and (B) aluminum
hydroxides in equilibrium with total aqueous Fe(III) and Al(III) species, respectively.
Solubility curves were constructed using Ksp values for Fe-hydroxides from Stumm and
Morgan (1981), and for gibbsite and diaspore from Peryea and Kittrick (1988). Shaded
areas in panel B show Al concentrations in our synthetic mixed oxide nanoparticles, with
numerical ranges indicating the corresponding mol%Al as a fraction of [Fe + Al].

76

Figure 3-7. Optimized structural models for goethite with (A) 2 clustered Al atoms, and
(B) 2 isolated Al atoms. The distance between two Al atoms is 3.15 and 6.73 in
clustered and isolated structures, respectively.

77

Figure 3-8. Optimized structural models for goethite with (A) 6 clustered Al
atoms, and (B) 6 isolated Al atoms. The distances among Al atoms are 3-3.5 and
4.5-5.5 in clustered and isolated structures, respectively.

78

Fe(II)
Fe(III)
Go-exp
Go-calc
2Al-cl
2Al-iso
6Al-cl
6Al-iso

150

10.4
10.3
10.2
10.1

140

10.0

a, A

Unit cell volume,

145

135

9.9
9.8

130

9.7
9.6

125

9.5
-5

120
-5

10

15

20

25

30

35

3.10

15

20

25

30

35

4.70
4.65

3.05

4.60

3.00

c, A

b, A

10

% Al substitution

% Al substitution

4.55

2.95
4.50
2.90
4.45
-5

10

15

20

25

% Al substitution

30

35

-5

10

15

20

25

30

% Al substitution

Figure 3-9. Experimental and calculated unit cell volume (A) and cell parameters a (B), b
(C) and c (D) for goethite and Al-substituted goethites as a function of mol% Al
substitution and type of substitution (clustered vs. isolated). Model calculations for
clustered substitution (upward pointing triangles) yield parameters that are closer to
measured values. Experimental unit cell parameters were obtained from Szytula et al.
(1968); Schulze (1984); Kosmas et al. (1986); Fazey and O'Connor (1991); Piszora and
Wolska (1998); Gualtieri and Venturelli (1999); Scheinost et al. (2001); Ruan et al.
(2002); Majzlan and Navrotsky (2003); Wells et al. (2006); Alvarez et al. (2007); Blanch
et al. (2008). Thin solid lines indicate theoretical dependence of goethite unit cell
parameters on mol% of Al substitution in goethite (Vegards law)

79

35

-600
Goethite

Energy TOTEN, ev

-605

2Al-cluster

-610

2Al-isolated

-615

6Al-cluster
6Al-isolated
diaspore

-620
-625
-630
-635
-10

10

20

30

40

50

60

70

80

90

100

% Al

Figure 3-10. VASP calculated free energies for iso-structural goethite and diaspore and
their Al-containing solid solutions.

80

Chapter 4
SPECTROSCOPIC AND MICROSCOPIC INVESTIGATIONS OF THE
COMPOSITION OF SOIL WATER AND ACCUMULATIONS FORMED IN
SITU IN A SPODOSOL

Abstract
Mobilization and immobilization of soil colloids play an important role in
elemental transport and distribution vertically within soil profiles. We investigated the
composition of soil waters from a well-drained Spodosol soil. The Spodosol had five
identifiable horizons: A (partially decomposed organic matter, OM); E (strongly leached,
light colored); Bh (enriched in Fe, Al and OM); Bhs (also enriched in Fe, Al and OM);
and C (slightly altered, similar to parent rock). Soil waters, extracted from each horizon
by vacuum filtration, were separated into two operationally defined fractions: (1) a
dissolved (< 0.45 m or soil solution) and (2) a colloidal (0.45 m to 1.2 m) size
fractions. In addition, we placed quartz wafers which represent a model quartz surface in
soils, at the bottom of each Spodosol horizon to observe formation of the coatings on soil
minerals in situ. It was found that more than 95% of the dissolved Fe and Al were
complexed with dissolved organic matter in all horizons, as suggested by thermodynamic
calculations. Organic and inorganic phases were detected in the colloidal fraction of soil
water; however, inorganic colloids (i.e., amorphous colloidal silica, kaolinite, and poorlycrystalline aluminosilicates) were dominant in the Bh horizon. Crystalline Fe oxides were
not detected in the colloidal fraction of soil waters. Fe, Al, and Si have the highest
mobility in organic-rich A and Bh horizons which suggests that the presence of organic
matter facilitates the transport of these elements by stabilizing inorganic colloids. It was
81

found that Fe and Si were mostly accumulated as inorganic colloids, while Al showed
closer association with organic matter. The two major mechanisms of immobilization of
Fe, Al, Si and OM are (1) polymerization of metal-OM complexes and (2) charge
neutralization of OM-inorganic colloid aggregates. Both of these processes seem to occur
when (Fe+Al) to C ratios increase in the colloidal fraction of the Bh horizon.

4.1. Introduction
The mobilization and immobilization of soil colloids play an important role in the
development of soil profiles (Riise et al., 2000; Pokrovsky et al., 2005; Tang et al., 2009;
Topeshta and Sokolova, 2009). Chemically, and as defined by Sposito (1989), the
characteristic property of colloids is that they do not dissolve in water to form solutions,
but instead remain as an identifiable solid phase in suspension. Typically, colloids are
defined as inorganic or organic particles with sizes ranging from 1 nm to 1 m (Buffle et
al., 1998). Thus, colloids are considered small enough and well-dispersed in soil waters
not to sediment within substantial periods of time (from hours to days). The stability of
soil colloids is determined by factors controlling or preventing their aggregation (Buffle
et al., 1998). The colloidal fraction is the most mobile in soil waters, while macroparticles (or large aggregates) are rapidly removed from soil waters by gravity (Wen et
al., 2008). Fe, Al and Si containing solids can undergo dissolution, migration, and
accumulation in soil profiles. These elements can migrate through soil as dissolved
species (Al(III), Fe(II), Fe(III), monomeric silica; including their complexes with
dissolved organic matter), and/or colloidal inorganic (e.g. Fe or Al hydroxide) and
colloidal organic (metal-organic complexes) phases. The dominance of each of these

82

transport forms depends on the type of soil and soil properties such as pH, organic matter
(OM) content, and drainage. In Spodosol soils, where soil water contains high amounts of
dissolved and colloidal organic substances, OM is expected to play a critical role in the
dissolution, transport and stabilization of inorganic colloids.
Spodosol soils (Soil Survey Staff, 1999) are characterized by intensive transport
of organic matter, Fe and Al hydroxides and clay minerals through the soil profile,
accumulation in subsurface horizons and formation of coatings on sand grains (Wilding
et al., 1983). However, the specific forms in which Fe, Al, and Si migrate to subsurface
horizons and their immobilization mechanisms are still uncertain. Two major competitive
theories of Spodosol formation have been proposed: (1) formation and downward
transport of complexes of organic acids with Al and Fe (McKeague et al., 1978; van Hees
and Lunstrom, 2000), and (2) silicate weathering followed by downward transport of Fe,
Al and Si in the form of polymeric Fe-Al-Si inorganic colloids (Farmer and Frazer, 1982;
Kartlun et al., 2000). The mobilization and translocation mechanisms presumed to be
involved in Spodosol formation implicate the following chemical processes: (1)
formation of stable water-soluble complexes of organic acids with Fe, Al, and Si (e.g., De
Coninck, 1980; Duchaufour, 1982; Buurman and van Reeuwijk, 1984); (2) reduction of
Fe in aluminosilicate clays by organic acids and migration of stable Fe(II)-organic
complexes (Skjemstad et al., 1992; Bloomfield, 1953); (3) hydrolysis of Al, Si, and Fe to
form stable inorganic colloidal sols (e.g., Farmer, 1981, 1982). The accumulation
(immobilization) of Al, Si and Fe in lower illuvial horizons has been explained by: (1)
precipitation of Fe and Al with silica as allophanic material; (2) precipitation of metalorganic complexes due to increase in (Fe +Al) to C ratios in organic complexes; and (3)
microbial decomposition of metal-organic complexes with further precipitation of Fe83

and Al-hydroxides (Schnitzer, 1969; Petersen, 1976, De Coninck, 1980). The dominant
mobilization, translocation, and immobilization mechanism will depend on the dynamics
of organic matter within the Spodosol profile (Buurman and Jongmans, 2004).
We studied the composition of the dissolved (< 0.45 m) and colloidal (0.45 m
to 1.2 m) size fractions of natural soil waters to understand the processes involved in the
translocation of Al, Fe, and Si in Spodosols. The goals of this study were: (1) to
determine the total concentration and composition of Fe, Al, and Si compounds in these
two soil water fractions; and (2) to investigate the role of organic C in the transport and
immobilization of these elements. In addition, field experiments were performed to study
the composition of soil materials accumulated in situ. These field investigations, which
involved placing quartz wafers within Spodosol horizons, were used for the chemical and
morphological evaluation of these accumulations.
Furthermore, we attempt to reconstruct the mechanisms of Fe, Al, and Si
migration and immobilization in Spodosols based upon the information obtained from
complex studies of soil water, and analyses of in situ accumulations and soil solids. We
propose that in Spodosols Fe, Al, and Si migrate from surface horizons mostly in the
form of inorganic colloids (Fe and Al hydroxides, short-range alumosilicates) stabilized
in solution by OM and accumulate in lower horizons due to increase in (Fe+Al) to C
ratios.

4.2. Methods
4.2.1. Field site description and samples
The Spodosol site chosen for this study is located in Black Moshannon State Forest
(Central Pennsylvania) on the level upland position (0-3% slopes) of the Allegheny
84

Plauteu. It is covered with coniferous vegetation (hemlock, Tsuga Canadensis) and


underlined by sandstones of Mississippian Ponoco formation. At this location, the mean
annual air temperature (MAAT) is approximately 10 C, and the mean annual
precipitation (MAP) is about 1145 mm/year (Ciolkosz and Thurman, 1992). Five soil
horizons (A, E, Bh, Bhs, and C), described in Table 4-1, were identified at this site. Soil
samples were collected from each horizon, transported to the laboratory, sieved through
2 mm sieve and air-dried.

4.2. 2. Analyses of soil solids


4.2.2.1. Soil pH and selective extractions
A 75 ml aliquot of purified water was added to 25 g of soil (3:1 water/soil ratio)
and equilibrated for 10 minutes. The pH of the sample was then measured using an lpha
pH 200 pH-meter by placing the electrode in the soil suspension with constant stirring.
Soil from each horizon was treated using a selective dissolution procedure to
estimate the amount of Fe, Al, and Si in various solid-phase pools or fractions. These are
operationally defined fractions and include treatment of the soil with Na-pyrophosphate
(for organically-bound Fe and Al), ammonium oxalate in the dark (for Fe, Al and Si in
poorly-crystalline or amorphous inorganic materials), and dithionate-citrate (for free,
non-silicate, crystalline Fe). For the dithionite-citrate (DC) extraction (Soil Conservation
Service, US Department of Agriculture, 1972), homogenized soil samples (< 2mm) were
ground to pass a 35-mesh sieve, weighed (0.5 g) into 50-ml centrifuge tubes, and 25 ml
of 0.68 M sodium citrate solution and 0.4 g of dithionite powder were added to the tubes.
The tubes were shaken overnight, and then centrifuged for 20 min at 10,000 RPM. The
filtered (0.2 m) supernatants were analyzed for Fe by Atomic Absorption Spectroscopy
(AAS). For the acid ammonium oxalate (AO) extraction (McKeague & Day, 1966), soil
85

samples (0.250 g) were weighed into foil-covered centrifuge tubes, 10 ml of the oxalate
solution (700 ml of 0.2 M oxalate solution and 535 ml of 0.2 M oxalic acid solution,
adjusted to pH 3) were added to the tubes, and the suspensions were shaken for 4 hours.
The suspensions were then centrifuged at 10,000 RMP for 20 min, and the filtered (0.2
m) supernatants were analyzed for Fe, Al, and Si by AAS. For the sodium
pyrophosphate (SP) extraction (Mckeague, 1967), soil samples were ground to pass a 100
mesh sieve, weighed (0.3 g) into 50 ml centrifuge tubes, and 30 ml of 0.1 M Napyrophosphate solution added. The tubes were shaken overnight, centrifuged at 20,000
RPM for 10 min, and the filtered (0.2 m) supernatants analyzed for Fe and Al by AAS.
It should be noted that pyrophosphate extracts organically-bound Fe and Al by
peptization (Parfitt and Childs, 1988), however, it may also extract some inorganic
amorphous Fe and Al (Bertsch and Bloom, 1996). All extraction procedures were done in
triplicate. The extracts were analyzed for Fe, Al, and Si by Inductively Coupled Plasma
Atomic Emission Spectroscopy (ICP-AES). The content of crystalline Fe was estimated
as the difference between CD-extractable Fe and AO-extractable Fe (Fed - Feo). The
difference between AO-extractable Fe (Feo) and Al (Alo) and SP-extractable Fe (Fep) and
Al (Alp) was used to estimate the contents of poorly-crystalline (amorphous) inorganic Fe
and Al, respectively. The weight percentage of amorphous materials in soils was
calculated as the sum of the amorphous fractions of Fe, Al and Si expressed in oxide
form (Fe2O3+ Al2O3 + SiO2) (Ciolkosz et al., 1989).
4.2.2.2. Total Organic Carbon (TOC) and Total Nitrogen
Soil samples (< 2 mm) were finely ground (80 mesh sieve) to ensure sample
homogeneity and oven-dried overnight at 50 C. An appropriate amount of sample (7-10
mg) was introduced in a soil solid module (SSM-5000) coupled to a Shimadzu TOC86

5000A Total Organic Carbon Analyzer. It was assumed that all carbon in the soils was
organic carbon, and that no mineral carbon (calcite, dolomite) was present. This
assumption is adequate due to the low pH of the soils (Doner and Lynn, 1989).
4.2.2.3. X-ray diffraction (XRD) of clay size fraction (< 2 m) from Spodosol horizons
X-ray diffraction (XRD) patterns were collected using a SCINTAG X2 theta2theta powder diffractometer with a liquid nitrogen cooled germanium solid state
detector and CuK radiation ( = 1.54059 ). Scans were collected from 2 to 30 o2 with
2/min step size. Details of sample preparation for X-ray diffraction analyses and
interpretation of the diffractograms can be found in Appendix C.

