Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES
Characterization of Natural and Affected Environments
Linking Thermodynamics to Pollutant Reduction Kinetics by Fe Bound to Iron Oxides 2+
Sydney M. Stewart, Thomas B. Hofstetter, Prachi Joshi, and Christopher A. Gorski Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 29 Mar 2018 Downloaded from http://pubs.acs.org on March 29, 2018
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 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 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.
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 35
Environmental Science & Technology
Linking Thermodynamics to Pollutant Reduction Kinetics by Fe2+ Bound to Iron Oxides Sydney M. Stewart,† Thomas B. Hofstetter,‡ Prachi Joshi,† and Christopher A. Gorski∗,† 1
†Department of Civil & Environmental Engineering, Pennsylvania State University, 212 Sackett Building, University Park, PA 16802 ‡Eawag, Swiss Federal Institute of Aquatic Science and Technology, CH-8600 D¨ ubendorf, Switzerland E-mail:
[email protected] Phone: 814-865-5673. Fax: 814-863-7304
2
Abstract
3
Numerous studies have reported that pollutant reduction rates by ferrous iron (Fe2+ )
4
are substantially enhanced in the presence of an iron (oxyhydr)oxide mineral. Devel-
5
oping a thermodynamic framework to explain this phenomenon has been historically
6
difficult due to challenges in quantifying reduction potential (EH ) values for oxide-
7
bound Fe2+ species. Recently, our group demonstrated that EH values for hematite-
8
and goethite-bound Fe2+ can be accurately calculated using Gibbs free energy of for-
9
mation values. Here, we tested if calculated EH values for oxide-bound Fe2+ could be
10
used to develop a free energy relationship capable of describing variations in reduction
11
rate constants of substituted nitrobenzenes, a class of model pollutants that contain
12
reducible aromatic nitro groups, using data collected here and compiled from the lit-
13
erature. All the data could be described by a single linear relationship between the
1
ACS Paragon Plus Environment
Environmental Science & Technology
14
logarithms of the surface-area-normalized rate constant (kSA ) values and EH and pH
15
values [log(kSA ) = −EH /0.059V − pH + 3.42]. This framework provides mechanistic
16
insights into how the thermodynamic favorability of electron transfer from oxide-bound
17
Fe2+ relates to redox reaction kinetics.
18
Introduction
19
Ferrous iron (Fe2+ ) bound to iron (oxyhydr)oxides reduces oxidized pollutants, 1–14 molecular
20
oxygen, 15–18 and nitrite 19 far more quickly than aqueous Fe2+ alone. Early work studying
21
this phenomenon proposed that Fe2+ forms surface complexes on iron oxide surfaces that
22
have higher electron densities than aqueous Fe2+ , due to surface hydroxyl ligands serving as
23
electron donors. 20,21 Consequently, oxide-bound Fe2+ was thought to have a more negative
24
reduction potential (EH ) value than aqueous Fe2+ in the same solution. 20,22–24 More recently,
25
however, two lines of reasoning have questioned the validity of this explanation. First,
26
spectroscopic and microscopic analyses demonstrated that an Fe2+ atom rarely forms a
27
stable adsorbed species on iron oxides, but rather it becomes oxidized to an Fe3+ atom
28
and donates an electron into the bulk semiconducting iron oxide lattice. 24–31 Second, the
29
explanation implicitly assumes that relevant solutions must be in a state of thermodynamic
30
disequilibrium, as all redox couples in a system at equilibrium have the same EH value. Using
31
electrochemical measurements, our group recently confirmed that suspensions containing
32
aqueous Fe2+ and goethite or hematite do reach thermodynamic equilibrium, 32 indicating
33
that oxide-bound Fe2+ and aqueous Fe2+ reached the same EH value.
34
An alternative hypothesis to explain this phenomenon is that the presence of an iron oxide
35
alters the Fe2+ oxidation product that forms and therefore the Gibbs free energy of the reac-
36
tion. 32–35 When aqueous Fe2+ oxidizes in the absence of an oxide, it produces an aqueous Fe3+
37
complex or a metastable iron oxide (e.g., ferrihydrite). 36 In contrast, when Fe2+ oxidizes in
38
the presence of a crystalline iron oxide surface, it commonly produces a thermodynamically-
39
stable iron oxide (e.g., goethite or hematite) via epitaxial growth. 15,37,38 Thermodynamically2
ACS Paragon Plus Environment
Page 2 of 35
Page 3 of 35
Environmental Science & Technology
40
stable iron oxides (e.g., goethite and hematite) have more negative Gibbs free energies of
41
formation than metastable iron oxides (e.g., ferrihydrite), and therefore the iron oxide is
42
thought to make Fe2+ oxidation and concomitant epitaxial oxide growth more exergonic
43
(i.e., thermodynamically favorable).
44
This hypothesis is consistent with characterizations of oxides made before and after pollu-
45
tant reduction reactions, 33,34,38 but it remains unclear if it can be invoked to quantitatively
46
explain trends in pollutant reduction kinetics. In theory, EH values of oxide-bound Fe2+
47
should relate to observed reaction rate constant (kobs ) values in a predictable manner if an
48
electron transfer event occurs during or before the rate-limiting step of the reaction. 39–42
49
The goal of this study was to test the validity of this argument for the reduction of a model
50
oxidized pollutant by oxide-bound Fe2+ . We characterized the free energy relationship that
51
exists between reaction rate constant values for nitrobenzene reduction as a function of the
52
EH values of oxide-bound Fe2+ and pH. We subsequently tested if the free energy relation-
53
ship could describe data compiled from the literature collected with other iron oxides and
54
substituted nitrobenzenes.
55
We focused on substituted nitrobenzenes for three reasons. First, there are many studies
56
in which substituted nitrobenzenes have been reacted with oxide-bound Fe2+ , which provided
57
us with relevant data to reinterpret. 3,5,6,8,9,43 Second, the elementary steps in the reduction
58
reaction of nitrobenzene to aniline have been extensively characterized, 44–48 which allowed us
59
to interpret kobs values in the context of elementary reaction steps. Third, nitrobenzene does
60
not appreciably adsorb to iron oxides, which means that disappearance rate of nitrobenzene
61
can be assumed to be equal to the transformation rate. 3
62
Materials and Methods
63
All experiments were conducted under anaerobic conditions in a glovebox (MBraun Unilab
64
Workstation, 100% N2 , O2 < 0.1 ppm). All lab supplies that came in contact with Fe2+ were
3
ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 35
65
left under vacuum for at least 12 hours in the exchange chamber and allowed to equilibrate
66
with the N2 atmosphere before use. Deionized water (Millipore Milli-Q system, resistivity
67
>18 MΩ·cm) was used to make all aqueous solutions. All liquid solutions were purged
68
with N2 (99.99%) for at least three hours before being cycled into the glovebox, except
69
for nitrobenzene. Approximately 2 mL of reagent grade nitrobenzene (Sigma-Aldrich) was
70
cycled through the vacuum chamber in a small sealed container with little headspace and
71
was used to make 5 mM and 25 mM stock solutions in deoxygenated methanol. Ferrous
72
chloride (FeCl2 , Acros, anhydrous, 99%) from a sealed ampule was used to make Fe2+ stock
73
solutions that were kept acidic using HCl. All measurements of aqueous Fe2+ were done
74
using the 1,10-phenanthroline method. 49
75
All experiments were done in aqueous solutions containing 25 mM KCl (EM Science)
76
and 25 mM pH buffer. The pH buffer used at pH values 6.0 and 6.5 was MES (2-N-
77
morpholinoethanesulfonic acid acid monohydrate, pKA 6.1, Aresco, ≥99%), and the pH
78
buffer used at pH 7.0 was MOPS (3-N-morpholinopropanesulfonic acid, pKA 7.2, EMD
79
Chemicals Inc., 100%). Note that past works have argued that the buffers used here may
80
affect iron oxide–aqueous Fe2+ interactions. 50,51 This possibility was unfortunately unavoid-
81
able, as we needed to maintain a constant pH in the experiments. While carbonate has
82
been recommended as an alternative buffer, 51 our interpretation of published data (below)
83
suggests that carbonate dramatically inhibits nitrobenzene reduction rates by oxide-bound
84
Fe2+ .
85
Goethite Synthesis and Characterization
86
An established protocol was used to synthesize two batches of micron-scale goethite. 52 We
87
synthesized two batches of goethite because the first batch was exhausted prior to completing
88
all experiments. After synthesis, the goethite was freeze-dried (Labconco benchtop freeze-
89
drier), ground with a mortar and pestle, and sifted through a 200 mesh sieve (Humboldt).
