Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)
Article
Synergistic Effect of Reductive and Ligand Promoted Dissolution of Goethite Zimeng Wang, Walter D. C. Schenkeveld, Stephan M. Kraemer, and Daniel E. Giammar Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 12 May 2015 Downloaded from http://pubs.acs.org on May 12, 2015
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 39
Environmental Science & Technology
Synergistic Effect of Reductive and Ligand Promoted Dissolution of Goethite Zimeng Wang†*, Walter D.C. Schenkeveld‡, Stephan M. Kraemer‡, and Daniel E. Giammar§ † Department of Civil and Environmental Engineering, Stanford University, Stanford, CA, United States ‡ Department of Environmental Geosciences, University of Vienna, Vienna, Austria §Department of Energy, Environmental and Chemical Engineering, Washington University in St. Louis, St. Louis, MO, United States
*Corresponding author Jerry Yang and Akiko Yamazaki Environment & Energy Building 473 Via Ortega Stanford, California 94305, United States Email:
[email protected] Web: www.stanford.edu/~wangzm
Manuscript Submitted to Environmental Science & Technology May 2015
0
ACS Paragon Plus Environment
Environmental Science & Technology
1
Abstract
2
Ligand promoted dissolution and reductive dissolution of iron (hydr)oxide minerals control
3
the bioavailability of iron in many environmental systems and have been recognized as
4
biological iron acquisition strategies. This study investigated the potential synergism between
5
ligands (desferrioxamine B (DFOB) or N,N'-Di(2-hydroxybenzyl)ethylenediamine-N,N'-diacetic
6
acid (HBED)) and a reductant (ascorbate) in goethite dissolution. Batch experiments were
7
performed at pH 6 with ligand or reductant alone and in combination and under both oxic and
8
anoxic conditions. Goethite dissolution in the presence of reductant or ligand alone followed
9
classic surface controlled dissolution kinetics. Ascorbate alone does not promote goethite
10
dissolution under oxic conditions due to rapid reoxidation of Fe(II). The rate coefficients for
11
goethite dissolution by ligands are closely correlated with the stability constants of the aqueous
12
Fe(III)-ligand complexes. A synergistic effect of DFOB and ascorbate on the rate of goethite
13
dissolution was observed (total rates greater than the sum of the individual rates), and this effect
14
was most pronounced under oxic conditions. For HBED, macroscopically the synergistic effect
15
was hidden due to the inhibitory effect of ascorbate on HBED adsorption. After accounting for
16
the concentrations of adsorbed ascorbate and HBED, a synergistic effect could still be identified.
1
ACS Paragon Plus Environment
Page 2 of 39
Page 3 of 39
Environmental Science & Technology
17
The potential synergism between ligand and reductant for iron (hydr)oxide dissolution may have
18
important implications for iron bioavailability in soil environments.
19 20
Introduction
21
Iron is an essential micronutrient for plants and microorganisms. In calcareous soils and
22
marine waters, the solubility of ferric (hydr)oxides can only provide iron concentrations that are
23
far below the needs of most plants, fungi, and microorganisms. Organisms have developed
24
strategies to cope with conditions of low iron availability and to enhance iron solubility. Given
25
that the proton-promoted dissolution of iron (hydr)oxides is very slow at circumneutral pH,1, 2
26
mobility of iron in aquatic environments can be enhanced by (1) reduction of Fe(III) to the more
27
soluble form of Fe(II)3-8 and (2) formation of soluble Fe(III)-complexes with chelating ligands.
28
In oxic environments at circumneutral pH, Fe(III) (hydr)oxides are the thermodynamically
29
stable forms of iron. Fe(II), even if generated by a surface-acting reductant, cannot persist in the
30
oxic solution9, 10 unless associated with certain specific Fe(II) binding ligands11 Iron is more
31
effectively mobilized through complexation with high affinity organic ligands. In agriculture,
32
corrective measures against iron limitation involve the application of iron chelated with synthetic
33
ligands such as EDDHA (ethylenediamine-N,N'-bis2-hydroxyphenylacetic acid).12,
2
ACS Paragon Plus Environment
13
Such
Environmental Science & Technology
Page 4 of 39
34
ligands can mobilize Fe and other metals from soil,14, 15 but the properties of environmental
35
surfaces can strongly affect the kinetics of their reactions with metals.16 Similarly, in natural
36
environments, an efficient biological iron acquisition strategy of bacteria, fungi, and
37
graminaceous plants involves the exudation of siderophores.17-20 Siderophores (e.g.,
38
desferrioxamine B, DFOB) are biogenic iron-binding ligands that are released under iron-
39
deficient conditions. For typical hydroxamate siderophore concentrations in soil (10 to 100
40
nM),21, 22 iron oxide solubility increases over a wide pH range. Above pH 4, ligand-promoted
41
dissolution is faster than proton-promoted dissolution.23-26 Previous studies have provided
42
spectroscopic evidence of siderophore surface complexes on iron oxides27,
43
application of the ligand promoted dissolution mechanism to describe siderophore-iron oxide
44
interaction. Because an Fe(III)-siderophore surface complex is a precursor of the intermediate
45
species in the rate-determining step in ligand controlled dissolution, the dissolution rate is a
46
linear function of the adsorbed ligand concentration29 and of the solution saturation state30,
47
∆
( )
= ∆ = 1 − exp = !1 − "$ ' * #
%&
28
, allowing the
(1)
48
where RL[nmol g−1h−1] is the ligand controlled dissolution rate, kL [h−1] is the first order rate
49
coefficient, [L]ads [nmol g−1] is the adsorbed ligand concentration, ∆G is the Gibbs free energy of
50
reaction, R is the gas constant, T is the absolute temperature, Q is the activity quotient, Keq is the 3
ACS Paragon Plus Environment
Page 5 of 39
Environmental Science & Technology
51
equilibrium constant of the overall reaction, and σ is Temkin’s coefficient31 (usually 1 to 3 for
52
metal oxide dissolution) to account for the non-elementary nature of a reaction (for an
53
elementary reaction the value is 1).31, 32
54
In addition to having low concentrations in the environment, hydroxamate siderophores also
55
do not adsorb strongly to iron oxide surfaces.23 Therefore, it has been proposed that siderophores
56
act in conjunction with other molecules to increase iron bioavailability.33-35 Low molecular
57
weight organic acids (many of which bind Fe(III) relatively weakly) are ubiquitous in soil, and
58
can have synergistic effects on siderophore-promoted dissolution of iron oxides, where the net
59
dissolution rates in the presence of both ligands are beyond the sum of the net dissolution rates
60
observed in single ligand systems.36 This synergistic effect has been documented for various
61
organic acids, different types of siderophores, and several iron(III) minerals,24, 26, 34, 37-40 with
62
several possible mechanisms proposed. First, the strong affinity of the siderophore for Fe(III)
63
increases the Fe(III) solubility and shifts the solution saturation state to maintain far from
64
equilibrium conditions (i.e., f(∆G) →1).34, 38 Second, organic ligands can form kinetically labile
65
iron complexes on the mineral surface, which can be rapidly extracted by the siderophore at
66
higher rates than pristine Fe(III) sites.24, 37 For experiments with oxalate as a complexing ligand,
67
the presence of a labile pool of Fe(III)-oxalato ternary surface complex was validated by isotope
4
ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 39
68
dilution experiments24, kinetic modeling40 and spectroscopy28, 41. Third, enhanced adsorption of
69
certain positively charged siderophores, like DFOB, may result from a surface charge shift
70
induced by adsorption of negatively charged compounds. This mechanism was used to interpret
71
the synergistic effect of fulvic acid39 and anionic surfactant35 on DFOB-promoted dissolution of
72
goethite.
