Rapid Iron Reduction Rates Are Stimulated by ... - ACS Publications

Feb 13, 2017 - Rapid Iron Reduction Rates Are Stimulated by High-Amplitude. Redox Fluctuations in a Tropical Forest Soil. Brian Ginn,. †. Christof M...
1 downloads 0 Views 3MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Rapid iron reduction rates are stimulated by highamplitude redox fluctuations in a tropical forest soil Brian R. Ginn, Christof Meile, Jared Wilmoth, Yuanzhi Tang, and Aaron Thompson Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b05709 • Publication Date (Web): 13 Feb 2017 Downloaded from http://pubs.acs.org on February 19, 2017

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 29

Environmental Science & Technology

1 2 3

Rapid iron reduction rates are stimulated by high-amplitude redox fluctuations in a tropical

4

forest soil

5

Brian Ginn1, Christof Meile2, Jared Wilmoth1, Yuanzhi Tang3 and Aaron Thompson1,*

6 7

1

2

8 9

University of Georgia (Crop and Soil Sci.)

3

University of Georgia (Marine Sci.)

Georgia Institute of Technology (Earth and Atmospheric Sci)

10 11 12

For submission to Environmental Science and Technology

13 14 15

*Corresponding author E-mail: [email protected], (01) 706-410-1293

16

1

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 29

17 18 19

Abstract:

20

altered by redox-driven processes. This study examined the influence of temporal oxygen variations on

21

Fe speciation in soils from the Luquillo Critical Zone Observatory (Puerto Rico). We incubated soils

22

under cycles of oxic-anoxic conditions (τoxic:τanoxic = 1:6) at three frequencies with and without

23

phosphate addition. Fe(II) production, P availability, and Fe mineral composition were monitored using

24

batch analytical and spectroscopic techniques. The rate of FeII reduction increased from ~ 3 to > 45

25

mmol Fe(II) kg-1 soil d-1 over the experiment with a concomitant increase of an Fe(II) concentration

26

plateau within each anoxic period. The apparent maximum in reducible Fe is similar in all treatments,

27

but was hastened by P-amendment. Numerical modelling suggests the Fe(II) dynamics can be

28

explained by the formation of rapidly reducible Fe(III) phases derived from the progressive dissolution

29

of native Fe(III) oxides accompanied by minor increases in Fe reducer populations. The shift in Fe(III)

30

reactivity is evident from Fe-reducibility assays using Shewanella sp., however was undetectable by

31

chemical extractions, Mössbauer or X-ray Absorption spectroscopies. More broadly, our findings

32

suggest Fe reduction rates are strongly coupled to redox dynamics of the recent past, and that frequent

33

shifts in redox conditions can prime a soil for rapid Fe-reduction.

Iron oxides are important structural and biogeochemical components of soils that can be strongly

Rapidly&Reducible&Fe(III)& Reducible&Fe(III)&

Oxida1on&

34

Fe(II)&

Reduc1on&

Reduc1on&

TOC ART

2

ACS Paragon Plus Environment

Page 3 of 29

35

Environmental Science & Technology

Introduction

36 37

Although iron comprises only 1-3% of the total mass in soils1, it has a profound influence on

38

the properties of soils. The high reactivity and large surface area of Fe-(oxyhydr)oxide minerals

39

influence many biogeochemical processes through the sorption of key nutrients or contaminants2-10.

40

When exposed to low oxygen conditions, Fe minerals serve as important electron acceptors for

41

microbial respiration11, leading to the dissolution of solid phases and release of sorbed and

42

incorporated constituents12, 13. This reductive dissolution of Fe has significant implications for the

43

global C cycle in soils where a substantial fraction of the total soil organic C oxidation is coupled to the

44

microbial reduction of Fe-(oxyhydr)oxides (e.g., 44% in humid tropical forest soils14), which can occur

45

in upland soils with Fe(II) production rates ranging from 0.5 to 50 mmol kg-1 d-1 15, 16. The availability

46

of Fe minerals toward reduction is strongly correlated with their surface area and particle size17, 18. The

47

crystallinity of Fe-(oxyhydr)oxides also strongly influences the kinetics of microbial Fe(III)

48

reduction19. Microbes can rapidly reduce ferrihydrite and other short-range ordered (SRO) Fe minerals

49

(typically within hours), but reduce well-ordered hematite (α-Fe2O3), goethite (α-FeOOH), and

50

lepidocrocite (γ-FeOOH) at slower rates (e.g., several months)18, 20.

51

In upland soils, Fe reduction takes place within microsites where high-levels of organic matter

52

combine with ephemeral water saturation and low O2 diffusion rates to yield anoxic conditions21, 22.

53

Since O2 often re-infuses such microsites as the soils dry, Fe(II) produced during the anoxic period via

54

microbial Fe reduction is often re-oxidized to Fe(III) solid phases. This Fe redox cycling can influence

55

the rapid cycling of other elements (e.g., C, N, and metals23). Redox fluctuations can drive repeated

56

dissolution and precipitation of Fe (oxyhydr)oxide minerals with potential influence on the crystallinity

57

of those phases21, 24-27. During anoxic periods, the sorption of the Fe2+(aq) ions to Fe mineral surfaces

3

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 29

58

can accelerate the process of Ostwald ripening from SRO to secondary crystalline mineral phases (e.g.,

59

goethite, magnetite) that are less chemically reactive and bioavailable than ferrihydrite26, 28, 29.

60

However, there are also studies illustrating SRO phases form or persist in fluctuating redox

61

environments30-34, indicating that the dynamics of the fluctuations are likely important in determining

62

the trajectory of mineral transformations.

63

Changes in the crystallinity of Fe (oxyhydr)oxides may also impact the fate of phosphorous (P)

64

in natural systems because Fe (oxyhydr)oxide surfaces have a great capacity for adsorbing P4, 5.

65

Ferrihydrite has a significantly higher sorption capacity than the crystalline (oxyhydr)oxides because of

66

its large surface area, but can form weaker surface complexes with aqueous species than the more

67

crystalline iron phases35-37. As a consequence, microbial Fe(III) reduction may simultaneously reduce

68

the overall capacity of soils to adsorb metal/metalloids while strengthening the remaining sorbate-

69

sorbent interactions.

70

Given the potential importance of Fe redox cycling for soil systems, the influence of redox

71

oscillations on Fe reduction rates needs to be quantified. A redox fluctuation can be characterized by

72

(a) the frequency of the redox oscillation; (b) the oscillation amplitude; and (c) the time spent under

73

oxic (τoxic) or anoxic (τanoxic) conditions. Here we hypothesize that shorter frequency oscillations will

74

lead to faster Fe reduction rates with associated changes in Fe-(oxyhydr)oxide mineral composition and

75

the distribution of phosphate. We test this by exposing a tropical soil to different redox oscillation

76

frequencies while holding the oscillation amplitude and ratio of τoxic: τanoxic constant. In order to

77

minimize spatial heterogeneity, soils were continuously mixed in slurries—separating the effects of

78

redox fluctuations from transport to the extent possible. We then measure changes in HCl-extractable

79

Fe(II), Fe mineral composition and P distribution, and simulate the experimental data in a numerical

80

model to calculate Fe reduction rates.

4

ACS Paragon Plus Environment

Page 5 of 29

Environmental Science & Technology

81 82

Methods

83 84 85

Field site and soil preparation Intact soil blocks were collected from the upper 10 cm of soil at a site similar to that described

86

by Peretyazhko and Sposito 38 in the Bisley watershed of the Luquillo Experimental Forest (LEF)

87

Puerto Rico (approximate GPS location 18.31553N,-65.74574W), which is part of the Luquillo Long-

88

term Ecological Research site and the Critical Zone Observatory. Samples were placed in 4-mil

89

polyethylene ziplock bags and transported without cooling inside an insulated container to the

90

University of Georgia. Within 24 h of sampling, all moist soils were processed by air-drying at 30°C

91

for 24 h followed by 2-mm sieving/homogenization (to remove large roots and gravel) under field soil

92

moisture conditions within a 95%:5% N2:H2 glovebox (Coy). All soils were stored at 22°C under oxic

93

conditions in the dark. Ginn et al. 39 showed these soils maintained the highest levels of microbial Fe-

94

reduction when air-dried prior to long-term storage. Soils in the LEF are classified predominately as

95

Ultisols formed from volcanic parent material with quartz diorite intrusions40, 41. The major primary

96

minerals consist of quartz and plagioclase feldspars with lesser amounts of biotite, hornblende, K-

97

feldspar, and accessory magnetite, sphene, apatite, and zircon42. Prior work by Silver et al.21 has

98

identified periodic fluctuations in soil O2 concentrations sufficient to generate anoxic conditions within

99

the soil. Further information on this soil collection and characterization are described elsewhere39, 43.