90
The purities of the goethite batches were confirmed using cryogenic 4
ACS Paragon Plus Environment
57
Fe Mo¨ssbauer spec-
Page 5 of 35
Environmental Science & Technology
91
troscopy. The BET surface area of the first goethite batch was 30.9±0.3 m2 /g (ASAP 2020
92
Automated Surface Area and Porosimetry System). The BET surface area of the second
93
batch was assumed to be equal to that of the first based on previous experience with this
94
synthesis procedure. Unfortunately, we exhausted the second batch of goethite before the
95
BET surface area was measured. Based on our experience, this synthesis method reliably
96
produces goethite with a specific surface area of ≈ 30 m2 /g.
97
Reduction Potential Measurements
98
We used mediated potentiometry to determine standard reduction potential (EH0 ) values for
99
each batch of goethite using a previously published protocol. 32 We prepared triplicate re-
100
actors with varied initial Fe2+ concentrations (100, 200, 400, 800, 1000, 1500 or 3000 µM)
101
and pH (6.0 or 7.0) values. After the addition of aqueous Fe2+ , the pH was readjusted to
102
the initial value using 1 N KOH and/or HCl stock solutions because the slight changes on
103
the order of approximately 0.1 pH units significantly affected measured and calculated EH
104
values. After measuring the initial Fe2+ concentration, we added goethite to the solutions
105
to achieve a mass loading of 1 or 2 g/L. Reactors were covered with aluminum foil to avoid
106
inadvertent photochemical reactions and mixed on a magnetic stir plate. After one hour of
107
mixing, a mediator compound (cyanomethyl viologen, EH0 = –0.14 V) 53,54 was added from a
108
10 mM stock solution to reach a final concentration between 10 and 20 µM. The reactors were
109
mixed for 24 hours, then a final aqueous Fe2+ measurement was made by filtering the solution
110
through a 0.45 µm nylon syringe filter. We then performed a potentiometric measurement
111
of the solution at two second intervals for one hour with a Pt ring combined redox electrode
112
(Metrohm) inside the glovebox. The EH value reported is the lowest value recorded during
113
the one hour measurement. Between measurements, the electrode was stored in 3 M KCl for
114
≥3 hours to ensure the internal KCl solution was not diluted. The electrode was cleaned pe-
115
riodically with 0.1 M HCl, and calibrated approximately monthly using quinhydrone buffers
116
prepared at pH 4.0 and 7.0 (Tokyo Chemical Industry Co., LTD). All measurements were 5
ACS Paragon Plus Environment
Environmental Science & Technology
117
made in reference to Ag/AgCl, and were subsequently converted to values in reference to
118
the standard hydrogen electrode (SHE) based on the calibrations. To convert from con-
119
centrations to activities, activity coefficients for Fe2+ were calculated according to reactor
120
conditions using Visual MINTEQ, and values ranged from 0.48 to 0.52.
121
Nitrobenzene Reduction Experiments
122
We set up reactors containing aqueous Fe2+ and goethite using the same approach as de-
123
scribed for the reduction potential measurements, except that no mediator was added. In
124
the reactors, we varied the total initial Fe2+ concentration, goethite loading, and pH value.
125
After an equilibration period of 24 hours in which the goethite and Fe2+ reacted, a sample
126
was taken to measure the final aqueous Fe2+ concentration. Next, the reaction was started
127
by adding an aliquot of nitrobenzene stock solution to the reactor using a glass syringe
128
to reach an initial concentration of approximately 10 µM. The reactor was sealed with a
129
Teflon-lined septa and aluminum crimp cap. Samples were taken at pre-determined time
130
points using a needle and syringe and filtered through a 0.2 µM PTFE syringe filter into
131
amber sample vials with PTFE-lined screw caps. Nitrobenzene and aniline were measured
132
using high-pressure liquid chromatography (HPLC) with a Supelcosil LC-18 column using
133
established methods. 32
134
Derivation of Free Energy Relationship
135
We derived a rate expression to relate the disappearance of nitrobenzene to changes of re-
136
duction potential and pH using a quasi-equilibrium assumption, which assumes that the
137
reaction steps preceding the rate-limiting step reach equilibrium. 40,42,55 The reduction path-
138
way of nitrobenzene is shown in Scheme 1, with the relevant iron oxidation reactions shown
139
in Figure S1. The overall rate of nitrobenzene (compound 1 in Scheme 1) disappearance
6
ACS Paragon Plus Environment
Page 6 of 35
Page 7 of 35
Environmental Science & Technology
NO2
NO
NH2
NHOH
+2e- +2H+
+2e- +2H+
+2e- +2H+
–H2O
–H2O
1 O
N
6 O
O
N
O
O
N
OH
O
+e–
+H+
+e–
–e–
–H+
–e–
1
3
2
N
OH
4
HO +H+ –H+
N
OH
O
N
– H2O 5
6
Scheme 1: Reaction scheme of nitrobenzene reduction. A. Overall steps of the reduction of nitrobenzene (1) to aniline. B. Elementary reaction steps of the reduction of nitrobenzene reduction to nitrosobenzene (6) that occurs through a series of electron and proton transfer reactions and ultimately a dehydration of N,N -dihydroxyaniline (5) to 6. 45,46 140
during a surface-catalyzed pseudo-first-order reduction reaction can be described by: −kobs · [1] d[1] = = −kSA · [1] dt A
141 142
(1)
143
1 ); A is the reactive where kobs is the observed pseudo-first-order reduction rate constant ( hr
144
surface area of an iron oxide ( mL ), which is calculated from the measured BET specific surface
145
area; and kSA is the surface-area-normalized reaction rate constant ( m2L·hr ).
2
146
Studies on the
15
N kinetic isotope effects 45–47 have indicated that the rate-limiting ele-
147
mentary step of the reduction of nitrobenzene (or substituted nitrobenzene) is the hydrolysis
148
of N,N -dihydroxyaniline (5) to nitrosobenzene (6) (5 → 6 in Scheme 1) when the reductant
149
is present in excess. If this is the case, the overall rate of nitrobenzene (1) disappearance
150
would depend on the rate constant for reaction 5 → 6, k5→6 , and the concentration of
151
N,N -dihydroxyaniline (5):
152 153
d[1] = −k5→6 · [5] dt
7
ACS Paragon Plus Environment
(2)
Environmental Science & Technology
154
Page 8 of 35
Setting equations 1 and 2 equal to each other relates kSA to k5→6 :
kSA = k5→6 ·
155 156
[5] [1]
(3)
157
For this relationship to have practical utility, the concentration of 5 must be expressed as
158
a function of the concentration of 1. Using the quasi-equilibrium method, the concentration
159
of 5 could be mathematically derived based on the concentrations of species 1 to 4 using
160
equilibrium expressions for reactions sequence 1 * )2* )3* )4* ) 5 in Scheme 1 and the
161
oxidation of Fe2+ to goethite (KFe2+ * )αFeOOH in eq. 4e derived from the half reaction shown
162
in eq. 12 below): [2] [1]
(4a)
163
K1* )2 =
164
K2* )3 =
(4b)
165
K3* )4
(4c)
166
K4* )5
167
KFe2+ * )αFeOOH
168
[3] [2] · [H+ ] [4] = [2] [5] = [4] · [H+ ] [H+ ]3 = [Fe2+ ]
(4d) (4e)
169
Through a sequence of substitutions for transient species 1 to 4 for 5 and the assumption
170
that electrons come from the oxidation of Fe2+ to α−FeOOH, eq. 3 can be transformed to
8
ACS Paragon Plus Environment
Page 9 of 35
171
172
Environmental Science & Technology
an interpretable relationship:
kSA · [1] = k5→6 · [5]
(5a)
173
+ = k5→6 · K4* )5 · [4] · [H ]
174
= k5→6 · K4* ) 5 · K3* )4 · KFe2+ * )αFeOOH · [3] ·
175
= k5→6 · K4* ) 5 · K3* )4 · KFe2+ * )3 · [2] · )αFeOOH · K2*
178
[Fe2+ ] [H+ ]2
(5c) [Fe2+ ] [H+ ]
[Fe2+ ]2 [H+ ]4 [Fe2+ ]2 2 = k5→6 · K4* ) · K · K · * * 2 3 1 2 ) 5 · K3* )4 · (KFe2+ * ) ) αFeOOH ) [H+ ]4 2 = k5→6 · K4* ) 5 · K3* )4 · (KFe2+ * ) 3 · K1* )2 · [1] · )αFeOOH ) · K2*
176
177
(5b)
kSA
(5d) (5e) (5f)
179
The observable reaction rate constant for nitrobenzene reduction will therefore depend on
180
the pH and Fe2+ concentration as well as the equilibrium constants for reactions 1 * ) 2, 2
181
* ) 3, 3 * ) 4, and 4 * ) 5.