73
It has been suggested that extracellular reductants also contribute to iron bioavailability.33, 42
74
Under iron-limited conditions, microorganisms and plant roots are known to release compounds
75
with reducing capacity, even when the surrounding environments are oxic19, 42, 43. Exogenous
76
reductants include pyridine-2,6-bis(monothiocarboxylate) (PDTC),44 ascorbate33 and various
77
phenolic compounds (e.g., caffeic acid).19 However, the geochemical role of reductants in
78
mediating iron bioavailability in oxic environments is not clear. In the presence of O2 at
79
circumneutral pH, any reduced Fe(II) on the iron oxide surface is rapidly reoxidized.1,
80
Nonetheless, several microbiological studies demonstrated a substantial role of a chemical
81
reductant in iron acquisition by aerobic microorganisms.33,
82
bioavailability based on cell growth, and their interpretation of the role of a reductive pathway
83
was largely based on cell membrane and intracellular processes such as (1) the direct uptake of
84
transient Fe(II) from the iron oxide surface by cells in very close proximity or direct contact47-49
45, 46
5
ACS Paragon Plus Environment
9, 10
These studies assessed iron
Page 7 of 39
Environmental Science & Technology
85
and (2) ascorbate acting as a reductant to “assist iron passage through the membrane
86
transporter”33.
87
The interplay of reductant and ligand on iron oxide dissolution is a classical research topic
88
that began with pioneering studies by Stumm and colleagues. They suggested that ascorbate and
89
oxalate/malonate in combination could dissolve iron oxides at rates beyond the sum of the
90
individual rates.3,
91
together with oxalate and postulated a catalytic role of Fe(II)-oxalate complexes in acting as a
92
surface-acting reductant for Fe(III).4 This pathway was observed to be faster than non-reductive
93
oxalate-promoted iron oxide dissolution. As opposed to soil relevant oxic and neutral conditions,
94
these investigations were performed under anoxic conditions at pH 3 or lower, where the
95
reductive dissolution pathway itself can generate substantial soluble Fe. Dhungana et al. studied
96
ferrihydrite dissolution at pH 7.5 by dipicolinic acid (a ligand) and H2S (a reductant), proposing
97
a catalytic role of the in situ generated Fe(II).50 However, it is still not clear whether oxygen
98
would compromise the reductive pathway and how it would affect the dissolution kinetics. A
99
synergistic effect between photoreductive and ligand promoted dissolution of iron (hydr)oxides
100
was reported in the context of oxic marine systems.51-53 Under UV irradiation, adsorbed oxalate
101
can act as the electron donor to form surface Fe(II), and the siderophore then acts as a shuttle to
7
They later observed increased dissolution rates when Fe(II) was added
6
ACS Paragon Plus Environment
Environmental Science & Technology
102
transfer surface Fe(II) into solution. This synergistic effect has been demonstrated in irradiated
103
systems but not in experiments performed in the dark.
104
Building on the existing knowledge of iron oxide dissolution mechanisms, the objective of
105
the present study was to systematically evaluate the interplay of reductive and ligand-promoted
106
dissolution of goethite under pH conditions relevant to soils. This interplay can have implications
107
for iron bioavailability to bacteria, fungi and plants. The central hypothesis was that a synergistic
108
effect can enhance iron mobility, especially under oxic conditions. To test the hypothesis,
109
dissolution experiments were performed under anoxic and oxic conditions with varied reductant
110
and ligand concentrations.
111 112
Materials and Methods
113
Materials. Goethite (α-FeOOH) was prepared according to the standard synthesis procedure
114
in Cornell and Schwertmann.2 Briefly, KOH was added to Fe(NO3)3 solution with vigorous
115
stirring. The initial suspension of red-brown ferrihydrite was heated at 70 °C for 60 h to produce
116
goethite. Excess reactants were removed by repeated dialysis against deionized water until the
117
conductivity of the wash water was below the detection limit of 1 µS/cm. After the final dialysis
118
step, the product was centrifuged and stored as a concentrated slurry at 4 °C. Polyethylene
7
ACS Paragon Plus Environment
Page 8 of 39
Page 9 of 39
Environmental Science & Technology
119
containers were used for preparation and storage. X-ray powder diffraction (Rigaku
120
GeigerflexD-MAX/A) confirmed the identity of the material as goethite. The specific surface
121
area was determined to be 38 m2/g by a multipoint N2-BET adsorption method (Autosorb-1-
122
C,Quantachrome). The L-ascorbic acid (C6H8O6, Sigma Aldrich, >95%), desferrioxamine B
123
mesylate salt (C25H48N6O8·CH4O3S, DFOB, Sigma Aldrich, >92.5%) and N,N'-Di(2-
124
hydroxybenzyl)ethylenediamine-N,N'-diacetic
125
(C20H24N2O6•HCl•H2O, HBED, Strem Chemical Inc. >98%) were used as received without
126
further purification. Ascorbate is a common biogenic reductant in soils, and ascorbate is a very
127
weak ligand for iron that cannot mobilize Fe through ligand promoted dissolution (Figure S1b of
128
the Supporting Information). DFOB is a microbial hydroxamate siderophore, and HBED is used
129
in synthetic iron fertilizers for calcareous agricultural soils. Given the concern of possible
130
degradation of these compounds in solutions, all the chemicals were kept in dry powder forms
131
(as received) until the goethite dissolution experiments were performed. Ultrapure water
132
(resistivity >18.2 MΩ·cm, Milli-Q, Millipore) was used in all experiments.
acid
monohydrochloride
hydrate
133
Goethite Dissolution Experiments. The experimental conditions are summarized in Table 1.
134
All experiments were performed at a constant temperature (22 °C) with reactors shielded with
135
aluminum foil to avoid potential photochemical redox reactions. Dissolution experiments were
8
ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 39
136
conducted in completely mixed batch reactors (100 mL) using magnetic stirring. A fixed ionic
137
strength (10 mM) was provided by sodium chloride. A pH of 6.0 that is relevant to soil
138
environments was set for all experiments. The pH was buffered by 1 mM 2-(N-
139
morpholino)ethanesulfonic acid (MES, as sodium salt, pKa = 6.15), a widely used pH buffer
140
verified to be applicable for goethite dissolution studies.39, 54 The solution pH was largely stable
141
during the experiments (6.0 ± 0.05), and the pH was adjusted to this range using concentrated
142
HCl or NaOH. Oxic experiments were performed using ultrapure water sparged with air (PO2 =
143
0.21 bar) and with the headspaces of the reactors connected to the atmosphere. Anoxic
144
experiments were performed in an anaerobic chamber (Coy Laboratory MI), where the gas
145
composition was controlled to less than 1 ppm for O2 and between 2% and 5% (v/v) for H2. The
146
ultrapure water used in these experiments was sparged by H2 with a Pd catalyst in the water. The
147
reactors and solutions were equilibrated with the anoxic atmosphere in the chamber overnight or
148
longer prior to the goethite dissolution experiments under anoxic conditions. All the experiments
149
started with 1 g/L goethite added to freshly prepared solutions containing aliquots of ligand
150
and/or reductant and with preadjusted and buffered pH. Samples were periodically collected,
151
filtered with 0.1 µm poly(ether sulfone) (PES) syringe filters (Millipore), and preserved in 2%
152
HNO3 for dissolved Fe analysis.
