Subscriber access provided by UNIV AUTONOMA DE COAHUILA UADEC
Characterization of Natural and Affected Environments
Biochar-mediated Anaerobic Oxidation of Methane Xueqin Zhang, Jun Xia, Jiaoyang Pu, Chen Cai, Gene W. Tyson, Zhiguo Yuan, and Shihu Hu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b01345 • Publication Date (Web): 17 May 2019 Downloaded from http://pubs.acs.org on May 19, 2019
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 28
Environmental Science & Technology
Biochar-mediated Anaerobic Oxidation of Methane
1 2 3
Xueqin Zhang1, Jun Xia1, Jiaoyang Pu1, Chen Cai1, Gene W Tyson2, Zhiguo Yuan1 & Shihu
4
Hu1,*
5 6
1Advanced
7
Technology, The University of Queensland, St. Lucia, Queensland 4072, Australia
8
2Australian
9
University of Queensland, Brisbane, QLD, Australia
Water Management Centre, Faculty of Engineering, Architecture and Information
Centre for Ecogenomics, School of Chemistry and Molecular Biosciences, The
10 11
Corresponding Author
12
* Tel: +61 7 3346 3230. Fax: +61 7 336 54726. Email:
[email protected] 1
ACS Paragon Plus Environment
Environmental Science & Technology
13
Table of Contents (TOC) Art
14
2
ACS Paragon Plus Environment
Page 2 of 28
Page 3 of 28
Environmental Science & Technology
15
ABSTRACT: Biochar was recently identified as an effective soil amendment for CH4 capture.
16
Corresponding mechanisms are currently recognized to be from physical properties of biochar
17
providing a favourable growth environment for aerobic methanotrophs which perform aerobic
18
methane (CH4) oxidation. However, our study shows that the chemical reactivity of biochar
19
can also stimulate anaerobic oxidation of CH4 (AOM) by anaerobic methanotrophic archaea
20
(ANME) of ANME-2d, which proposes another plausible mechanism for CH4 mitigation by
21
biochar amendment in anaerobic environments. It was found that by adding biochar as the sole
22
electron acceptor in an anaerobic environment, CH4 was biologically oxidised, with CO2
23
production of 106.3 ± 5.1 μmol g-1 biochar. In contrast, limited CO2 production was observed
24
with chemically reduced biochar amendment. This biological nature of the process was
25
confirmed by mcr gene transcript abundance as well as sustained dominance of ANME-2d in
26
the microbial community during microbial incubations with active biochar amendment.
27
Combined FTIR and XPS analyses demonstrated that the redox activity of biochar relates to
28
its oxygen-based functional groups. Based on microbial community evolution as well as
29
intermediates production during incubation, different pathways in terms of direct or indirect
30
interactions between ANME-2d and biochar were proposed for biochar-mediated AOM.
3
ACS Paragon Plus Environment
Environmental Science & Technology
31
INTRODUCTION
32
Environmental and social concerns over global methane (CH4) emissions relate to its high
33
global warming potential (GWP) which is 25 times that of CO2 over a period of 100 years.1
34
Most CH4 released to the atmosphere (69%) is microbially derived, from natural sources
35
including oceans, wetlands, sediments, soils, and from anthropogenic sources such as rice
36
paddies, sewage treatment plants, landfill.2 In an attempt to mitigate the adverse effects on
37
climate caused by increased CH4 emission, efforts have targeted low-level CH4 emissions,
38
mainly from anthropogenic sites. A popular mitigation strategy is to enhance microbial CH4
39
oxidation in situ by applying amendments including inorganic materials such as sand3 and slag4,
40
and organic materials such as sludge5, compost6, and wood pellets7 to fields which would
41
otherwise emit significant CH4.
42
Biochar was recently proposed as an alternative amendment for CH4 mitigation, and it has
43
been shown to decrease CH4 emission from landfill7, 8 and paddy soils.9, 10 Biochar is produced
44
via pyrolysis or gasification of organic substances under oxygen-limited conditions.11 It
45
features merits of microbial-decomposition recalcitrance to organic amendments9 and
46
microbial affinity to inorganic amendments,1 which catalyse renewed interests in mechanisms
47
of biochar amendment decreasing CH4 emissions. Biochar’s high surface area and porosity
48
enable a high water-retention capacity, creating favourable conditions for the colonization of
49
CH4-affiliated organisms.7,
50
microbial CH4 oxidation kinetics by increasing gas diffusion.9, 10
9
In addition, biochar’s porosity and adsorbability facilitate
51
Biochar is generally considered to be chemically inert toward adsorbates,12 and all
52
aforementioned mechanisms are based on our common understanding of biochar’s physical
53
properties in terms of favourable surface properties, porosity and adsorbability. However,
54
recent studies have demonstrated biochar is also chemically reactive. Characteristics of biochar
55
such as a graphitic structure and/or the presence of redox-active functionalities such as (hydro)
4
ACS Paragon Plus Environment
Page 4 of 28
Page 5 of 28
Environmental Science & Technology
56
quinones, facilitate abiotic and biological redox reactions.12-14 Therefore, biochar-amended soil
57
can potentially act as a carbon sink by stimulating microbial CH4 respiration, or microbial
58
metabolism of CH4 via other pathways, as yet undetermined.
59
Moreover, the current understanding of mechanisms by which biochar impacts CH4 fluxes
60
is limited to its influences on aerobic CH4 oxidation by methanotrophs.7,
61
anaerobic oxidation of methane (AOM) by anaerobic methanotrophs also plays a significant
62
role in global CH4 emission mitigation.15, 16 Due to the unfavourable gas accessibility of soil,
63
biochar amended into deep fields may involve itself into active mediation of anoxic CH4
64
oxidation, the feasibility and mechanism of which have yet to be explored.
9, 10
However,
65
Therefore, the aim of the present study was to investigate the chemically reactive role of
66
biochar amendment on anoxic CH4 oxidation for the mitigation of CH4 emission. A mixed
67
microbial culture dominated by anaerobic methanotrophs was coupled to biochar amendment,
68
and CH4 turnover and AOM gene transcription were monitored. In addition, biochar as a direct
69
electron sink for mitigating global CH4 emissions was evaluated based on globally available
70
biomass-feedstocks which could be converted into biochar, and as an electron shuttle indirectly
71
facilitating AOM coupled to other ubiquitous electron acceptors was assessed based on the
72
standard reduction potential of biochar.
73
MATERIALS AND METHODS
74
Biochar. The commercial Triple R Biochar (Triple R Co, AUS) used here was produced by
75
pyrolysis at 600-650 oC from sustainably sourced plant material. The detailed properties of the
76
biochar are summarized in Table S1 of the Supporting Information. The biochar was ground,
77
and a fraction was then sieved (40 µm). The powder was then rinsed with Milli-Q water three
78
times to remove organics before being oven-dried at 80 oC. Total organic carbon (TOC)
79
analysis suggested the washed biochar contained little dissolved organic carbon (Figure S1).