182
While the electron transfer reactions involving the oxidation of Fe2+ to an iron oxide
183
(in this case, goethite, α−FeOOH) can be expressed in terms of an equilibrium constant
184
(KFe2+ * )αFeOOH ), representing them in terms of EH values is more familiar and practical in
185
the context of redox reactions. The Nernst equation for the goethite-Fe2+ (aq) half reaction can
186
be written in the following form (see eqs. 12 and 13 below for more details):
EH = EH0 −
187 188
RT [Fe2+ ] ln + 3 F [H ]
(6)
189
The Nernst equation can be re-expressed in terms of KFe2+ * )αFeOOH , which is inverted because
190
the equilibrium constant is for the oxidation reaction:
191
192 193
RT 1 = ln F K 2+ * Fe )αFeOOH RT [Fe2+ ] RT 1 EH = ln − ln + 3 F KFe2+ * F [H ] )αFeOOH EH0
9
ACS Paragon Plus Environment
(7) (8)
Environmental Science & Technology
194
Page 10 of 35
Solving for KFe2+ * )αFeOOH yields:
KFe2+ * )αFeOOH
195 196
F [H+ ]3 = exp − · EH · RT [Fe2+ ]
(9)
197
Substituting equation 9 into equation 5f yields a relationship in which kSA can be related to
198
EH : 2 F [Fe2+ ]2 [H+ ]3 = k5→6 · K4* exp − · K · E · K · · K · * * * 3)4 H 2)3 1)2 )5 RT [H+ ]4 [Fe2+ ] 2F = k5→6 · K4* · EH · [H+ ]2 ) 5 · K3* )4 · K2* )3 · K1* )2 · exp − RT
199
kSA
200
kSA
201
202
log (kSA ) = log (k5→6 · K4* · K2* )5 · K3* ) 3 · K1* )2 ) − {z )4 } |
2F · EH 2.303RT
− 2 · pH
(10c)
d
204
205
(10b)
Taking the log of both sides of the equation yields:
203
(10a)
which at room temperature (25◦ C) simplifies to:
log(kSA ) = −2 ·
206 207
EH − 2 · pH + d 0.059V
(11)
208
Note that this derivation can also be done in terms of the aqueous Fe3+ activity based on
209
the solubility of the iron oxide present. 32
210
Results and Discussion
211
0 Quantification of Goethite-Aqueous Fe2+ EH Values
212
We initially confirmed that EH values of suspensions containing goethite (α−FeOOH(s) ) and
213
aqueous Fe2+ (Fe2+ (aq) ) were consistent with the following half-reaction and corresponding
214
Nernst equation:
215
α−FeOOH(s) + e− + 3 H+ * ) Fe2+ (aq) + 2 H2 O 10
ACS Paragon Plus Environment
(12)
Page 11 of 35
Environmental Science & Technology
EH = EH0 −
216
217
RT RT ln[Fe2+ ln[H+ ] (aq) ] + 3 F F
(13)
where brackets denote activities. At 25◦ C, eq. 13 simplifies to:
EH = EH0 − 0.059V · log[Fe2+ (aq) ] − 0.177V · pH
218
(14)
219
We measured EH values of goethite suspensions over a range of total Fe2+ concentrations
220
(100 to 3000 µM) at two pH values (6.0 and 7.0) using mediated potentiometry, a technique
221
in which a soluble redox mediator is used to facilitate redox equilibration between a solution
222
and a redox electrode performing a potentiometric measurement (Figure 1). Data for goethite
223
from our previous work using the same method is shown for reference. 32 We characterized
224
two goethite batches made using identical synthesis routes because the first goethite batch
225
was exhausted before all nitrobenzene experiments were completed. We fit the data for each
226
goethite batch according to eq. 14 by floating the EH0 value and slope terms for Fe2+ (aq) activity
227
and pH using a least-squares multiple linear regression (Igor Pro software v7, Wavemetrics).
228
The activity coefficient for Fe2+ (aq) was calculated according to reactor conditions. There
229
was good agreement between the theoretical slope terms in eq. 14 and those from the
230
fitted regression (Fe2+ (aq) term: theoretical: 0.059 V, batch 1: 0.071±0.008 (±σ) V, batch 2:
231
0.065±0.005 V; pH term: theoretical: 0.177 V, batch 1: 0.177±0.008 V, batch 2: 0.165±0.005
232
V), confirming that the half-reaction in eq. 12 accurately described the EH values of the
233
suspensions.
234
These data were fit to calculate EH0 values for each batch of goethite. Since the deviations
235
between the theoretical and measured slope terms were most likely due to experimental error,
236
we used the theoretical slope values in eq. 14 to fit the data using a linear regression. The fits,
237
shown as dashed lines in Figure 1, yielded EH0 values of 0.801±0.004 (±σ) V (r2 = 0.98) for
238
batch 1 and 0.800±0.003 V (r2 = 0.99) for batch 2, which were within the range of previously
239
reported values (0.72 to 0.84 V). 56,57 The EH0 values were both approximately 0.03 V higher
240
than the value measured in our previous work using the same technique (0.768±0.001 V), 32 11
ACS Paragon Plus Environment
Reduction potential, EH [V vs. SHE]
Environmental Science & Technology
Page 12 of 35
0.0 pH = 6.0 -0.1
-0.2 pH = 7.0 Batch 1 Batch 2 Gorski et al., 2016
-0.3 4
6 8
100
2
4
6 8
1000
Aqueous Fe
2+
2
4
6
[µM]
Figure 1: Measured reduction potential values of the goethite-Fe2+ (aq) redox couple for the two goethite batches (red circles for batch 1, blue triangles for batch 2). Reduction potential values are shown as a function of solution pH and Fe2+ (aq) concentration after 24 hours of equilibration. Solution conditions: 25 mM buffer (MOPS for pH 7.0 and MES for pH 6.0), 25 mM KCl, 17 µM mediator (cyanomethyl viologen). Error bars show standard deviation between triplicate reactors and are contained within some markers. 241
which was likely due slight differences between the goethite batches used in this study and
242
our previous work. Since the EH0 values of the two goethite batches used here were identical
243
within error to each another, we could directly compare data collected for each batch. For all
244
subsequent nitrobenzene reduction experiments, we calculated EH values of each suspension
245
0 based on the pH, Fe2+ (aq) activity, and measured EH value (0.800 V) according to eq. 14.
246
Role of EH on Nitrobenzene Reduction Kinetics at pH 6.0
247
We initially performed nitrobenzene reduction experiments at a single pH value (6.0) to
248
simplify interpretation by eliminating pH-dependent effects on goethite-Fe2+ (aq) EH values (eq.
12
ACS Paragon Plus Environment
Page 13 of 35
Environmental Science & Technology
249
14) and nitrobenzene reduction kinetics. To determine the relationship between kobs values
250
and initial EH values of the goethite-Fe2+ (aq) redox couple, we measured nitrobenzene reduction
251
rate constants as a function of goethite concentration (0.5 to 4.0 g/L) and total Fe2+ (aq) con-
252
centration (100 to 5000 µM). All the nitrobenzene reduction experiments could initially be
253
described using a pseudo-first-order kinetic model (eq. 1), confirming that the reaction was
254
first-order with respect to nitrobenzene (example data shown in Figure S2). In experiments
255
that had relatively low initial Fe2+ (aq) concentrations, reaction kinetics eventually deviated
256
from what could be fit with a pseudo-first-order kinetic model, and therefore reported kobs
257
values were fit using only the portion of the data that yielded a linear fit between ln(kobs )
258
vs. time. In the experiments that went to completion, the mass balance of nitrobenzene
259
reduced to aniline produced ranged from 98 to 120% (average = 107±5%), and the molar
260
ratio of Fe2+ (aq) lost from solution to nitrobenzene reduced was 6.2(±2.8):1 (data shown in
261
Figure S3), consistent with the expected ratio of 6:1 from the electron balance of the overall
262
reaction (i.e., six electrons are needed to reduce one aromatic NO2 group to an NH2 group).