9
ACS Paragon Plus Environment
Page 11 of 39
Environmental Science & Technology
153
Adsorption experiments. Adsorption experiments quantified the adsorption of reductant and
154
ligands to the goethite surface. All adsorption experiments were conducted under the ambient
155
atmosphere using batch reactors (50 mL) that were completely mixed on a shaker. The goethite
156
concentration, ionic strength, pH, and buffer were selected to be consistent with those of the
157
dissolution experiments. Wide ranges of ascorbate, DFOB and HBED loadings were equilibrated
158
with 1 g/L goethite for 2 hours, which was found to be sufficient for reaching adsorption
159
equilibrium while having minimal goethite dissolution. Filtered (0.1 µm) samples were analyzed
160
within 1 hour after filtration for the dissolved reductant and ligand concentrations.
161
Aqueous Analysis. The pH of solutions and suspensions was measured with a glass pH
162
electrode (Accumet). Dissolved Fe concentration was measured by inductively coupled plasma
163
mass spectrometry (ICP-MS, PerkinElmer ELAN DRC II). The detection limit for Fe was 10 µg
164
L-1 (0.18 µM). It should be noted that ICP-MS is not able to differentiate between the oxidation
165
states of dissolved Fe.
166
In adsorption experiments, the dissolved concentrations of ascorbate, HBED and DFOB were
167
analyzed using a UV-Vis spectrophotometer (PerkinElmer-Lambda XLS). Absorbances of Fe-
168
DFOB and Fe-HBED complexes were measured at 439 and 493 nm, respectively. These peaks
169
were confirmed to be free of interference with the ascorbate peak. For ascorbate, the original
10
ACS Paragon Plus Environment
Environmental Science & Technology
170
filtrate samples were measured for absorbance at 243 nm.3,
171
DFOB or HBED in the filtrate was quantified by measuring the absorbance of Fe-DFOB or Fe-
172
HBED complexes upon addition of a fixed concentration of Fe(III) that is in excess of the
173
ligand.55 This method was successfully used in previous studies on ligand promoted goethite
174
dissolution34,
175
described in the Supporting Information). The modified method was verified to be effective; the
176
measured concentration correlated remarkably well (R2> 0.999) with the values obtained using
177
the classical method (Figure S2 of the Supporting Information).
37, 56
4
Page 12 of 39
The concentration of dissolved
. We modified this method to overcome the ascorbate interference (details
178
The MES/NaCl/pH 6 solution was used as the blank and its spectrum referenced to DI water
179
does not have peaks overlapping those of interest. The absorbance of the original filtrate samples
180
without added Fe(III) was also measured at 439 or 493 nm to account for the DFOB in the
181
solution that was already complexed to Fe(III) released during the two-hour adsorption
182
experiments. Due to the short contact time for the adsorption experiments, the absorbance for the
183
original filtrate was verified to be negligible relative to that after the addition of excess Fe(III).
184
The adsorbed ascorbate concentration, DFOB and HBED were calculated by dividing their
185
concentration loss by the goethite concentration.
11
ACS Paragon Plus Environment
Page 13 of 39
Environmental Science & Technology
186
Rate Calculation and Modeling. The dissolution rates of goethite were estimated by linear
187
regression of the dissolved Fe concentration versus time for the first three samples unless
188
otherwise noted. The regression was not forced through the origin, and a positive intercept was
189
due to the fast initial dissolution, which is commonly observed in batch experiments of iron
190
(hydr)oxide dissolution, possibly from small amounts of labile Fe species.24, 51, 57 After the rapid
191
initial dissolution, the regression of dissolved iron versus time is linear as was anticipated for
192
dissolution reactions far from equilibrium.58, 59 Caution was taken in the data selection for rate
193
calculations to make sure that the dissolution rate was captured without being affected by
194
significant decrease of the free ligand concentration; this data selection allows the solution
195
saturation term in Eq. 1 to be ignored in the subsequent analysis. The fraction of the initial
196
goethite that was dissolved during the experiment was so small that the mass normalized
197
dissolution rate was calculated using the initial goethite concentration. Rate coefficients were
198
obtained by linear correlation of dissolution rates versus adsorbed reductant or ligand
199
concentrations. Chemical equilibrium modeling was performed using MINEQL+ in which
200
activity coefficients were calculated by the Davies equation.60
201 202
Results and Discussion 12
ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 39
203
Adsorption Experiments. The adsorption isotherm of ascorbate suggests that a 1 mM total
204
ascorbate concentration already approached surface saturation (Figure S3a of the Supporting
205
Information). The addition of up to 80 µM DFOB or HBED could displace at most 26% of the
206
total adsorbed ascorbate (Figure S3b,c of the Supporting Information). All the ligand
207
concentrations used in this study (up to 20 µM for DFOB, and up to 10 µM for HBED) fell
208
within the initial steep range of their adsorption isotherms (Figure 1). Based on the initial slopes
209
of the isotherm data, adsorption of DFOB and HBED appears less favorable than adsorption of
210
ascorbate.
211
Ascorbate and ligands were expected to compete for surface adsorption sites, and they can
212
also impact the adsorption equilibrium of each other by altering the surface charge of goethite.
213
The addition of 1 mM ascorbate decreased the adsorption of HBED to goethite, while it did not
214
cause a substantial change in the DFOB adsorption. The difference between the effect of
215
ascorbate on DFOB and HBED adsorption may be caused by differences in the charges of their
216
surface species.
217
(H4DFOB+), while the dominant HBED species is negatively charged (H3HBED−) (Figure S4). If
218
the adsorbed forms of DFOB and HBED follow a similar trend, with the DFOB surface
219
complexes being either neutral or positively charged, then adsorption of ascorbate on goethite at
At pH 6, the dominant dissolved DFOB species is positively charged
13
ACS Paragon Plus Environment
Page 15 of 39
Environmental Science & Technology
220
this pH (where the goethite surface has a positive surface charge in the absence of any adsorbing
221
species), would decrease the positive charge of the goethite surface, making adsorption of the
222
negatively charged HBED less favorable and adsorption of the positively charged DFOB more
223
favorable.
224
Reductant or Ligand in Isolation: Control Experiments. The results from the control
225
experiments (Exp. 1-12) in which either a reductant or a ligand was added are consistent with
226
classical metal (hydr)oxide dissolution theories. As expected from solubility calculations, in the
227
absence of reductant and ligand, goethite dissolution (Exp. 1, 4) was too slow to be detectable,
228
regardless of whether O2 was present. Ascorbate could promote goethite dissolution under
229
anoxic condition (Exp. 2, 3). The established reductive dissolution mechanism, in which the
230
detachment of Fe(II) is the rate determining step, describes the dissolution rate as a linear
231
function of adsorbed reductant.4 The dissolution rates are linearly correlated with the surface
232
concentration of adsorbed ascorbate (Figure 2a, kascorbate = 5.7×10−4 h−1). However no net
233
dissolution of goethite by ascorbate was observed under oxic conditions since the dissolved iron
234
concentrations never exceeded the the detection limit (Exp. 5, 6). When dissolved O2 is available,
235
the surface Fe(II) generated by ascorbate does not enter the solution since the heterogeneous
236
oxidation of Fe(II) is very fast.