5
ACS Paragon Plus Environment
Environmental Science & Technology
80
Inoculum. The mixed culture which was used as inoculum was taken from a 5 L parent
81
reactor performing nitrate reduction coupled to AOM.17 The mixed culture is most dominated
82
by an anaerobic methanotrophic archaeon, known as Candidatus ‘Methanoperedens
83
nitroreducens’ or ANME-2d.18
84
For microbial inoculation preparation, biomass was collected from the parent reactor and
85
centrifuged 6000 rpm for 15 min. The biomass pellet was then washed three times
86
anaerobically with fresh medium to eliminate residual nitrate, followed by resuspension in
87
fresh medium with double volume for downstream inoculations. The medium consisted of
88
ingredients modified from that used in the parent reactor17 (0.15 g L-1 NH4Cl, 0.015 g L-1
89
CaCl2·2H2O, 0.1 g L-1 MgCl2·6H2O, trace elements,17 and 25 mM PBS to buffer pH variation
90
caused by biochar amendment), and was anaerobically prepared by sparging with nitrogen for
91
30 min.
92
Test Incubations for Biochar-mediated CH4 Oxidation. Incubations were performed in
93
4.5-mL Exetainer vials (Labco, UK) with sacrificial sampling. Biochar with exact quantities
94
(0.04 or 0.08 g) was added to vials as the potential electron acceptor. Before inoculation, vials
95
containing biochar were placed in a vacuum chamber overnight to eliminate absorbed oxygen.
96
Biomass suspension was then distributed to aliquots of 2 mL in Exetainers, resulting in
97
biochar concentrations of 40 g L-1 and 20 g L-1, respectively. The abiotic control test was
98
prepared by adding 2 mL fresh medium into 0.08 g biochar-amended vials.
99
To test our hypothesis that biochar can play a role as an electron sink to medicate microbial
100
CH4 oxidation, another control incubation was prepared with chemically reduced biochar.
101
Vials containing 0.08 g biochar were filled with 4 mL 50 mM sodium bisulfite solution, and
102
were shaken overnight in an anaerobic chamber (Coy Laboratory Products Inc., USA). The
103
chemically reduced biochar was then washed anaerobically with fresh medium three times
104
before inoculating with 2 mL of biomass.
6
ACS Paragon Plus Environment
Page 6 of 28
Page 7 of 28
Environmental Science & Technology
105
To confirm the role of CH4 as electron donor, a control test was performed without CH4
106
feeding. These vials contained 0.08 g biochar and biomass inoculation, and were pressurized
107
with 2.0 mL argon gas.
108
All other Exetainer vials were sealed and pressurized by injecting 2.0 mL CH4 (also provided
109
as electron donor). All inoculation procedures were carried out anaerobically in an anaerobic
110
chamber. Inoculated vials were then transferred to a shaker-incubator (30 oC and 120 rpm,
111
Multitron, INFPRS HT, Switzerland).
112
The CH4 oxidation activity was quantified by determining inorganic carbon (CO2)
113
production as well as CH4 consumption. To quantify CO2 accumulation, three vails were
114
harvested every 3 days and acidified with 0.2 mL anaerobically prepared HCl stock solution (1
115
M) to extract dissolved CO2. Vials were then kept for at least 1 h for headspace equilibration
116
before CO2 was analysed by gas chromatography. CH4 consumption was determined by CH4
117
concentration change in the both headspace and liquid of incubation vails. CH4 concentration
118
in the headspace was determined by direct gas chromatography measurement and that in the
119
liquid was calculated form headspace concentration based on Henry’s Low. Vials were also
120
harvested on a regular basis from duplicated incubations (with 0.08 g biochar amendment) to
121
test for acetate production test. To test for biological CH4 oxidation, parallel vails containing
122
0.08 g biochar or 0.08 g chemically reduced biochar were harvested at time intervals for mcr
123
gene transcriptomic analysis by reverse transcription PCR (RT-PCR), and microbial
124
community analysis by 16S rRNA amplicon pyrosequencing.
125
Electrochemical Characterizations of Biochar. Cyclic voltammetry (CV) was used to
126
characterize the redox properties of biochar. Biochar-based electrodes were prepared by
127
coating conductive carbon cloth (AvCarb 1071 HCB, FuelCellStore, USA) with biochar
128
powder (prepared as previously described). Carbon cloth was cut into a square piece (2cm
129
2cm), and then cleaned by rinsing in 1 M HCl followed by rinsing in pure acetone, each for 24
7
ACS Paragon Plus Environment
Environmental Science & Technology
130
h. Afterwards, cloth pieces were washed thoroughly with Milli-Q water and dried at room
131
temperature before use. For the coating process, biochar powder (5 mg) was mixed with 1000
132
uL absolute ethanol and 100 uL Nafion (5%, Sigma, USA) under sonication to obtain a coating
133
ink. The ink was then spread evenly on both sides of the carbon cloth, and dried completely.
134
Finally, the biochar-coated cloth was connected to titanium wires for use as electrodes for CV
135
tests. A PBS solution (50 mM) with 1 M KCl was used as the electrolyte for CV tests, with
136
scans recorded from -0.8 V to + 0.8 V, and varied sweep rates from 1 mV s-1 to 50 mV s-1.
137
The biochar’s oxidative state was further evaluated and quantified by direct electrochemical
138
reduction, performed in a sealed electrolysis cell (50 mL). Biochar-based electrodes with mass
139
loadings of 2.5 mg, 5.0 mg, 7.5 mg, 10 mg, 12.5 mg, respectively, were used as working
140
electrodes, and a titanium wire was applied as the counter electrode. The electrolysis cell was
141
filled with 40 mL electrolyte solution (as above) that had been sparged for 30 min to eliminate
142
dissolved oxygen before sealing the vessel. Electrochemical reduction was achieved with a
143
potentiostat (BioLogic VSP, Claix, France) by applying a reductive potential of -0.4 V versus
144
an Ag/AgCl reference electrode (RE-5B, BASi, USA). Reductive current responses were
145
recorded. Each electrochemical reduction was equilibrated once reductive current was
146
decreased to baseline (current response of electrochemical reduction of carbon cloth without
147
biochar loading). The resulting current peaks were integrated to determine electron transfer (Q)
148
for the reduction of biochar with different masses: 𝑡
149
Q=
∫𝑡0𝐼𝑑𝑡 𝐹
150
Where I is the baseline-calibrated reductive currents recorded, t is time, and F (=96485 C mol-1
151
e-) is the Faraday constant. The electron storage capacity (ESC) of biochar was defined by
152
integrated electrons transfer per unit biomass mass, which was derived by fitting the slope of
153
the regression line of reductive electron transfer versus biochar mass.