100 101

Redox oscillation experiments

102

Three oscillation treatments were performed simultaneously. In each experiment, 4.5 g (dry-

103

weight) of live soil (not sterilized) was suspended in a buffer solution containing 2 mM KCl and 10

104

mM solution of 2-(N-morpholino)ethanesulfonic acid (MES) buffer with a 45 g final suspension mass. 5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 29

105

The suspension was supplemented with 200 mg kg-1 of P (i.e. 6.46 mmol kg-1 soil and 0.58 mmol L-1

106

suspension). The slurries were incubated for 56 d in a glovebox (95% N2, 5% H2) on an end-over-end

107

shaker at 250 rpm. Although these experiments likely had sufficient labile organic carbon to support Fe

108

reduction rates (see below), the presence of H2 in the glovebox may have served as an additional

109

electron donor for Fe reduction44. In each of the experiments, the soil slurries were exposed to

110

alternating oxic-anoxic conditions at frequencies of 3.5, 7, and 14 d with the ratio of time under

111

oxic/anoxic conditions (τoxic:τanoxic) maintained at 1:6 for all treatments (i.e., the 3.5 d oscillation had 3

112

d of anoxic time and 0.5 d of oxic time; the 7 d period had 6 d anoxic and 1 d of oxic; and the 14 d

113

period had 12 d anoxic and 2 d oxic time). Samples were transferred between shaker tables inside and

114

outside of the glovebox to create anoxic and oxic conditions, respectively. Control soil slurries were

115

maintained at constant anoxic or oxic conditions for 56 d.

116

Samples were extracted at the beginning and end of each oscillation cycle using wide-orifice

117

pipette tips that allowed complete collection of soil particles in the slurry. Aqueous Fe(II) was

118

extracted from the soil slurries by centrifuging the samples at 11,000 rcf for 30 min, which will remove

119

particles > 40 nm based on Stokes’ law with an assumed mean particle density of 2.72 g cm-3 for

120

basalt-influenced soils45. Acid-extractable Fe(II) was solubilized by suspending the remaining pellet in

121

1.0 mL of 0.5 M HCl and shaking it for 2 h on a horizontal shaker at 150 rpm. The extracts were then

122

centrifuged at 11,000 rcf for 30 min and the supernatants analyzed for Fe(II) using a modified ferrozine

123

protocol24 on a Shimadzu-1700 UV-vis spectrophotometer at λ=562 nm. The dry weight of each pellet

124

from the centrifugation step was measured to correct for minor differences in suspension density

125

between treatments and allow accurate dry-weight normalization. All values are reported as averages of

126

triplicate measurements with error bars indicating standard deviation.

127 128

FeIII-reduction rate experiment 6

ACS Paragon Plus Environment

Page 7 of 29

129

Environmental Science & Technology

To quantify Fe(III) reduction rates midway through the experiment, additional samples were

130

oscillated at the three frequencies given above for 28 d and then exposed to 7 d of continuously anoxic

131

conditions during which they were sampled for aqueous and acid-extractable Fe(II) every 0.5 d as

132

described above. This data captured the iron reduction dynamics during a reducing period and was used

133

to parameterize the reduction kinetics in the numerical model described below.

134 135 136

Citrate-ascorbate extractable Fe To quantify the short-range-ordered solid Fe phases in the samples, we conducted citrate-

137

ascorbate extractions46 on freeze-dried soils from the unreacted samples or at the end of the oscillation

138

experiment (freeze-dried immediately upon completion of the experiment). Briefly, the samples were

139

suspended in a solution containing 0.2 M Na-citrate and 0.05 M ascorbic acid (pH = 6) at a 60:1

140

reagent:sample weight ratio. The suspensions were shaken for 16 h on a horizontal shaker and

141

centrifuged at 11,000 rcf for 30 min. The supernatants were analyzed for total Fe on an atomic

142

absorption (AA) spectrometer (Perkin Elmer, AAnalyst 200).

143 144

Mössbauer spectroscopy and X-ray absorption spectroscopy

145

Fe speciation of the soil samples was characterized via 57Fe Mössbauer spectroscopy and X-ray

146

absorption spectroscopy (XAS). Details on data collection, analysis, and interpretation are given in the

147

Supporting Information (SI) Sections 2 and 3.

148 149 150

Fe reduction assay: Shewanella incubations We quantified the availability of Fe(III) for microbial reduction by incubating the soil with a

151

culture of Shewanella oneidensis MR-1. The purpose of this inoculation experiment was to determine

152

whether redox fluctuations altered the iron oxide availability for microbial Fe-reduction. S. oneidensis 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 29

153

MR-1 was used because it is a good model Fe-reducers, capable of fast Fe-reduction and can live in

154

both oxic and anoxic conditions. This experiment is performed solely to assess Fe availability across

155

treatments, and is not designed to quantify the in situ iron reduction rate. The assay was performed on

156

each replicate of soils freeze-dried from the end of the three different oscillation frequencies, the

157

anaerobic and aerobic controls, and unreacted soil. Post-oxidation, freeze-dried soil was used to

158

maintain consistency across treatments. Shewanella oneidensis MR-1 was grown to stationary phase in

159

media containing (per liter): 0.5 g KH2PO4, 1.0 g Na2SO4, 2.0 g NH4Cl, 1.0 g yeast extract, 0.5 mM

160

CaCl2, 0.1 mM MgSO4, 10 mM Na-lactate, and 50 mM Fe-citrate. The cells were washed by

161

centrifugation at 3,000 rcf for 30 min and resuspended in fresh media without yeast extract. The

162

washing step was repeated twice to thoroughly remove spent media, after which 0.9 g of washed cell

163

culture was added to 2-mL centrifuge tubes containing 0.1 g of freeze-dried soil so each sample

164

contained the same ratio of bacterial culture to soil mass. The cultures were incubated for a week under

165

anoxic conditions. Freeze-dried soils with no S. oneidensis inoculation were used as controls. The

166

cultures were sampled for acid-extractable Fe(II) by pipetting 0.1 mL of sample from each culture and

167

centrifuging at 11,000 rcf for 30 min. The supernatants were poured off and the pellets were re-

168

suspended in 0.1 mL of 0.5 M HCl. The suspensions were shaken for 2 h on a horizontal shaker. The

169

extracts were then centrifuged at 11,000 rcf for 30 min. Samples were analyzed for acid-extractible

170

Fe(II) using the ferrozine protocol described above.

171 172 173

Phosphate extractions Sequential extractions were performed according to Hedley et al.47. A soil mass of 0.01 g was

174

sequentially extracted for 16 h (each) on a horizontal shaker with 0.6 mL of ultrapure water, 0.5 M

175

NaHCO3, 0.1 M NaOH, and 1 M HCl. After each extraction, the tubes were centrifuged at 11,000 rcf

176

for 30 min. In addition, we quantified microbial biomass P following Hedley and Stewert48. Briefly, 1 8

ACS Paragon Plus Environment

Page 9 of 29

Environmental Science & Technology

177

mL of CHCl3 was added to 0.01 g of soil in a centrifuge tube and capped. The solution was shaken

178

periodically over 30 min and then uncapped so that the CHCl3 could evaporate overnight. In all cases,

179

supernatants were analyzed for P using a molybdate blue protocol49 on a Shimadzu-1700

180

spectrophotometer at λ = 880 nm.