14
ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 39
237
Goethite dissolution by DFOB and HBED was not impacted by the presence or absence of
238
oxygen in the absence of a reductant (Exp. 7 and 9, 11 and 12). The lack of an effect of O2 on the
239
goethite dissolution rate in the presence of DFOB or HBED indicates that the process did not
240
involve any substantial release of Fe(II) as in a reductive dissolution mechanism. Some previous
241
research has seen evidence for hydrolysis and oxidation of siderophores on the goethite surface28,
242
50
243
enough extents to influence the measured Fe release. The dissolution rates were linearly
244
proportional to the adsorbed DFOB concentration (Figure 2b), which was consistent with the
245
ligand-promoted dissolution mechanism. HBED is expected to follow the same mechanisms,
246
although we only measured dissolution at one HBED loading. The rate coefficient (as in Eq. 1)
247
for DFOB (0.0089 h−1) was smaller than that for HBED (0.027 h−1) by a factor of three. At pH 6,
248
HBED is a stronger ligand for Fe(III) than DFOB (Figure S5 of the Supporting Information).
; however, if such reactions occurred in the present experiments, then they were not to great
249
The rate coefficients for goethite dissolution promoted by the ligands studied in this work
250
and in previous studies that were carried out under similar conditions (pH 6, etc.) show a
251
remarkable correlation with the stability constants of the 1:1 fully deprotonated Fe(III)-ligand
252
complex (Figure 3). It has been previously reported that dissolution rates increase with
253
increasing complex stability constants.61, 62 Our present analysis compiled the rate coefficients
15
ACS Paragon Plus Environment
Page 17 of 39
Environmental Science & Technology
254
instead of the rates as the other studies did, which allowed us to exclude the variability in ligand
255
concentration and the different adsorption affinities of the ligands and possible thermodynamic
256
constraints (See the terms in Eq. 1). The correlation of rate coefficients with stability constants
257
suggests that the binding strength of the ligand to Fe in solution is related to the rate-limiting step
258
in ligand-promoted dissolution. Deeper interpretation of this correlation is hampered by the
259
unknown surface speciation of the adsorbed ligand in general and the coordinative environment
260
of the precursor of the rate determining step in particular.
261
Goethite Dissolution: Ascorbate + DFOB. Comparing the macroscopic dissolution rates of
262
goethite in the presence of ligand and reductant together with the rates in the presence of ligand
263
or reductant individually, a synergistic effect of ascorbate and DFOB on goethite dissolution was
264
observed. This effect was supported by goethite dissolution rates obtained from various
265
combinations of ascorbate and DFOB concentrations, both under oxic and anoxic conditions
266
(Figures 4 and 5). With the adsorption data and kinetic model, the synergistic effect could not be
267
explained merely by the change of the concentrations of adsorbed ligand or reductant (Table 1).
268
Under anoxic condition, iron was released as both Fe(III) (complexed by DFOB) and Fe(II)
269
(either as a free ion or complexed by DFOB with weaker affinity than to Fe(III)63). While
270
previous studies indicated that it is possible for Fe(II)-DFOB complexes to decompose with the
16
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 39
271
DFOB’s hydroxamate group reduced to amide group and Fe(II) oxidized to Fe(III),64, 65 any such
272
auto-decomposition either did not occur or was negligible in the time scale of the present
273
experiments.
274
The synergistic effect was particularly strong under oxic condition (Figure 5). Ascorbate
275
alone did not lead to any goethite dissolution, but its addition together with DFOB resulted in
276
higher dissolution rates than with DFOB alone. The rates under oxic conditions were still lower
277
than those under anoxic conditions. During the experiment, ascorbate was consumed by
278
oxidation at the iron oxide surface and possibly by a reaction with oxygen. However, considering
279
the high excess of ascorbate in experiments relative to observed net iron reduction under anoxic
280
conditions, this may not significantly affect surface concentrations of ascorbate, particularly at 1
281
mM total ascorbate concentrations where adsorption densities appear close to surface saturation
282
(Figure S3a of the Supporting Information). Indeed, near constant dissolution rates throughout
283
the experiments indicate that this effect is small enough to not influence dissolution (Figure 5a
284
and b). Non-linear dissolution was only observed at the lowest DFOB concentration (5 µM).
285
Here we see that the dissolution occurred so fast that it approached equilibrium before the end of
286
the experiment. Theoretically 5 µM DFOB could solubilize 5 µM Fe(III) assuming that there is
287
no adsorption of Fe-DFOB complexes, but the dissolution rate decreased drastically as the
17
ACS Paragon Plus Environment
Page 19 of 39
Environmental Science & Technology
288
system approached equilibrium and the concentration of uncomplexed free DFOB in solution
289
decreased (Eq. 1). Therefore, only the results from the first two samples were used for rate
290
calculations.
291
Experiments with varying concentrations of ascorbate and DFOB had two noteworthy trends.
292
First, for a fixed DFOB concentration, the synergistic effect in Fe mobilization increased with
293
increasing ascorbate concentration. Second, initial goethite dissolution rates in the presence of
294
both ascorbate and DFOB were independent of DFOB concentrations when the ascorbate
295
concentration was at a fixed value (Figure 5a to c). The second phenomenon was intriguing as it
296
suggests that the synergy may occur in the concentration ranges relevant in soils, where
297
reductants and siderophore are in submilli-molar and submicro-molar ranges, respectively.22, 66-68
298
Goethite Dissolution: Ascorbate + HBED. The interplay between ascorbate and HBED in
299
goethite dissolution was controlled by their interactive adsorption behavior. Under anoxic
300
conditions, the extent of goethite dissolution in the presence of both 1 mM ascorbate and 10 µM
301
HBED was marginally higher than the sum of the dissolution from the experiments with
302
reductant and ligand in isolation (Figure 6a). Taking into account the competitive adsorption
303
effect (i.e., ascorbate substantially displacing adsorbed HBED; Figure 1b) and using the
304
established surface-controlled dissolution rate model, the measured dissolution rate was a factor
18
ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 39
305
of two higher than the predicted one. This suggests an effect that is more than simply addition of
306
the two independent surface-controlled dissolution mechanisms. It may indicate that there is a
307
synergistic effect of ascorbate and HBED on the goethite dissolution rate coefficient, even
308
though the macroscopic effect is obscured by the influence of the competitive adsorption of
309
ascorbate and HBED on the net dissolution rate. Under oxic conditions, the addition of 1 mM
310
ascorbate even decreased the goethite dissolution rate compared with the rate for 10 µM HBED
311
only (Figure 6b). The adsorption data (Figure 1b) provide a reasonable explanation as the
312
reductant-ligand-surface interaction was dominated by the strong inhibitory effect of ascorbate
313
on HBED adsorption.