263
The kobs values among experiments were surface-area-normalized to account for differences
264
in goethite loadings among experiments according to eq. 1, as previously described in the
265
literature. 3,5,6
266
We plotted a free energy relationship in which log(kSA ) was related to the reaction’s
267
thermodynamic driving force (i.e., the Gibbs free energy of the reaction, ∆Grxn ). For redox
268
reactions, ∆Grxn is typically expressed in terms of EH . 41,58,59 Here, we used the EH values of
269
◦ the goethite-Fe2+ (aq) redox couple divided by 0.059 V (i.e., by 2.303 · RT /F when T =25 C).
270
The data, shown in Figure 2 and tabulated in Table S1, exhibited a linear relationship (r2 =
271
0.96) with a slope of –0.99±0.06. Note that EH /0.059 V scales with log{Fe2+ (aq) } in a 1:1 ratio
272
(eq. 14), and therefore a slope of ≈ −1 indicated that nitrobenzene reduction was first-order
273
with respect to the Fe2+ (aq) activity:
274
kSA = kFe · [Fe2+ (aq) ]
13
ACS Paragon Plus Environment
(15)
Environmental Science & Technology
Page 14 of 35
-1.5
-1
-2
log([kSA/L·hr ·m ])
-1.0
-2.0 -2.5 -3.0 -1.5
-1.0 -0.5 EH/0.059 V
0.0
Figure 2: Log transformed surface-area-normalized reaction rate coefficient [log(kSA )] values vs. EH values divided by 0.059 V of the goethite-Fe2+ (aq) redox couple at pH 6.0. The black line shows the linear fit with slope = −0.99 ± 0.06, intercept = −2.79 ± 0.06, r2 = 0.96. 275
More generally, the fact that the slope was not zero was consistent with an electron
276
transfer step being involved in the rate-limiting step(s) of nitrobenzene reduction. 40,42 The
277
mechanistic implications of the slope value are discussed below. The practical significance of
278
the slope is that EH values of the goethite-Fe2+ (aq) redox couple may be used as a parameter
279
to predict trends in reduction rate constants of nitrobenzene, and likely other nitroaromatic
280
compounds, for a single pH value. To understand if this relationship of kSA with reaction
281
free energy trend was also valid at different pH values, we performed additional experiments.
282
Effect of pH on Nitrobenzene Reduction Rates
283
We measured nitrobenzene reduction kinetics at pH values of 6.5 and 7.0 for a subset of
284
the reactor conditions tested at pH 6.0 (goethite concentrations of 1 and 2 g/L; total Fe2+ (aq) 14
ACS Paragon Plus Environment
Page 15 of 35
Environmental Science & Technology
285
concentrations of 100 to 1000 µM; data tabulated in Table S1). Since the pH value influences
286
both the EH value of the goethite-Fe2+ (aq) redox couple (eq. 14) and the thermodynamic driving
287
force dependence of nitrobenzene reduction rate constants, which involves proton transfer
288
steps, we expected that log(kSA ) values would vary as a function of both EH and pH:
289
log(kSA ) = a ·
EH + b · pH + c 0.059V
(16)
290
where a, b, and c are constants. Since eq. 16 contains two independent variables, its ability
291
to describe the data cannot be easily visualized in a 2-dimensional plot. We fit the data
292
using a multiple linear regression according to eq. 16 while floating b and c and holding a
293
equal to –1 (based on the results in Figure 2). The least-squares fit generated a pH slope
294
term (b) of –1.06±0.09 and an intercept term (c) of 3.6±0.6 (χ2 = 0.67; n = 23). The b value
295
was ≈ −1, which was the value that we used in subsequent analyses. To visually represent
296
the quality of this fit, we plotted log(kSA ) versus a · EH /0.059V + b · pH with a = −1 and
297
b = −1 (Figure 3, c = 3.12 ± 0.05, r2 = 0.96). The observation that the data was best
298
described with a pH slope term that differed from 0 indicated that nitrobenzene reduction
299
kinetics by goethite-bound Fe2+ were indeed pH-dependent, consistent with the reduction
300
pathway of nitroaromatic compounds and the Fe2+ oxidation reaction. 45,46
301
Note that representing the data this way masks the actual pH-dependency of log(kSA ),
302
since EH is itself a function of 3 · pH (eqs. 13 and 14). When the EH term is written out
303
according to eq. 14 and substituted into eq. 16 with a and b equal to –1, it is clear that the
304
reaction is second-order with respect to pH:
305
log(kSA ) = −
EH0 + log[Fe2+ (aq) ] + 2 · pH 0.059V
(17)
306
The mechanistic significance that can be drawn from the pH-dependency is discussed below.
307
To test the robustness of this free energy relationship, we plotted previously reported
308
log(kSA ) values for the reduction of nitrobenzene or substituted nitrobenzene compounds 15
ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 35
-1
-2
log([kSA/L·hr ·m ])
0
-1
-2 pH 7.0 pH 6.5 pH 6.0
-3 -6
-5
-4
-3
-EH/0.059 V - pH Figure 3: Relationship between log(kSA ), pH, and EH of the goethite-Fe2+ (aq) redox couple. The line shown is the fit with only the y-intercept term floated (c in eq. 16 = 3.12±0.05, r2 = 0.96). 309
by Fe2+ bound to goethite (Figure 4). 3,6,8,43,60 Since the rate of substituted nitrobenzene
310
compounds depends on the substituent groups present, we normalized EH values using a
311
published linear trend between log(kobs ) values and the compounds’ EH values for the first
312
electron transfer step (i.e., EH1 ). 59 The difference in EH1 values between nitrobenzene and
313
the substituted nitrobenzenes (i.e., ∆EH1 ) was calculated as:
0
0
0
0
314
0
0
∆EH1 = EH1 (nitrobenzene) − EH1 (substituted nitrobenzene)
(18)
0
315
The reported correlation between log(kobs ) and ∆EH1 used to normalize the data was: 59 0
316
log(kSA ) = 0.53 · ∆EH1 + c0
16
ACS Paragon Plus Environment
(19)
Page 17 of 35
Environmental Science & Technology
0
-1
-2
log([kSA/L·hr ·m ])
1
This study Strehlau et al., 2017 (nano-goethite) Fan et al., 2016 Jones et al., 2016 (nano-goethite) Jones et al., 2016 Colón et al, 2006 Elsner et al., 2004
-1 -2 -3 -4 -6
-5
-4
-3
-EH/0.059 V - pH Figure 4: log(kobs ) values as a function of pH and the calculated EH value of the goethite3,5,6,8,43 Fe2+ The method used to cal(aq) redox couple for data compiled from the literature. culate EH values is provided in Section S2. Reactor conditions varied between reported experiments: total Fe2+ (aq) ranged between 0.375 and 1.24 mM, goethite loading ranged between 0.025 and 1.54 g/L, BET specific surface areas of goethite ranged between 16.2 and 51 m2 /g, and pH values range between 6.1 and 7.97. The line shown is the fit from Figure 3. 317
where c0 is the y-intercept. The calculated EH values were corrected for this factor prior to
318
plotting them (data in Table S2).
319
Remarkably, virtually all the data complied from these studies performed by five different
320
groups of research teams between 2004 and 2016 fell upon the line derived from this work.
321
The observation that this simple free energy relationship could explain data collected for
322
goethite with different substituted nitrobenzene concentrations was particularly compelling
323
given the fact that the studies used different goethite batches prepared with a variety of
324
synthesis routes.
325
Interestingly, the two outlier points in Figure 3 were collected using carbonate as a pH
17
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 35
326
buffer, 43 whereas all the other studies including ours used Good’s buffers. Note that we
327
accounted for the fact that a portion (∼7%) of the Fe2+ (aq) was complexed with carbonate
328
(FeHCO3+ ) when calculating the activity of free Fe2+ (aq) from their data (Visual MINTEQ).
329
Carbonate is known to decrease the amount of available reactive surface area on an iron
330
2+ oxide and decrease the extent of Fe2+ uptake (aq) uptake. Since nitrobenzene reduction by Fe
331
is a surface-catalyzed reaction, the presence of carbonate explains the significantly lower
332
log(kSA ) values observed in this study.