4.2.3. Analyses of soil waters


4.2.3.1. Soil water extraction
Water extraction experiments were designed for sandy soils which do not
normally hold large amounts of water when field-moist. To sample soil horizons with
minimum disturbance of soil structure, we used in-house-built round sample holders
(height, 44 mm; diameter, 64 mm) cut from PVC tubes. The sample holders were pushed
into each horizon to obtain an undisturbed soil sample. The soil-containing rings were
sealed with parafilm to prevent drying and stored at 4C. Water extractions were carried
out using a filtration apparatuses with a 0.45 m cellulose acetate pore size filters and
with 1.2 m glass fiber (grade 696) pore size filters. A dissolved fraction (i.e., soil
solution, < 0.45 m) and a colloidal fraction (< 1.2 m, but also includes the dissolved
fraction) was therefore obtained. All filters were prewashed 3 times with de-ionized
purified water. The soils field capacity, determined in preliminary experiments, was 50,
20, 25, 25 and 20 % water for the A, E, Bh, Bhs and C horizons, respectively. The soils
(in PVC sample holders) were rewetted to field capacity, and allowed to equilibrate for
87

24 hours at 4 C. The PVC sample holders with field-collected soils were then placed
inside the filtration apparatuses and connected to a vacuum line with 5-10 psig of vacuum
pressure for 30 minutes. The extracted soil waters were stored at 4 C or freeze-dried (for
infrared analyses). A portion of the soil waters (< 0.45 m and < 1.2 m filtrates) was
acidified by addition of HCl and total Fe, Al and Si determined by inductively coupled
plasma atomic emission spectroscopy (ICP-AES).
4.2.3.2. Dissolved organic carbon (DOC) and Colloidal organic carbon (COC)
Dissolved organic carbon (DOC) is operationally defined as the organic carbon in
soil water that passes through a 0.45 m pore size filter (Thurman, 1985). DOC was
determined in the 0.45 m filtrates (section 2.3.1) using a SHIMADZU TOC-5000A
Total organic carbon analyzer. Visual Minteq software (Gustafsson, 2009) was used to
calculate the speciation of Fe and Al in the presence of DOC (0.45 m filtrates) based on
Fe, Al, and DOC concentrations, pH, and thermodynamic stability constants.
A measure of colloidal organic carbon (COC) was obtained by filtration of the <
1.2 m soil water fraction through a 0.45 m pore size Millipore silver filter. The carbon
retained on (collected onto) the 0.45 m silver filter is defined as colloidal organic carbon
(COC). Silver filters were used because of their low carbon concentration and availability
of pore size diameter (0.45 m). The silver filters were pre-treated by: (1) heating in a
furnace at 300 C for 30 min to remove any trace amounts of carbon, and (2) oven drying
at 60 C for 1 hour to remove moisture. The weights of the pre-treated filters were
recorded. Before filtration of soil waters, each filter was washed 3 times with 5 ml of
purified water. To collect the colloidal organic carbon fraction, 1 ml of the < 1.2 m soil
water fraction was passed though a 0.45 m pore size silver filter, which was then dried
at 60 C for 1 hour. The weight of the filter was recorded again. The amount of material
88

retained on the filter was estimated as the difference between the weights of the dried
filter after and before filtration. The carbon content of the filters (colloidal organic
carbon, COC) was determined using an EA 1110 CHNS-O analyzer (CE Instruments).
4.2.3.3. Attenuated Total Reflectance Fourier Transform Infrared ATR-FTIR
Infrared spectroscopy was used to detect the presence of organic and inorganic
compounds in soil water extracts (both <0.45 m and <1.2 m size fractions). To
increase the intensity of the infrared spectra, 1 ml of soil water was placed in 1.6 ml glass
vials, freeze-dried, and reconstituted by adding 0.05 ml of Milli-Q water to the dry
residue. Three laboratory replicates for each sample set and blanks were prepared. ATRFTIR measurements were performed using a Bruke Tenon 27 FTIR instrument with a
diamond-Zn-Se ATR sensor. A 0.5 L aliquot of the reconstituted soil water extract was
placed on the sensor and dried to form a thin coating before data collection. Each sample
was measured in duplicate. Drying and data collection were done under an argon gas
atmosphere. The spectra were collected in the 4500-600 cm-1 range with a 8 cm-1
resolution and a sample scan time equal to 128 scans. The ATR-FTIR spectra of mineral
and organic standards were also collected and used for the identification of components
in the spectra of soil waters.
4.2.3.4. Scanning electron microscopy - Energy dispersive spectroscopy (SEM-EDS)
The material collected on the 0.45 m silver filters (section 2.3.2), that is, colloids
ranging in size between 0.45 m and to 1.2 m, were observed using a Hitachi SEM,
model S-3500N. Measurements were conducted in low vacuum (20 kPa) to avoid
electron charging of non-coated samples. We chose not to coat the silver filters to
preserve them for COC analyses (as described in section 2.3.2). The elemental
composition of the colloids present on the filter was detected from secondary electrons
89

using a Princeton Garmon Tech EDS unit. For each filter, 5 to 8 areas (~100-m size)
with 2-4 spot measurements in each area were analyzed by EDS (Appendix A).
4.2.3.5. Transmission electron microscopy (TEM)
Samples for TEM were prepared by placing a single drop of soil water (<1.2 m
size fraction) onto a 200 mesh (~74 m pore size) holey carbon-coated copper grid and
allowed to air-dry. No sample dilution was performed. TEM images and selected area
electron diffraction (SAED) patterns were collected using a JEOL EM-2010 with a LaB6
filament operated at 200 kV.

4.2.4. In situ coating formation within the Spodosol profile


4.2.4.1. Experimental setup
We excavated a 10 m (length) x 1 m (width) x 0.5 m (depth) trench and placed
cleaned quartz wafers (single crystal chips, 20x20x1 mm in size) at the bottom of the A,
Bh, and Bhs horizons of our Spodosol site. These quartz wafers, which represent a model
quartz surface in soils, were placed in specially designed sample holders (Fig. 4-1). The
sample holder containing the quartz wafer was positioned in direct contact with the soil
above. The trench was carefully re-filled with the soil, and the quartz wafers were
recovered after 1 and 3 years. The recovered quartz wafers were rinsed several times with
purified water to remove loosely attached soil particles from the surface and stored in
parafilm-sealed wafer holders. The material that accumulated on these wafers represents
the material that migrated through the profile(s) above and was immobilized on the
quartz surface (model the formation of coatings).
4.2.4.2. Scanning Electron Microscopy - Energy Dispersive Spectroscopy (SEM-EDS)
The quartz wafers were observed under a Hitachi SEM, model S-3500N.
Measurements were conducted in low vacuum (20 kPa) to avoid electron charging of
90

non-coated samples. Elemental composition was detected from secondary electrons using
a Princeton Garmon Tech EDS unit. The detection limit for EDS is about 0.1 wt%. For
each wafer (A, Bh, and Bhs), 5-10 areas (100 m size) were investigated with 1-7 points
(spots) measured in each area (Appendix B). Since the wafers consisted of quartz (SiO2),
it was not possible to determine Si concentration and percent elemental composition of
the coatings. However, the relative intensity of C, P, S, Al, Fe, and Mg, as defined from
the EDS spectra, can be used to categorize accumulations as organic-rich and/or
inorganic-rich coatings.
4.2.4.3. Fe-GIXAS (Gracing Incidence X-ray Absorption Spectroscopy)
The chemical forms of Fe in surface coatings were characterized using EXAFS
(extended X-ray absorption fine structure) spectroscopy in Grazing Incidence mode
(GIXAS). This surface sensitive technique was used to distinguish between inorganic
(Fe-O-Fe) and organic (Fe-O/N-C) bonding environments of Fe in in-situ accumulations
on quartz wafers placed at the bottom of the Bh and Bhs horizons. GIXAS and EXAFS
experiments were conducted at beamline 11-2 of the Stanford Synchrotron Radiation
Facility (SSRL). The energy was calibrated to the Fe K-edge (7112 eV) using a double
crystal Si(111) monochromator. The quartz wafers were placed on an adjustable table and
the angle of the incident beam adjusted to 5 degrees. The Fe-GIXAS spectra were
collected in fluorescence mode using a Ge detector in the energy range of 200 eV to
+800 eV relative to the Fe K-edge energy. The Fe-EXAFS spectra of Fe-containing
organic and inorganic/mineral standards were collected and used for comparison with the
spectra of field samples. About 10 to 20 spectra depending on scan quality were taken for
each sample. The averaged spectra were analyzed using the Athena interface to the
IFEFFIT program package (Ravel and Newville, 2005). Normalization of the spectra
91

included: (1) subtraction of a polynomial pre-edge function and (2) background removal
(above the absorption edge) using a cubic polynomial spline function. The data were k2weighted to enhance higher k values. The k2-weighted curves were Fourier-transformed
over the 2.3 to 10 -1 range to generate radial structure functions (RSF).

4.3. Results
4.3.1. Characteristics of Soil Solids
Soils from the Black Moshannon Spodosol site are highly acidic (pH range from
3.4 to 4.3), and the soil pH increases with depth (Table 4-1). As expected, the carbon and
nitrogen contents are highest in the A horizon and lowest in the E horizon (Table 4-1). In
subsurface horizons, the maximum accumulation of organic matter occurs in the Bh
horizon as indicated by its C and N concentrations and characteristic dark brown color.
Net immobilization of N is expected to occur at C/N ratios above 30, while net
mineralization is expected at C/N ratios below 20 (Stevenson and Cole, 1999). The C/N
ratios in Spodosol horizons were 28, 22, and 9 in the A, Bh, and Bhs horizons,
respectively. Thus, the C/N ratios suggest that organic matter in the A horizon is N
depleted when compared to organic matter in the Bh and Bhs horizons, and intensive
decomposition of OM with N mineralization and leaching could take place in the Bhs
horizon. XRD analyses of the clay fraction (< 2mm) indicate quartz is the dominant
mineral phase in all soil horizons, followed by kaolinite and illite. In addition, the illuvial
Bh and Bhs horizons, and the C horizon show broad XRD peaks between 10 and 14 for
Mg-saturated clays which indicate the presence of inter-stratified vermiculate-illite
minerals (Table 1, Appendix C). The presence of inter-stratified vermiculate-illite
minerals suggests a higher degree of weathering in the Bh, Bhs and C horizons.
92

The Fe, Al and Si obtained from selective extractions are shown in Fig. 4-2. In
general, the greatest accumulations of crystalline, amorphous and organic forms of Fe, Al
and Si occurred in the Bh and Bhs horizons. Crystalline forms of Fe were 2-4 times
higher than poorly-crystalline Fe in all horizons, except Bh-horizon, where
concentrations of these two forms were similar (Fig. 4-2). Inorganic forms of Fe (both
crystalline and poorly crystalline) were clearly dominant over organically-bound Fe. In
contrast, the organically-bound Al was higher then poorly-crystalline Al in A, Bh and
Bhs,C horizons. In Bh horizon, the amount of poorly-crystalline Al was ~ 4.5 times
higher than that of poorly-crystalline Si, but in Bhs horizon there was twice more poorlycrystalline Si than Al. The highest content of amorphous material (up to 2.7%) was in Bh
horizon (Fig. 4-2). Such high content of amorphous material can be explained by the
presence of large accumulations of organic matter in Bh horizon (C and N, Table 4-1),
which could retard crystallization of Fe and Al hydroxides (Martinez and McBride, 1999;
Martinez and Martinez-Villegas, 2008).

4.3.2. Soil water: dissolved and colloidal fraction


Soil water was separated by filtration into two fractions based on particle size: (1)
<0.45 m and (2) <1.2 m particle size. These fractions were operationally defined as
dissolved and colloidal, respectively. It is important to note that the passage of
colloidal material less than 0.45 microns in size may result in the overestimation of
dissolved concentrations. However, our preliminary studies using 0.2 m pore size filters
showed similar Fe, Al and Si concentrations in both 0.2 m and 0.45 m filtrates. This
suggests that the amount of colloids with particle sizes from 0.2 to 0.45 m was not
significant in <0.45 m size fraction.