314
While the specific mechanisms of how adsorption of a reductant inhibits that of a ligand (or
315
vice versa) remain to be elucidated, the results from the ascorbate-HBED-goethite system
316
highlighted the possibility of intrinsic synergistic effects being obscured on a macroscopic scale
317
by adsorption interactions. In a slightly different but also environmentally relevant system
318
(oxalate-siderophore), Akafia et al. found that dissolution rates could increase, decrease, or
319
remain unchanged as compared to the rates in single-ligand systems, depending on the pH,
320
siderophore present, and the identity of the metal oxyhydroxide.62 Although no adsorption results
19
ACS Paragon Plus Environment
Page 21 of 39
Environmental Science & Technology
321
were presented in that study, the authors suggested a possible role of decreased ligand adsorption
322
in accounting for the diminishing synergistic effect under certain conditions.
323
Possible Mechanisms for the Synergistic Effect. Excluding the effect of enhanced
324
adsorption of the ligands or reductants, there are at least three possible mechanisms that may
325
contribute to the observed synergistic effect, although our results based on macroscopic
326
measurements of dissolution rates and adsorption equilibrium are inconclusive regarding the
327
specific mechanism. The first potential mechanism is that the generation of surface Fe(II)
328
labilizes the goethite structure. Although ascorbate cannot lead to goethite dissolution by itself
329
when O2 is present, it can still facilitate redox reactions generating surface Fe(II) (probably short
330
lived). Knowledge of surface Fe(II) and goethite interactions has recently been advanced by the
331
discovery of electron transfer and atom exchange (ETAE) between aqueous Fe(II) and the Fe(III)
332
(hydr)oxide surface via a mechanism involving simultaneous Fe(II) oxidative adsorption coupled
333
to Fe(III) reductive dissolution.69, 70 The active recrystallization process may kinetically labilize
334
the goethite structure for faster ligand-promoted dissolution. Such an increase of the dynamics of
335
the goethite lattice would be consistent with previous observations of increased rates of iron
336
isotope exchange in the presence of Fe(II).70 The second possible mechanism is that DFOB may
337
detach a fraction of the short lived surface Fe(II) at a rate that outcompetes the reoxidation of
20
ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 39
338
surface Fe(II) back to Fe(III). Borer et al. presented a similar mechanism for the synergy
339
between a siderophore and a photoreductant for goethite dissolution.51 Although the Fe(II) that
340
was detached from the surface into solution by complexation with DFOB would be rapidly
341
oxidized by O2 to form an Fe(III)-DFOB complex, reoxidation would have been delayed and
342
relocated from the surface to the aqueous phase. In the presence of excess and highly accessible
343
O2, there might only be small amounts of surface Fe(II) due to its short lifetime. Our
344
experimental results indicated that the surface Fe(II) sites might be saturated by DFOB at
345
concentrations as low as 5 µM. When provided with 0.1 or 1 mM ascorbate, increasing the
346
DFOB concentration from 5 to 20 µM only marginally increased the dissolution rates (Figure S6
347
of the Supporting Information). The third possible mechanism is related to the radical scavenging
348
capability of DFOB71. As a one-electron-transfer process, reduction of surface Fe(III) is
349
accompanied by the oxidation of ascorbate to a radical species, which can re-oxidize surface
350
Fe(II). DFOB is a known radical scavenger that has been demonstrated to inhibit the reoxidation
351
of Fe(II).52, 72 It is not clear if HBED and other ligands could affect the redox reaction between
352
ascorbate and iron surface sites by a similar mechanism. Overall, our current results are not able
353
to attribute the observed synergistic effect to any specific mechanism. Labeled stable isotope
354
experiments and molecular dynamic simulations may provide more insights into the process.
21
ACS Paragon Plus Environment
Page 23 of 39
Environmental Science & Technology
355
Environmental Implications. Knowledge of iron (hydr)oxide dissolution in the presence of
356
biogenic ligands is the foundation of understanding the natural strategies of iron acquisition by
357
microorganisms and plants as well as engineered interventions to improve bioavailability. Co-
358
exudation of reductants and ligands under iron limitation has been observed for both plants and
359
microorganisms. Our results demonstrate that reductants and ligands can act synergistically to
360
promote iron (hydr)oxide dissolution at soil pH values and concentrations that are relevant to
361
soils. This synergy may serve as an efficient biological iron acquisition strategy under oxic
362
conditions. Certain types of plants exude phenolics (e.g., caffeic acid) to mobilize Fe, many of
363
which are both ligands and reductants for Fe(III) with moderate potency.8, 73-75 The reductant-
364
ligand synergy may shed light on refined kinetic models for iron (hydr)oxide dissolution by
365
phenolics. Previous studies focusing on microbial growth proposed that iron reduction generated
366
a pool of iron that is readily bioavailable to aerobic microorganisms.33, 42, 45 The present results
367
further confirmed that iron reduction works in conjunction with ligands to increase the rate of
368
mobilization of Fe from poorly soluble iron sources.
369
While the reductant-ligand synergistic effect is supported by convincing evidence in abiotic
370
goethite model systems, translation of the results to real soil systems has to consider a number of
371
soil processes. Our HBED-ascorbate-goethite results illustrate that the synergistic effect may be
22
ACS Paragon Plus Environment
Environmental Science & Technology
Page 24 of 39
372
masked by competitive adsorption (e.g., by phosphate, sulfate, humic and fulvic acids), which is
373
modulated
374
iron(hydr)oxides commonly contain several mol% substituted aluminum.2 An inhibitory effect of
375
Al was observed for the reductive dissolution pathway77-79 but not for ligand promoted
376
dissolution26 by DFOB. The complexation of aluminum with ligands (e.g., siderophores) may
377
help Al detach from iron (hydr)oxide surfaces.80 Therefore, the ligand-reductant synergy may be
378
more complex when considering the impurities in natural iron oxides. Moreover, further research
379
is required to evaluate the role of the ligand-reductant synergy in biological iron uptake in the
380
presence of microorganisms and plants, whose releases of biogenic exudates are regulated by the
381
surrounding iron bioavailability.
by
solution
pH
and
electrostatic
interactions.39,
76
Naturally occurring
382 383
Acknowledgements 384
The authors thank Jeffrey Catalano for his valuable inputs throughout this study. Chao Pan,
385
Manvitha Marni, Vrajesh Mehta provided laboratory assistance. This research was supported by
386
the International Center for Advanced Renewable Energy and Sustainability (ICARES) at
387
Washington University in St. Louis. Comments and suggestions of three anonymous reviewers
388
and Associate Editor Timm Strathmann helped us improve an earlier version of this paper.
23
ACS Paragon Plus Environment
Page 25 of 39
Environmental Science & Technology
389 390
Supporting Information Available Additional information (including 2 tables, 6 figures and additional discussion about the
391 392
analytical methods) is available free of charge via the Internet at http://pubs.acs.org.
393 394
References
395
(1)
Stumm, W.; Morgan, J. J., Aquatic Chemistry. 3rd ed.; Wiley Interscience: New York, 1996.
396
(2)
Cornell, R. M.; Schwertmann, U., The Iron Oxides: Structure, Properties, Reactions,
397
Occurrences and Uses. Wiley: 2003.
398
(3)
399
dissolution of iron oxides by ascorbate: The role of carboxylate anions in accelerating reductive
400
dissolution. J. Colloid Interface Sci. 1990, 138 (1), 74-82.
401
(4)
402
mechanisms. Langmuir 1991, 7 (4), 809-813.