333
Free Energy Relationship for Nitrobenzene Reduction
334
by Fe2+ and Different Iron Oxides
335
The consistent trend between log(kobs ) values and pH and EH values for goethite compelled us
336
to explore the extent to which this relationship could be applied to reported data collected for
337
other iron oxides, including hematite, 3,60 lepidocrocite, 3,8,60 ferrihydrite, 3,8 and magnetite. 9
338
Note that since the Gibbs free energy of formation of each oxide differs, the EH0 values for
339
2+ each oxide-Fe2+ (aq) redox couple also differs. We calculated EH of the oxide-Fe(aq) redox couples
340
according to the following half reactions and their corresponding Nernst equations:
341
Hematite :
3 1 α−Fe2 O3(s) + 3H+ + e− * ) Fe2+ (aq) + H2 O 2 2
(20a)
342
Lepidocrocite :
α−FeOOH(s) + 3H+ + e− * ) Fe2+ (aq) + 2H2 O
(20b)
343
Ferrihydrite :
Fe(OH)3(s) + 3H+ + e− * ) Fe2+ (aq) + 3H2 O
(20c)
Magnetite :
1 3 * Fe2+ Fe3 O4(s) + 4H+ + e− ) + 2H2 O 2 2 (aq)
(20d)
344 345
346
The Nernst equations for all these reactions, except magnetite, are identical to the one in
347
eq. 13. The Nernst equation for the magnetite half-reaction is:
348
EH = EH0 −
3 RT RT ln[Fe2+ ln[H+ ] (aq) ] + 4 2 F F 18
ACS Paragon Plus Environment
(21)
Page 19 of 35
Environmental Science & Technology
349
32 We previously measured the EH0 value for the hematite-Fe2+ (aq) half reaction (0.768±0.002 V).
350
We calculated EH0 values for the half reactions containing lepidocrocite (0.846 V), ferrihydrite
351
(0.937 V), and magnetite (1.067 V) using existing thermodynamic data. 61,62 While solutions
352
containing lepidocrocite and ferrihydrite are not at thermodynamic equilibrium (i.e., the ox-
353
ides should transform to more stable phases), the EH0 values can be useful approximations if
354
the oxides remain stable over the course of the reaction. Values of EH were calculated accord-
355
0 ing to Fe2+ (aq) activity, pH, and EH values (data in Tables S2 and S3). Calculated EH values
356
were also corrected to account for differences of reactivities of the substituted nitrobenzenes
357
by normalizing reaction rate constants according to their EH1 values to nitrobenzene using
358
eq. 19.
0
359
All the reduction rate constants compiled from the literature for goethite, hematite,
360
lepidocrocite, and ferrihydrite exhibited a clear linear trend consistent with the linear free
361
energy relationship developed for goethite (Figure 5). In the case of magnetite, the data
362
initially did not fall on the line when we calculated EH values according to eq. 21. This
363
was likely due to the incorrect assumption that magnetite was the oxidation product that
364
formed. A recent study that characterized the Fe2+ oxidation product that formed when
365
4-chloro-nitrobenzene was reduced by suspensions containing magnetite and Fe2+ (aq) observed
366
the formation of lepidocrocite, not magnetite, as the dominant oxidation product. 38 When we
367
calculated EH values assuming that lepidocrocite was the oxidation product, the magnetite
368
data fell in line with the data for the other oxides (Figure 5).
369
Fitting all the data with a line yielded a slope of 0.76, a y-intercept of –3.57, and an
370
r2 value of 0.68 (n = 74, fit not shown). Excluding three outlier points, which are denoted
371
with asterisks inside the markers, significantly improved the fit (r2 = 0.92, n = 71) and
372
yielded a slope of 1.02±0.03 and a y-intercept of 3.42±0.17 (line shown in Figure 5). Two
373
of the outliers were points collected using a carbonate pH buffer, 43 as discussed above. The
374
third outlier point was excluded because it clearly deviated from the overall trend, although
375
we could not identify an explanation as to why. 60 The 95% prediction band (encompassed
19
ACS Paragon Plus Environment
Environmental Science & Technology
0
-1
-2
log([kSA/L·hr ·m ])
1
-1
Page 20 of 35
This study Elsner et al., 2004 Jones et al., 2016 Colón et al., 2006 Fan et al., 2016 Strehlau et al., 2016 Klausen et al., 1995 Goethite Hematite Lepidocrocite Ferrihydrite Magnetite
* *
-2 -3
*
-4 -7
-6
-5
-4
-3
-2
-EH/0.059 V - pH Figure 5: Compilation of log(kobs ) values as a function of pH and the calculated EH values for the reduction of nitrobenzene by Fe2+ in the presence of goethite, hematite, lepidocrocite, ferrihydrite, and magnetite. The solid line is the least-squares linear regression excluding the three outlier points (denoted with asterisks). The fit yielded a slope of 1.02±0.03, a y-intercept of 3.42±0.17, and an r2 value of 0.92 (n = 71; 24 from this study, 47 from the literature, 3 outlier points excluded). The shaded area represents the 95% confidence band, and the dashed lines encompass the 95% prediction band. Data were taken from refs. 3,6,8,9,43,60. 376
by the dashed lines in Figure 5) was approximately ±1 log unit on the y-axis, indicating
377
that 95% of the reported kSA values were within an order of magnitude of the values that
378
would be predicted by the free energy relationship. Some of the scatter among the datasets
379
was likely due to variations in the buffer type and concentrations used in each study. Prior
380
work found that commonly used pH buffers caused iron oxide particle aggregation, which
381
influenced kSA values by up to 0.5 log units. 51 This ability for the free energy relationship
382
to describe the literature data indicates that pH values and EH values for oxide-Fe2+ (aq) redox
383
couples can be used to estimate nitrobenzene, and likely nitroaromatic, transformation rate
20
ACS Paragon Plus Environment
Page 21 of 35
384
Environmental Science & Technology
constants within an order of magnitude.
385
A critical assumption implicit to this approach is that Fe2+ oxidation results in homoepi-
386
taxial growth of the underlying oxide (e.g., goethite grows on goethite), with the exception
387
of magnetite. In the case of goethite, several studies have indeed found that Fe2+ oxida-
388
tion leads to goethite particle growth. 37,38,43 Ferrihydrite, hematite, and lepidocrocite are
389
more complicated. In the case of hematite, Fe2+ oxidation can lead to both the growth of
390
hematite and the nucleation of goethite particles. 38 Interestingly, this does not influence the
391
thermodynamic calculations because the half reactions for the oxidation of Fe2+ to hematite
392
or goethite have virtually identical EH0 values and identical Nernst equations. 32 Very little
393
work has examined what Fe2+ oxidation products form in the presence of lepidocrocite or
394
ferrihydrite. 8 Both oxides are known to undergo secondary mineralization reactions to more
395
63–67 stable phases in the presence of aqueous Fe2+ It is therefore (aq) at circumneutral pH values.
396
likely that Fe2+ may oxidize to form oxides other than original phase, which could explain
397
why the ferrihydrite and lepidocrocite points in Figure 5 are generally higher than those
398
of goethite and hematite. With this caveat in mind, assuming that nitrobenzene reduction
399
is coupled to homoepitaxial oxide growth works from a practical perspective for all oxides
400
except magnetite.
401
Two recent works have reported similar free energy relationships in which the authors
402
plotted log(kSA ) versus EH /0.059 V values for multiple iron oxides and substituted nitroben-
403
zene compounds at a single pH value. 6,8 The studies differed from the present work in that
404
the authors measured EH values for suspensions by preparing parallel reactors. In both
405
studies, the fitted slope values (0.16 8 or 0.49±0.15 6 ) were significantly smaller than the one
406
reported here (i.e., ≈1). In the study that observed a slope of 0.16, a portion of the reactors
407
contained dissolved Si that blocked iron oxide surface sites and therefore affected calculated
408
kSA values. 8 In the other study, the authors speculated that measured EH values may have
409
contained experimental artifacts, as the measurements depended on the concentrations of
410
Fe2+ and iron oxide in a way that deviated from the Nernst equation. 6 Note that data taken
21
ACS Paragon Plus Environment
Environmental Science & Technology
411
from both these studies fell on the trend line in Figure 5 when we calculated EH values based
412
on the pH and Fe2+ (aq) activity and that the calculated values significantly differed from the
413
measured ones in most cases. We are unsure as to why the calculated and measured EH
414
values from these studies differed.