93

4.3.2.1. Dissolved concentration of Fe, Al, Si and C


For < 0.45 m fraction (Fig. 4-3A) Si concentrations were the highest following
Al and then Fe. All elements reached their maximum contents in Bh horizon. The
concentration of Si decreased in the Bhs compared to upper Bh horizon. Fe distribution in
the profile followed the same trend as Si, but was near the detection limit in C horizon. In
contrast to Si and Fe, Al concentrations were the same in Bh and Bhs horizon. The DOC
concentration was the highest in A horizon (740 mg/L), depleted in E horizon, and
accumulated in Bh horizon (380 mg/L) with further decrease down the profile (Fig. 43B). Note, that DOC concentrations were much higher that that of Al, Fe, and Si in soil
solution (0.45 m filtrates). Visual Minteq calculations predict that 95% of the Fe and Al
exist as metal-organic complexes (Fig. 4-3C), i.e. only very small amounts of free Fe(III)
or Al cations are expected to be in solution.
4.3.2.2. Concentrations of Fe, Al, Si and C in colloidal fraction of soil waters
For the 0.45 to 1.2 m fraction (Fig. 4-4A,B), the highest Fe, Al, and Si
concentrations were found in the Bh horizon. The concentration of Fe was generally
lower than that of Al. The Si concentrations were 100-400 higher than of Al
concentrations. If significant amounts of imogolite or allophone were present, the Si:Al
ratio would be close to 2:1 (Paterson et al. 1991), however such excess of Si over Al was
probably due to the presence of amorphous silica (SiO2) or small quartz particles.
In the colloidal fraction (0.45m to 1.2 m), colloidal organic carbon (COC)
concentrations were high in the A horizon and depleted in the E horizon, with further
release in the lower profile (Fig. 4-4C). Nitrogen concentrations were highest in the
colloids from A and E horizons (~5%) and decreased two times the Bh, Bhs, and C
horizons. The C/N ratios in the colloids were 10, 1, 16, 7, and 11 in A, E, Bh, Bhs, and C
94

horizons, respectively. In general, the C/N ratios between 10 and 17 are characteristic for
well-developed humic acids, while C/N ratios of soil fulvic acids range from 18 to 38
(Tan, 2003). Therefore, our C/N ratios suggest the dominance of humic acids in the soil
colloids.
Dissolved metal-organic complexes become colloids when (Fe+Al)/C ratios reach
a certain value. For example, metal-organic complexes start to form colloids >0.45 m at
(Fe+Al)/C ratios of ~0.03-0.04 at pH=4-4.5 (Nierop et al., 2002). In this study metal to
carbon ratios, (Fe+Al)/C, in the soil colloids reached up to 0.055 in the Bh horizon,
which is more than 10 times higher compared to other horizons (Fig. 4-4D). Such ratios
suggest immobilization of metal-organic complexes, which is in agreement with highest
accumulation of Fe, Al, and OM in Bh horizon.
4.3.2.3. ATR-FTIR analysis of soil water
The ATR-FTIR spectra of the dissolved (<0.45 m) and colloidal (<1.2 m)
fraction of soil water were obtained for all horizons to identify mineral and organic
components with depth. In general, the absorption peaks for the E, Bhs and C horizons
were about half the intensity or less compared to those for the A and Bh horizons. This
observation suggests the highest amount of colloidal material is in the A and Bh horizons
which is in the agreement with the data given in Fig. 4-4A,B.
The ATR-FTIR spectra of the soil water samples were compared to the spectra of
reference mineral and organic suspensions (Fig. 4-5) and literature data (Table 4-2). An
infrared spectrum is usually divided into two regions: group frequency region (4000-1300
cm-1), where each band reflects vibration of two or more atoms or molecules, and
fingerprint region (1300-650 cm-1) with single bond stretching and bending vibrations
(Tan, 2003). It could be seen that group frequency region for soil water (Fig. 4-6) yielded
95

broader bands than for reference suspensions due to high heterogeneity of the water
composition. No crystalline Fe minerals were identified in any soil extracts, and
ferrihydrite was not detectable by IR because it produces only broad peak in the 36003200 cm-1 region (Fig. 4-5). The features of both organic and inorganic compounds were
distinguished in the group frequency region.
For the water fraction <1.2 m, the two sharp peaks at 2920 and 2850cm-1 in the
horizons A, E and Bh, (Fig. 4-6) were assigned to the C-H stretch of aliphatic groups.
The intensity of these peaks, characteristic to humic acid (Table 4-2), decreased with
depth which suggests the decrease of relative fraction of humic acid compared to
inorganic mineral colloids. These peaks were accompanied by the doublet at about 16001400 cm-1, which was interpreted as an indication of carboxylate functional groups bound
with metal ions (Allard, 2002). An additional shoulder located at around 1700 cm-1 was
assigned to the C=O bonds of carboxylic groups (Table 4-2). Note that the relative
intensity of the 1700 cm-1 peak (COOH) decreased with depth, especially for water
fraction <0.45 m (Fig. 4-5), due to possible complexation of carboxylic groups with
metals in lower horizons.
The presence of mineral colloids in soil water was revealed by 3700 cm-1 and
3600 cm-1 peaks in E and Bh spectra (Fig. 4-6A), which were attributed to O-H stretching
in kaolinite (Table 4-2). Smectite would also have 3600 cm-1 peak, but less sharp than the
well-crystallized kaolinite (Farmer, 1974). The spectra for the A and Bh horizons
exhibited a 3200 cm-1 peak in the OH-stretching region, which was possibly due to the
presence of colloidal silica. According to the ATR-FTIR analysis of water structure in
colloidal silica (Bailey and McGuire, 2007), liquid water is composed of a mixture of
structures with different degree of hydrogen bonding. A fully tetrahedrally coordinated
96

hydrogen bonded water molecule contributes to the intensity at low frequencies (~3200),
while free, weakly hydrogen-bonded water appears at higher frequencies. Our reference
spectra of colloidal silica also exhibited two bands at 3400 and 3200 cm-1 (Fig. 4-6). The
significant amount of such tetrahedrally coordinated water molecules in soil water has
been explained by (1) the presence of strongly polarizing cations, such as Al3+, since
water molecule coordinated to them form stronger hydrogen bonds to water in outer
spheres of coordination (Farmer, 1974), (2) decrease of the amount of adsorbed water
and/or increase of the number of hydrogen bonds during the suspension drying on the
ATR detector (Burneau and Barres, 1990), or (3) low colloidal density of suspension
when silica surfaces cause an increase in the ordering of water molecules through
hydrogen bonding (Bailey and McGuire, 2007). The latter explanation was supported by
the presence of even more prominent 3200 cm-1 peak in the soil water fraction <0.45 m,
which presumably had much lower colloidal density compared to the <1.2 m water
fraction.
The presence of mineral colloids in the soil water was also confirmed by IR
vibrations in the fingerprint region in the A and Bh horizons (Fig. 4-6). The peaks at
1100-900 cm-1 were assigned to Si-O stretching vibrations of colloidal silica, kaolinite, or
allophane. For example, Bailey and McGuire (2007) noticed that the Si-O-Si stretching
band of bulk silica was prominent at 1116 cm-1 and the Si-O stretch of silanol group at
the silica surface could be seen at 980 cm-1, while Kartlun et al. (2000) assigned the
doublet peak at 994 and 947 cm-1 to Si-O-Si vibrations in imogolite. Kaolinite also
contributed to the intensity of ~1000-900 peaks in the <1.2 m water fraction, however
the height of the 3700-3600 cm-1 peaks in Bh horizon (Fig. 4-6B) was much smaller
compared to ~1000 cm-1 peak in pure kaolinite spectra (Fig. 4-5). Therefore, we can
97

conclude that 1000 cm-1 vibrations in the samples from the Bh horizon were arisen not
only due to kaolinite presence, but also due to the effect of colloidal silica and, possibly,
poorly-crystalline alumosilicates. Note that in the water fraction <0.45 m (Fig. 4-6A)
the characteristic kaolinite peaks (3700-3600 cm-1) were absent, however Si-O-Si
stretching vibrations were present in the A and Bh horizons. This observation suggests
the presence of non- or poorly-crystalline Si mineral nanoparticles passed through 0.45

m pore size filter. In the water fraction <1.2 m, relative proportion of mineral colloids
(1000 cm-1 peak) compared to organic compounds (~1630 cm-1 ) increased from the A to
Bh horizons, while for the water fraction <0.45 m, mineral-to-organic proportion
remained relatively constant (Fig. 4-6).
4.3.2.4. SEM-EDS analyses of the colloidal fraction (0.45-1.2 m particles size) of soil
waters
SEM-EDS was used to obtain images (Fig. 4-8 and Appendix A, Figs. A-1 A-3)
and chemical composition (Fig. 4-8, Appendix A, Table A-1) of the soil colloids (0.451.2 m). The colloids extracted from the A horizon exhibited amorphous mass with some
fibrous features typical of humic acids (Fig. 4-7B, Fig. A-1). In the Bh horizon a
significant part of the colloids consisted of 0.5-1.5 m round particles, containing C, Al,
Si, Fe (and Mg) embedded into darker C-rich matrix (Fig. 4-7C, Fig. A-2). Such
observation is in agreement with Ugolini and co-authors (1977), who observed 0.5 to 1.5
micron oblong, ellipsoidal or spherical organo-mineral particles in the soil solution of
Spodosol. The colloids obtained from Bhs horizon looked featureless (Fig. A-3), however
some areas contained occasional particles with elevated Fe content (~6% Fe versus 2%
Fe average) (Fig. 4-7D).
The histograms of Al, Si, C, Mg and Fe distribution in soil colloids are shown in
98

Fig. 4-8. All elements (Al, Fe, Si and Mg) showed a wide range of concentrations in
colloids extracted from the Bhs horizon. In contrast, colloidal carbon concentrations were
highest in A and Bh horizons (Table A-1). Both range and values of the (Al+Fe)/C
atomic ratios increased with depth reaching the highest values in the Bhs horizon (Fig. 48). In A horizon, only 4 out of 22 spots (18%) had detectable Fe, while in Bh horizon,
more than 65% (15 out of 23 spots) contained detectable Fe. In the Bhs horizon, 12 of 29
spots (41%) had detectable Fe. The Si:Al ratio was very high in the A horizon and ranged
from 2.4 to 324 with the average of 184, which suggests the presence of colloidal silica.
In the Bh horizon, the Si:Al ratios were less than 1.5 for all spots but one, so the presence
of 2:1 layer alumosilicates was not supported. However, kaolinite could have been
present because 10 of 23 spots had the Si:Al ration close to 1. This observation is in
agreement with ATR-FTIR data which also confirm the presence of kaolinite in the Bh
horizon. Also, low Si:Al ratios (under 0.5) with the average Si:Al=0.8 suggested the
presence of Al-hydroxides. Bhs colloids showed more heterogeneity in Si:Al ratio, which
ranged from 0.1 to 9.9 with the average of 1.8. At least 24% of spots had Si:Al ratio > 2
which suggested the presence of 2:1 clay.
4.3.2.5. TEM-EDS of the colloidal fraction of soil waters
TEM-EDS analyses revealed morphological features and chemical composition of
the colloidal fraction of soil waters from Bh and Bhs horizons (Figs. 4-9 and 4-10). It
should be noted that composition and morphology of the Bh and Bhs colloids were quite
different. In general, three types of morphological features were distinguished in the Bh
horizon: (1) small spherical amorphous SiO2 particles (~100 nm) as inferred from EDS
and electron diffraction (Fig. 4-9A); (2) round organic-rich amorphous colloids, ~2-3 m
in size with traces of Fe, Al, P and Cl (Fig. 4-9B); and (3) organic-rich areas with
99

embedded inorganic crystallites (Si>Fe>Ca>Al) (Fig. 4-9C). The presence of amorphous


silica particles in the Bh horizon is in agreement with our ATR-FTIR and chemical data.
Also, the presence of Fe, Al, and Si within an organic matrix (Fig. 4-9B and C)
apparently illustrates an active role of the organic materials in the transferring of these
elements through the soil profile. The Bh colloids exhibited intensive Si, Ca, and Fe
peaks, while the intensity of the C peak was lower than that of Si.
A typical TEM image and chemical composition for Bhs colloids showed a more
homogenious organic matrix consisting of Al, Si, Ca, and Fe, that contained discrete
crystallites (Fig. 4-10). The Bhs colloids were enriched in Al and C, and had lower Si,
Ca, and Fe concentrations. Our TEM-EDS data is in general agreement with our SEMEDS results, however a greater number of observations is needed for a more
comprehensive interpretation of the TEM-EDS results (Buffle et al., 1998).