403
(5)
404
Langmuir 1997, 13 (6), 1835-1839.
405
(6)
406
goethite. Geochim. Cosmochim. Acta 2001, 65 (9), 1367-1379.
407
(7)
408
of hematite (α-Fe2O3) by ascorbate. Colloids Surf. 1989, 39 (2), 303-309.
409
(8)
410
Cosmochim. Acta 1989, 53 (5), 961-971.
411
(9)
412
Struct. Bonding (Berlin), Jørgensen, C. K.; Neilands, J. B.; Nyholm, R.; Reinen, D.; Williams, R. J. P.,
413
Eds. Springer Berlin Heidelberg: 1969; Vol. 6, pp 116-156.
414
(10)
Stumm, W.; Lee, G. F., Oxygenation of ferrous iron. Ind. Eng. Chem. 1961, 53 (2), 143-146.
415
(11)
Suzuki, Y.; Kuma, K.; Kudo, I.; Hasebe, K.; Matsunaga, K., Existence of stable Fe (II) complex
416
in oxic river water and its determination. Water Res. 1992, 26 (11), 1421-1424.
Dos Santos Afonso, M.; Morando, P. J.; Blesa, M. A.; Banwart, S.; Stumm, W., The reductive
Suter, D.; Banwart, S.; Stumm, W., Dissolution of hydrous iron (III) oxides by reductive Deng, Y., Effect of pH on the reductive dissolution rates of iron (III) hydroxide by ascorbate. Larsen, O.; Postma, D., Kinetics of reductive bulk dissolution of lepidocrocite, ferrihydrite, and Banwart, S.; Davies, S.; Stumm, W., The role of oxalate in accelerating the reductive dissolution LaKind, J. S.; Stone, A. T., Reductive dissolution of goethite by phenolic reductants. Geochim. Spiro, T. G.; Saltman, P., Polynuclear complexes of iron and their biological implications. In
24
ACS Paragon Plus Environment
Environmental Science & Technology
417
(12)
418
(EDDHA/Fe3+, EDDHMA/Fe3+ and the novel EDDHSA/Fe3+) to correct iron chlorosis. Eur. J. Argon.
419
2005, 22 (2), 119-130.
420
(13)
421
H., Performance of soil-applied FeEDDHA isomers in delivering Fe to soybean plants in relation to the
422
moment of application. J. Agric. Food Chem. 2010, 58 (24), 12833-12839.
423
(14)
424
biodegradability of EDDHA chelates under calcareous soil conditions. Geoderma 2012, 173, 282-288.
425
(15)
426
Considerations on the shuttle mechanism of FeEDDHA chelates at the soil-root interface in case of Fe
427
deficiency. Plant Soil 2014, 379 (1-2), 373-387.
428
(16)
429
parameters on the kinetics of the displacement of Fe from FeEDDHA chelates by Cu. J. Phys. Chem. A
430
2012, 116 (25), 6582-6589.
431
(17)
432
Proc. Natl. Acad. Sci. U. S. A. 2003, 100 (7), 3584-3588.
433
(18)
434
roots of grasses. Plant Physiol. 1986, 80 (1), 175-180.
435
(19)
436
iron acquisition by plants. Adv. Agron. 2006, 91, 1-46.
437
(20)
438
2009, 109 (10), 4580-4595.
439
(21)
440
rhizosphere: biochemistry and organic substances at the soil–plant interface, 2nd ed.; Pinton, R.;
441
Varanini, Z.; Nannipieri, P., Eds. CRC Press: Boca Raton, 2007; pp 173-200.
442
(22)
443
chelators in soils. Nature 1980, 287 (5785), 833-834.
444
(23)
445
2004, 66 (1), 3-18.
446
(24)
447
of siderophores and organic acids. Geochim. Cosmochim. Acta 2007, 71 (23), 5635-5650.
448
(25)
449
ligand-promoted dissolution of goethite [α-FeOOH (s)]. Geochim. Cosmochim. Acta 1996, 60 (22), 4403-
450
4416.
Page 26 of 39
Álvarez-Fernández, A.; Garcı́a-Marco, S.; Lucena, J. J., Evaluation of synthetic iron(III)-chelates
Schenkeveld, W. D.; Reichwein, A. M.; Bugter, M. H.; Temminghoff, E. J.; van Riemsdijk, W.
Schenkeveld, W.; Hoffland, E.; Reichwein, A.; Temminghoff, E.; Van Riemsdijk, W., The Schenkeveld, W. D.; Reichwein, A. M.; Temminghoff, E. J.; van Riemsdijk, W. H.,
Schenkeveld, W. D.; Reichwein, A. M.; Temminghoff, E. J.; van Riemsdijk, W. H., Effect of soil
Raymond, K. N.; Dertz, E. A.; Kim, S. S., Enterobactin: An archetype for microbial iron transport. Römheld, V.; Marschner, H., Evidence for a specific uptake system for iron phytosiderophores in Kraemer, S.; Crowley, D.; Kretzschmar, R., Geochemical aspects of phytosiderophore-promoted Sandy, M.; Butler, A., Microbial iron acquisition: marine and terrestrial siderophores. Chem. Rev. Crowley, D. E.; Kraemer, S. M., Function of siderophores in the plant rhizosphere. In The
Powell, P.; Cline, G.; Reid, C.; Szaniszlo, P., Occurrence of hydroxamate siderophore iron Kraemer, S. M., Iron oxide dissolution and solubility in the presence of siderophores. Aqua. Sci. Reichard, P.; Kretzschmar, R.; Kraemer, S., Dissolution mechanisms of goethite in the presence Holmén, B. A.; Casey, W. H., Hydroxamate ligands, surface chemistry, and the mechanism of
25
ACS Paragon Plus Environment
Page 27 of 39
Environmental Science & Technology
451
(26)
Cervini-Silva, J.; Sposito, G., Steady-state dissolution kinetics of aluminum-goethite in the
452
presence of desferrioxamine-B and oxalate ligands. Environ. Sci. Technol. 2002, 36 (3), 337-342.
453
(27)
454
study of the adsorption of desferrioxamine B and aerobactin to the surface of lepidocrocite (γ-FeOOH).
455
Geochim. Cosmochim. Acta 2009, 73 (16), 4661-4672.
456
(28)
457
on α-FeOOH particles. J. Phys. Chem. C 2011, 115 (43), 21191-21198.
458
(29)
459
Al2O3 and BeO. Geochim. Cosmochim. Acta 1986, 50 (9), 1847-1860.
460
(30)
461
controlled dissolution of oxide phases. Geochim. Cosmochim. Acta 1997, 61 (14), 2855-2866.
462
(31)
463
S. L.; Kubicki, J. D.; White, A. F., Eds. Springer New York: 2008; pp 151-210.
464
(32)
465
Geochim. Cosmochim. Acta 2008, 72 (1), 99-116.
466
(33)
467
ascorbate in mobilization of iron from hematite by the aerobic bacterium Pseudomonas mendocina. Appl.
468
Environ. Microbiol. 2010, 76 (7), 2041-2048.
469
(34)
470
goethite in the presence of desferrioxamine B and oxalate ligands: implications for the microbial
471
acquisition of iron. Chem. Geol. 2003, 198 (1), 63-75.