8
ACS Paragon Plus Environment
Page 8 of 28
Page 9 of 28
Environmental Science & Technology
154
Chemical Analyses. Quantities of CH4 and CO2 in the headspace were determined by gas
155
chromatography. 100 µL of headspace gas sample was collected by gas-tight syringe (1710SL,
156
Hamilton, USA) and the concentration was quantified with an Agilent Gas Chromatograph
157
(GC, 7890A, Agilent, USA) equipped with a HayeSepQ column (2440 2.0mm) and a thermal
158
conductivity detector. Argon was used as carrier gas at a flow rate of 28 mL min-1. The injector,
159
column and detector temperatures were maintained at 110, 45 and 170 ℃ respectively during
160
tests. Gases were identified according to their retention times, and their concentrations were
161
calculated against standards.
162
Liquid samples were taken from vails using a syringe, and then filtered immediately through
163
a 0.22 μM disposable sterile millipore filter (Merck). Acetate concentration was measured
164
using a Shimadzu HPLC system with a Bio-Rad HPLC Column (300×7.8 mm) and a refractive
165
index detector. Peaks corresponding to those of organic acids were confirmed by retention time,
166
co-elution with standards, and by comparing absorbance spectra with those of the standards.
167
Total organic carbon (TOC) was measured by using a TOC Analyzer (Shimadzu 5000 A).
168
Fourier Transform Infrared Spectroscopy (FTIR). FTIR analysis was performed using
169
an infrared spectrometer (Nicolet 6700, Thermo Scientific, USA) equipped with a Diamond
170
ATR crystal (Attenuated Total Reflection). The spectrum was collected in the frequency
171
range 2,000–600 cm−1.
172
X-ray Photoelectron Spectroscopy (XPS). XPS spectra were collected by a Kratos Ultra
173
Axis XPS with Al Kα X-ray. Spectra calibrated on the 538.1 (O1s) and 296.1 (C1s) eV peaks
174
were analyzed using XPSPEAK.
175
RT-PCR. Biomass was collected from each incubation vial (2 mL), and total cell RNA was
176
extracted using RNeasy Mini kit (Qiagen, Germany) according to manufacturer’s protocol. An
177
additional DNase treatment was performed (provided within the Qiagen RNase-Free DNase
178
Set). RNA quality was checked by agarose-gel electrophoresis, and the RNA concentration 9
ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 28
179
was measured with a spectrophotometer Nanodrop (ND-1000, Thermo Fisher, USA). RNA
180
was reverse transcribed to cDNA using the QuantiTect Reverse Transcription Kit (Qiagen) via
181
mcrA gene primer 345R. Quantitative real-time PCR (qPCR) amplification was performed by
182
using a QuantStudio real-time PCR system (Thermo Fisher Scientific, USA) with the following
183
composition: 0.1 μL mix of 20 μM of the mcrA gene primers 159F and 345R, 5 μL of SYBR
184
Green PCR Master Mix (Thermo Fisher Scientific), 2 μL DNA (1ng/ μL) and 3 μL of Milli-Q
185
water. The PCR cycle conditions were as follows: 95 °C for 10 min, followed by 40 cycles of
186
95 °C for 15 s, 60 °C for 1 min, then a final melt curve stage at 95 °C for 15 s, 60 °C for 1 min
187
and 95 °C for 15s. The standard curve was constructed from a series of 10-fold dilution series
188
of plasmid DNA of a known copy number, with R2 values of at least 0.99 for all assays.
189
16S rRNA Gene Sequencing. At the end of the incubation course, biomass (1 mL) was
190
collected from incubation vials, and total cell DNA was extracted using FastDNA SPIN for
191
Soil kit (MP Biomedicals, USA) according to the manufacturer’s instructions. The extracted
192
DNA concentration was quantified with NanoDrop 2000 (Thermo Fisher Scientific, USA). The
193
16S
194
AAACTYAAAKGAATTGACGG-3 ′ ) and 1392R (5 ′ -ACG- GGCGGTGTGTRC-3 ′ ). A
195
QIAquick PCR Purification Kit (Qiagen) and a Quant-iT dsDNA HS assay kit (Invitrogen)
196
were employed to purify and quantify the PCR products, respectively. Amplicons were pooled
197
in equimolar concentration and sequenced with an Illumina sequencer (Illumina, USA). Raw
198
sequencing data were quality-filtered and demultiplexed using Trimmomatic, with poor-
199
quality sequences trimmed and removed. Subsequently, high-quality sequences at 97%
200
similarity were clustered into operational taxonomic units (OTUs) using QIIME with default
201
parameters, and representative OTU sequences were taxonomically aligned against
202
Greengenes 16S rRNA database. An OTU table consisting of the taxonomic classification and
rRNA
gene
was
amplified
using
10
universal
ACS Paragon Plus Environment
primer
set
926F
(5 ′ -
Page 11 of 28
Environmental Science & Technology
203
OTU representative sequences was generated as the main analysis results, based on which a
204
microbial community heatmap was based (Rstudio).
205
RESULTS AND DISCUSSION
206
CH4 Oxidation Activity Assessed by CO2 Accumulation and CH4 Consumption. As shown
207
in Figure 1a, , there was no CO2 production detected in abiotic incubation tests (without
208
biomass inoculation). This suggested that chemical catalysis of CH4 by biochar was not
209
occurring. Limited CO2 was produced in the presence of biomass without CH4 feeding, with
210
an accumulation of 1.26 ± 0.05 µmol CO2 over 21 days. Other potential sources of CO2, for
211
example, from biochar intrinsic organics or biomass lysis, were excluded as biochar rinse
212
solutions contained no detectable TOC after 3 washes (Figure S1). In addition, the stable
213
ammonium concentration observed over time (data not shown) indicated negligible microbial
214
lysis was occurring. Interestingly, acetate accumulation was traced during the initial incubation
215
period (within 4 days) regardless of the presence of CH4, presumably as a result of
216
decomposition of microbial intracellular storage compounds that were generated whilst in the
217
parent reactor (Figure S2).18 A gradual decrease in acetate concentration after day 5 indicated
218
that acetate may be released as an intermediate supporting biochar-mediated microbial-
219
respiration and CO2 production (Figure S2).