4.3.3. Characteristics of in situ coatings formed in the Spodosol profile


The term surface coatings refer to precipitation or accumulation of a different
phase or phases on the surface of a mineral (Hodson, 2003). According to our
experimental design, the accumulations on the quartz wafer reflect chemically-bonded
material not be easily removed by water (Section 2.4.1) and defined in this work as
coatings. These coatings were presumably formed by the reaction of colloidal material
with the quartz surface.
4.3.3.1. SEM-EDS of coatings of wafers
SEM-EDS analyses of in-situ coatings formed in the A horizon consisted of
organic and inorganic phases in various proportions (Appendix B, Figs. B-1-e B-1-e).
Most coatings were monomorphic, consisting of organo-Al(-Fe) compounds (Fig. B1-b),
although Fe concentrations were close to the detection limit. Areas with elevated/high Al
100

concentration were found within various monomorphic coatings (Figs. B-1-c and .B-1-d).
Elevated Al concentrations may reflect supersaturation of organo-Al complexes with
respect to Al, resulting in the precipitation of discrete inorganic Al-hydroxide phases
within the organic matrix. Polymorphic organic matter coatings of irregular thickness,
containing both Al and Fe were also found (Fig. B-1-e). In general, the coatings obtained
from the A-horizon were mostly organic (C-rich), always contained Al, and sometimes
Ca and Fe.
Two distinct types of coatings were obtained from the Bh horizon (Appendix B):
(1) organic(C)-rich isomorphic 50-100 m dark accumulations, and (2) mineral which
consisted of agglomerates of variously-sized particles. An example of a typical organicrich coating is shown in Fig. B-2-a, which had Na, Cl and K, but no Al or Fe was
detected. Typical mineral coatings had pronounced Al, K, Fe and Ti peaks (Fig. B-2-b
Fig. B2-e), suggesting the presence of aluminosilicate clays and Fe-(Ti-) oxides. Some
coatings were presented by mixtures of organic and inorganic (no C) phases with highly
heterogeneous composition (Figs. 4-3-f and 4-3-i). In general, however, the majority of
the coatings from the Bh horizon were inorganic. It is important to note that Al was
present in both organic and inorganic accumulations, while detectable concentrations of
Fe (> 1 wt.%) were only associated with inorganic phases (Figs. B-2-f and B-2-i). This
suggests that at least some Fe accumulated from the Bh horizon in mineral form, perhaps
as Fe-hydroxides or Fe alumosilicates ( e.g., chlorite or illite). Only 7 out of the 23 areas
analyzed (30%) had C, while Al and K were present in almost all spots, and more than
50% of analyzed areas had detectable Fe concentrations (Table B-1).
Silicate clay and iron oxide rich accumulations were dominant in the Bhs horizon
(Table B-1, Figs. B-3-a B-3-c, and B-3-e). Al and Fe were present in all areas analyzed,
101

and 11 out of 23 areas contained K, which together with Al and Si suggests the presence
of illite (Seaman et al., 1997). Only one area had C-rich coating similar to ones observed
in the Bh horizon, but larger in size (~200 m) (Fig. B-3-d). This coating contained C, S,
Cl, K and Ca, but no Fe and Al were detected. Overall, 21 out of 23 spots contained the
elements characteristic of inorganic phases, which suggest in situ coatings in the Bhs
horizon were mostly inorganic.
In summary, in A horizon the morphology and composition of natural
accumulations in A horizon suggest organo-Al-(Fe)-complexes form predominantly
monomorphic coatings. Two different types of coatings were observed in the Bh horizon,
namely organic-rich with Al and inorganic phases with Al and Fe. In the Bhs horizon we
observed that Fe and Al are associated with silicate clay rich coatings.
4.3.3.2. Fe-EXAFS analysis of quartz wafers
Fe-GIXAS measurements were performed for quartz wafers retrieved from Bh
and Bhs horizons after 1 year and 3 years of exposure within the soil profile. The Fourier
transformed Fe-EXAFS spectra (radial structure function, RSF) of in-situ coated quartz
wafers are shown in Figure 4-11. The RSF show the average distance from the Fe atom to
its first and second nearest neighbors. The highest peak at about 1.5 corresponds to the
first nearest neighbors (O/N). The spectra for Fe-organic (Fe-citrate, Fe-catechol, FeEDTA) and mineral (goethite and hematite) reference compounds show distances not
corrected for phase shift to second nearest neighbors, which are ~2.3 for Fe-C and 2.5,
and 3.2 for Fe-Fe contributions to the spectra (Gaustaffson, 2007; Joris et al., 2008).
Accumulations from the Bhs horizon showed spectral features (peaks at 2.6 and 3.3 )
characteristic of mineral reference materials (goethite and hematite) after 1 and 3 years.
This result agrees with SEM-EDS data that showed association of Fe with
102

inorganic/mineral phases in the Bhs horizon. After one year accumulations in the Bh
horizon showed features of iron minerals, however after 3 years accumulations showed
features of both inorganic and organic compounds.

4.4. Discussion
4.4.1. Transport and accumulation forms of Fe, Al, and Si in Spodosols
4.4.1.1. Aluminum
Aluminum can be transported in soil profile as aqueous species, mineral colloids
(such as Al-hydroxides, proto-imogolite and allophane, and 1:1 or 2:1 alumosilicate
clays) and organo-mineral complexes. In the solution fraction (<0.45 m) water fraction,
DOC concentrations were high enough to complex up to 90% of all Al present in this
water fraction, i.e. only small amounts of ionic Al species could exist in soil water. ATRFTIR spectra and the Si:Al ratios obtained from SEM-EDS data showed the presence of
kaolinite, and, possibly, 2:1 clay minerals in the Bh horizon. The possibility for Al to
move as 2:1 alumosilicates was also confirmed by the presence of vermiculite-illite in the
soil clay fraction, as determined by XRD analysis. The possibility of Al migration as
amorphous alumosilicates (proto-imogolite) was unlikely. The pH values of soil water
suspension were lower than the pH required for proto-imogolite colloids to be stable (pH
>4.8) (Farmer and Lumsdon, 2001) although the obtained Si:Al ratios in soil colloids
were appropriate for imogolite (0.7 1.1) (Harsh et al., 2002). While a significant part of
Al could migrate as inorganic colloids, TEM-EDS data indicated the association of Al
with organic matter in the Bh and Bhs horizons. Composition of coatings formed on
quartz wafers, i.e. immobilization of the constituents of soil waters, showed clear
association of Al with OM. Al could be transported in the form of Al-organic complexes
103

that have been immobilized in the Bh horizon, with subsequent escape of Al to form Alhydroxides or alumosilicates, as evidenced by accumulation of significant concentrations
of Al in the lower Bhs horizon (Fig. 4-9).
4.4.1.2. Iron
Fe solubility in well aerated soils is largely controlled by Fe-oxyhydroxides and
organic Fe-chelates (Lindsay, 1991). Unlike Al, the solubility of inorganic Fe phases is
much lower than that of Al at pH=4-5 (Stumm and Morgan, 1981). Kinetics will also
favor formation of Fe-hydroxides. The rate constant of Fe(III) precipitation at
circumneutral pH and 25 C was estimated to be 16 1.5 106 M1 s1 (Pham et al.,
2005), while the rate constants for formation of organic Fe(III) complexes can range from
3.3 104 to 3.2 106 M1 s1 (Fujii et al., 2008). Also, Gustaffson and co-authors (2007)
found that a significant part of Fe bound to organic matter was already hydrolyzed
forming small iron hydroxide clusters. In our samples, for <1.2 m soil water fraction, Fe
concentrations were already supersaturated with respect to iron hydroxide, and at any
concentration of organic acids in soil water iron hydroxides could form. However, unlike
conventional beliefs, Fe hydroxide is not immobilized, but can persist in suspension and
migrate in soil profile due to sorption of organic substances onto Fe hydroxide
nanoparticles (Seijo et al., 2009). Organic matter may play dual role here. First, by
adsorption to the mineral surfaces, it prevents further growth of Fe-oxides or hydroxides
to macro-size particles, which would settle down from solutions due to their large size.
Second, at the pH of natural water (4 to 5), positively charged Fe-hydroxides would be
electrostatically attractive to negatively charged quartz and clay minerals, and Fe
mobility can be reduced. However, OM (ligands, macromolecules, or nanoparticles),
which also bear negative surface charge can be preferentially adsorbed to the Fe104

hydroxide particles and create negatively charged clusters, which would prevent further
aggregation via electrostatic repulsion forces. While there is no sufficient evidence of this
mechanism from our soil water data, both TEM-EDS and SEM-EDS data showed the
aggregates containing organic-rich matrix with embedded inorganic particles. According
to the SEM-EDS data for soil colloids, there were increased accumulations of Fe in lower
Bhs horizon, probably due to precipitation of Fe-hydroxides. Another supportive piece of
information is the composition of accumulations on quartz wafers, i.e material that
migrated though soil profile to be immobilized on quartz surface. Based on Fe-EXAFS
data, association of Fe with inorganic minerals was found in the Bhs horizon for both 1year in the 3-year accumulations. The SEM_EDS analysis of 3-year soil coatings showed
direct Fe association with inorganic mineral particles in both Bh and Bhs horizons. Also,
Fe-EXAFS for the Bh horizon suggested the presence of inorganic components for 1-year
accumulations and both organic and inorganic components for the 3-year coatings. Such
difference in coating composition with time could have been due to the seasonal/annual
variations in the soil water composition when migration of either organic or inorganic
compounds was dominant.
4.4.1.3. Silica
Silica had high concentrations in the A horizon and apparently migrated as
colloidal silica (SiO2) particles as evidenced by SEM-EDS, TEM-EDS and infrared data.
The colloidal silica could have been a precursor for the formation of the secondary
alumosilicates in the Bh horizon, where Al could liberate from Al-organo complexes and
coupled with silica to form alumosilicates. Silica colloids could play an important role in
adsorption of organic matter onto its surface (Jada et al., 2006). Various mechanisms
have been proposed to explain the sorption of humic substances, including electrostatic
105

and hydrophobic attractions (Juhna et al., 2003) and bridging mechanism, when
polyvalent cations (such as Fe3+ and Al3+) act as bridges between humic molecules and
silica or clay minerals (Tan, 2003, Tombacz et al., 2004).

4.4.2. Role of organic matter in podzolization process


4.4.2.1. Transport forms
In Spodosol water (in both <0.45 m <1.2 m fractions), organic compounds
were dominant over inorganic compounds, as it follows from metal/C ratios. There are
several important roles OM can play in Fe, Al and Si mobilization, transport and
immobilization in Spodosols process. First, it liberates Fe, Al, and Si from weathered
minerals (Lindsay, 1991). Afterwards, two competitive reactions may take place in soil
water: (1) OM sorption onto inorganic particles, that electrostatically and sterically
stabilize these colloids in soil water and allow them to move down soil profile (Juhna et
al., 2003; Jada et al., 2006), and (2) chelation of metal-humic acid complexes where
carboxyl groups are predominant in OM-metal complexation (Schnitzer, 1984) (Fig.413). The SEM-EDS and TEM-DS data obtained in this study support either of these
transport mechanisms, showing inorganic particles in organic-matrix in Bh horizon (Fig.
4-8) and exhibiting organic matter containing Si, Al and Fe (Figs. 4-10 and 4-11).. In
should be noted that in the Spodosol literature, the traditional term organic matter-metal
complex may include both forms discussed above (Fig. 4-13) and intergrades between
them. However, our data show that inorganic colloidal movement took place in Spodosol
profile, but not exclusively in the form of Fe-Al-Si sols as was suggested by Farmer and
Frazer (1982).

106

4.4.2.2. Immobilization
It was demonstrated (Mckeague et al., 1971) that for sedimentation to occur the
(Fe+Al)/C atomic ratio in water should be more than 0.06. The (Fe+Al)/C ratios in the A
horizon were much lower ranging from 0.001 to 0.03 (average of 0.01), which suggests
high stability of organo-minerals colloids. In contrast, in the Bhs horizon (Fe+Al)/C
ratios were high indicating that immobilization could occured. Bh colloids showed
intermediate values showing that both immobilization and stability of colloids were
possible. According to Riise et al (2000), in soil water the metal-humic acid complexes
remained soluble when metal/carbon ratios were low, but the increase of (Al+Fe)/C leads
to neutralization of negatively charged colloidal particles by polyvalent cations and,
therefore, increased polymerization, aggregation, and sedimentation of colloids.
Saturation of organic molecule with metal ions causes the increased number of metal
bridges, which leads to polymerization of such complexes in Bh horizon. When metalorganic compounds were immobolized in Bh horizon, the presence of organic ligands
inhibited formation of crystalline Fe and Al minerals, which explains significant amount
of amorphous material (up to 27%) in that part of soil profile.
In the colloidal fraction of the soil water, humic matterinorganic colloids
interactions are also very important. For example, in most natural systems, the small (few
nanometers) fulvic compounds will stabilize the inorganic colloids, while rigid polymers
(0.1-1m) will destabilize them (Buffle et al., 1998; Seijo et al., 2009). The stability of
such colloidal aggregates is defined by electrostatic interaction between mineral particles
and organic matter. In the pH range typical for Spodosol soil (3.5-4.5), humic acids are
negatively charged and will be electrostatically repelled by other negatively charged soil
minerals, such as 2:1 alumosilicates and colloidal silica. However, Fe and Al hydroxides
107

are positively charged at pH< 6, so we can expect their strong attractive electrostatic
interaction with humic acids, which may result in heterocoagulation of colloids (Seijo et
al., 2009). When such colloids are neutralized (metal/C increases), they tend to aggregate
and immobilize. The actual mechanisms involved in the flocculation of organic matter
may depend on the type of cation and its hydrolysis products.