415
Interpretation of the Slope Terms in the Free Energy Relationship
416
The trends between log(kSA ) and EH and pH (i.e., the slope terms a and b in eq. 16)
417
provide insights into the rate limiting step(s) in the reduction of reducible aromatic nitro
418
groups by oxide-bound Fe2+ . To interpret these trends, we briefly discuss the current state
419
of knowledge regarding the reduction pathway of nitroaromatics, using nitrobenzene as a
420
surrogate (Scheme 1). The reduction of nitrobenzene (species 1 in Scheme 1) to aniline is
421
initiated by a series of two proton and two electron transfer steps before nitrosobenzene (6) is
422
formed as first measurable product through a dehydration step, which controls the rate of ni-
423
trobenzene disappearance. Since the reaction from 1 to 6 is irreversible under environmental
424
conditions, 41 we can confine the kinetic interpretation of the free energy relationship to these
425
steps. Early studies examining the reduction of nitroaromatic compounds by oxide-bound
426
Fe2+ observed linear correlations between log(kobs ) and the one-electron reduction poten-
427
tial values of nitroaromatic compounds, EH1 . 3,7,9,41 These observations led to the hypothesis
428
that the first electron transfer step to nitrobenzene, 1 * ) 2, was a suitable descriptor of the
429
relative reduction rates and potentially the rate-limiting step of reactions 1 to 6 (Scheme
430
1B). 41,58 More recent evidence from large
431
duction of substituted nitrobenzenes 45–47,68 implied, however, that the dehydration step of
432
the N,N -dihydroxyaniline (5 → 6) was the rate-liming step, consistent with electrochemical
433
experiments examining the reduction of aromatic NO2 groups. 69,70
0
15
N kinetic isotope effect associated with the re-
434
We examined if the free energy relationship observed here was consistent with (i) reaction
435
5 → 6 being the rate-limiting step and (ii) one-electron reduction potentials being used as
436
descriptors for variations in reduction rate constants. We derived a rate expression for the 22
ACS Paragon Plus Environment
Page 22 of 35
Page 23 of 35
Environmental Science & Technology
437
reduction of nitrobenzene by Fe2+ assuming that 5 → 6 was rate-limiting with the quasi-
438
equilibrium method, which assumes that the preceding reaction steps are in equilibrium (the
439
derivation is in the subsection “Derivation of Free Energy Relationship” above). 40,42,55 This
440
method is thought to be valid if the rate constant of the rate-limiting step is at least 100
441
times smaller than those of the preceding steps. 42 With eq. 11, we are able to reconcile
442
potentially contradicting views with regard to the rate-limiting step of aromatic NO2 group
443
reduction. Note that even though we assume that reaction 5 → 6 is rate-limiting, log(kSA )
444
scales linearly with changes of EH1 /0.059 V in constant d, as shown in the derivation.
0
445
The derived slope terms (a for the EH /0.059 term and b for the pH term) in eq. 11 are
446
both –2. Note that the b term does not reflect the overall pH-dependency of the reaction
447
since the EH value is itself pH-dependent. The derived equation can also be expressed in
448
terms of equilibrium constants instead of EH , in which case the derived rate expression
449
indicates that reaction kinetics would be second-order with respect to [Fe2+ ] and forth-order
450
with respect to 1/[H+ ] (eq. 5f). Clearly, the derived a and b terms differed from the observed
451
terms, which were both ≈ –1. The observed values instead indicate that the reduction rate
452
constants depend only on the free energy changes associated with transferring two protons
453
and one electron. There are two possible explanations for the discrepancy. The first is that
454
the quasi-equilibrium assumption used to derive the theoretical relationship is invalid for this
455
system because the preceding reaction steps do not reach equilibrium. Such a case can arise
456
when rate constants have similar orders of magnitude. 55 The second possibility is that the
457
second electron transfer step (3 * ) 4) and second proton transfer step (4 * ) 5) in Scheme
458
1B have sufficiently lower activation free energies than the preceding ones, making the steps
459
kinetically irrelevant. In either case, this exercise demonstrates that free energy relationship
460
for log(kSA ) depends on the transfer of one proton and one electron, not two of each.
23
ACS Paragon Plus Environment
Environmental Science & Technology
461
Environmental Implications
462
In this work, we observed that EH values of oxide-bound Fe2+ could be used to accurately
463
describe variations in nitrobenzene reduction rates. This finding is significant because it
464
validates the use of thermodynamic data for oxide-bound Fe2+ to explain trends in nitroaro-
465
matic reduction rates. The data indicates that it is unnecessary to describe “adsorbed Fe2+ ”
466
as a discrete species when explaining trends in pollutant transformation rates. Instead, the
467
61,62 EH of Fe2+ can (aq) , which can be calculated from readily available thermodynamic data,
468
be used effectively. The extent to which this free energy relationship can be applied to other
469
classes of pollutants containing different reducible functional groups remains unclear at this
470
point due simply to the lack of data, but it seems highly probable that that EH values
471
will be useful if the reaction is completely or partially rate-limited by one or more electron
472
transfer steps. Of course, in real environmental systems, pollutant reduction rates will also
473
be affected by other species, such as carbonate, natural organic matter, and silicate, which
474
compete for reactive sites and/or passivate iron oxide surfaces.
475
The findings from this work can also explain a puzzling observation. Past works found
476
negligible reduction of nitrite, molecular oxygen, and nitrobenzene occurred when an iron
477
2+ 2+ oxide was exposed to Fe2+ (aq) and either the residual Fe(aq) was removed or most of the Fe(aq)
478
was taken up the oxide. 16,19,29 This observation was difficult to reconcile with the hypothesis
479
that oxide-bound Fe2+ is a strong reductant. 21 The observation can, however, be explained
480
based on the observations from this study. Removing Fe2+ (aq) from the system makes the EH
481
2+ values of both Fe2+ more positive. This suggests that the concen(aq) and oxide-bound Fe
482
tration of Fe2+ (aq) in groundwater systems can be used an important predictive parameter for
483
pollutant fate, as it is a proxy for oxide-bound Fe2+ EH values in the presence of iron oxides.
24
ACS Paragon Plus Environment
Page 24 of 35
Page 25 of 35
Environmental Science & Technology
484
Acknowledgement
485
This work was supported by the U.S. National Science Foundation (grant EAR-1451593 to
486
CAG) and by the Swiss National Science Foundation (grant 200021 149283 to TBH). We
487
thank William Burgos for reviewing the manuscript prior to submission.
488
Supporting Information Available
489
Example kinetic data of nitrobenzene reduction and tabulated experimental and literature
490
data.
25
ACS Paragon Plus Environment
Environmental Science & Technology
Graphical TOC entry This study Literature log(kSA)
491
1.02 1
-EH/0.059 V - pH
26
ACS Paragon Plus Environment
Page 26 of 35
Page 27 of 35
492
Environmental Science & Technology
References
493
(1) Amonette, J. E.; Workman, D. J.; Kennedy, D. W.; Fruchter, J. S.; Gorby, Y. A.
494
Dechlorination of carbon tetrachloride by Fe(II) associated with goethite. Environ. Sci.
495
Technol. 2000, 34, 4606–4613.
496
497
498
499
500
501
(2) Buerge, I. J.; Hug, S. J. Influence of mineral surfaces on chromium(VI) reduction by iron(II). Environ. Sci. Technol. 1999, 33, 4285–4291. (3) Col´on, D.; Weber, E. J.; Anderson, J. L. QSAR study of the reduction of nitroaromatics by Fe(II) species. Environ. Sci. Technol. 2006, 40, 4976–4982. (4) Cui, D.; Eriksen, T. E. Reduction of pertechnetate by ferrous iron in solution: Influence of sorbed and precipitated Fe(II). Environ. Sci. Technol. 1996, 30, 2259–2262.
502
(5) Elsner, M.; Schwarzenbach, R.; Haderlein, S. Reactivity of Fe(II)-bearing minerals to-
503
ward reductive transformation of organic contaminants. Environ. Sci. Technol. 2004,
504
38, 799–807.
505
(6) Fan, D.; Bradley, M. J.; Hinkle, A. W.; Johnson, R. L.; Tratnyek, P. G. Chemical
506
reactivity probes for assessing abiotic natural attenuation by reducing iron minerals.
507
Environ. Sci. Technol. 2016, 50, 1868–1876.
508
(7) Hofstetter, T. B.; Heijman, C. G.; Haderlein, S. B.; Holliger, C.; Schwarzenbach, R. P.
509
Complete reduction of TNT and other (poly)nitroaromatic compounds under iron re-
510
ducing subsurface conditions. Environ. Sci. Technol. 1999, 33, 1479–1487.
511
(8) Jones, A. M.; Kinsela, A. S.; Collins, R. N.; Waite, T. D. The reduction of 4-
512
chloronitrobenzene by Fe(II)-Fe(III) oxide systems - correlations with reduction po-
513
tential and inhibition by silicate. J. Hazard. Mater. 2016, 320, 143–149.