472
(35)
473
enhance siderophore-promoted dissolution of goethite. Environ. Sci. Technol. 2007, 41 (10), 3633-3638.
474
(36)
475
Catalysis, Grassian, V. H., Ed. CRC Press: Boca Raton, 2005; pp 61-81.
476
(37)
477
phytosiderophores: rates, mechanisms, and the synergistic effect of oxalate. Plant Soil 2005, 276 (1-2),
478
115-132.
479
(38)
480
phosphate in the presence of desferrioxamine-B and oxalate ligands at pH = 4 - 6 and T = 24 ± 0.6 °C.
481
Chem. Geol. 2012, 320–321 (0), 1-8.
482
(39)
483
the presence of desferrioxamine B and Suwannee River fulvic acid at pH 6.5. Geochim. Cosmochim. Acta
484
2013, 115, 1-14.
Borer, P.; Hug, S. J.; Sulzberger, B.; Kraemer, S. M.; Kretzschmar, R., ATR-FTIR spectroscopic
Simanova, A. A.; Loring, J. S.; Persson, P., Formation of ternary metal-oxalate surface complexes Furrer, G.; Stumm, W., The coordination chemistry of weathering: I. Dissolution kinetics of δKraemer, S. M.; Hering, J. G., Influence of solution saturation state on the kinetics of ligandBrantley, S. L., Kinetics of mineral dissolution. In Kinetics of Water-Rock Interaction, Brantley, Yang, L.; Steefel, C. I., Kaolinite dissolution and precipitation kinetics at 22 °C and pH 4. Dehner, C. A.; Awaya, J. D.; Maurice, P. A.; DuBois, J. L., Roles of siderophores, oxalate, and
Cheah, S.-F.; Kraemer, S. M.; Cervini-Silva, J.; Sposito, G., Steady-state dissolution kinetics of
Carrasco, N.; Kretzschmar, R.; Pesch, M.-L.; Kraemer, S. M., Low concentrations of surfactants Martin, S. T., Precipitation and dissolution of iron and manganese oxides. In Environmental Reichard, P.; Kraemer, S.; Frazier, S.; Kretzschmar, R., Goethite dissolution in the presence of
Cervini-Silva, J.; Kearns, J.; Banfield, J., Steady-state dissolution kinetics of mineral ferric
Stewart, A. G.; Hudson-Edwards, K. A.; Dubbin, W. E., Mechanisms of goethite dissolution in
26
ACS Paragon Plus Environment
Environmental Science & Technology
485
(40)
Reichard, P. U.; Kretzschmar, R.; Kraemer, S. M., Rate laws of steady-state and non-steady-state
486
ligand-controlled dissolution of goethite. Colloids Surf., A 2007, 306 (1), 22-28.
487
(41)
488
readsorption process. Langmuir 2008, 24 (14), 7054-7057.
489
(42)
490
reduction by Pseudomonas mendocina in response to iron deprivation. Geomicrobiol. J. 2000, 17 (4),
491
261-273.
492
(43)
493
Pseudomonas sp. Chem. Geol. 1996, 132 (1–4), 25-31.
494
(44)
495
chelating properties of pyridine-2, 6-bis (thiocarboxylic acid) produced by Pseudomonas spp. and the
496
biological activities of the formed complexes. BioMetals 2002, 15 (2), 103-120.
497
(45)
498
mobilization from ferrihydrite by the Pseudomonas mendocina ymp strain. Appl. Environ. Microbiol.
499
2007, 73 (10), 3428-3430.
500
(46)
501
L., Ferritin and ferrihydrite nanoparticles as iron sources for Pseudomonas aeruginosa. J. Biol. Inorg.
502
Chem. 2013, 18 (3), 371-381.
503
(47)
504
hematite (α-Fe2O3) nanoparticles to a common aerobic bacterium. Environ. Sci. Technol. 2010, 45 (3),
505
977-983.
506
(48)
507
Fe-bearing montmorillonite clay. Geochim. Cosmochim. Acta 2013, 117, 191-202.
508
(49)
509
nanoparticles: Effects of crystalline order, siderophores, and alginate. Environ. Sci. Technol. 2014.
510
(50)
511
bis(monothiocarboxylic acid) and hydrolysis products. Geochim. Cosmochim. Acta 2007, 71 (23), 5651-
512
5660.
513
(51)
514
induced dissolution of colloidal iron(III) (hydr)oxides. Mar. Chem. 2005, 93 (2–4), 179-193.
515
(52)
516
of iron(III) (hydr)oxides in the absence and presence of organic ligands: Experimental studies and kinetic
517
modeling. Environ. Sci. Technol. 2009, 43 (6), 1864-1870.
Loring, J. S.; Simanova, A. A.; Persson, P., Highly mobile iron pool from a dissolution− Hersman, L. E.; Huang, A.; Maurice, P. A.; Forsythe, J. H., Siderophore production and iron
Hersman, L.; Maurice, P.; Sposito, G., Iron acquisition from hydrous Fe(III)-oxides by an aerobic Cortese, M. S.; Paszczynski, A.; Lewis, T. A.; Sebat, J. L.; Borek, V.; Crawford, R. L., Metal
Dhungana, S.; Anthony, C. R.; Hersman, L. E., Effect of exogenous reductant on growth and iron
Dehner, C.; Morales-Soto, N.; Behera, R. K.; Shrout, J.; Theil, E. C.; Maurice, P. A.; Dubois, J.
Dehner, C. A.; Barton, L.; Maurice, P. A.; DuBois, J. L., Size-dependent bioavailability of
Kuhn, K. M.; DuBois, J. L.; Maurice, P. A., Strategies of aerobic microbial Fe acquisition from Kuhn, K.; DuBois, J. L.; Maurice, P. A., Aerobic microbial Fe acquisition from ferrihydrite Dhungana, S.; Anthony III, C. R.; Hersman, L. E., Ferrihydrite dissolution by pyridine-2,6-
Borer, P. M.; Sulzberger, B.; Reichard, P.; Kraemer, S. M., Effect of siderophores on the lightBorer, P.; Sulzberger, B.; Hug, S. J.; Kraemer, S. M.; Kretzschmar, R., Photoreductive dissolution
27
ACS Paragon Plus Environment
Page 28 of 39
Page 29 of 39
Environmental Science & Technology
518
(53)
Borer, P.; Sulzberger, B.; Hug, S. J.; Kraemer, S. M.; Kretzschmar, R., Wavelength-dependence
519
of photoreductive dissolution of lepidocrocite (γ-FeOOH) in the absence and presence of the siderophore
520
DFOB. Environ. Sci. Technol. 2009, 43 (6), 1871-1876.
521
(54)
522
hydroxamate siderophores on Fe release and Pb(II) adsorption by goethite. Geochim. Cosmochim. Acta
523
1999, 63 (19–20), 3003-3008.
524
(55)
525
Humic Substance in the Global Environment and Implications on Human Health, Senesi, N.; Miano, T.
526
M., Eds. Elsevier: Amsterdam, 1994; pp 1183-1188.
527
(56)
528
G., Temperature dependence of goethite dissolution promoted by trihydroxamate siderophores. Geochim.
529
Cosmochim. Acta 2002, 66 (3), 431-438.
530
(57)
531
bearing minerals in marine systems. Rev. Mineral. Geochem. 2005, 59 (1), 53-84.