4.4.5. Environmental significance of this study


In this study we explored methods to study colloidal movement in short-term
processes in Spodosols. Understanding short-term element migration is important for
practical applications such as land use or environmental pollution. Future work aims to
quantify relative proportion of organic and inorganic forms of Fe, Al, and Si in soil
colloids and to study seasonal variations in soil water composition. Because interactions
between inorganic and organic colloids in soil solutions involve complex mechanisms,
there is a high need to explore charging behavior, colloidal stability, and migration of
mineral nanoparticles in the presence of dissolved and particulate organic matter.

4.5. Conclusions
In both dissolved (<0.45 m) and colloidal (<1.2 m) fractions of soil water the
highest concentrations of Fe, Al, and Si were found in organic-rich A and Bh horizons,
which suggests that the presence of organic matter facilitates transport of these elements.
Thermodynamic calculations predicted that more than 95% of Fe and Al exist as metalorganic complexes in the dissolved fraction of soil water. In soil colloids both organic
and inorganic compounds were detected, however, inorganic colloids, such as colloidal
silica, kaolinite and poorly-crystalline alumosilicates, were dominant in the Bh horizon.
There is a strong evidence that Si migrates through the Spodosol profile in the form of
108

amorphous silica and, to a lesser extent, in the form of poorly crystalline or crystalline
(kaolinite) alumosilicates. Al showed strong association with organic matter. No
crystalline Fe oxides were detected in the soil water. The two major mechanisms of
immobilization of Fe, Al, Si, and OM could take place (1) polymerization of metal-OM
complexes and (2) charge neutralization of OM-inorganic colloids aggregates. Both of
these processes presumably occur when (Fe+Al) to C ratios in colloidal fraction increase
in the Bh horizon.

Acknowledgements
This material is based upon work supported by the National Science Foundation
under Grant No. CHE-0431328. TEM work was performed in electron microscopy
facility of the Materials Characterization Laboratory at Penn State University. Dr.
Ciolkosz is greatly appreciated for the discussions of filed experiment setup and for the
valuable comments to the manuscript. Authors thank Dr. Trevor Clark for the help with
TEM measurements. SEM-EDS and XRD measurements were made at Material
Research Laboratory at the Pennsylvania State University. EXAFS measurements were
performed at the beamline 11-2 at Stanford Synchrotron Radiation Laboratory (SSRL),
Stanford. Use of the SSRL was supported by the U.S. Department of Energy, Office of
Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-98CH10886.

References
1. Allard, T., Ponthieu, M., and Weber, T. (2002) Nature and properties of
suspended solids in the Amazon Basin. Bul. Soc. Geo. France, 173, 67-75
2. Bailey, J.R. and McGuire, M. M. (2007) ATR-FTIR observations of water
structure in colloidal silica: implications for the hydration force mechanism.
Langmuir, 23, 10996-10999
3. Bertsch, P.M. and Bloom, P.R. (1996) Aluminum. In: Methods of soil analysis.
Part 3. Chemical methods. Ed. Bigham SSSA
4. Bloomfield, C. (1953) A study of podzolization: Part I. The mobilization of iron
and aluminum by Scots pine needles. J. Soil Sci., 4, 5-16
5. Buffle, J., Wilkinson, K.J., Stoll, S., Filella, M., and Zhang, J. (1998) A
generalized description of aquatic colloidal interactions: the threecolloidal
interaction approach. Env. Sci.Tech., 32, 2887-2899
109

6. Buurman, P., and Jongmans, A.G. (2004) Podzolization an additional paradigm.


Edafologia, 9, 107-114
7. Buurman, P., and van Reeuwijk, L.P. (1984) Protoimogolite and the process of
Spodosol formation: a critical note. J. Soil Sci., 35, 447-452
8. Burneau, A. and Barres, O. (1990) Comparative study of the surface hydroxyl
groups of fumed and precipitated silicas. 2. Characterization by infrared
spectroscopy of the interactions with water. Langmuir 1990, 1364-1372
9. Ciolkosz, E.J., Cronce, R. C and Dobos, R.R. (1989) Amorphous material in
Pennsylvania soils. Agronomy Series Number 102, Agronomy department, The
Pennsylvania State University, University park
10. Ciolkosz, E.J., and Thurman, N.C. (1992) Geomorphology and Soils of the
Northeastern United States and Pennsylvania: A series of Reprints. Agronomy
Series Number 116, Agronomy department, The Pennsylvania State University,
University Park
11. De Coninck, F. (1980) Major mechanisms in formation of spodic horizons.
Geoderma, 24, 101-128
12. Doner, H.E. and Lynn, W.C. (1989) Carbonates, halide, sulfate, and sulfide
minerals. In: Minerals in soil environments. Eds: Dixon, J.B. and Weed, S.B.
SSSA book series, 1. Sec. Ed., pp. 279-330.
13. Duchaufour, P. 1982. Pedology: Pedogenesis and Classification. George Allen
and Unwin, London
14. Farmer, V.C. (1974). The layer silicates. In: The infrared spectra of minerals. pp.
340-349 London: Mineralogical Society.
15. Farmer, V.C. (1982) Significance of the presence of allophone and imogolite in
podzol Bs horizons for podzolization mechanisms: A review. Soil Sci. Plant
Nutr., 28, 571-578
16. Farmer, V.C. and Frazer, A.R. (1982) Chemical and colloidal stability of sols in
the Al2O3 Fe2O3-SiO2-H2O system: their role in podzolization. J.Soil Sci., 33,
737-742
17. Farmer, V.C. and Lumsdon, D.G. (2001) Interactions of fulvic acid with
aluminum and a proto-imogolite sol: the contribution of E-horizon eluates to
podzolization. Eur. J. of Soil Sci. 52, 177-188
18. Fritsch E., Allard T., and Benedetti M.F. (2009) Organic complexation and
translocation of ferric iron in podzols of the Negro River watershed. Separation of
secondary Fe species from Al species. Geochim Cosmochim Acta, 73, 18131825
19. Fujii, M., Rose, A.L., Waite, T.D., and Omura, T. (2008) Effect of divalent
cations on the kinetics of Fe(III) complexation by organic ligands in natural
waters. Geochim Cosmochim Acta, 72, 1335-1349
20. Giesler, R., Illvesniemi, H., Nyberg, I., van Hees, P., Starr, M., Bishop, K., and
Lundstrom, U.S. (2000) Mobilization of Al, fe, Si and base cations in three
110

Spodosols. Geoderma, 94, 247-261


21. Gustafsson, J.P. (2009) MINTEQ 2.53
http://www.lwr.kth.se/english/OurSoftWare/Vminteq/index.htm
22. Gustafsson, J.P., Persson, I., Kleja, D.B., and van Schaik, J.W. (2007) Binding of
Fe(III) to organic soils: EXAFS spectroscopy and chemical equilibrium modeling.
Environ. Sci. Tech. 41, 1232-1237
23. Jada, A., Akbour, R. A., Douch, J. (2006) Surface charge and adsorption from
water onto quartz sand of humic acid. Chemosphere, 64, 1287-1295
24. Jansen, B., Nierop, G.J., Verstraten, J.M. (2005) Mechanisms controlling the
mobility of dissolved organic matter, aluminum and iron in podzol horizons. Eur.
J. Soil Sci. 56, 537-550
25. Juhna, T., Klavins, M., Eglite, L. (2003) Sorption of humic substances on aquifer
materialat artificial recharge of ground water. Chemosphere, 51, 861-868
26. Kartlun, E., Bain, D., Gustafsson, J.P., Mannerkoski, H., Murad, E., Wagner, U.,
Fraser, T., McHardy, B. and Starr, M. (2000) Surface reactivity of poorly ordered
minerals in podzol B horizons. Geoderma, 94, 236-286
27. Lindsay, W.L. (1991) Iron oxide solubilization by organic matter and its effect on
iron availability. Plant and Soil, 130, 27-34
28. Loeppert and Inskeep (2001) Iron In: Methods of soil analysis. Part 3. Chemical
methods. Ed. Bigham SSSA
29. Martnez-Villegas, N. and Martnez, C.E. (2008). Solid- and solution- phase
organics dictate copper distribution and speciation in multi-component systems
containing ferrihydrite, organic matter, and montmorillonite. Environmental
Science & Technology. 42, 28332838
30. Martnez, C.E. and McBride, M.B. (1999) Dissolved and labile concentrations of
Cd, Cu, Pb, and Zn in aged ferrihydrite-organic matter systems. Environmental
Science & Technology, 33, 745750
31. McKeague, J.A., and Day, J.H. (1966) Dithionite and oxalate extractable Fe and
Al as aids in differentiating various classes of soils. Soil Sci, 46, 13-22
32. McKeague, J.A. (1967) An evaluation of 0.1 pyrophosphate and pyrophosphatedithionite in comparison with oxalate as extractants of the accumulation products
in Podzols and some other soils. Can. J. Soil Sci., 47, 95-99
33. Mckeague, J.A., Brydon, J.E., and Miles, N.M. (1971) Differentiation of forms of
extractable iron and aluminum in soils. Soil Sci. Soc. Am. Proc. 35, 33-38
34. McKeague, J.A., Ross, G.J., and Gamble, D.C. (1978) Properties, criteria of
classification and genesis of Podzolic soils in Canada. pp. 27-60. In: Quaternary
Soils, Ed.: Mahaney, W.C. Geo Abstracts, Norwich.
35. Mokma, D.L, and Buurman, P. (1982) Podzols and podzolization in temperate
regions, ISM Monograph 1. International Soil Museum, Wageningem, The
Netherlands.
111

36. Nierop, K.G.J., Jansen, B., and Verstraten, J.A. (2002) Dissolved organic matter,
aluminium and iron interactions: precipitation induced by metal/carbon ratio, pH
and competition. Sci. Total Env. 300, 201-211
37. Parfitt R. L., and C. W. Childs (1988). Estimation of forms of Fe and Al: A
review and analysis of contrasting soils by dissolution and Moessbauer methods.
Aust. J. Soil Res. 26, 121144.
38. Petersen, L. (1976) Podzols and podzolization, DSR Forlag, Copenhagen
39. Paterson, E., Goodman, B.A., and Farmer, V.C. (1991) The chemistry of
aluminum, iron, and manganese oxides in acid soils. In: Soil acidity, Eds. Uhlrich,
B. and Sumner, M.E. Springer, Berlin, pp. 97-124
40. Ravel, B. and Newville, M. (2005) ATHENA, ARTEMIS, HEPHAESTUS: data
analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Rad.
12, pp.537-541.
41. Pokrovsky, O.S., Dupre, B. and Schott, J. (2005) Fe-Al-organic colloids control
of trace elements in peat soil solutions: results of ultrafiltration and dialysis.
Aquatic Geochemistry, 11, 241-278
42. Righi, D., and De Coninck, F. (1974) Micromorphological aspects of Humods and
Haplaquods of the Landes du Medoc, France. In. G.K. Rutherford (ed.). Soil
Microscopy, pp.567-588. Limestone Press, Kingston, Ontario, 975 pp
43. Riise, G., Van Hees, P., Lundstrom, U., and Strand, L.T. (2000) Mobility of
different size fractions of organic carbon, Al, Fe, Mn and Si in podzols.
Geoderma, 94, 237-247
44. Pham, A.N., , Rose, A.L., Feitz, A.J., and Waite, T.D. (2005) Kinetics of Fe(III)
precipitation in aqueous solutions at pH 6.09.5 and 25 C. Geochim Cosmochim
Acta, 70, 640-650
45. Schnitzer,M. (1969) Reactions between fulvic acid, a soil humic compound and
inorganic soil consistuents. Soil Sci. Soc. Am. Proc., 26, 362-365
46. Schnitzer, M. (1984) Characterization of organic matter extracted from podzol B
horizon. In: Podzols. Ed. Buurman, P. pp 35-41
47. Schoenberger, P.J., Wysock, D.A., Benham, E.C., and Broderson, W.D. (Editors)
(2002). Field book for describing and sampling soils, Version 2.0. Natural
Resources Conservation Service., NE.
48. Schwartz, D. (1988) Some podzols of bateke sands and their origins. Geoderma,
43, 229-247
49. Seaman, J.C., Bertsch, P.M., Strom, R.N. (1997) Characterization of colloids
mobilized from southeastern coastal plain sediments. Env. Sci. Tech., 31, 27822790
50. Seijo, M., Ulrich, S., Filela, M., Buffle, J., and Stoll, S. (2009). Modeling the
adsorption and coagulation of fulvic acids on colloids by Brownian dynamics
simulations. Env. Sci. Tech. In press