514
(9) Klausen, J.; Troeber, S. P.; Haderlein, S. B.; Schwarzenbach, R. P. Reduction of substi-
27
ACS Paragon Plus Environment
Environmental Science & Technology
515
tuted nitrobenzenes by Fe(II) in aqueous mineral suspensions. Environ. Sci. Technol.
516
1995, 29, 2396–2404.
517
518
(10) Liger, E.; Charlet, L.; Van Cappellen, P. Surface catalysis of uranium(VI) reduction by iron(II). Geochim. Cosmochim. Acta 1999, 63, 2939–2955.
519
(11) Pecher, K.; Haderlein, S. B.; Schwarzenbach, R. P. Reduction of polyhalogenated
520
methanes by surface-bound Fe(II) in aqueous suspensions of iron oxides. Environ. Sci.
521
Technol. 2002, 36, 1734–1741.
522
(12) R¨ ugge, K.; Hofstetter, T. B.; Haderlein, S. B.; Bjerg, P. L.; Knudsen, S.; Zraunig, C.;
523
Mosbæk, H.; Christensen, T. H. Characterization of predominant reductants in an
524
anaerobic leachate-contaminated aquifer by nitroaromatic probe compounds. Environ.
525
Sci. Technol. 1998, 32, 23–31.
526
527
(13) Strathmann, T. J.; Stone, A. T. Mineral surface catalysis of reactions between Fe(II) and oxime carbamate pesticides. Geochim. Cosmochim. Acta 2003, 67, 2775–2791.
528
(14) Vikesland, P. J.; Valentine, R. L. Iron oxide surface-catalyzed oxidation of ferrous iron
529
by monochloramine: Implications of oxide type and carbonate on reactivity. Environ.
530
Sci. Technol. 2002, 36, 512–519.
531
(15) Jones, A. M.; Griffin, P. J.; Collins, R. N.; Waite, T. D. Ferrous iron oxidation under
532
acidic conditions–The effect of ferric oxide surfaces. Geochim. Cosmochim. Acta 2014,
533
145, 1–12.
534
(16) Park, B.; Dempsey, B. A. Heterogeneous oxidation of Fe(II) on ferric oxide at neutral
535
pH and a low partial pressure of O2 . Environ. Sci. Technol. 2005, 39, 6494–6500.
536
(17) Sung, W.; Morgan, J. J. Kinetics and product of ferrous iron oxygenation in aqueous
537
systems. Environ. Sci. Technol. 1980, 14, 561–568.
28
ACS Paragon Plus Environment
Page 28 of 35
Page 29 of 35
538
539
540
541
542
543
544
545
546
547
Environmental Science & Technology
(18) Tamura, H.; Goto, K.; Nagayama, M. The effect of ferric hydroxide on the oxygenation of ferrous ions in neutral solutions. Corros. Sci. 1976, 16, 197–207. (19) Tai, Y.-L.; Dempsey, B. A. Nitrite reduction with hydrous ferric oxide and Fe(II): Stoichiometry, rate, and mechanism. Water Res. 2009, 43, 546–552. (20) Stumm, W. Aquatic surface chemistry: Chemical processes at the particle-water interface; Wiley: New York, 1987. (21) Stumm, W. Chemistry of the solid-eater interface: Processes at the mineral-water and particle-water interface of natural systems; Wiley: New York, 1992; p 428. (22) Hering, J.; Stumm, W. Oxidative and reductive dissolution of minerals. Reviews in Mineralogy and Geochemistry 1990, 23, 427–465.
548
(23) Orsetti, S.; Laskov, C.; Haderlein, S. B. Electron transfer between iron minerals and
549
quinones: estimating the reduction potential of the Fe(II)-goethite surface from AQDS
550
speciation. Environ. Sci. Technol. 2013, 47, 14161–14168.
551
(24) Silvester, E.; Charlet, L.; Tournassat, C.; Gehin, A.; Greneche, J. M.; Liger, E. Redox
552
potential measurements and M¨ossbauer spectrometry of Fe(II) adsorbed onto Fe(III)
553
(oxyhydr)oxides. Geochim. Cosmochim. Acta 2005, 69, 4801–4815.
554
555
556
557
558
559
(25) Gorski, C. A.; Scherer, M. M. Aquatic Redox Chemistry; American Chemical Society, 2011; Chapter 15, pp 315–343. (26) Gorski, C. A.; Scherer, M. M. Influence of magnetite stoichiometry on FeII uptake and nitrobenzene reduction. Environ. Sci. Technol. 2009, 43, 3675–3680. (27) Larese-Casanova, P.; Scherer, M. M. Fe(II) sorption on hematite: New insights based on spectroscopic measurements. Environ. Sci. Technol. 2007, 41, 471–477.
29
ACS Paragon Plus Environment
Environmental Science & Technology
560
(28) Rosso, K. M.; Yanina, S. V.; Gorski, C. A.; Larese-Casanova, P.; Scherer, M. M. Con-
561
necting observations of hematite (α−Fe2 O3 ) growth catalyzed by Fe(II). Environ. Sci.
562
Technol. 2010, 44, 61–67.
563
564
565
566
(29) Williams, A.; Scherer, M. M. Spectroscopic evidence for Fe(II)-Fe(III) electron transfer at the Fe oxide-water interface. Environ. Sci. Technol. 2004, 38, 4782–4790. (30) Yanina, S. V.; Rosso, K. M. Linked reactivity at mineral-water interfaces through bulk crystal conduction. Science 2008, 320, 218–222.
567
(31) Zarzycki, P.; Kerisit, S.; Rosso, K. M. Molecular dynamics study of Fe(II) adsorption,
568
electron exchange, and mobility at goethite (α−FeOOH) surfaces. J. Phys. Chem. C
569
2015, 119, 3111–3123.
570
(32) Gorski, C. A.; Edwards, R.; Sander, M.; Hofstetter, T. B.; Stewart, S. Thermodynamic
571
characterization of iron oxide - aqueous Fe2+ redox couples. Environ. Sci. Technol.
572
2016, 50, 8538–8547.
573
(33) Felmy, A. R.; Ilton, E. S.; Rosso, K. M.; Zachara, J. M. Interfacial reactivity of radionu-
574
clides: emerging paradigms from molecular-level observations. Mineral. Mag. 2011, 75,
575
2379–2391.
576
(34) Felmy, A. R.; Moore, D. A.; Rosso, K. M.; Qafoku, O.; Rai, D.; Buck, E. C.; Ilton, E. S.
577
Heterogeneous reduction of PuO2 with Fe(II): Importance of the Fe(III) reaction prod-
578
uct. Environ. Sci. Technol. 2011, 45, 3952–3958.
579
(35) Krumina, L.; Lyngsie, G.; Tunlid, A.; Persson, P. Oxidation of a dimethoxyhydro-
580
quinone by ferrihydrite and goethite nanoparticles: Iron reduction versus surface catal-
581
ysis. Environ. Sci. Technol. 2017, 51, 9053–9061.
582
583
(36) Schwertmann, U.; Carlson, L.; Fechter, H. Iron oxide formation in artificial ground waters. Aquat. Sci. 1984, 46, 185–191. 30
ACS Paragon Plus Environment
Page 30 of 35
Page 31 of 35
Environmental Science & Technology
584
(37) Chun, C. L.; Penn, R. L.; Arnold, W. A. Kinetic and microscopic studies of reductive
585
transformations of organic contaminants on goethite. Environ. Sci. Technol. 2006, 40,
586
3299–3304.
587
(38) Larese-Casanova, P.; Kappler, A.; Haderlein, S. B. Heterogeneous oxidation of Fe(II) on
588
iron oxides in aqueous systems: Identification and controls of Fe(III) product formation.
589
Geochim. Cosmochim. Acta 2012, 91, 171–186.
590
(39) Deakin, M. R.; Kovach, P. M.; Stutts, K.; Wightman, R. M. Heterogeneous mechanisms
591
of the oxidation of catechols and ascorbic acid at carbon electrodes. Anal. Chem. 1986,
592
58, 1474–1480.
593
(40) Ram, M.; Hupp, J. T. Linear free energy relations for multielectron transfer kinetics:
594
A brief look at the Broensted/Tafel analogy. J. Phys. Chem-US 1990, 94, 2378–2380.
595
(41) Schwarzenbach, R. P.; Gschwend, P. M.; Imboden, D. M. Environmental organic chem-
596
istry; John Wiley & Sons, 2016.