220
In contrast, with CH4 feeding, vials containing 0.08 g biochar rapidly accumulated CO2
221
reaching a maximum concentration at 15 days (Figure 1a, biochar + biomass). The amount of
222
CO2 in the reactor became stable after 15 days, implying that all biochar added have been
223
reduced and the reaction stopped. The supply of CH4 resulted in faster CO2 production rate as
224
well as much higher accumulation (8.50 ± 0.41 µmol) to that without CH4 feeding, indicating
225
that the CO2 was produced from CH4. As all other potential electron acceptors (including
226
oxygen, nitrate and sulphate) had been eliminated from incubations, biochar was likely the sole
11
ACS Paragon Plus Environment
Environmental Science & Technology
227
candidate mediating CH4 oxidation. This hypothesis is supported by the limited CO2
228
accumulation (1.34 ± 0.28 µmol) in the incubation containing chemically reduced biochar
229
(Figure 1a, 0.08 g reduced biochar + biomass). The small amount of CO2 production in this
230
case probably originated from incomplete chemical reduction of biochar, and the CO2
231
accumulation gap mediated by reduced biochar demonstrated that reactive biochar played a
232
role as electron sink for AOM. Furthermore, the incubation with 0.04 g biochar demonstrated
233
a similar CO2 production trend to the 0.08 g biochar incubation, but with approximately half
234
the amount of CO2 accumulation (4.09 ± 0.41 µmol), which was a quantitative support of
235
biochar’s active role mediating AOM processes. The CH4 oxidation capacities of biochar
236
calculated based on the CO2 production rates are 106.3 ± 5.1 μmol g-1 and 102.3 ± 10.3 μmol
237
g-1, separately, during these two tests.
238 239
Figure 1. CO2 production (a) and CH4 consumption (b) in incubations with biochar only, 0.08
240
g biochar with biomass, 0.04 g biochar with biomass, 0.08 g chemically-reduced biochar with
241
biomass, and 0.08 g biochar with biomass without CH4. Data represent the mean values of
242
triplicate incubations ± standard deviation of the mean.
243
CH4 consumption was monitored as a more direct way to describe CH4 oxidation activity
244
mediated by biochar. It should be noted that due to the technical difficulty of accurately
245
controlling manual feeding of 2 mL CH4 to incubation vials, CH4 concentrations were
12
ACS Paragon Plus Environment
Page 12 of 28
Page 13 of 28
Environmental Science & Technology
246
measured with a relatively wide confidence interval (Figure 1b). Even so, an increasing trend
247
in CO2 production was observed with a concomitant decreasing trend in CH4 consumption in
248
biochar amended biotic incubations. A total of 6.35 ± 0.52 µmol and 4.12 ± 0.96 µmol CH4
249
was consumed in 0.08 g and 0.04 g biochar amendments, respectively. CH4 consumption was
250
related to CO2 production with a corresponding balance error of 25.3% and 0.7%, respectively
251
(calculation details in Supporting Information). In contrast, incubations with biochar only
252
(without microbial inoculum) or with chemically reduced biochar with microbial inoculum
253
(Figure 1b, reduced biochar + biomass) demonstrated limited CH4 consumption over time.
254
Together, these results support the hypothesis that biochar can mediate anaerobic CH4
255
oxidation.
256
Tracking Microbial CH4 Oxidation through mcr Gene Transcription and Community
257
Analysis. The inoculum used here contains ANME-2d which anaerobically activates CH4
258
through reverse methanogenesis mediated by the mcr gene.18 Thus we used mcr transcript
259
abundance to indicate the bio-activity of AOM. We measured mcr expression in microbial
260
incubations amended with biochar and chemically reduced biochar, respectively, to determine
261
how the redox capacity of biochar affects AOM activity.
262
As shown in Figure 2a, actively growing biomass demonstrated high transcript levels for the
263
mcr gene (ca. 5.1 108 copies, day 0). A significant drop in transcript abundance was observed
264
on day 3 in 0.08 g biochar-amended incubations. As the biomass used here was originally
265
cultured with nitrate as the electron acceptor, the transcript drop was likely due to the less
266
favourable electron transfer to solid biochar rather than to soluble nitrate. Nevertheless, it still
267
remained at a considerable level (4.6 108 copies) in comparison with the reduced biochar
268
amendment (6.0 107 copies), demonstrating the contribution of biochar to preservation of
269
AOM. Consistent with the gradual decrease in CO2 accumulation (Figure 1a), the mcr gene
270
transcript abundance decreased steadily over time (Figure 2a), which was likely due to 13
ACS Paragon Plus Environment
Environmental Science & Technology
271
saturation of the biochar electron-accepting capacity. In contrast, microbial incubation with
272
reduced biochar showed relatively limited mcr transcription from 3 days of incubation onwards,
273
indirectly demonstrating the dependence of AOM on the biochar redox capacity.
274 275
Figure 2. (a) Transcript abundances of the mcr gene for microbial incubations with 0.08 g
276
biochar and 0.08 g chemically reduced biochar, respectively. Error bars represent one standard
277
deviation from three triplicate incubations; (b) microbial community diversity in incubations
278
sampled on day 21: A, inoculum; B, 0.08 g biochar amendment; C, 0.08 g chemically reduced
279
biochar amendment.
280
To further interpret the relationship between AOM and biochar chemical reactivity,
281
microbial community evolution was compared in non-reduced and chemically reduced biochar
282
amended incubations, respectively (Figure 2b). In the non-reduced biochar amendment,
283
methanotrophic activator ANME-2d dominated the community (Figure 2b, B, 31.0%),
284
comparable to that of the inoculum (32.2%). In contrast, ANME-2d represented only 6.6% of
285
the microbial community in the reduced biochar amendment, a decrease of almost 80% from
286
the inoculum (Figure 2b, C). A legitimate explanation for this is that because reduced biochar
287
has limited electron-accepting capacity, ANME-2d could not gain enough energy to survive
288
due to interruption of the electron transport chain. Thus microbial community evolution results
289
further confirmed that chemically reactive biochar contributed to anaerobic CH4 oxidation.
290
It is worth noting that Geobacteraceae were enriched in incubations with both non-reduced 14
ACS Paragon Plus Environment
Page 14 of 28
Page 15 of 28
Environmental Science & Technology
291
(7.9%) and reduced biochar (4.3%) (Figure 2b, B and C, respectively), likely stimulated by the
292
production of substrate, acetate (Figure S2). The extracellular electron transfer capability of
293
Geobacteraceae has been widely recognized and reported,19 and redox-active biochar can
294
mediate microbial respiration via extracellular electron transfer.20 Combined, these
295
observations indicate that biochar can also stimulate Geobacteraceae to metabolize acetate.
296
The presence of Geobacteraceae implied that biochar mediated AOM perhaps occurs through
297
syntrophic interaction between ANME-2d and Geobacteraceae.
298
Redox Properties of Biochar Characterized by Electrochemical Analysis. Combined,
299
the results suggest that biochar is able to mediate AOM depending on its oxidation state, with
300
biochar losing the capacity when reduced. Other studies have also attributed the ability of
301
biochar to mediate biological reactions to its redox properties.12, 13 In our study, the redox
302
properties of biochar were assessed by cyclic voltammetry (Figure 3a). The bare carbon cloth
303
showed a smooth curve with a weak current response, while the biochar-coated carbon cloth
304
displayed a remarkable current increase with obvious redox peaks. Cyclic voltammetry
305
performed at high scan rates above 5 mV s-1 displayed a dramatic reductive current increase
306
once the potential negatively reached -0.25 V (Figure 3a, inset), caused by interactions on the
307
interface of biochar and electrolyte. Given the existence of overpotential for practical H2
308
evolution by electrical reduction, reduction current turnovers shown on CV curves were not
309
likely from H2 evolution as the reduction turnover potential observed here (around -0.45 V vs.