112

51. Skjemstad, J.O, Fitzpatrick, R.W. and Zarcinas, B.A. (1992) Genesis of podzols
on coastal dunes in southern queensland. 2. Geochemistry and forms of elements
as deduced from various soil extraction procedures. Aust. J. Soil Res, 30, 615-644
52. Soil Survey Staff (1975). Soil Taxonomy. A basic system of soil classification for
making and interpreting soil surveys. U.S. Dept. Agr., Handbook. No. 436, pp.
754.
53. Soil Conservation Service, US Department of Agriculture (1972) Soil survey
laboratory methods and procedures for collecting soil samples. Soil Survey
Invetigations, Report No.1 (revised). U.S. Government Printing Office,
Washington DC
54. Stevenson, F.J., and Cole, M.A. (1999) Cycles of soil: carbon, nitrogen,
phosphourus, sulfur, micronutrients. New York, Willey, 2nd Ed
55. Stumm, W., and Morgan, J.J. Aquatic Chemistry: An Introduction Emphasizing
Chemical Equilibria in Natural Waters, 2nd ed. New York: John Wiley & Sons,
1981
56. Tan, K. H. (2003) Humic Matter in Soil and Environment. Principles and
Controversies. Marcel Dekker, New York
57. Tang, Z., Wu, L. and Luo, Y. (2009) Size fractionation and characterization of
nanocolloidal particles in soils. Environ. Geochem. Health, 3, 1-10
58. Tombacz, E., Libor, Z., Illes, E., Majzik, A. and Klumpp, E. (2004). The role of
reactive surface sites and complexation by humic acids in the interaction of clay
mineral and iron oxide particles. Org Geochem , 35, 257-267
59. Topeshta, I.I. and Sokolova, T.A. (2009) Aluminum compounds in soil solutions
and their migration in podzolic soils on two-layered deposits. Soil chemistry, 42,
24-35
60. Thurman, E.M. (1985) Organic geochemistry of natural waters. Martinus
Nijhoff/Dr. W. Junk Publishers, 497 p
61. Ugolini, F.C., Dawson, H., and Zachara, J. (1977). Direct evidence of particle
migration in the soil solution of a podzol. Science, 198, 603-605
62. Van Hees, P., and Lunstrom, U., Starr, M.and Giesler, R. (2000) Factors
influencing aluminum in soil solution of podzolic soils. Geoderma, 94, 287-308
63. Wen, L.S., Warnken, K.W., and Santschi, P.H.(2008) Comparison between the
Trinity River and the Trinity River Estuary (Galveston Bay, Texas). Marine
Chemistry 112, 2037
64. Wigginton, N.S., Haus, K.L., and Hochella, M.F. (2007) Aquatic environmental
nanoparticles. J. Env. Monitoring, 9, 1306-1316
65. Wilding, L.P., Smeck, N.E., and Hall, G.F. (1983) Spodosols. In: Pedogenesis and
soil taxonomy. II. Soil orders. Developments in Soil Science 11B Elsevier, p. 217248

113

66. Yuan, G., Soma, M., Seyama, H., Theng, B.K.G., Lavkulich, L.M., and
Takamatsu, T. (1998) Assessing the surface composition of soil particles from
some podzolic soils by X-ray photoelectron spectroscopy. Geoderma, 86, 169-181

114

Tables
Table 4-1. Description of soil horizons of Black Moshannon Spodosol site. Morphology standards used ones given in Schoenberger et
al.,2002.
Horizon

Depth,
cm

0-15

15-25

Bh

25-30

Bhs

30-50

50-84

Properties
Black (10YR 2/1), loamy sand, weak granular
structure, non-sticky, abrupt boundary, many
fine and thick roots from the trees, partially
decomposed organic residues
Light gray (10YR 6/1), loamy sand, singlegrained, non-sticky, abrupt wavy boundary
Strong brown (7.5 YR 5/6), sandy loam, weak,
fine granular structure, friable, non-sticky,
diffuse wavy boundary
Reddish-yellow (5YR 6/8), sandy loam,
medium sub-angular blocky structure, slightly
sticky, gradual boundary
Olive-yellow (2.5Y 6/6), loamy sand, singlegrained, medium to fine platy structure , nonsticky, gradual boundary

Total C
%

Total N
%

Minerals in
clay fraction
(< 2 m)

3.4

23.282.46

0.840.25

Q, K, Ilt**

3.9

n.d*

0.010.02

Q, K, Ilt

3.8

3.010.23

0.140.01

Q, K, Ilt, Ve

4.1

0.360.03

0.040.01

Q, K, Ilt, Ve

4.3

0.010

0.020.02

Q, K, Ilt, Ve

Soil pH
(H2O:soil=3:1)

*n.d. not detected;


** Minerals notation: Q quartz, K kaolinite, Ilt illite, Ve - vermiculate

115

Table 4-2. Infrared band assignments in the 4000-600 cm-1 region


Component
name

Absorption band,
in cm-1
3435-3400

OH and N-H stretch

Aliphatic C-H:
asymmetric stretch, -CH3
symmetric stretch, -CH2

2930-2920
2852-2850
1750-1715

C=O stretch of COOH groups and ketones

Humic acid
1650-1580

Fulvic acid

1040

C-O stretching of polysaccharides

1640 1630
1120-1110

SiOH deformation of SiOH on surface

Si-O stretch of silanol groups at silica


surface

1600

OH bending mode

1160

Si-O-Si stretching of silica structure

Si-O stretch of silanols

1040-1010

3630
1030,920
3380, 3430, 3520,
3620

Si-O stretch of silanols


OH bending: surface OH, inner OH
OH- stretching
OH-bending mode
OH stretching

1,2, 3, 4,
5, this
work
1

6, 7, 8, 9

10, 11,
this work

S; 3 sharp bands
S

12, this
work

S
S
S
S, 4 sharp peaks

OH bending

OH translational vibration

OH stretching

1060, 1020, 970, 915


740-730
3200

Goethite

S, broad

936-930, 915-910

Gibbsite

OH bending mode

OH- stretching

Referenceb

Broad

H-bonded silanols

3700-3600

Montmorrilonite

OH stretching of SiOH or sorbed water

3570-3500

1000-950

Kaolinite

Salts of COOH

980-930

Allophane,
imogolite

Asymmetric stretching of COO

1390

3800-3200
Colloidal silica
(SiO2)

Intensitya

Band assignment

12, this
work

11, 12, 13,


this work
12, this
work

900, 800
OH bending vibrations
S
Indication of intensity: S = strong, M = medium, and W = weak.
b
References: 1 Tan (2003); 2 Gonzalez-Perez et al. (2008); 3 Bonifacio et al. (2006); 4 Bonifacio et al. (2008), 5
Ellerbrock et al. (2005), 6 - Takamura et al. (1963), 7 - Guiton and Pantano (1993), 8 Zhdanov (1987), 9 Burneau and
Barres (1990), 10 Kartlun et al. (2000), 11 Gustaffson et al. (2000), 12 Farmer (1974), 13 Frost et al. (1998).
c
Other IR bands of fulvic acid have similar positions with humic acid, but may have different relative intensities (Tan, 2003)

116

Figures

5 cm
Sample holder

Profile view

Top view

Figure 4-1. Placement of quartz wafers into soil profile

117

Si forms

Fe forms
A1

amorphous
E2

Horizons

E2

Horizons

A
1

crystalline
amorphous
organic

Bh

3
Bh

4
Bhs

Bhs4

C5
0

0.0

10

0.1

0.2

Fe (g kg-1)

Si (g

A1

A1

amorphous
organic

Horizons

Horizons

E2

Bh3

E2

Bh3

Bhs4

Bhs
4

C5

C5
1

0.4

Percentage of amorphous material

Al forms

0.3

kg-1)

0.0

0.5

Al (g kg-1)

1.0

1.5

2.0

2.5

3.0

% amorphous material

Figure 4-2. Fe, Al and Si forms in soil obtained from selective extractions

118

Al

Horizons

Si-045
Al-045
Fe-045

Si

Bh

Al

Fe

Horizons

AlOH+2
Al+3
Al DOM1

Bh

Bhs

Bhs

C
0

Concentration,
mg/L
Concentration,

90

92

94

96

98

100

% of total Al concentration

mg/kg

Fe

200

Dissolved Carbon, mg/L


400
600

800

Bh
Bhs
C

Horizon

Horizons

Fe(OH)2+
FeOH+2
Fe DOM1

Bh

Bhs

C
90

92

94

96

98

100

% of total Fe concentration

Figure 4-3. Dissolved (<0.45 m soil water filtrates) concentrations of (A) Fe, Al, Si, and (B) carbon. Also shown are the results from
thermodynamic calculations of (C) Al and (D) Fe speciation. Fe DOM1 and Al DOM1 denote Fe and Al complexes with dissolved
organic matter, respectively.

119

Horizons

% Carbon and Nitrogen in colloids

Al

Fe

E Al
Fe
Al

Bh

Fe

Bhs Al
Fe
C Al
Fe
0

10

20

30

40

50

A
Horizons
Horizons

E
Carbon
Bh

Nitrogen

Bhs
C

10

20

60

65

70

75

Concentration (mg/kg)

Si

Ratio: (Fe+Al) / C in colloids

E
Bh
Bhs
C
0

0.000

0.010

0.020

0.030

0.040

0.050

0.060

Horizons
Horizon

Horizons

500 1000 1500 2000 2500 3000 3500

E
Bh
Bhs
C

Concentration, mg/kg

Figure 4-4. Concentrations of (A) Fe and Al, (B) Si, and (C) carbon and nitrogen in the colloidal fraction (particles within the 0.45 m
to 1.2 m size range) of soil water filtrates. (D) Atomic ratios of (Fe+Al) to carbon present in the colloidal fraction of soil waters.

120

800
3140

900

Fe-hydroxide
crystalline

goethite

Fe-hydroxide
amorphous

ferrihydrite
3450
1020

3520

740-730
Al-hydroxide
crystalline

3620

1030

gibbsite

1000
3690

910

3670
3620

1:1
alumosilicate

1110
1030

kaolinite

3630

920

montmorrilonite

1120

3550
3500

imogolite

1160

1630

2:1
alumosilicate

amorphous
alumosilicate

3200
amorphous
SiO2

silica
3400

1580

2920

1390

2850

organic

Humic acid
4000

3600

3200

2800

2000

1600

1200

800

wavenumber, cm-1

Figure 4-5. ATR-FTIR spectra of mineral and organic standards

3200

B
1190 1100
1410

3200

2920

1630
1715

840

2850

1400
1630
1700

Absorbance

3600
1630

1090

1627

2920

3700

1400

1085
1395

1100
1195
1415
850
1630

Bh
1420 1115
1650
1120
1420

Bhs
C
4000

3200
3500

3000

1600

1200

cm-1

Wavenumber,
wavenumber, cm-1

800

922 800
700

3600
2918

3700

Bh

4000

1630
1700

2930
3600

3200

2800

1400

920
800
700

1390 1040
1640

2920

Bhs
C

2500
2000

1032

1030

3200
3200

1030

1640
2000

1390 1040

1600

1200

800

wavenumber
Wavenumber,
cm-1(cm-1)

Figure 4-6. ATR-FTIR spectra of soil water extracted from A, E, Bh, Bhs and C horizons. A - <0.45 m size fraction, and B <1.2 m size fraction.

Figure 4-7. Typical SEM images of soil colloids (0.45-1.2 m) on 0.45 m pore size
silver filter. A. Blank filter; B. Colloids extracted from A-horizon; C. Colloids from Bh
horizon; insert are magnified particles from larger image; D. Colloids from Bhs horizon
with Fe-enriched particle.

18
16
14
12
10
8
6
4
2
0

0
18
16
14
12
10
8
6
4
2
0
-5
18
16
14
12
10
8
6
4
2
0
-5

9
8
7
6
5
4
3
2
1
0
9
8
7
6
5
4
3
2
1
0

A-horizon

10

15

20

25

30

35

40

10

12

14

16

15
8
6
4
2
0
17

Bh-horizon
6

A-horizon

16

10 20 30 40 50 60 70 80 90

Bh-horizon

16

Bh-horizon

15
8

6
2

10

15

20

25

30

35

40

4
2
0

10

12

14

16

Bhs-horizon
Bhs-horizon

10

15

20

10 20 30 40 50 60 70 80 90

25

30

35

40

10

12

Bhs-horizon

16

0
17

14

16

Fe concentration, atomic %

Al concentration, atomic %

9
8
7
6
5
4
3
2
1
0

17

A-horizon

15
8
6
4
2
0

10

20

30

A-horizon

Bh-horizon

Bh-horizon

0
5

Bhs-horizon

1
0
7

0.00

0.04

0.08

0.12

0.16

0.20

Bh-horizon

5
4
3
2

1
3

A-horizon

0 10 20 30 40 50 60 70 80 90 100

90

80

1
4

70

60

7
6

50

A-horizon

40

Si concentration, atomic %

10 20 30 40 50 60 70 80 90 100

Bhs-horizon

0
0.00
7
6

0.04

0.08

0.12

0.16

0.20

Bhs-horizon

Mg concentration, atomic %

10 20 30 40 50 60 70 80 90 100

Carbon concentration, atomic %

0
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40

Atomic (Fe+Al)/C ratios

Figure 4-8. Chemical composition of soil colloids (0.45 m< size <1.2 m) obtained
from SEM-EDS measurements for A, Bh, and Bhs horizons. Shown are the distribution
of concentrations (atomic %) of major elements and (Fe+Al)/C ratios in Spodosol
124

horizons. Y-axis values indicate frequencies of occurrence versus concentration range.

Figure 4-9. TEM-EDS data of colloids (0.45 to 1.2 m) from Bh horizon. A.


Amorphous silica particles ~100 nm size; B. Carbon-rich amorphous round structures
with some traces of Fe and Al; C. Organic matter with some crystalline inorganic

125

particles containing Si, Fe, Ca and Al as suggested by diffraction pattern and EDS
spectra.

Figure 4-10. TEM-EDS data of colloids (0.45 to 1.2 m) from Bhs horizon containing
amorphous organo-mineral colloids containing C, Al (major elements) and traces of Si,
Ca, Fe, P, and S.