597
(42) Guidelli, R.; Compton, R. G.; Feliu, J. M.; Gileadi, E.; Lipkowski, J.; Schmickler, W.;
598
Trasatti, S. Defining the transfer coefficient in electrochemistry: An assessment (IUPAC
599
Technical Report). Pure Appl. Chem. 2014, 86, 245–258.
600
(43) Strehlau, J. H.; Stemig, M. S.; Penn, R. L.; Arnold, W. A. Facet-dependent oxidative
601
goethite growth as a function of aqueous solution conditions. Environ. Sci. Technol.
602
2016, 50, 10406–10412.
603
(44) Bylaska, E. J.; Salter-Blanc, A.; Tratnyek, P. G. Aquatic Redox Chemistry; ACS Sym-
604
posium Series; American Chemical Society, 2011; Vol. 1071; Chapter 3, pp 37–64.
605
(45) Hartenbach, A.; Hofstetter, T. B.; Berg, M.; Bolotin, J.; Schwarzenbach, R. P. Using
606
nitrogen isotope fractionation to assess abiotic reduction of nitroaromatic compounds.
607
Environ. Sci. Technol. 2006, 40, 7710–7716. 31
ACS Paragon Plus Environment
Environmental Science & Technology
608
(46) Hartenbach, A. E.; Hofstetter, T. B.; Aeschbacher, M.; Sander, M.; Kim, D.; Strath-
609
mann, T. J.; Arnold, W. A.; Cramer, C. J.; Schwarzenbach, R. P. Variability of nitrogen
610
isotope fractionation during the reduction of nitroaromatic compounds with dissolved
611
reductants. Environ. Sci. Technol. 2008, 42, 8352–8359.
612
(47) Hofstetter, T. B.; Neumann, A.; Arnold, W. A.; Hartenbach, A. E.; Bolotin, J.;
613
Cramer, C. J.; Schwarzenbach, R. P. Substituent effects on nitrogen isotope fractiona-
614
tion during abiotic reduction of nitroaromatic compounds. Environ. Sci. Technol. 2008,
615
42, 1997–2003.
616
(48) Salter-Blanc, A. J.; Bylaska, E. J.; Johnston, H. J.; Tratnyek, P. G. Predicting reduction
617
rates of energetic nitroaromatic compounds using calculated one-electron reduction
618
potentials. Environ. Sci. Technol. 2015, 49, 3778–3786.
619
(49) Tamura, H.; Goto, K.; Yotsuyan.T,; Nagayama, M. Spectrophotometric determination
620
of iron(II) with 1,10-phenanthroline in presence of large amounts of iron(III). Talanta
621
1974, 21, 314–318.
622
(50) Buchholz, A.; Laskov, C.; Haderlein, S. B. Effects of zwitterionic buffers on sorption of
623
ferrous iron at goethite and its oxidation by 0 CCl4 . Environ. Sci. Technol. 2011, 45,
624
3355–3360.
625
(51) Stemig, A. M.; Do, T. A.; Yuwono, V. M.; Arnold, W. A.; Penn, R. L. Goethite
626
nanoparticle aggregation: Effects of buffers, metal ions, and 4-chloronitrobenzene re-
627
duction. Environ. Sci: Nano 2014, 1, 478–487.
628
629
(52) Schwertmann, U.; Cornell, R. Iron oxides in the laboratory: Preparation and characterization, 2nd ed.; Wiley-VCH: New York, 2000; p 188.
630
(53) Sander, M.; Hofstetter, T. B.; Gorski, C. A. Electrochemical analyses of redox-active
631
iron minerals: A review of nonmediated and mediated approaches. Environ. Sci. Tech-
632
nol. 2015, 49, 5862–5878. 32
ACS Paragon Plus Environment
Page 32 of 35
Page 33 of 35
Environmental Science & Technology
633
(54) Gorski, C. A.; Kl¨ upfel, L.; Voegelin, A.; Sander, M.; Hofstetter, T. B. Redox properties
634
of structural Fe in clay minerals. 2. Electrochemical and spectroscopic characterization
635
of electron transfer irreversibility in ferruginous smectite, SWa-1. Environ. Sci. Technol.
636
2012, 46, 9369–9377.
637
(55) Marshall, A. T.; Vaisson-B´ethune, L. Avoid the quasi-equilibrium assumption when
638
evaluating the electrocatalytic oxygen evolution reaction mechanism by Tafel slope
639
analysis. Electrochem. Commun. 2015, 61, 23 – 26.
640
641
642
643
(56) Fischer, W. R. Standard potentials (E 0 ) of iron(III) oxides under reducing conditions. J. Plant. Nutr. Soil Sci. 1987, 150, 286–289. (57) Roden, E. E. Geochemical and microbiological controls on dissimilatory iron reduction. C. R. Geosci. 2006, 338, 456–467.
644
(58) Tratnyek, P. G.; Hoign´e, J.; Zeyer, J.; Schwarzenbach, R. QSAR analyses of oxidation
645
and reduction rates of environmental organic pollutants in model systems. Sci. Total
646
Environ. 1991, 109, 327–341.
647
(59) Tratnyek, P. G.; Weber, E. J.; Schwarzenbach, R. P. Quantitative structure-activity
648
relationships for chemical reductions of organic contaminants. Environ. Toxicol. Chem.
649
2003, 22, 1733–1742.
650
(60) Elsner, M.; Haderlein, S. B.; Kellerhals, T.; Luzi, S.; Zwank, L.; Angst, W.; Schwarzen-
651
bach, R. P. Mechanisms and products of surface-mediated reductive dehalogenation of
652
carbon tetrachloride by Fe(II) on goethite. Environ. Sci. Technol. 2004, 38, 2058–2066.
653
(61) Robie, R.; Hemingway, B.; Fisher, J. Thermodynamic properties of minerals and related
654
substances at 298.15 K and 1 Bar (105 Pascals) pressure and at higher temperatures;
655
Report, 1995.
33
ACS Paragon Plus Environment
Environmental Science & Technology
656
657
658
659
(62) Navrotsky, A.; Mazeina, L.; Majzlan, J. Size-driven structural and thermodynamic complexity in iron oxides. Science 2008, 319, 1635–1638. (63) Hansel, C. M.; Benner, S. G.; Fendorf, S. Competing Fe(II)-induced mineralization pathways of ferrihydrite. Environ. Sci. Technol. 2005, 39, 7147–7153.
660
(64) Hansel, C. M.; Benner, S. G.; Neiss, J.; Dohnalkova, A.; Kukkadapu, R.; Fendorf, S.
661
Secondary mineralization pathways induced by dissmilatory iron reduction of ferrihy-
662
drate under advective flow. Geochim. Cosmochim. Acta 2003, 67, 2977–2992.
663
664
(65) Yee, N.; Shaw, S.; Benning, L. G.; Nguyen, T. H. The rate of ferrihydrite transformation to goethite via the Fe(II) pathway. Am. Mineral. 2006, 91, 92–96.
665
(66) Yang, L.; Steefel, C. I.; Marcus, M. A.; Bargar, J. R. Kinetics of Fe(II)-catalyzed trans-
666
formation of 6-line ferrihydrite under anaerobic flow conditions. Environ. Sci. Technol.
667
2010, 44, 5469–5475.
668
(67) O’Loughlin, E. J.; Gorski, C. A.; Scherer, M. M.; Boyanov, M. I.; Kemner, K. M. Effects
669
of oxyanions, natural organic matter, and bacterial cell numbers on the bioreduction
670
of lepidocrocite (γ−FeOOH) and the formation of secondary mineralization products.
671
Environ. Sci. Technol. 2010, 44, 4570–4576.
672
(68) Gorski, C. A.; Nurmi, J. T.; Tratnyek, P. G.; Hofstetter, T. B.; Scherer, M. M. Redox
673
behavior of magnetite: implications for contaminant reduction. Environ. Sci. Technol.
674
2010, 44, 55–60.
675
(69) Laviron, E.; Roullier, L. The reduction mechanism of aromatic nitro compounds in
676
aqueous medium Part 1. Reduction to dihydroxylamines between pH 0 and 5. J. Elec-
677
troanal. Chem. 1990, 288, 165–175.
678
(70) Laviron, E.; Meunier-Prest, R.; Lacasse, R. The reduction-mechanism of aromatic nitro-
34
ACS Paragon Plus Environment
Page 34 of 35
Page 35 of 35
Environmental Science & Technology
679
compounds in aqueous-medium Part 4. The reduction of p-nitrobenzophenone between
680
pH=-5 and pH 14. J. Electroanal. Chem. 1994, 375, 263–274.
35
ACS Paragon Plus Environment