310
SHE) was highly close to theoretical H2 evolution potential (-0.421 V vs. SHE; pH = 7).
311
Accordingly, oxidative peaks were observed at all scan rates (Figure 3a, inset). Together, these
312
results imply that redox-active sites are present on the biochar contributing to its chemical
313
reactivity. More specifically, the CV conducted at the lowest rate of 1 mV s-1 revealed an
314
obvious redox couple centred at a formal potential of -0.045 V vs. Ag/AgCl. This formal
315
potential is more positive than the midpoint potentials of general cytochromes known to be
15
ACS Paragon Plus Environment
Environmental Science & Technology
316
involved in microbial extracellular electron transfer (EET).21 Combined, these CV results
317
indicate that biochar is chemically active as an electron sink, and its standard reduction
318
potential range enables biochar to mediate microbial EET.
319
To further evaluate and quantify the capacity of biochar to act as an electron sink for
320
mediating AOM, direct electrochemical reductions of biochar were carried out (Figure 3b). All
321
biochar samples initially generated a strong reductive current response, and then approached
322
equilibrium, as the biochar became saturated with electrons. Increasing biochar mass loading
323
resulted in correspondingly stronger current peaks and a linear increase in respective reduction
324
charge, all of which demonstrate the positive oxidation state of biochar and its capacity to
325
accept electrons. The linear relationship between biochar reduction charge and the biochar
326
mass revealed the electron storage capacity of biochar to be 0.74 mmol e- g-1 (Figure 3b, inset),
327
and this was comparable to the value (0.85 mmol e- g-1) calculated from biochar reduction
328
coupled to microbial CH4 oxidation (calculation detailed in Supporting Information).
329 330
Figure 3. Electrochemical characterizations of biochar (a) Cyclic voltammetry (CV) of carbon
331
cloth alone (CC support), and biochar-coated (5 mg biochar loading) carbon cloth (Biochar),
332
at scan rates ranging from 1 mV s-1 to 50 mV s-1 (as indicated). (b) Current responses to
333
electrical reduction (at -0.4 V vs. Ag/AgCl) with different biochar masses (as indicated), and
334
the relationship between respective biochar mass and reduction charges (insert). The slope of
335
the linear regression line corresponds to the reduction charge of per biochar mass, defined as 16
ACS Paragon Plus Environment
Page 16 of 28
Page 17 of 28
336 337
Environmental Science & Technology
electron storage capacity (ESC) of biochar. Characterization of Biochar by XPS and FTIR. Previous studies have showed that
338
biochar’s chemical reactivity is normally related its redox-active moieties.12-14,
339
biochar’s functional groups were characterized through XPS and FTIR to understand the active
340
source that potentially drives AOM.
341
22
Thus,
The qualitative survey detected carbon and oxygen (Figure S3), the latter being commonly
342
attributed in literature to biochar oxygenic moieties.13,
343
resolution core-line scan for C elements (C1s) revealed that the most prominent peak occurred
344
at 285.1 eV for the C-C bond.24, 25 The C1s XPS spectra also demonstrated peaks at 286.5 eV
345
and 288.9 eV for C-O and C=O bonds, respectively.25,
346
containing bonds was also verified by high-resolution O1s XPS spectra (Figure S4).24, 25
23
As shown in Figure 4a, the high
26
The existence of these oxygen-
347 348
Figure 4. (a) High-resolution C1s XPS spectra of biochar; and (b) FTIR profile of biochar
349
Further FTIR characterization revealed two C-C absorbance peaks stretching from 800-
350
1,200 cm-1 (Figure 4b).27, 28 Meanwhile, notable vibrations of C-O at 1,440 cm-1 and C=O at
351
1,570 cm-1 were detected, which was indicative of oxygenic moieties in terms of carboxylic
352
acid groups and quinone structures.13,
353
including biochar is widely recognized to be derived from electron-unsaturated oxygenic
27, 28
The chemical reactivity of carbon materials
17
ACS Paragon Plus Environment
Environmental Science & Technology
354
moieties such as (hydro)quinones,
355
especially C=O moiety, revealed by XPS and FTIR characterizations suggests that these
356
functional groups could actively mediate AOM. CH4 oxidation mediated by biochar could be
357
expressed by the following stoichiometric equation:
358
CH4 + 8 C=O + 2H2O 8 C-OH + CO2
12-14, 22
thus the existence of oxygen-based functionalities,
359
Mechanisms for Biochar-mediated AOM. The current study shows that biochar feasibly
360
mediates AOM due to inherent redox-active functionalities. Both abiotic- and biological
361
reduction of biochar have long been reported,12, 29 while our study, to the best our knowledge,
362
is the first report of microbial reduction of biochar by anaerobic methanotrophs. Importantly,
363
our study also provides evidence that anaerobic methanotrophic archaea (ANME) utilize an
364
EET mechanism, a concept which has been proposed but not yet fully confirmed.30, 31 The
365
ANME-2d used in our incubations encodes numerous (38 species) multiheme c-type
366
cytochromes providing a genetic basis for biochar respiration via EET,18 as c-type cytochromes
367
can function as electron shuttles from within the cell to solid electron acceptors external to the
368
cell.32, 33 It was noteworthy that Geobacteraceae also evolved as an important member of the
369
microbial consortium with biochar potentially playing a role as sole electron acceptor (Figure
370
2b). Geobacteraceae is well-recognized for its EET ability,32 and it has been reported to be
371
able to establish syntrophic associations with other microbes via direct interspecies electron
372
transfer (DIET).34 Therefore, we propose that direct EET occurs from ANME-2d to biochar,
373
but direct interspecies electron transfer as a syntrophic coupling mechanism between ANME-
374
2d and Geobacteraceae may also be occurring during biologically mediated biochar reduction.