126

Fe-(O,N)
Fe-Fe
Fe-Fe
Fe-C
Bh_2008
Bh_2006
Bhs_2008
Bhs_2006
Hematite
Goethite
Fe-catechol
Fe-EDTA
Fe-citrate

R(A)
Figure 4-11. Fourier transforms of Fe-EXAFS spectra of Fe-organic standards (Fecitrate, Fe-EDTA and Fe-catechol), Fe-mineral standards (goethite and hematite) and
field samples. These plots show inter atomic distances from Fe atom to its nearest
neighbors (not corrected for phase shifts).

127

C-H- Fe-hyd

Al-hyd

C-H O--Fe(III)--O C-H


O

Al-hyd

Fe-hyd

Al

O
O

C-H O

Fe(II)

Figure 4-12. Schematic diagram depicting suggested composition of metal-organocomplexes in soil water. Two-end members are shown: mineral particles with organic
groups attached to their surface (left) and organic particles bridged by Fe and Al ions
(right). The relative proportion of these two end-members and composition of the
intergrades between them depends on the supply of metals and organics to the soil water.

128

Chapter 5
CONCLUSIONS AND FUTURE WORK

All three projects described above used state-of-the-art approaches to study the
effect Al-substitution and time may have on the structure, and potential reactivity, of
complex Fe-Al- nanoparticles. In all cases, we linked laboratory investigations to
processes that occur in natural systems. We found that: (1) the presence of minor
concentrations of Al (2%) in solution significantly decreased the rate constant of nanogoethite formation (Chapter 2); (2) mixed amorphous Fe-Al hydroxide nanoparticles
precipitate forming indefinitely stable suspensions both with respect to colloidal stability
and phase transformation (Chapter 3); (3) Al incorporates into goethite by forming Alclusters, which is energetically more favorable compared to isolated substitution of Al for
Fe; however, even in clustered arrangement, Al creates strain on the goethite structure
which delays its further crystallization (Chapter 3); and (4) in soil environments (i.e.,
spodosol soil) Fe, Al, and Si migrate predominantly as inorganic colloids, however,
organic matter plays an important role in colloidal immobilization (Chapter 4).
Overall, the results from this investigation enhanced our knowledge and can help
predict the reactivity of soil colloids and coatings and therefore the retention and
transport of trace elements in soil systems.
Possible venues for future research are as follows:

Determination of the rates of nano-goethite formation in the presence of organic

ligands using the MCR approach we developed.

129

Further development of colloidal sampling and size separation techniques for soil
water.

Additional TEM investigations of soil colloids is promising for the differentiation


among inorganic nanoparticles, organic colloids (humic matter), and their
aggregates.

130

APPENDIX A. Composition and morphology of soil colloids (0.45m to 1.2 m) by


SEM-EDS
Table A-1. Elemental composition of colloids (0.45m <size<1.2 m) on silver filters as
derived from SEM-EDS analysis. Elements are expressed in atomic percentages. All
areas analyzed were morphologically similar (see Figures A4-1-1, A4-1-2, and A4-1-3).

Al+Fe
C

Horizon

Spot #

Si

Al

Mg

Fe

Cl

Si:Al

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22

60.39
66.19
74.29
15.47
70.54
71.47
66.23
69.02
43.53
74.86
80.57
80.42
64.32
69.11
33.11
81.4
73.22
63.16
80.75
77.27
80.95
77.67

20.51
16.59
4.89
62.38
1.17
10.46
13.38
20.19
36.74
6.56
3.51
8.51
14.41
17.12
57.35
3.38
6.8
29.58
6.56
3.43
2.75
5.86

0.12
0.99
2.04
bd
0.29
1.25
0.77
bd
0.14
1.54
1.03
1.06
1.03
0.05
bd
0.77
1.79
0.38
0.62
1.02
0.71
1.49

0.57
0.58
0.98
bd
bd
2.25
1.89
0.6
0.44
1.87
0.81
0.51
2.36
1.17
bd
0.56
0.91
bd
bd
bd
bd
bd

bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
0.61
bd
bd
bd
bd
1.94
bd
bd
bd
1.46
1.14

18.38
15.67
17.84
22.14
27.95
14.63
12.4
7.54
17.99
9.55
9.51
7.6
13.63
12.5
9.52
9.72
11.54
4.87
8.48
16.54
12.48
12.21

bd
bd
bd
bd
bd
bd
4.12
2.63
1.16
3.38
2.45
0.81
4.22
bd
bd
bd
bd
bd
bd
bd
0.43
0.87

bd
bd
bd
bd
bd
bd
bd
bd
bd
2.59
0.64
0.22
bd
bd
bd
2.28
1.97
1.54
2.58
1.18
0.64
bd

bd
bd
bd
bd
bd
bd
1.17
bd
bd
0.95
1.48
0.22
bd
bd
bd
1.91
1.8
0.45
1
0.54
0.56
0.77

170.9
16.8
2.4
na
4.0
8.4
17.4
na
262.4
4.3
3.4
8.0
14.0
342.4
na
4.4
3.8
77.8
10.6
3.4
3.9
3.9

average

67.00

16.01

0.78

0.70

0.23

13.30

0.91

0.62

0.49

50.6

Horizon

Spot #

Si

Al

Mg

Fe

Cl

Si:Al

0.016
Al+Fe
C

Bh

1
2
3
4
5
6
7
8

79.11
73.6
75.59
61.27
59.78
73.61
87.42
93.17

4.76
1.72
1.93
0.46
10.47
1.39
0.67
0.67

4.29
4.31
2.13
3.97
6.76
1.77
1.18
1.2

0.63
2.72
2.02
4.72
2.71
2.50
1.06
0.97

9.84
1.35
0.29
bd
2.34
2.18
bd
bd

8.57
16.25
18.03
29.58
17.96
18.55
9.68
3.98

bd
bd
bd
bd
bd
bd
bd
bd

bd
bd
bd
bd
bd
bd
bd
bd

bd
bd
bd
bd
bd
bd
bd
bd

1.1
0.4
0.9
0.1
1.5
0.8
0.6
0.6

0.179
0.077
0.032
0.065
0.152
0.054
0.013
0.013

131

0.002
0.015
0.027
0.004
0.017
0.012
0.003
0.021
0.013
0.021
0.016
0.001
0.009
0.051
0.006
0.008
0.013
0.027
0.034

9
10
11
12
13
14
15
16
17
18
19
20
21
22
23

79.1
81.74
76.16
91.24
86.73
93.31
88.95
82.44
92.6
81.1
75.02
93.98
91.59
91.37
89.26

1.31
7.39
0.7
0.93
1.65
1.64
1.14
7.03
1.01
1.78
4.31
2.1
1.4
1.47
1.42

1.81
2.29
3.84
2.46
4.58
2.6
3.29
5.65
2.38
2.08
5.12
1.85
3.79
4.23
1.79

1.17
2.15
0.68
1.76
0.89
bd
1.9
0.9
0.91
bd
bd
bd
bd
bd
2.69

0.54
bd
bd
bd
2.78
bd
3.48
2.46
1.59
1.02
3.09
0.86
1.94
1.34
bd

16.06
6.44
19.62
bd
bd
bd
bd
bd
bd
12.39
10.53
bd
bd
bd
bd

bd
bd
bd
3.68
1.95
1.68
1.3
1.46
1.5
0.87
0.62
0.39
0.37
0.42
4.89

bd
bd
bd
bd
0.55
bd
bd
bd
bd
0.24
0.57
0.38
0.24
0.35
bd

bd
bd
bd
bd
0.88
0.74
bd
bd
bd
0.52
0.77
0.43
0.66
0.83
bd

0.7
3.2
0.2
0.4
0.4
0.6
0.3
1.2
0.4
0.9
0.8
1.1
0.4
0.3
0.8

0.030
0.028
0.050
0.027
0.085
0.028
0.076
0.098
0.043
0.038
0.109
0.029
0.063
0.061
0.020

average

82.53

2.49

3.19

1.32

1.53

8.16

0.83

0.10

0.21

0.8

Horizon

Spot #

Si

Al

Mg

Fe

Cl

Si:Al

0.060
C
(Al+Fe)

Bhs

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
average

72.17
89.88
83.8
74.21
53.6
90.95
57.11
77.45
71.7
25.82
42.34
60.49
78.73
80.22
bd
bd
bd
bd
80.38
56.55
59.77
3.02
33.73
2.71
bd
13.45
10.03
69.49
74.39
46.97

1.67
0.44
11.82
2.75
3.78
3.63
0.85
2.16
3.47
40.77
27.97
16.79
2.52
2.61
70.21
85.29
57.78
56.92
7.6
15.34
11.5
57.25
12.33
54.09
8.84
9.71
6.67
23.23
1.07
20.66

2.09
8.05
1.20
5.7
14.28
10.15
10.07
8.41
11.55
25.64
22.27
13.23
7.53
5.43
24.4
11.23
37.58
36.56
7.78
20.3
19.29
32.13
34.38
38
3.07
3.13
1.74
5.49
2.59
14.60

1.91
1.67
3.17
1.06
0.65
4.11
1.14
1.37
bd
4.91
2.54
3.15
3.64
6.46
bd
bd
bd
1.32
bd
3.61
4.18
bd
7.49
bd
1.42
bd
bd
bd
1.08
2.89

bd
bd
bd
bd
2.29
bd
bd
0.26
bd
bd
bd
bd
bd
bd
1.08
1.19
0.88
1.69
bd
bd
bd
1.53
bd
0.89
14.7
10.47
12.23
bd
6.14
4.45

22.17
bd
bd
16.24
25.36
bd
30.85
10.29
13.28
2.98
4.81
5.97
7.73
5.89
4.3
2.35
3.83
3.56
4.23
4.56
5.61
6.08
5.87
4,34
14.66
14.97
10.04
1.79
14.73
9.69

bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
27.84
30.32
39.47
bd
bd
bd

bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd

bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd
bd

0.8
0.1
9.9
0.5
0.3
0.4
0.1
0.3
0.3
1.6
1.3
1.3
0.3
0.5
2.9
7.6
1.5
1.6
1.0
0.8
0.6
1.8
0.4

bd
5.79
bd
bd
bd
bd
bd

bd
23.68
17.2
19.84
bd
bd
bd

1.4
2.9
3.1
3.8
4.2
0.4
1.8

132

0.029
0.090
0.014
0.077
0.309
0.112
0.176
0.112
0.161
0.993
0.526
0.219
0.096
0.068
na
na
na
na
0.097
0.359
0.323
10.639
1.019
14.022
na
1.011
1.393
0.079
0.117
1.335

Figure A-1. SEM images of colloids (0.45-1.2 m) from A horizon

133

134

Figure A-2. SEM images of colloids (0.45-1.2 m) from Bh horizon

135

136

137

Figure A-3. SEM images of colloids (0.45-1.2 m) from Bhs horizon

138

139

140

141

142

APPENDIX B. SEM-EDS analyses of quartz wafers implanted in the Spodosol soil


Table B-1. Elemental composition of accumulations on quartz wafers. Only the presence
and relative intensity (tr trace, x - medium, X - high) is presented.

Horizon
A

Horizon
Bh

Area /
Spot

Al

1/1
1/2
1/3
2/1
3/1
4/1
4/2
4/3
4/4

x
x
x
x
x
X
X
X
X

tr
x

x
x
X
x
X
x
x

Area /
Spot
1/1
2/1
3/1
3/2
3 /3
4/1
5/1
5/2
6/1
6/2
6/3
6/4
6/5
7/1
7/2
7/3
8/1
8/2
8/3
8/4

C
x

tr
tr

Fe

x
tr
x
x
x

Ca

x
x

Al

Fe

X
X
X
X
X

x
x
x

x
X
tr
x
x
tr

x
x
x
x
x

Cl

x
x

x
x

Ca
x

X
tr
x

x
x

x
x
X
X

tr
x

x
tr

X
x

Mg
x

X
X

Mg

O
x
x
X
X
X
x
x
X
X
X
x
x
x
x
x
X

x
X
x

tr
tr

K
x
X
X
x
X
X

Cl

x
X
x
x
x

x
x
x

x
x
x
X
x
x
x

x
x
x
x
x
x
tr

x
x
x

143

8/5
9/1
9/2

Horizon
Bhs

Area /
Spot
1/1
1/2
1/3
1/4
1/5
1/6
1/7
2/1
2/2
2/3
2/4
3/1
3/2
4/1
4/2
5/1
5/2
5/3
7/1
7/2
7/3
8 /1
8 / 2*
8/3

x
X

X
X
x

x
X
x

Al
X
X
X
x
X
X
X
X
X
X
X
x
x

Fe
X
X
tr
tr
x
x
X
x
x
x
x
x
x

Ca

x
X
x

x
x

Cl

tr

X
X
X
X
x
x
X

X
X
X
x
tr
tr
x

O
X
X
X
X
X
X
X
X
X
X
X
X
X
x
x
x
X
x
x
x
x
x

x
x

Mg

X
X
x

tr
x
tr
tr

tr
x
x
x
X

x
x

* Example of the quartz wafer (SiO2 ) background

144

Figure B-1-a. SEM image and EDS spectra of A-horizon, area1.

145

Figure B-1-b. SEM image, EDS spectra and elemental map of A-horizon, area 2.

146

2 m

Figure B-1-c. SEM image and elemental map of A-horizon, area 3.

147

15 um

100 um

Figure B-1-d. SEM image, EDS spectra and elemental map of A-horizon, area 4.