181 182

Phosphate adsorption isotherms

183

To assess the adsorption capacity of the soil, we constructed phosphate adsorption isotherms on

184

unamended soils. Briefly, 1 mL of 0, 80, 160, 320, and 640 ppm phosphate was added to 0.1 g of each

185

freeze-dried soil in a centrifuge tube. The suspensions were shaken on a horizontal shaker for 2 h and

186

then centrifuged for 30 min at 11,000 rcf. The supernatants were collected and measured as described

187

above.

188 189

Numerical model

190

Experiments were simulated with MATLAB (version R2015b) using a simple, parsimonious,

191

process-based kinetic model to describe and test mechanisms of microbial Fe(III) reduction, growth of

192

the Fe-reducing microbial community, abiotic Fe(II) oxidation by O2, and solid phase partitioning of Fe

193

minerals into distinct pools. Redox oscillations are imposed by setting O2 concentrations to 200 μM

194

during oxic periods (slightly below saturation) and 0 during anoxic periods, with Fe(II), Fe- reducing

195

microbial community, and a fast and a slowly reacting Fe(III) solid phase (FeOxfast and FeOxslow,

196

respectively) as state variables. During anoxic periods, the rate of microbial Fe(III) reduction is given

197

as:

198

Rred = ∑ ki [ BM ] i

199

[ FeOxi ] K m, i + [ FeOxi ]

Eq. 1

where ki is a cell specific rate constant, [BM] is the concentration of Fe reducers, and Km(i) is a half 9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 29

200

saturation constants and i represents the fast and slowly reacting Fe(III) solid phases, respectively. The

201

net growth rate of Fe-reducing microbes (RBM) is set proportional to the Fe reduction rate, RBM = γ Rred

202

, with a growth efficiency γ set to 0.06 (based on growth yields from Kostka et al.50). As the oxic

203

periods are relatively short in our experiments, no net growth or death of the Fe reducers is considered

204

during oxic periods. Under oxic conditions, Fe reduction is set to 0, and the oxidation of Fe(II) is

205

implemented following Stumm and Morgan51: 𝑅!" = 𝑘!" 𝐹𝑒 !! 𝑂!

206

Eq. 2

207

where the Fe(II) oxidation rate constant kox is set to 3 × 102 M-1 min-1, 1000 fold higher than the value

208

reported for homogeneous oxidation kinetics at the pH of 5.5 in our the experiments to accommodate

209

heterogeneous reaction rates52. Fast oxidation can promote the production of less ordered and more

210

reactive Fe oxides53, 54, thus we partition all the oxidized Fe(II) to the fast pool given the rapid

211

oxidation rate observed in our experiment. During our experiments, we routinely observed that a

212

portion of the HCl-extractable Fe(II) formed during the anoxic half-cycle is resistant to oxidation,

213

resulting in an apparent accumulation of Fe(II) over the course of the experiment. We represented this

214

in the numerical model by removing a small portion of the Fe(II) from the reactive pool, which we

215

hypothesize results from physical protection (encasement) against oxidation. The rate of Fe(II) encased

216

was set to 2% (q) of the rate of iron oxidation to fit the experimental data. Thus, the dynamics of fast

217

and slowly reacting Fe(III) solid phases are given by:

218 219

!!"#$!"#$ !" !!"#$!"#$ !"

!"!"

= 1 − 𝑞 𝑅!" − 𝑅!"#

Eq. 3

!"#$ = −𝑅!"#

Eq. 4

220

Initial reducible Fe(III) solid concentrations were based on the observed values of 10 and 140 mmol L-1

221

for fast and slowly reacting Fe oxides, respectively, with zero Fe(II). The fast pool represents Fe that

222

can be reduced in less than 3-d39; while the slow pool Fe is set to reproduce the Fe(II) plateau levels 10

ACS Paragon Plus Environment

Page 11 of 29

Environmental Science & Technology

223

observed in the oscillation experiments. To match the Fe reduction rates across all treatments, the two

224

model parameters representing kfast *[BM]t=0 and kslow*[BM]t=0 were set to 6 and 1.5 mmol kg-1 d-1 in

225

experiments without P additions, and 10 and 2.3 mmol kg-1 d-1 in experiments with P addition,

226

reflecting alleviation of P limitation. The fast and slow rate parameters were determined by fitting the

227

data from FeIII-reduction rate experiment conducted after 28-days of incubation (see above). All other

228

model parameters were identical across all simulations.

229 230

Results and Discussion

231 232 233

Fe reduction dynamics Fe(II) concentrations responded to changing redox conditions (Figure 1) with peak values at the

234

end of the anoxic periods and the lowest values after an oxic period. Peak Fe(II) concentrations

235

increased from 10 to 148-165 mmol kg-1 soil in the P-amended treatments as the experiment progressed

236

through successive redox cycles (Figure 1). The peak Fe(II) concentrations plateaued near 150 mmol

237

FeII kg-1 soil after about 1 month and increased only slowly for the remainder of the anoxic half-cycles.

238

The phosphate-amended treatments reached this ~150 mmol FeII kg-1 soil plateau at earlier oscillation

239

cycles than the unamended treatments, but by the end of the experiment all treatments exhibited similar

240

amplitudes of Fe(II) oscillation, except the 3.5 d unamended control soil (Figure 1).

241

The increase in peak Fe(II) concentrations was similar across most oscillation frequency

242

treatments, despite different anoxic incubation lengths (Figure SI-1). Thus, the time-averaged iron

243

reduction rate is lower in treatments with shorter oscillation frequencies, but the similar peak Fe(II)

244

concentrations between treatments suggests total Fe reduction extent is independent of oscillation

245

frequency. To assess how the frequency of redox oscillations affected the reduction kinetics, we

246

repeated our oscillation experiments (for the phosphorus addition treatments only) initiating a 11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 29

247

continuous 2-week anoxic incubation after one month of redox oscillations (Figure 2). By tracking the

248

accumulation of Fe(II) over time (Figure 2), we surmise there are two phases in the production of

249

Fe(II) during anoxic incubations: (1) an initial, rapid Fe(II) production interval followed by (2) an

250

interval with much slower Fe(II) production that brings the Fe(II) concentrations into proximity of the

251

peak Fe(II) concentrations at the end of each anoxic cycle (Figure 1). Irrespective of oscillation

252

frequency, a slow Fe-reduction phase (or plateau) occurs after 4 days of anoxic incubation. This plateau

253

could explain why all of the phosphate-amended soil slurries had similar peak Fe(II) concentrations

254

(even the 3.5-day treatment), as 90% of the plateau FeII concentration is achieved within 3 days of

255

incubation (Figure 1). Although the second experiment was only conducted with the P-amended soils,

256

we suspect this plateau may not manifest as prominently in the 3.5-d unamended treatment because the

257

overall slower Fe reduction rates in the unamended controls.

258

Our experiments were conducted on re-wetted soils that were oxidized without subsequent

259

drying during the redox oscillations. This experimental design minimized heterogeneity during the

260

experiment and tightened reproducibility between replicates, however it may have impacted the iron

261

dynamics in two important ways. First, the re-wetting of soils is well known to produce a flush of

262

organic carbon55, and thus our treatments likely contain more soluble, labile organic carbon than if we

263

had initiated the experiments from a field moist state. This likely provided an excess of electron donor

264

in the form of soluble organic carbon and it may have increased iron availability through complexation

265

or by facilitating electron shuttling relative to the field moist state56-60. Second, oxidizing the samples

266

under saturated conditions represents the common situation in moist soils or near-surface sediments61

267

in which pulses of O2 accompany the infiltration of rainwater—which has high dissolved O2 levels;

268

whereas is our design does not represent as well soil-drying events, where oxidation fronts follow the

269

major fluid flow paths and create a heterogeneous distribution of FeIII phases62.