532
(58)
Sposito, G., Chemical Equilibria and Kinetics in Soils. Oxford University Press: 1994.
533
(59)
Lasaga, A. C., Kinetic Theory in the Earth Sciences. Princeton University Press: 1998.
534
(60)
Schecher, W. D.; McAvoy, D. C. MINEQL+: A chemical equilibrium modeling system, version
535
4.6, Environmental Research Software: Hallowell, ME, 2007.
536
(61)
537
(hydr)oxides and their dissolution at pH 3.0 to 6.0. Soil Sci. Soc. Am. J. 2008, 72 (6), 1557-1562.
538
(62)
539
dissolution as promoted by structurally diverse siderophores and oxalate. Geochim. Cosmochim. Acta
540
2014, 141 (0), 258-269.
541
(63)
542
contaminants. In Aquatic Redox Chemistry, Tratnyek, P. G.; Grundl, T. J.; Haderlein, S. B., Eds.
543
American Chemical Society: Washington DC, 2011; Vol. 1071, pp 283-313.
544
(64)
545
(desferrioxamine B) complexes with flavin mononucleotide in the absence of strong iron (II) chelators.
546
Geochim. Cosmochim. Acta 2010, 74 (5), 1513-1529.
547
(65)
548
acids: oxidation of iron(II) to iron(III) by desferrioxamine B under anaerobic conditions. J. Inorg.
549
Biochem. 2001, 83 (2–3), 107-114.
550
(66)
551
soil solution. Biogeochemistry 2004, 71 (2), 247-258.
Kraemer, S. M.; Cheah, S.-F.; Zapf, R.; Xu, J.; Raymond, K. N.; Sposito, G., Effect of
Solinas, V., Cation effects on the adsorption of desferrioxamine B (DFOB) by humic acid. In
Cocozza, C.; Tsao, C. C.; Cheah, S.-F.; Kraemer, S. M.; Raymond, K. N.; Miano, T. M.; Sposito,
Kraemer, S. M.; Butler, A.; Borer, P.; Cervini-Silva, J., Siderophores and the dissolution of iron-
Cervini-Silva, J., Adsorption of trihydroxamate and catecholate siderophores on α-iron Akafia, M. M.; Harrington, J. M.; Bargar, J. R.; Duckworth, O. W., Metal oxyhydroxide
Strathmann, T. J., Redox reactivity of organically complexed iron(II) species with aquatic
Kim, D.; Duckworth, O. W.; Strathmann, T. J., Reactions of aqueous iron–DFOB
Farkas, E.; Enyedy, É. A.; Zékány, L.; Deák, G., Interaction between iron(II) and hydroxamic
Holmström, S. J. M.; Lundström, U. S.; Finlay, R. D.; Hees, P. A. W. V., Siderophores in forest
28
ACS Paragon Plus Environment
Environmental Science & Technology
Page 30 of 39
552
(67)
Harter, R. D.; Naidu, R., Role of metal-organic complexation in metal sorption by soils. Adv.
553
Agron. 1995, 55, 219-263.
554
(68)
555
environments. In Chemistry of Aquatic Systems: Local and Global Perspectives, Bidoglio, G.; Stumm, W.,
556
Eds. Springer Netherlands: 1994; Vol. 5, pp 337-374.
557
(69)
558
conduction. Science 2008, 320 (5873), 218-222.
559
(70)
560
Fe (II) and goethite: An Fe isotope tracer study. Environ. Sci. Technol. 2009, 43 (4), 1102-1107.
561
(71)
562
nitroxide free radical. FEBS Lett. 1987, 222 (2), 246-250.
563
(72)
564
lepidocrocite (γ-FeOOH) in the presence of desferrioxamine B and aerobactin. Geochim. Cosmochim.
565
Acta 2009, 73 (16), 4673-4687.
566
(73)
567
caffeic acid, chlorogenic acid, sinapic acid, ferulic acid and naringin. J. Inorg. Biochem. 2004, 98 (8),
568
1457-1464.
569
(74)
570
iron(III) reduction by caffeic acid. Soil Sci. Soc. Am. J. 1995, 59 (5), 1301-1307.
571
(75)
572
model for rhizosphere processes. J. Plant Nutr. 1989, 12 (5), 581-592.
573
(76)
574
and sulfate on iron oxide surfaces. Geochim. Cosmochim. Acta 2015, 158, 130-146.
575
(77)
576
reductive dissolution of Al-substituted goethites by Clostridium butyricum. Soil Biol. Biochem. 2002, 34
577
(8), 1147-1155.
578
(78)
579
substitution on microbial reduction of Fe(III) (hydr)oxides. Geochim. Cosmochim. Acta 2010, 74 (24),
580
7086-7099.
581
(79)
582
hematite in dithionite. Clay Miner. 1987, 22 (3), 329-337.
583
(80)
584
negatively charged oxygen donor groups. Complexation of iron (III), gallium (III), indium (III),
585
aluminum (III), and other highly charged metal ions. Inorg. Chem. 1989, 28 (11), 2189-2195.
Stone, A.; Godtfredsen, K.; Deng, B., Sources and reactivity of reductants encountered in aquatic
Yanina, S. V.; Rosso, K. M., Linked reactivity at mineral-water interfaces through bulk crystal Handler, R. M.; Beard, B. L.; Johnson, C. M.; Scherer, M. M., Atom exchange between aqueous Morehouse, K. M.; Flitter, W. D.; Mason, R. P., The enzymatic oxidation of Desferal to a Borer, P.; Kraemer, S. M.; Sulzberger, B.; Hug, S. J.; Kretzschmar, R., Photodissolution of
Hynes, M. J.; O'Coinceanainn, M., The kinetics and mechanisms of reactions of iron(III) with
Deiana, S.; Gessa, C.; Marchetti, M.; Usai, M., Phenolic-acid redox properties - pH influence on Boyer, R. F.; Clark, H. M.; Sanchez, S., Solubilization of ferrihydrite iron by plant phenolics - a Hinkle, M. A. G.; Wang, Z.; Giammar, D. E.; Catalano, J. G., Interaction of Fe(II) with phosphate Dominik, P.; Pohl, H. N.; Bousserrhine, N.; Berthelin, J.; Kaupenjohann, M., Limitations to the
Ekstrom, E. B.; Learman, D. R.; Madden, A. S.; Hansel, C. M., Contrasting effects of Al
Torrent, J.; Schwertmann, U.; Barron, V., The reductive dissolution of synthetic goethite and Evers, A.; Hancock, R. D.; Martell, A. E.; Motekaitis, R. J., Metal ion recognition in ligands with
29
ACS Paragon Plus Environment
Page 31 of 39
Environmental Science & Technology
586 587
30
ACS Paragon Plus Environment
Environmental Science & Technology
Page 32 of 39
Table 1. Summary of the batch experiments for goethite dissolution at pH 6. (Part a) Reductant/Ligand alone Exp.IDa
Ascorbate (mM)
HBED (µM)
DFOB (µM)
PO2 (bar)
Dissolution rate (nmol·g-1h-1)
End time of regression (h)b
1 2 3 4 5 6 7 8 9 10 11 12
0 1 0.1 0 1 0.1 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 10 10
0 0 0 0 0 0 10 20 10 5 0 0