148

Figure B-1-e. SEM image, EDS spectra of A-horizon, area 5.

149

Figure B-2-a. SEM image and EDS spectra of Bh-horizon, area 1.

Figure B-2-b. SEM image, EDS spectra of Bh-horizon, area 2.

150

Figure B-2-c. SEM image and EDS spectra of Bh-horizon, area 3.

Figure B-2-d. SEM image and EDS spectra of Bh-horizon, area 4.

151

Figure B-2-e. SEM image and EDS spectra of Bh-horizon, area 5.

152

Figure B-2-f. SEM image and EDS spectra of Bh-horizon, area 6.

153

Figure B-2-f. SEM image and EDS spectra of Bh-horizon, area 6 (magnified).

154

Figure B-2-g. SEM image and EDS spectra of Bh-horizon, area 7

155

Figure B-2-h. SEM image and EDS spectra of Bh-horizon, area 8

Figure B-2-i. SEM image and EDS spectra of Bh-horizon, area 9


156

Figure B-3-a SEM image and EDS spectra of Bhs-horizon, area 1


157

Figure B-3-b SEM image and EDS spectra of Bhs-horizon, area 2


158

Figure B-3-c SEM image and EDS spectra of Bhs-horizon, area 3

159

Figure B-3-d SEM image and EDS spectra of Bhs-horizon, area 4


160

Figure B-3-e SEM image and EDS spectra of Bhs-horizon, area 5

161

Figure B-3-f SEM image and elemental maps of Bhs-horizon, area 6

162

Figure B-3-g SEM image and EDS spectra of Bhs-horizon, area 7

Figure B-3-h. SEM image and EDS spectra of Bhs-horizon, area 8


163

APPENDIX C. X-ray diffraction (XRD) of Spodosol clay fraction (< 2 m)

Sample preparation
Soil pre-treatment
Physical and chemical processes of a soil are controlled to a large extent by the
minerals that compose the clay size fraction (McBride, 1994). This fraction is usually
obtained by dispersion of the soil in water, however, soil constituents such as organic
matter (OM) and iron oxides prevent clay dispersion and separation or may act as matrix
interferences causing peak shift during XRD measurements (Soukup et al., 2008). Airdried soils (< 2mm) were pretreated for XRD by removing organic matter using sodium
peroxide (H2O2) (Thurman, et al., 1992) followed by removal of sesquioxides according
to the procedure of Mehra and Jackson (1960). Although these pretreatments may result
in degradation of certain minerals (smectite) or cause loss of precision (Hughes et al.,
1994), we decided to remove OM and sesquioxides due to their high content in spodosol
horizons and their likelihood to prevent clay dispersion.
Soil dispersion
Ten grams of air-dried pre-treated soil (<2mm) were weighed into a 100 ml
plastic centrifuge tube and 5 ml of sodium hexametaphosphate were added as a
dispersion agent. The tubes were filled to 9.5 cm height with reverse osmosis (RO) water,
closed with a stopper, placed in a mechanical shaker and shaken overnight. The stopper
was then removed and rinsed with enough RO water to bring the volume to the 10 cm

164

mark. The suspensions were centrifuged at 760 rpm for 3 min, and the supernatant
(containing the < 2 m clay fraction) decanted.
Sample preparation for XRD analysis: cation saturation and heating
Layer spacings of ~7, 10 and 14 are commonly observed in clay minerals, and
they are diagnostic for their characterization. However, ambiguity of soil clay
determination arises from the fact that there is usually a mixture of several clay minerals.
Ambiguities in clay mineral identification may be resolved by specific sample treatments,
using characteristic chemical and thermal properties of clay minerals. Four sample
treatments were used to resolve the mineral composition of spodosol horizons: (1) clay
saturation with K+, (2) clay saturation with Mg2+, (3) heating of K+-saturated samples at
550 C, and (4) ethylene-glycol solvation of Mg2+-saturated samples.
Cation saturated samples were prepared by adding a small amount (5 ml) of the
clay suspension to a centrifuge tube and washing twice with 5 ml of 1 N KCl (K+saturation) or 1 N MgCl2 (Mg2+-saturation) solutions. The cation saturated suspensions
were washed twice with RO water to remove excess cations. After the last washing and
decantation, a few drops of suspension were transferred to the surface of a clean glass
slide, allowed to dry at room temperature, and then oven dried at 50 C for 1 hour to
ensure uniform conditions of dryness (Theisen and Harward, 1961). After the XRD
patterns for the cation-saturated samples were obtained, the K+-saturated samples were
placed in an oven at 550 C for 1 hour. The Mg2+-saturated samples were placed face-up
in a desiccator containing ethylene-glycol in the reservoir beneath the plate. The
desiccator was placed in an oven at 60 C overnight. XRD measurements were performed
immediately after sample removal from the desiccator.
165

Basis for interpretation of XRD results


A 7 spacing indicates the presence of kaolinite, a non-expandable 1:1 mineral
with low cation exchange capacity. Although chlorite or serpentine may present a
second-order maximum at ~7 , a peak at 14 must be present for their identification.
Heat treatment at 550 C destroys the lattice of kaolinite and therefore the 7 peak
should disappear thus confirming its presence.
A 10 spacing is indicative of unexpanded mica-type layers. An example of this
type of 2:1 mineral commonly found in soils is illite, a non-swelling hydrous mica with
interlayer space mainly occupied by potassium cations. Halloysite (1:1 clay mineral) will
also have a 10 peak, however, hydrated halloysite readily dehydrates to yield a 7
peak upon heating above 100 C. Therefore, the absence of a 7 peak after heating rules
out the presence of halloysite.
A 14 spacing is usually due to chlorites, vermiculites and smectites. Chlorite
has a 14 peak with all treatments, including heating at 550 C. Vermiculite exhibits a
14 peak with Mg2+ saturation, however, it collapses to 10 with K+ saturation.
Smectite shows a 14-16 peak with Mg2+ saturation, but expands to 17-18 with Mg2+glycerol saturation and shifts to 10 with K+ saturation (Kimpe, 1993). According to Lin
et al (2002), broad XRD peaks between 10 and 14 for Mg-saturated clays indicate the
presence of inter-stratified vermiculate-illite minerals. These peaks collapse to 10 after
heating at 550 C.
REFERENCES
166

1. McBride, M.B. (1994) Environmental Chemistry of Soils. Oxford University


Press, Inc.
2. Mehra, O.P., and Jackson, M.L. (1960) Iron oxide removal from soils clays by a
dithionite citrate system buffered with sodium bicarbonate, Clays Clay Miner. 7,
313317.
3. Soukup, D.A., Buck, B.J., and Harris, W. (2008) Preparing soils for mineralogical
analysis. In: Methods of soil analysis. Part 5 Mineralogical Methods, SSSA
Book Series. Eds.: Ulery., A.L., and Drees, L.R.
4. Thurman, N.C., and Ciolkosz, E.J., and Dobos, R.R. (1992) Pennsylvania State
University Soil Characterization Laboratory Methods Manual. Publication #117.

167

Quartz
3.34 A

A - horizon

Quartz
4.25 A
Illite
10A

Kaolinite
7.15 A

Mg2+-Glycol

Illite
10A

Kaolinite
7.15 A
Illite
4.9 A

Quartz
4.25 A

Mg2+ only
Illite
10A
Illite
4.9 A

Quartz
4.25 A

no kaolinite
at ~7 A

at 550 C

Illite
10A

+
K only

Kaolinite
7.15 A

Quartz
4.25 A
Kaolinite
Chlorite
3.5 A

10
d-space, A

Figure C-1. XRD spectra of different treatments for the clay fraction (< 2 m) from Ahorizon
168

Quartz
3.34 A

E - horizon

Kaolinite
3.5 A
Quartz
4.25 A
Kaolinite
7.15 A

Mg2+-Glycol

Kaolinite
7.15 A

Mg

2+

Quartz
4.25 A

Kaolinite
3.5 A

only
Quartz
4.25 A

Illite
10A

no kaolinite
at ~7 A

at 550 C
Kaolinite
7.15 A
Illite
10A

+
K only

Quartz
4.25 A
Kaolinite
Chlorite
3.5 A

10
d-space, A

Figure C-2. XRD spectra of different treatments for the clay fraction (< 2 m) from Ehorizon

169

Quartz
3.34 A

Bh - horizon

Illite
10A

2+

Mg -Glycol

Kaolinite
7.15 A

Kaolinite
3.5 A
Illite Quartz
4.9 A 4.25 A

Vermiculite
Chlorite
Smectite
14 A

Quartz
3.34 A

Kaolinite
7.15 A
Illite
10A

Illite
4.9 A

Mg2+ only

Vermiculite
Chlorite
Smectite
14 A

no kaolinite
at ~7 A

Illite
10A

Quartz
4.25 A

Quartz
4.25 A

Illite
10A

+
K at 550 C

Kaolinite
3.5 A

Kaolinite
7.15 A

Illite
4.9 A

Quartz
4.25 A
Kaolinite
Chlorite
3.5 A

only

10
d-space, A
Figure C-3. XRD spectra of different treatments for the clay fraction (< 2 m) from Bhhorizon

170

Quartz
3.34 A

Bhs - horizon
Vermiculite
Chlorite
Smectite
14 A

Kaolinite
3.5 A

Kaolinite
7.15 A

Illite
10A

Illite
4.9 A

Quartz
4.25 A

2+

Mg -Glycol
Vermiculite
Chlorite
Smectite
14 A
Kaolinite
7.15 A
Illite
10A

Mg

2+

Illite
4.9 A

Quartz
4.25 A
no kaolinite
at ~7 A

Illite
4.9 A

at 550 C
Vermiculite
Chlorite
Smectite
14 A

Quartz
4.25 A

only

Illite
10A

Kaolinite
3.5 A

only

Illite
10A

Kaolinite
7.15 A

Quartz
4.25 A
Kaolinite
Chlorite
3.5 A

10
d-space, A

Figure C-4. XRD spectra of different treatments for the clay fraction (< 2 m) from Bhshorizon

171

Quartz
3.34 A

C - horizon
Vermiculite
Chlorite
Smectite
14 A

Illite
10A

Kaolinite
3.5 A

Kaolinite
7.15 A
Illite
4.9 A

Quartz
4.25 A

Mg2+-Glycol

Vermiculite
Chlorite
Smectite
14 A

Illite
10A

Kaolinite
3.5 A

Kaolinite
7.15 A
Illite
4.9 A

Quartz
4.25 A

Mg2+ only
Illite
10A

no kaolinite
at ~7 A

Illite
4.9 A

Quartz
4.25 A

+
K at 550 C
Vermiculite
Chlorite
Smectite
14 A

+
K only

Illite
10A
Kaolinite
7.15 A
Illite
4.9 A

Quartz Kaolinite
4.25 A Chlorite
3.5 A

10
d-space, A

Figure C-5. XRD spectra of different treatments for the clay fraction (< 2 m) from Chorizon

172

CURRICULUM VITAE
Ekaterina Bazilevskaya
Education:
Ph.D. in Soil Sciences and Biogeochemistry, PSU, Fall 2009
M.Sc. in Geosciences, PSU, Fall 2004
B.S. in Geochemistry, Moscow State University, Russia, Spring 1993
Professional skills:
Extensive analytical instrument experience:
Infrared spectroscopy (ATR-FTIR), SEM, TEM, ZetaPALs, NanoSizer, Elemental
Analyzer (EA) for C and N.
Synchrotron work:
EXAFS and XRD at BNL; GIXAS at Stanford Synchrotron National Laboratory
Lab experience: synthesis of nanoparticles, glovebox experiments, basic soil analysis
Field experience: developing experiment on in situ formation of muneral coatings in forest soils
Work experience
2004-present: Research Assistant, Department of Crop and Soil Sciences,
2001-2004: Research Assistant, Department of Geosciences
1997-1999: Tax Assistant, PricewaterhouseCoopers, Moscow, Russia
1995-1997: Research Assistant, Vernadsky Institute of Geochemistry and Analytical
Chemistry, Moscow, Russia
1991-1993: Lab Assistant, , Moscow State University, Moscow, Laboratory of
Analytical Geochemistry
1991: Project Assistant Kamchatka Geological Crew,. Summer Internship.
Teaching Experience:
Teaching Assistant, Spring 2008; Introduction to soils, SOILS 101; Spring 2007;
Environmental Soil Chemistry, SOILS 419; Fall, 2002; Earth Materials, Geosc 201
Mentor for REU students, Summer 2006, 2007
Awards
Dixon Soil Mineralogy Award for the best oral presentation the ASA-CSSA-SSSA
International Meeting, November, 2007, New Orleans, Louisiana.
2006-2007 CAS&DCSS/PSU Grant for the proposal A Novel Statistical Approach to
Study Phase Transformation of Environmental Colloids, $2000
Center of Environmental Kinetics Analysis (CEKA) Fellow, 2005 2008
9th Annual ECSS, PSU, 2005 - second place for poster presentation
7th Annual ECSS, 2004 second place for poster presentation
Community Activities and Public Services
Environmental Chemistry Student Symposium (ECSS) committee, 2007, 2008
Department of CSS, Coffee Room Committee 2006-2007
Judge for Pennsylvania Junior Academy of Science (PJAS), 2006, 2007

You might also like