270 12

ACS Paragon Plus Environment

Page 13 of 29

271 272

Environmental Science & Technology

Fe mineral composition To quantify the changes in the availability of Fe for microbial reduction63, we inoculated the

273

same freeze-dried soils (before and after treatment) with S. oneidensis MR-1 in a defined media to

274

quantify the rapidly reducible Fe(III) content (Figure 3). The purpose of this experiment was not to

275

simulate Fe-reduction that occurs in the field, but to determine the bioavailability of Fe after exposure

276

to fluctuating redox conditions. The S. oneidensis cultures produced FeII more rapidly following the

277

3.5-day, 7-day, 14-day, and anoxic control slurry incubations than in the oxic control and slurries with

278

soils prior to the oscillation experiments. We found Fe(III) reduction in the non-inoculated soils was

279

negligible compared to that occurring in the inoculated soils. Thus, outside of a possible stimulatory

280

effect of Shewanella on the growth of the native soil bacteria32, we assume the Fe(III)-reduction assays

281

were not biased by differences in soil microbial population or activity.

282

The 3.5-day, 7-day, 14-day, and anoxic control soils exhibited similarly higher rates of Fe(III)

283

reduction in the S. oneidensis assays than the oxic and initial soils (Figure 3, SI section 4). Although

284

each treatment had different redox oscillation periods, all are likely to have similar abundances of

285

recently oxidized Fe(III), as each oscillation treatment reached similar Fe(II) concentration plateaus in

286

their last anoxic cycle, whereas the oxic control and the starting soil had much lower HCl-extractable

287

Fe(II) concentrations. Since we terminated each treatment following an oxidation half cycle (including

288

the anoxic control), the rapid introduction of O2 oxidizes nearly all of the recently produced Fe2+(aq),

289

likely forming a very reactive, short-range-ordered (SRO) phase54.

290

However, we did not find significant changes in Fe mineral composition between treatments

291

using citrate-ascorbate chemical extractions (Figure S2), 57Fe Mössbauer spectroscopy (MBS, SI

292

section 2), or X-ray adsorption spectroscopy (XAS, SI section 3). This suggests one or more of the

293

following: (1) the proportional change in the SRO pool is too small relative to the total background soil

294

Fe in the soil for these chemical and spectroscopic measures; (2) the re-oxidized Fe is more proximal to 13

ACS Paragon Plus Environment

Environmental Science & Technology

295

the microbial iron reducers—perhaps precipitating as thin surface coatings on the original SRO-Fe

296

phases, while retaining similar local structure or by forming complexes with extracellular polymeric

297

substances on the bacterial surface64-66; (3) the re-oxidized Fe co-precipitated with organic electron

298

shuttling agents that can increase the availability for microbial reduction without yielding detectable

299

changes in mineral composition.

300

Page 14 of 29

Prior work from redox fluctuation experiments illustrates that Fe mineral crystallinity can

301

increase or decrease24, 32, 33, 67, or remain undetected68, although the specific conditions driving the

302

trajectory of mineral change are unclear. Documented increases in Fe mineral crystallinity might be

303

associated with slow or incomplete Fe oxidation. Liptzin and Silver67 found increases in Fe

304

crystallinity when exposing intact, un-homogenized soil samples to redox fluctuations and Thompson

305

et al.24 similarly found Fe crystallinity increases when redox fluctuating soil slurries were exposed to

306

progressive oxidation intended to approximate conditions within an aggregate. Likewise, if the initial

307

phase has very low crystallinity, then exposure to Fe2+(aq) during the anoxic periods may promote

308

increases in crystallinity as Tomaszewski et al.26 showed when exposing ferrihydrite to a redox

309

fluctuating environment. However, decreases in Fe crystallinity have also observed during redox

310

fluctuations, especially when accompanied with soil drying69, 70, or when the starting material is of

311

higher crystallinity32. Furthermore, there are other reports, akin to our study, where changes in Fe

312

crystallinity cannot be detected following redox fluctuations68.

313 314

Process identification

315

We identified process dynamics consistent with the observational data by implementing

316

different processes in a numerical model. Our implementation is able to reproduce the observed

317

increase in Fe(II) concentrations (aqueous plus 0.5M HCl extractable) during the anoxic cycles, as well

318

as the plateau phase revealed in our 2nd experiment (Figure 1). The basic structure of the model 14

ACS Paragon Plus Environment

Page 15 of 29

Environmental Science & Technology

319

includes two pools of iron with different cell-specific Fe reduction rates (kfast and kslow). This was

320

critical for reproducing the high Fe(II) production rates at the beginning of each anoxic period,

321

followed by slower Fe(II) production rates associated with reduction of the less reactive iron pool.

322

A key feature of the whole data set (P-amended and non-amended treatments) is that net Fe(II)

323

production rates increase over the course of the experiment (Figure 1). The initial rates of Fe reduction

324

were ~ 3 mmol Fe(II) kg-1 soil d-1, which are similar to rates measured previously on this soil39.

325

However, during the later half of the experiment, the Fe reduction rates increased to 17-66 mmol FeII

326

kg-1 soil d-1 across all treatments (Figure 4). This increase in peak Fe(II) concentrations over the course

327

of the experiment can be explained by the combined effect of an increased pool of reactive iron

328

oxide—formed during rapid oxidation of Fe(II) accumulated during the anoxic periods—and a minor

329

increase in the activity of the iron reducing community. We measured Fe reduction population densities

330

at the beginning and end of the experiment (Table S1) and found no difference within the log unit

331

precision of our methods, which—consistent with the model simulations—implies the Fe reducer

332

population increased less than tenfold over the course of the experiment. Fe(II) concentrations also

333

increase slightly during the oxic half-cycles over the course of the experiment, amounting to what

334

appears as a upward drifting Fe(II) baseline (Figure 1). This may reflect the formation of a mixed-

335

valance iron mineral, a stable Fe(II)-organic matter complex or an Fe sulfide phase, represented in the

336

model by the partitioning a portion of Fe(II) into a non-oxidizable form that is still extractable via 0.5M

337

HCl (see methods). Thus, in our final model, we assumed that the “active Fe”—defined as Fe that

338

could be mobilized via Fe reduction or other processes within the timescale of our experiment—moved

339

between four reservoirs: (1) a rapidly reducible Fe(III) solid; (2) a slowly reducible Fe(III) solid; (3)

340

dissolved or adsorbed Fe(II); and (4) a non-oxidizable Fe(II) phase. Both Fe(III) solid phases

341

contributed to the trend of increasing FeII peaks over the course of the oscillation experiment. During

342

each oxidation cycle, we partitioned all of the Fe(II) oxidized into the rapidly reducing Fe(III) solid 15

ACS Paragon Plus Environment

Environmental Science & Technology

343

phase, hence providing each subsequent simulated reduction cycle with a larger pool of rapidly

344

reducible Fe(III).

345

Page 16 of 29

An alternative explanation for the plateau in Fe(II) concentrations is that sorption of Fe(II) to Fe

346

mineral surfaces may pacify the Fe(III) toward further reductive dissolution. Roden and Zachara2

347

attributed a microbial Fe(III)-reduction plateau they observed in pure culture experiments to this effect.

348

However, in our experiment, peak Fe(II) concentrations clearly increase with subsequent oscillation

349

cycles, suggesting Fe reduction rates across all treatments increased as the experiment progressed

350

(Figure 4).

351

Imposing the above constraints and processes does not yield a model fit to all frequency

352

treatments unless the model parameters are tuned to each treatment. However, by imposing a short

353

suppression of simulated microbial growth at the beginning of each anoxic cycle—representing a

354

predominance of intracellular metabolic adjustments over the formation of new biomass—we can

355

achieve a consistent match with the experimental Fe(II) data. A single global parameter set matches the

356

observations across all oscillation frequencies (Figure 1), except that the P-amended treatments

357

required a higher cell-specific Fe reduction rate than the unamended treatments.

358

Considering first the non-amended treatments, the simulated peak rates of Fe reduction are

359

similar for the 3.5-d and 7-d treatments (peak rates of Fe(II) production ~ 30 mmol kg-1 soil d-1), but

360

substantially slower in the 14-d treatment (peak rates of Fe(II) production ~ 17 mmol kg-1 soil d-1).