375
Moreover, as previously mentioned the production of acetate during incubation provides
376
another potential electron transfer mechanism based on interspecies transfer via metabolic
377
intermediates. Based on results from the incubation control without CH4 (Figure 1), the
18
ACS Paragon Plus Environment
Page 18 of 28
Page 19 of 28
Environmental Science & Technology
378
presence of acetate was likely due to its release from intercellular storages of the inoculum.18
379
It is also possible that real-time acetate production occurred from direct oxidation of CH4 via
380
the reductive acetyl CoA pathway in ANME-2d.18 In our study, ANME-2d was also potentially
381
linked to Geobacteraceae through a diffusible metabolite with acetate acting as an intermediate,
382
rather than through direct electrical connections. For this mechanism, electrons originating
383
from CH4 are not necessarily transferred from the cytoplasm to the outer membrane in ANME-
384
2d. However, in this cases, the biosynthesis of acetate from CH4 oxidation thermodynamically
385
requires extra electron acceptors.35 As no other electron acceptors were added to our
386
incubations, the only pathway that could produce acetate would be via EET directly or
387
indirectly to biochar (Figure 5, dotted box), and an intermediate based electron transfer
388
mechanism could be applicable. Therefore, we propose that ANME-2d achieves AOM via an
389
EET dependent pathway, potentially through three mechanisms (Figure 5).
390 391
Figure 5. Proposed mechanisms for biochar-mediated AOM 1) Proposed direct electron
392
transfer from ANME-2d to biochar; 2) Indirect extracellular electron transfer from ANME-2d
393
to biochar based on direct interspecies electron transfer between ANME-2d and
394
Geobacteraceae; 3) indirect extracellular electron transfer from ANME-2d to biochar based on
395
interspecies electron transfer between Geobacteraceae and ANME-2d via metabolism of
396
intermediate, acetate (direct electron transfer from ANME-2d to biochar is the
397
thermodynamical premise for biosynthesis of acetate from CH4 oxidation in this case) 19
ACS Paragon Plus Environment
Environmental Science & Technology
398
Practical Implications. Biochar amendment has been reported to be an effective approach to
399
capturing CH4 emissions by stimulating aerobic methanotrophic activity.9, 10 Our study shows
400
that biochar amendment can mediate CH4 oxidation in anoxic environments, and offers
401
plausible mechanisms for stimulation of microbial CH4 oxidation by biochar based on
402
interaction with ANME. In our study, biochar acted as an electron sink for CH4 oxidation
403
providing a reservoir for anaerobic CH4 fixation with potential applications for global CH4
404
emission mitigation.
405
Large-scale applications of biochar are not prevalent, and are limited by the challenges of
406
commercial production necessary to achieve high biochar yield through biomass pyrolysis.36
407
The majority of biochar is currently sourced from non-commercial, non-engineered production,
408
that is known as ‘black carbon’, and produced from open biomass burning.37 Electron
409
equivalent estimations (see Supporting Information for details) indicate that combined, all
410
current biochar production provides a negligible electron sink for CH4 oxidation as compared
411
with global CH4 emission (Table 1), even assuming all black carbon is coupled to AOM in
412
anoxic sites of natural environments. So it is difficult to say biochar nowadays is contributing
413
a lot to decrease CH4 emission by mediating AOM via direct supply of electron sink.
414
However, our report may provide scientific support for the promotion of large-scale biochar
415
application and production.36 Biochar provides a sustainable strategy to decrease CO2, CH4 and
416
N2O emissions, sequestering organics through biomass pyrolysis38. Widespread application of
417
biochar to agricultural soils could further decrease greenhouse gas emissions by providing a
418
substantial electron sink for AOM. On a global scale, if the maximum sustainable amount of
419
available biomass feedstock is 2.27 Pg carbon per year, and approximately 49% C yield is
420
achievable for biochar production from biomass pyrolysis,38 we can expect that a year of
421
biochar production provides a potential electron reservoir of at least 1.11 1012 mol e-
422
(electron-accepting capacity of 0.85 mmol e- g-1 biochar) and up to 9.10 1012 mol e- based on 20
ACS Paragon Plus Environment
Page 20 of 28
Page 21 of 28
Environmental Science & Technology
423
the maximum biochar electron-accepting capacity of 7 mmol e- g-1 (Table 1; calculation details
424
in Supporting Information). In coupling to CH4 emission, an electron reservoir of this capacity
425
would effectively mitigate global CH4 emissions assuming that biochar is broadly applied to
426
CH4 sources such as landfill and rice paddies (Table S2).39 For example, based on electron-
427
accepting capacity of biochar in this study, we estimate that 341.8 t biochar amendment per ha
428
per year would be able to effectively control CH4 emission from rice paddy (calculation details
429
in Supporting Information). However, it is noteworthy that these are theoretical estimations,
430
with several premises: 1) maximum biochar yield from sustainable biomass resources,38 2) all
431
of the biochar produced can be used at landfill sites and paddy soils to reduce methane emission,
432
3) proper design and installation can ensure biochar-mediated AOM process in subsurface
433
CH4-producing environments. The effectiveness of biochar amendments on anaerobic CH4
434
mitigation at landfill sites should be further evaluated and relevant technical design during
435
practical applications warrant following investigations.
436
Through its chemical reactivity, biochar can shuttle electron transfer from microbes to
437
extracellular electron acceptors.40 The evidence here of extracellular electron transfer between
438
ANME-2d and biochar would suggest that AOM driven by other electron acceptors could be
439
promoted by biochar. For instance, Fe(III) minerals have been reported to be able to be directly
440
coupled to AOM.41 The redox potentials of iron minerals are strongly influenced by
441
environmental conditions, and the wide redox potential range over which different forms of
442
Fe(III) exist (Eo -314 mV to 770 mV vs. SHE) implies that biochar (Eo = 152 mV vs. SHE)
443
could act as an intermediary during AOM.42 A similar shuttling process may also be applicable
444
to manganese-driven AOM due to the highly positive standard reduction potential of MnO2 (Eo
445
= 1.23 mV vs SHE), although direct or indirect electron transfer from methanotrophs to MnO2
446
has not been confirmed.41
21
ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 28
447
Table 1. Evaluation of electron-donating capacity of global CH4 production and electron-
448
accepting capacity of biochar Annual biochar production ( g yr-1) current black carbon production (Mt)
large-scale commercial biochar production (Pg)
8.4012
1.31a Electron-accepting capacity of biochar (mol e- yr-1)
based on electron storage capacity in this study
based on biochar’s maximum electron storage capacity12
based on electron storage capacity in this study
based on biochar’s maximum electron storage capacity
6.46 1010
5.32 1011
1.11 1012
9.10 1012
Annual global CH4 emission (g yr-1) b emission from rice fields
emission from landfills
total global emission
9.50 1013
5.00 1013
5.20 1014
Electron-donating capacity (mol e- yr-1) 4.75 1013
2.50 1013
Calculation based on annual biomass-feedstock availability; available sites for large-scale biochar amendment
a
2.60 1014 b
assume rice paddies and landfills would be
449 450
Recent evidence indicates that sulfate-dependent AOM occurs through direct interspecies
451
electron transfer between ANME and sulfate-reducing bacteria (SRB).30, 31 Hence, the capacity
452
of biochar to promote interspecies electron transfer38, 43 suggests it could also promote sulfate-
453
driven AOM in sediment. Given the ubiquity of iron/manganese minerals and sulfate in anoxic
454
environments, an abundant electron reservoir exists for AOM. In this scenario, biochar is not
455
limited by its inherent electron-accepting capacity for direct mediation of CH4 oxidation; rather,
456
biochar is more likely to contribute to global CH4 emissions reduction as an electron shuttle to
457
promote AOM coupled to other electron acceptors. Logical next experiments are herein
458
desirable to affirm this propose.