361

Because the iron reduction rates are only high during the first ~ 4 days of each anoxic period (Figure

362

4), the 14-d treatment has longer period in which the microbes are largely inactive. With microbial

363

growth linked to metabolic activity, this leads to less growth. Over the course of the experiment, the

364

simulated biomass increased 4.3× in both the 3.5-d and 7-d treatments, compared with an increase of

365

only 2.7× in the 14-d treatment (Figure S3), which in turn leads to lower Fe(II) peak concentrations

16

ACS Paragon Plus Environment

Page 17 of 29

366

Environmental Science & Technology

(Figure 1).

367 368 369

Phosphorus considerations Augmenting the soil slurries with phosphate enhanced the production of Fe(II) during the

370

anoxic cycles (Figure 1), suggesting that microbial activity was stimulated. Consistent with this

371

hypothesis, the amount of phosphorus released following exposure to CHCl3 increased over the course

372

of the experiment (Figure S5), suggesting P accumulated in the microbial biomass. No other changes in

373

the phosphorus distributions (based on the Hedley protocol47) were observed in the treatments, with

374

nearly all of the P residing in the NaOH extractable pool (negligible P was found in the aqueous,

375

NaHCO3, and HCl extracts and thus were not reported). NaOH extractable P is considered to be

376

associated with Fe and Al oxide surfaces in the standard interpretation of the Hedley protocol47.

377

Previous measurements of Luquillo soils also found P to be predominately associated with mineral

378

surfaces38, 67. We found there was substantial potential for P uptake in these soils, but the amount and

379

shape of the phosphate adsorption isotherms (which followed typical Langmuir curves) did not differ

380

between treatments and controls (Figure S4, Table S2).

381

The phosphorus addition treatments had higher simulated rates of Fe reduction than the non-

382

amended controls, with peak rates of Fe(II) production ~66, ~52, and ~31 mmol kg-1 soil d-1 for the 3.5-

383

d, 7-d, and 14-d treatments, respectively. Higher cell-specific Fe reduction rate constants are required

384

to match the temporal evolution of Fe(II) oscillations in the P-amended treatments than are required in

385

the non-amended controls (an increase of 66.7% and 50.3% for fast and slowly reacting iron oxides,

386

respectively). Similar to the non-amended treatments, our simulations increase microbial biomass more

387

in the shorter frequency treatments as a direct consequence of their proportionally longer time with

388

access to rapidly reducible iron (Fe(III)RR) than the longer frequency treatments. Over the course of the

389

experiment, the simulated biomass in the P-amended treatments increased 7.0×, 5.1× and 3.2× in the 17

ACS Paragon Plus Environment

Environmental Science & Technology

390

Page 18 of 29

3.5-d, 7-d, and 14-d treatments, respectively.

391

This may be the first study illustrating that additions of P promote an increase in Fe(II)

392

production rates in native soils or sediments however, there is a large body of work based on synthetic

393

minerals. In these pure-system studies, addition of phosphate often alters the rate of Fe(II) production,

394

although the direction and magnitude of change is inconsistent, depending on P concentrations, Fe

395

mineral composition and other factors71-78. Our P-amended soils had 6.46 mmol kg-1 soil (or 0.58 mmol

396

L-1 suspension) of P added, which is lower than most of the P amendments used in prior synthetic iron

397

mineral studies—typical P amendments are 2 - 20 mmol L-1 in the suspension74, 75, with 0.4 mmol L-1

398

often considered a low P treatment73. Thus, given the low amount of P added in our treatments, we

399

suspect the increase in Fe(II) production rates results from an alleviation a P limitation on microbial

400

growth, which is consistent with P uptake to the biomass (Figure S5).

401 402

Ecosystem Implications

403

Upland ecosystems are subjected to short periods of anoxia that often lead to some ephemeral

404

iron reduction, which can influence C cycling and the fate of many nutrients and chemicals. We have

405

shown that successive periods of redox fluctuations over a month timescale with ample organic C can

406

increase soil Fe reduction rates, regardless of fluctuation frequency. Thus, Fe reduction rates during the

407

onset of anoxia are strongly coupled to physiochemical dynamics of the recent past. Terrestrial systems

408

exposed to rain events that yield iron reducing conditions and a build-up of Fe(II) thus might be primed

409

for more rapid iron reduction rates than systems emerging from a long drought or with continuously

410

wet soils.

411

The amount of Fe(II) produced following a prolonged anoxic incubation is often predicted

412

through selective extractions (oxalate, citrate-ascorbate, hydroxylamine, etc.) targeting the short-range-

413

ordered (SRO) phases. Our work is consistent with this in that the total amount of Fe(II) produced 18

ACS Paragon Plus Environment

Page 19 of 29

Environmental Science & Technology

414

during an anoxic cycle later in the experiment (or in the anoxic control) was similar to the citrate-

415

ascorbate extractable iron (~ 30% of total Fe). However, SRO abundance is a poor predictor of iron

416

reduction rate, because the SRO abundance does not change throughout the experiment, while iron

417

reduction rate changes dramatically. Instead, we propose the availability of SRO phases for Fe

418

reduction is enhanced through rapid oxidation events by atmospheric air (~21% O2), creating a sub-set

419

of rapidly reducible SRO-Fe [Fe(III)RR]. We found Shewanella cultures could reduce Fe(III) much

420

faster from soils exposed to redox fluctuations than from the oxic controls, however we could not

421

detect any corresponding shifts in mineral composition via selective extractions, Mössbauer

422

spectroscopy, or X-ray adsorption spectroscopy. Regardless, by the later half of the experiment nearly

423

the entire SRO Fe pool undergoes reduction within four days.

424

Soils that are frequently exposed to redox fluctuations have a large pool of rapidly reducible

425

iron to transfer electrons to during periods of oxygen depletion—a biogeochemical battery 79 that is

426

recharged by rapid oxidation of previously anoxic pore waters. Our work highlights the importance of

427

identifying the reaction timescales and limiting reactants/catalysts (microbes) in predicting iron

428

reduction rates in soils. Prior work suggests this soil does not have significant C- limitations 39, but we

429

did find the soil responded to P-amendments with more rapid increases in iron reduction rate than

430

unamended soils. The P-amended treatments required a faster cell-specific Fe reduction rate when P

431

was added and suggests iron reducer activity (population growth, CO2 production, etc.) is higher when

432

redox fluctuations occur on timescales similar to the process timescales (length of time to oxidize most

433

of the Fe(II) or the length of time to a Fe(II) concentration pseudo-plateau). Soils that are exposed to

434

near-weekly redox pulsing (due to variable rain events or water-table shifts) may therefore be more

435

dominated by iron biogeochemical dynamics than soils that experience redox shifts only seasonally.

436

19

ACS Paragon Plus Environment

Environmental Science & Technology

437

ASSOCIATED CONTENT

438

Supporting Information

439

Four supporting information sections are included: Section 1 contains two additional tables and five

440

additional figures cited in the main text; Section 2 contains a detailed description of the Mössbauer

441

(MBS) characterization of soil Fe; Section 3 contains a detailed description of the XAS

442

characterization of soil Fe and a comparison of the XAS and MBS findings; Section 4 contains a

443

description of the statistical test of Shewanella sp. assay. This material is available free of charge via

444

the Internet at http://pubs.acs.org.

445 446

AUTHOR INFORMATION

447

Corresponding Author

448

*E-mail: [email protected].

449

Notes

450

The authors declare no competing financial interest.

451 452 453 454

ACKNOWLEDGEMENTS We thank Whendee Silver for providing access to the Bisley site; Michelle Scherer, Drew Latta, and

455

Tim Pasakarnis for Mössbauer measurements; Chris Gorski and Prachi Joshi for preliminary

456

electrochemical analyses of these soils; Qing Ma (APS 5-BM-D) and Ryan Davis (SSRL 4-1) for

457

assistance with XAS analysis; and Nehru Mantripragada for laboratory assistance. This work was

458

funded by USDA - NIFA Soil Processes Program AFRI- NIFA Grant no. 2009-65107-05830 to AT

459

and CM; NSF grants EAR-1331841, EAR-1053470, EAR- 1451508, and DEB-1457761 to AT; and

460

NSF OCE-1559087 to YT. Portions of this research were conducted at the Stanford Synchrotron 20

ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29

461

Environmental Science & Technology

Radiation Lightsource (SSRL) and the Advanced Photon Source (APS).

462 463

21

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 29

464

References

465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517

(1) Loeppert, R. H.; Inskeep, W. P., Iron. In Methods of soil analysis: Part 3-Chemical methods, Sparks, D. L., Ed. Soil Science Society of America: 1996; pp 639-664. (2) Roden, E. E.; Zachara, J. M., Microbial reduction of crystalline iron(III) oxides: Influenceof oxide surface area and potential for cell growth. Env. Sci. Technol. 1996, 30, 1618-1628. (3) Frenzel, P.; Bosse, U.; Janssen, P. H., Rice roots and methanogenesis in a paddy soil: Ferric iron as an alternative electron acceptor in the rooted soil. Soil Biol. Biochem. 1999, 31, 421-430. (4) Ann, Y.; Reddy, K. R.; Delfino, J. J., Influence of redox potential on phosphorus solubility in chemically amended wetland organic soils. Ecol. Eng. 2000, 14, 169-180. (5) Miller, A. J.; Schuur, E. A. G.; Chadwick, O. A., Redox control of phosphorus pools in hawaiian montane forest soils. Geoderma 2001, 102, 219-237. (6) Fimmen, R.; Richter, D.; Vasudevan, D.; Williams, M.; West, L., Rhizogenic Fe–C redox cycling: A hypothetical biogeochemical mechanism that drives crustal weathering in upland soils. Biogeochemistry 2008, 87, 127. (7) Sulzberger, B.; Suter, D.; Siffert, C.; Banwart, S.; Stumm, W., Dissolution of Fe(III) (hydr)oxides in natural waters; laboratory assessment on the kinetics controlled by surface coordination. Mar. Chem. 1989, 28, 127-144. (8) Bruemmer, G. W.; Gerth, J.; Tiller, K. G., Reaction kinetics of the adsorption and desorption of nickel, zinc and cadmium by goethite. I. Adsorption and diffusion of metals. J. Soil Sci. 1988, 39, 37-52. (9) Sparks, D. L., Kinetics of soil chemical processes. Academic Press: 2013. (10) Spadini, L.; Manceau, A.; Schindler, P. W.; Charlet, L., Structure and stability of Cd2+ surface complexes on ferric oxides: 1. Results from EXAFS spectroscopy. J. Colloid Interf. Sci. 1994, 168, 73-86. (11) Canfield, D. E.; Thamdrup, B.; Hansen, J. W., The anaerobic degradation of organic-matter in danish coastal sediments: Iron reduction, manganese reduction, and sulfate reduction. Geochim. Cosmochim. Acta 1993, 57, 3867-3883. (12) Lovley, D. R., Organic-matter mineralization with the reduction of ferric iron - a review. Geomicrobiol. J. 1987, 5, 375-399. (13) McBride, M. B., Environmental chemistry of soils. Oxford University Press: NY, 1994. (14) Dubinsky, E. A.; Silver, W. L.; Firestone, M. K., Tropical forest soil microbial communities couple iron and carbon biogeochemistry. Ecology 2010, 91, 2604-2612. (15) Yang, W. H.; Liptzin, D., High potential for iron reduction in upland soils. Ecology 2015, 96, 2015-2020. (16) Hall, S. J.; Liptzin, D.; Buss, H. L.; DeAngelis, K.; Silver, W. L., Drivers and patterns of iron redox cycling from surface to bedrock in a deep tropical forest soil: A new conceptual model. Biogeochemistry 2016, 130, 177-190. (17) Bonneville, S.; Behrends, T.; Van Cappellen, P., Solubility and dissimilatory reduction kinetics of iron(III) oxyhydroxides: A linear free energy relationship. Geochim. Cosmochim. Acta 2009, 73, 5273-5282. (18) Bonneville, S.; Van Cappellen, P.; Behrends, T., Microbial reduction of iron(III) oxyhydroxides: Effects of mineral solubility and availability. Chem. Geol. 2004, 212, 255-268. (19) Roden, E. E.; Wetzel, R. G., Organic carbon oxidation and suppression of methane production by microbial Fe(III) oxide reduction in vegetated and unvegetated freshwater wetland sediments. Limnol. Oceanogr. 1996, 41, 1733-1748. (20) Roden, E. E., Geochemical and microbiological controls on dissimilatory iron reduction. C. R. Geosci. 2006, 338, 456-467. (21) Silver, W. L.; Lugo, A. E.; Keller, M., Soil oxygen availability and biogeochemistry along rainfall and topographic gradients in upland wet tropical forest soils. Biogeochemistry 1999, 44, 301-328. (22) Sexstone, A. J.; Revsbech, N. P.; Parkin, T. B.; Tiedje, J. M., Direct measurement of oxygen profiles and denitrification rates in soil aggregates. Soil Sci. Soc. Am. J. 1985, 49, 645-651. (23) Roden, E. E., Analysis of long-term bacterial vs. Chemical Fe(III) oxide reduction kinetics. Geochim. Cosmochim. Acta 2004, 68, 3205-3216. (24) Thompson, A.; Chadwick, O. A.; Rancourt, D. G.; Chorover, J., Iron-oxide crystallinity increases during soil redox oscillations. Geochim. Cosmochim. Acta 2006, 70, 1710-1727. (25) Mejia, J.; Roden, E. E.; Ginder-Vogel, M., Influence of oxygen and nitrate on Fe (hydr) oxide mineral transformation and soil microbial communities during redox cycling. Env. Sci. Technol. 2016, 50, 3580-3588. (26) Tomaszewski, E. J.; Cronk, S. S.; Gorski, C. A.; Ginder-Vogel, M., The role of dissolved Fe (II) concentration in the mineralogical evolution of Fe (hydr) oxides during redox cycling. Chem. Geol. 2016, 438, 163-170. (27) Vogelsang, V.; Kaiser, K.; Wagner, F. E.; Jahn, R.; Fiedler, S., Transformation of clay-sized minerals in soils exposed to prolonged regular alternation of redox conditions. Geoderma 2016, 278, 40-48. (28) Zachara, J. M.; Kukkadapu, R. K.; Fredrickson, J. K.; Gorby, Y. A.; Smith, S. C., Biomineralization of poorly