459 460
SUPPORTING INFORMATION
461
Calculation details and additional tables and figures, as mentioned in the text, are available in 22
ACS Paragon Plus Environment
Page 23 of 28
Environmental Science & Technology
462
the supporting information section available free of charge via the Internet at
463
http://pubs.acs.org.
464 465
AUTHOR INFORMATION
466
Corresponding Author
467
Tel: +61 7 3346 3230. Fax: +61 7 336 54726. Email:
[email protected] 468
Notes
469
The authors declare no competing financial interest.
470 471
ACKNOWLEDGMENTS
472
The Authors thank Dr Eloise Larsen for careful editing of the manuscript. We are grateful to
473
The AWMC Analytical Services Laboratory (ASL) for all chemical analysis. This work is
474
supported by the Australian Research Council (ARC) through the projects of Australian
475
Laureate Fellowship (FL170100086) and Discovery Project (DP170104038), and the U.S.
476
Department of Energy’s Office of Biological Environmental Research (DE-SC0010574). X.Z.
477
is supported by The University of Queensland International Scholarship (UQI); J.X. is
478
supported by the UQ Research Training Scholarship J.P. is supported by The University of
479
Queensland International Scholarship and China Scholarship Council Scholarship; S.H. is
480
supported by the Advanced Queensland Research Fellowship.
23
ACS Paragon Plus Environment
Environmental Science & Technology
481
REFERENCES
482
1.
483
systems: a review. Reviews in Environmental Science and Bio/Technology 2013, 13, (1), 79-
484
107.
485
2.
486
processes involved. Environmental microbiology reports 2009, 1, (5), 285-292.
487
3.
488
oxidation capacity of compacted soils intended for use as landfill cover materials. Waste
489
management 2011, 31, (5), 833-842.
490
4.
491
methane emission from flood water rice farming. Agriculture, ecosystems & environment 2008,
492
128, (1-2), 21-26.
493
5.
494
refuse bio-cover. Bioresource technology 2011, 102, (3), 2328-2332.
495
6.
496
Huber-Humer, M.; Spokas, K., Microbial methane oxidation processes and technologies for
497
mitigation of landfill gas emissions. Waste Management & Research 2009, 27, (5), 409-455.
498
7.
499
added to landfill cover soil on microbial methane oxidation: A laboratory column study. J
500
Environ Manage 2017, 193, 19-31.
501
8.
502
amendment on soil methane concentration profiles and diffusion in a rice-wheat annual rotation
503
system. Sci Rep 2016, 6, 38688.
504
9.
505
emission from Chinese paddy soils. Soil Biology and Biochemistry 2012, 46, 80-88.
Sadasivam, B. Y.; Reddy, K. R., Landfill methane oxidation in soil and bio-based cover
Conrad, R., The global methane cycle: recent advances in understanding the microbial
Rachor, I.; Gebert, J.; Gröngröft, A.; Pfeiffer, E.-M., Assessment of the methane
Ali, M. A.; Oh, J. H.; Kim, P. J., Evaluation of silicate iron slag amendment on reducing
Lou, Z.; Wang, L.; Zhao, Y., Consuming un-captured methane from landfill using aged
Scheutz, C.; Kjeldsen, P.; Bogner, J. E.; De Visscher, A.; Gebert, J.; Hilger, H. A.;
Yargicoglu, E. N.; Reddy, K. R., Effects of biochar and wood pellets amendments
Xu, X.; Wu, Z.; Dong, Y.; Zhou, Z.; Xiong, Z., Effects of nitrogen and biochar
Feng, Y.; Xu, Y.; Yu, Y.; Xie, Z.; Lin, X., Mechanisms of biochar decreasing methane
24
ACS Paragon Plus Environment
Page 24 of 28
Page 25 of 28
Environmental Science & Technology
506
10.
507
Oxidation in Landfill Cover Soil Amended with Biochar. Journal of Geotechnical and
508
Geoenvironmental Engineering 2014, 140, (9), 04014047.
509
11.
510
applications of biochar for environmental remediation: a review. Critical Reviews in
511
Environmental Science and Technology 2015, 45, (9), 939-969.
512
12.
513
microbial electron donor and acceptor. Environmental Science & Technology Letters 2016, 3,
514
(2), 62-66.
515
13.
516
M., Carbon black as an alternative cathode material for electrical energy recovery and transfer
517
in a microbial battery. Scientific reports 2017, 7, (1), 6981.
518
14.
519
trinitro-1, 3, 5-triazine in the presence of hydrogen sulfide and black carbon. Environmental
520
science & technology 2008, 42, (6), 2118-2123.
521
15.
522
F. S., Anaerobic oxidation of methane: mechanisms, bioenergetics, and the ecology of
523
associated microorganisms. Environmental science & technology 2008, 42, (18), 6791-6799.
524
16.
525
process. Annu Rev Microbiol 2009, 63, 311-334.
526
17.
527
denitrifying anaerobic methane oxidizing microorganisms. Environ Microbiol Rep 2009, 1, (5),
528
377-384.
Reddy, K. R.; Yargicoglu, E. N.; Yue, D.; Yaghoubi, P., Enhanced Microbial Methane
Xie, T.; Reddy, K. R.; Wang, C.; Yargicoglu, E.; Spokas, K., Characteristics and
Saquing, J. M.; Yu, Y.-H.; Chiu, P. C., Wood-derived black carbon (biochar) as a
Zhang, X.; Guo, K.; Shen, D.; Feng, H.; Wang, M.; Zhou, Y.; Jia, Y.; Liang, Y.; Zhou,
Kemper, J. M.; Ammar, E.; Mitch, W. A., Abiotic degradation of hexahydro-1, 3, 5-
Caldwell, S. L.; Laidler, J. R.; Brewer, E. A.; Eberly, J. O.; Sandborgh, S. C.; Colwell,
Knittel, K.; Boetius, A., Anaerobic oxidation of methane: progress with an unknown
Hu, S.; Zeng, R. J.; Burow, L. C.; Lant, P.; Keller, J.; Yuan, Z., Enrichment of
25
ACS Paragon Plus Environment
Environmental Science & Technology
529
18.
530
G. W., Anaerobic oxidation of methane coupled to nitrate reduction in a novel archaeal lineage.
531
Nature 2013, 500, (7464), 567-570.
532
19.