22

ACS Paragon Plus Environment

Page 23 of 29

518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573

Environmental Science & Technology

crystalline Fe(III) oxides by dissimilatory metal reducing bacteria (DMRB). Geomicrobiol. J. 2002, 19, 179-207. (29) Hansel, C. M.; Benner, S. G.; Neiss, J.; Dohnalkova, A.; Kukkadapu, R. K.; Fendorf, S., Secondary mineralization pathways induced by dissimilatory iron reduction of ferrihydrite under advective flow. Geochim. Cosmochim. Acta 2003, 67, 2977-2992. (30) Chadwick, O. A.; Gavenda, R. T.; Kelly, E. F.; Ziegler, K.; Olson, C. G.; Elliott, W. C.; Hendricks, D. M., The impact of climate on the biogeochemical functioning of volcanic soils. Chem. Geol. 2003, 202, 195-223. (31) Chorover, J.; Amistadi, M. K.; Chadwick, O. A., Surface charge evolution of mineral-organic complexes during pedogenesis in hawaiian basalt. Geochim. Cosmochim. Acta 2004, 68, 4859-4876. (32) Coby, A. J.; Picardal, F.; Shelobolina, E.; Xu, H.; Roden, E. E., Repeated anaerobic microbial redox cycling of iron. Appl. Environ. Microbiol. 2011, 77, 6036-6042. (33) Li, Y.; Yu, S.; Strong, J.; Wang, H., Are the biogeochemical cycles of carbon, nitrogen, sulfur, and phosphorus driven by the “FeIII–FeII redox wheel” in dynamic redox environments? J. Soil Sed. 2012, 12, 683-693. (34) Thompson, A.; Rancourt, D. G.; Chadwick, O. A.; Chorover, J., Iron solid-phase differentiation along a redox gradient in basaltic soils. Geochim. Cosmochim. Acta 2011, 75, 119-133. (35) Dzombak, D. A.; Morel, F. M. M., Surface complexation modeling: Hydrous ferric oxide. John Wiley & Sons: 1990. (36) Axe, L.; Anderson, P. R., Experimental and theoretical diffusivities of Cd and sr in hydrous ferric oxide. J. Colloid Interf. Sci. 1997, 185, 436-448. (37) Trivedi, P.; Axe, L., Modeling Cd and Zn sorption to hydrous metal oxides. Env. Sci. Technol. 2000, 34, 2215-2223. (38) Peretyazhko, T.; Sposito, G., Iron(III) reduction and phosphorous solubilization in humid tropical forest soils. Geochim. Cosmochim. Acta 2005, 69, 3643. (39) Ginn, B. R.; Habteselassie, M. Y.; Meile, C.; Thompson, A., Effects of sample storage on microbial Fe-reduction in tropical rainforest soils. Soil Biol. Biochem. 2014, 68, 44-51. (40) Beinroth, F. H., Some highly weathered soils of Puerto Rico, 1. Morphology, formation and classification. Geoderma 1982, 27, 1-73. (41) Frizano, J.; Johnson, A. H.; Vann, D. R.; Scatena, F. N., Soil phosphorus fractionation during forest development on landslide scars in the luquillo mountains, Puerto Rico. Biotropica 2002, 34, 17-26. (42) White, A. F.; Blum, A. E.; Schulz, M. S.; Vivit, D. V.; Stonestrom, D. A.; Larsen, M.; Murphy, S. F.; Eberl, D., Chemical weathering in a tropical watershed, luquillo mountains, Puerto Rico: I. Long-term versus short-term weathering fluxes. Geochim. Cosmochim. Acta 1998, 62, 209-226. (43) Tishchenko, V.; Meile, C.; Scherer, M. M.; Pasakarnis, T. S.; Thompson, A., Fe2+ catalyzed iron atom exchange and re-crystallization in a tropical soil. Geochim. Cosmochim. Acta 2015, 148, 191-202. (44) Achtnich, C.; Bak, F.; Conrad, R., Competition for electron donors among nitrate reducers, ferric iron reducers, sulfate reducers, and methanogens in anoxic paddy soil. Biol. Fertil. Soils 1995, 19, 65-72. (45) Thompson, A.; Chadwick, O. A.; Boman, S.; Chorover, J., Colloid mobilization during soil iron redox oscillations. Env. Sci. Technol. 2006, 40, 5743-5749. (46) Reyes, I.; Torrent, J., Citrate-ascorbate as a highly selective extractant for poorly crystalline iron oxides. Soil Sci. Soc. Am. J. 1997, 61, 1647-1654. (47) Hedley, M. J.; Stewart, J. W. B.; Chauhan, B., Changes in inorganic and organic soil phosphorus fractions induced by cultivation practices and by laboratory incubations. Soil Sci. Soc. Am. J. 1982, 46, 970-976. (48) Hedley, M. J.; Stewart, J. W. B., Method to measure microbial phosphate in soils. Soil Biol. Biochem. 1982, 14, 377385. (49) Murphy, J.; Riley, J. P., A modified single solution method for the determination of phosphate in natural waters. Analytica chimica acta 1962, 27, 31-36. (50) Kostka, J. E.; Dalton, D. D.; Skelton, H.; Dollhopf, S.; Stucki, J. W., Growth of iron(III)-reducing bacteria on clay minerals as the sole electron acceptor and comparison of growth yields on a variety of oxidized iron forms. Appl. Environ. Microbiol. 2002, 68, 6256-6262. (51) Stumm, W.; Morgan, J. J., Aquatic chemistry: Chemical equilibrium and rates in natural waters. 3rd ed.; John Wiley & Sons, Inc.: New York, NY, 1996; p 1022. (52) Van Beek, C.; Hiemstra, T.; Hofs, B.; Nederlof, M.; Van Paassen, J.; Reijnen, G., Homogeneous, heterogeneous and biological oxidation of iron (II) in rapid sand filtration. J. Water Supply Res. T. 2012, 61, 1-13. (53) Senko, J. M.; Dewers, T. A.; Krumholz, L. R., Effect of oxidation rate and Fe (II) state on microbial nitratedependent Fe (III) mineral formation. Appl. Environ. Microbiol. 2005, 71, 7172-7177. (54) Steefel, C. I.; Van Cappellen, P., A new kinetic approach to modeling water-rock interaction: The role of nucleation, precursors, and ostwald ripening. Geochim. Cosmochim. Acta 1990, 54, 2657-2677. (55) Lundquist, E. J.; Jackson, L. E.; Scow, K. M., Wet–dry cycles affect dissolved organic carbon in two california agricultural soils. Soil Biol. Biochem. 1999, 31, 1031-1038.

23

ACS Paragon Plus Environment

Environmental Science & Technology

574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629

Page 24 of 29

(56) Kleber, M.; Eusterhues, K.; Keiluweit, M.; Mikutta, C.; Mikutta, R.; Nico, P. S., Mineral–organic associations: Formation, properties, and relevance in soil environments. In Advances in agronomy, Donald, L. S., Ed. Academic Press: 2015; Vol. 130, pp 1-140. (57) Kappler, A.; Newman, D. K., Formation of Fe(III)-minerals by Fe(II)-oxidizing photoautotrophic bacteria. Geochim. Cosmochim. Acta 2004, 68, 1217-1226. (58) Taillefert, M.; Beckler, J. S.; Carey, E.; Burns, J. L.; Fennessey, C. M.; DiChristina, T. J., Shewanella putrefaciens produces an Fe(III)-solubilizing organic ligand during anaerobic respiration on insoluble Fe(III) oxides. J. Inorg. Biochem. 2007, 101, 1760-1767. (59) Royer, R. A.; Burgos, W. D.; Fisher, A. S.; Unz, R. F.; Dempsey, B. A., Enhancement of biological reduction of hematite by electron shuttling and Fe(II) complexation. Env. Sci. Technol. 2002, 36, 1939-1946. (60) Royer, R. A.; Burgos, W. D.; Fisher, A. S.; Jeon, B.-H.; Unz, R. F.; Dempsey, B. A., Enhancement of hematite bioreduction by natural organic matter. Env. Sci. Technol. 2002, 36, 2897-2904. (61) Couture, R.-M.; Charlet, L.; Markelova, E.; Madé, B. t.; Parsons, C. T., On–off mobilization of contaminants in soils during redox oscillations. Env. Sci. Technol. 2015, 49, 3015-3023. (62) Schulz, M.; Stonestrom, D.; Lawrence, C.; Bullen, T.; Fitzpatrick, J.; Kyker-Snowman, E.; Manning, J.; Mnich, M., Structured heterogeneity in a marine terrace chronosequence: Upland mottling. Vadose Zone J. 2016, 15. (63) Eusterhues, K.; Neidhardt, J.; Hädrich, A.; Küsel, K.; Totsche, K., Biodegradation of ferrihydrite-associated organic matter. Biogeochemistry 2014, 119, 45-50. (64) Warren, L. A.; Ferris, F. G., Continuum between sorption and precipitation of Fe (III) on microbial surfaces. Env. Sci. Technol. 1998, 32, 2331-2337. (65) Frankel, R. B.; Bazylinski, D. A., Biologically induced mineralization by bacteria. Rev. Mineral. Geochem. 2003, 54, 95-114. (66) Mikutta, C.; Mikutta, R.; Bonneville, S.; Wagner, F.; Voegelin, A.; Christl, I.; Kretzschmar, R., Synthetic coprecipitates of exopolysaccharides and ferrihydrite. Part I: Characterization. Geochim. Cosmochim. Acta 2008, 72, 11111127. (67) Liptzin, D.; Silver, W. L., Effects of carbon additions on iron reduction and phosphorus availability in a humid tropical forest soil. Soil Biol. Biochem. 2009, 41, 1696-1702. (68) Parsons, C. T.; Couture, R. M.; Omoregie, E. O.; Bardelli, F.; Greneche, J. M.; Roman-Ross, G.; Charlet, L., The impact of oscillating redox conditions: Arsenic immobilisation in contaminated calcareous floodplain soils. Environ. Pollut. 2013, 178, 254-263. (69) McGeehan, S. L.; Fendorf, S. E.; Naylor, D. V., Alteration of arsenic sorption in flooded-dried soils. Soil Sci. Soc. Am. J. 1998, 62, 828-833. (70) Mansfeldt, T.; Schuth, S.; Häusler, W.; Wagner, F.; Kaufhold, S.; Overesch, M., Iron oxide mineralogy and stable iron isotope composition in a gleysol with petrogleyic properties. J. Soil Sed. 2012, 12, 97-114. (71) Fredrickson, J. K.; Zachara, J. M.; Kennedy, D. W.; Dong, H. L.; Onstott, T. C.; Hinman, N. W.; Li, S. M., Biogenic iron mineralization accompanying the dissimilatoryreduction of hydrous ferric oxide by a groundwater bacterium. Geochim. Cosmochim. Acta 1998, 62, 3239-3257. (72) Zachara, J. M.; Fredrickson, J. K.; Li, S. M.; Kennedy, D. W.; Smith, S. C.; Gassman, P. L., Bacterial reduction of crystalline Fe3+ oxides in single phasesuspensions and subsurface materials. Am. Mineral. 1998, 83, 1426-1443. (73) Glasauer, S.; Weidler, P. G.; Langley, S.; Beveridge, T. J., Controls on Fe reduction and mineral formation by a subsurface bacterium. Geochim. Cosmochim. Acta 2003, 67, 1277-1288. (74) Kukkadapu, R. K.; Zachara, J. M.; Fredrickson, J. K.; Kennedy, D. W., Biotransformation of two-line silicaferrihydrite by a dissimilatory Fe(III)-reducing bacterium: Formation of carbonate green rust in the presence of phosphate. Geochim. Cosmochim. Acta 2004, 68, 2799-2814. (75) Borch, T.; Masue, Y.; Kukkadapu, R. K.; Fendorf, S., Phosphate imposed limitations on biological reduction and alteration of ferrihydrite. Env. Sci. Technol. 2007, 41, 166-172. (76) O’Loughlin, E. J.; Boyanov, M. I.; Flynn, T. M.; Gorski, C. A.; Hofmann, S. M.; McCormick, M. L.; Scherer, M. M.; Kemner, K. M., Effects of bound phosphate on the bioreduction of lepidocrocite (γ-FeOOH) and maghemite (γ-Fe2O3) and formation of secondary minerals. Env. Sci. Technol. 2013, 47, 9157-9166. (77) O’Loughlin, E. J.; Gorski, C. A.; Scherer, M. M.; Boyanov, M. I.; Kemner, K. M., Effects of oxyanions, natural organic matter, and bacterial cell numbers on the bioreduction of lepidocrocite (γ-FeOOH) and the formation of secondary mineralization products. Env. Sci. Technol. 2010, 44, 4570-4576. (78) Zachara, J. M.; Kukkadapu, R. K.; Peretyazhko, T.; Bowden, M.; Wang, C. M.; Kennedy, D. W.; Moore, D.; Arey, B., The mineralogic transformation of ferrihydrite induced by heterogeneous reaction with bioreduced anthraquinone disulfonate (AQDS) and the role of phosphate. Geochim. Cosmochim. Acta 2011, 75, 6330-6349. (79) Byrne, J. M.; Klueglein, N.; Pearce, C.; Rosso, K. M.; Appel, E.; Kappler, A., Redox cycling of Fe (II) and Fe (III) in

24

ACS Paragon Plus Environment

Page 25 of 29

630 631

Environmental Science & Technology

magnetite by Fe-metabolizing bacteria. Science 2015, 347, 1473-1476.

632

25

ACS Paragon Plus Environment

Environmental Science & Technology

633

Page 26 of 29

Figure 1

634

Unamended

635 636 637 638 639 640 641 642 643 644

40

0

14

7d

160

28

Days

120 80 40

200

0

7

Days

120 80 40 0

0

7

200

Phosphate amended 14 d

160

Figure 1: Fluctuations and model

120 80 40 200 56

0

14

7d

160

28

42

56

Days

120 80 40

200 14 21 28 35 42 49 56

3.5 d

160

42

2+

200

2+

80

0.5 M HCl extractable Fe -1 mmol kg

120

0.5 M HCl extractable Fe -1 mmol kg

2+

14 d

160

0.5 M HCl extractable Fe -1 mmol kg

0.5 M HCl extractable Fe

0.5 M HCl extractable Fe -1 mmol kg

2+

2+

(mmol kg )

0.5 M HCl extractable Fe 0.5 M HCl extractable Fe 2+ -1 -1 -1 mmol kg mmol kg

2+

200

0

160

7

14 21 28 35 42 49 56

3.5 d

Days

120 80 40

0 14 21 28 35 42 49 56

0

7

Days

14 21 28 35 42 49 56

Days

Figure 1. Fluctuations in FeII concentrations due to redox oscillations with periods of 3.5 d (bottom), 7 d (middle) or 14 d (top). All treatments have the same oxic:anoxic exposure time ratio of 6:1 and the oxic cycles involve exposure to 21% O2. Graphs on the left side are for un-amended slurries and those on the right are for treatments with added phosphorous. Symbols indicate measured values with 1 s.d. error bars based on three replicates. The solid lines represent the numerical model.

26

ACS Paragon Plus Environment

Page 27 of 29

645 646

Environmental Science & Technology

Figure 2

647

648 649 650 651 652 653 654

Figure 2. FeII-concentrations under anoxic conditions following 28 d of redox oscillations with periods of 3.5 d, 7 d, or 14 d in treatments amended with phosphorous. Error bars are 1 s.d. This data was used to fit the dynamics during the reduction period in the numerical model across all treatments.

27

ACS Paragon Plus Environment

Environmental Science & Technology

655

Page 28 of 29

Figure 3

Oscillations & anoxic control after oxidation

Unreacted & oxic control after oxidation

656 657 658 659 660 661 662 663 664 665

Figure 3 Figure 3. FeII concentrations in a liquid culture of S. oneidensis MR-1 incubated with the sole source of ferric iron deriving from freeze-dried soil samples from the indicated redox oscillation treatments and controls (all of which ended with oxic exposure, including the anoxic control). Each data point is the difference in acid-extractable FeII concentrations between inoculated soils and the non-inoculated controls. Error bars are 1 s.d. See SI-Section 4 for more details.

28

ACS Paragon Plus Environment

Page 29 of 29

666

Environmental Science & Technology

Figure 4

667

Phosphate Amended 80

Net Fe(II) Production Rate -1 mmol kg

40

14 d

30 20 10

Net Fe(II) Production Rate -1 -1 mmol kg d

668 669 670 671 672 673 674 675 676

Net Fe(II) Production Rate -1 -1 mmol kg d

Net Fe(II) Production Rate -1 -1 mmol kg d

40

7d

30 20 10 40

3.5 d

30 20 10 0

0

7

14 21 28 35 42 49 56

Net Fe(II) production rate -1 -1 mmol kg d

Net Fe(II) Production Rate -1 -1 mmol kg d

Unamended

14 d

60 40 20 80

7d

60 40 20

80

3.5 d

60 40 20 0

0

7

14 21 28 35 42 49 56

Days

Days

Figure 4 Figure 4. Net Fe(II) production rates calculated at each time point from the numerical model for the unamended (left graphs) and P-amended (right graphs) treatments. Values shown only for anoxic periods.

29

ACS Paragon Plus Environment