533
Aklujkar, M.; Butler, J. E.; Giloteaux, L.; Rotaru, A.-E., Geobacter: the microbe electric's
534
physiology, ecology, and practical applications. In Advances in microbial physiology, Elsevier:
535
2011; Vol. 59, pp 1-100.
536
20.
537
P.; Lovley, D. R., Promoting interspecies electron transfer with biochar. Scientific reports 2014,
538
4, 5019.
539
21.
540
conservation–the foundation for optimizing bioelectrochemical systems. Frontiers in
541
microbiology 2015, 6, 575.
542
22.
543
herbicides and explosives. Environmental toxicology and chemistry 2013, 32, (3), 501-508.
544
23.
545
of plant biomass-derived black carbon (biochar). Environmental science & technology 2010,
546
44, (4), 1247-1253.
547
24.
548
temperature adhesive and understanding the interfacial chemistry using XPS, ToF-SIMS and
549
Raman spectroscopy. Materials & Design 2016, 109, 622-633.
550
25.
551
to oxygen reduction reaction catalysts based on amine-modified metal-loaded carbons. An XPS
552
and ss-NMR investigation. Materials Chemistry and Physics 2015, 162, 234-243.
Haroon, M. F.; Hu, S.; Shi, Y.; Imelfort, M.; Keller, J.; Hugenholtz, P.; Yuan, Z.; Tyson,
Lovley, D. R.; Ueki, T.; Zhang, T.; Malvankar, N. S.; Shrestha, P. M.; Flanagan, K. A.;
Chen, S.; Rotaru, A.-E.; Shrestha, P. M.; Malvankar, N. S.; Liu, F.; Fan, W.; Nevin, K.
Kracke, F.; Vassilev, I.; Krömer, J. O., Microbial electron transport and energy
Oh, S. Y.; Son, J. G.; Chiu, P. C., Biochar‐mediated reductive transformation of nitro
Keiluweit, M.; Nico, P. S.; Johnson, M. G.; Kleber, M., Dynamic molecular structure
Kadiyala, A. K.; Sharma, M.; Bijwe, J., Exploration of thermoplastic polyimide as high
Marzorati, S.; Ragg, E. M.; Longhi, M.; Formaro, L., Low-temperature intermediates
26
ACS Paragon Plus Environment
Page 26 of 28
Page 27 of 28
Environmental Science & Technology
553
26.
554
single-walled carbon nanotubes by solution-phase ozonolysis. The journal of physical
555
chemistry B 2002, 106, (47), 12144-12151.
556
27.
557
of activated carbons. carbon 1999, 37, (9), 1379-1389.
558
28.
559
Critical reviews in analytical chemistry 2016, 46, (6), 502-520.
560
29.
561
derived black carbon (biochar). Environmental science & technology 2014, 48, (10), 5601-
562
5611.
563
30.
564
wiring enables electron transfer between methanotrophic archaea and bacteria. Nature 2015,
565
526, 587-590.
566
31.
567
acceptors decouple archaeal methane oxidation from sulfate reduction. Science 2016, 351,
568
(6274), 703-707.
569
32.
570
oxides by Shewanella and Geobacter: a key role for multihaem c‐type cytochromes. Molecular
571
microbiology 2007, 65, (1), 12-20.
572
33.
573
catalyze iron-dependent anaerobic oxidation of methane. Proceedings of the National Academy
574
of Sciences 2016, 113, (45), 12792-12796.
575
34.
576
review of microbiology 2017, 71, 643-664.
Banerjee, S.; Wong, S. S., Rational sidewall functionalization and purification of
Figueiredo, J.; Pereira, M.; Freitas, M.; Orfao, J., Modification of the surface chemistry
Ţucureanu, V.; Matei, A.; Avram, A. M., FTIR spectroscopy for carbon family study.
Klüpfel, L.; Keiluweit, M.; Kleber, M.; Sander, M., Redox properties of plant biomass-
Wegener, G.; Krukenberg, V.; Riedel, D.; Tegetmeyer, H. E.; Boetius, A., Intercellular
Scheller, S.; Yu, H.; Chadwick, G. L.; McGlynn, S. E.; Orphan, V. J., Artificial electron
Shi, L.; Squier, T. C.; Zachara, J. M.; Fredrickson, J. K., Respiration of metal (hydr)
Ettwig, K. F.; Zhu, B.; Speth, D.; Keltjens, J. T.; Jetten, M. S.; Kartal, B., Archaea
Lovley, D. R., Syntrophy goes electric: direct interspecies electron transfer. Annual
27
ACS Paragon Plus Environment
Environmental Science & Technology
577
35.
578
B.; Agrawal, S.; Nazem-Bokaee, H.; Gopalakrishnan, S., Reversing methanogenesis to capture
579
methane for liquid biofuel precursors. Microbial cell factories 2016, 15, (1), 11.
580
36.
Lehmann, J., A handful of carbon. Nature 2007, 447, 143-144.
581
37.
US EPA, Report to Congress on Black Carbon: Department of the Interior,
582
Environment, and Related Agencies Appropriations Act, 2010. National Service Center for
583
Environmental Publications 2012, 1-388.
584
38.
585
biochar to mitigate global climate change. Nature communications 2010, 1, 56.
586
39.
587
Journal of Soil Science 1997, 77, (2), 167-177.
588
40.
589
Biochar as an Electron Shuttle between Bacteria and Fe(III) Minerals. Environmental Science
590
& Technology Letters 2014, 1, (8), 339-344.
591
41.
592
Hu, S., A methanotrophic archaeon couples anaerobic oxidation of methane to Fe (III)
593
reduction. The ISME journal 2018, 12, 1929-1939.
594
42.
595
neutral pH. FEMS microbiology ecology 2001, 34, (3), 181-186.
596
43.
597
P.; Lovley, D. R., Promoting interspecies electron transfer with biochar. Sci Rep 2014, 4, 5019.
Soo, V. W.; McAnulty, M. J.; Tripathi, A.; Zhu, F.; Zhang, L.; Hatzakis, E.; Smith, P.
Woolf, D.; Amonette, J. E.; Street-Perrott, F. A.; Lehmann, J.; Joseph, S., Sustainable
Topp, E.; Pattey, E., Soils as sources and sinks for atmospheric methane. Canadian
Kappler, A.; Wuestner, M. L.; Ruecker, A.; Harter, J.; Halama, M.; Behrens, S.,
Cai, C.; Leu, A. O.; Xie, G.-J.; Guo, J.; Feng, Y.; Zhao, J.-X.; Tyson, G. W.; Yuan, Z.;
Straub, K. L.; Benz, M.; Schink, B., Iron metabolism in anoxic environments at near
Chen, S.; Rotaru, A. E.; Shrestha, P. M.; Malvankar, N. S.; Liu, F.; Fan, W.; Nevin, K.
598
28
ACS Paragon Plus Environment
Page 28 of 28