Mineral–Biochar Composites: Molecular Structure ... - ACS Publications

Ben Pace , Paul Munroe , Christopher E. Marjo , Paul Thomas , Bin Gong , Jessica Shepherd , Wolfram Buss , Stephen Joseph. Journal of Analytical and A...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF CAMBRIDGE

Article

Mineral-Biochar Composites: Molecular Structure and Porosity Aditya Rawal, Stephen D. Joseph, James M. Hook, Chee H. Chia, Paul R. Munroe, Scott W. Donne, Yun Lin, David Phelan, David Richard Graham Mitchell, Ben Pace, Joseph Horvat, and J. Beau W. Webber Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00685 • Publication Date (Web): 10 Jun 2016 Downloaded from http://pubs.acs.org on June 12, 2016

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 25

Environmental Science & Technology

1

Mineral-Biochar Composites: Molecular Structure and Porosity

2 3

Aditya Rawal a*, Stephen D. Josephb*, James M. Hook,a Chee H. Chiab, Paul R. Munroe b,

4

Scott Donnec, Yun Linc , David Phelan h, David R. G. Mitchell d, Ben Paceb, Joseph Horvat

5

d

,

J. Beau W. Webber e,f.

6 7

a

8

Australia 2052

9

b

Mark Wainwright Analytical Centre, University of New South Wales, Kensington, NSW,

School of Materials Science and Engineering, University of New South Wales,

10

Kensington, NSW, Australia 2052

11

c

12

Chemistry, Callaghan, New South Wales 2308, Australia

13

d

14

Australia

15

e

Lab-Tools Ltd., Canterbury Enterprise Hub, University of Kent, Canterbury, Kent. CT2 7NJ.

16

f

Institute of Petroleum Engineering, Heriot-Watt University, Edinburgh. EH14 4AS.

17

h

18

Wales 2308, Australia

University of Newcastle, School of Environmental and Life Sciences, Office C325,

Electron Microscopy Centre, AIIM, University of Wollongong, Wollongong NSW, 2522,

University of Newcastle, Electron Microscope and X-Ray Unit, Callaghan, New South

19 20

* Corresponding authors. Tel.: +612 9385 4693; fax: +612 9385 4663.

21

E-mail address: [email protected]; [email protected]

22 23 24 25

Aditya Rawal and Stephen Joseph had equal intellectual and experimental inputs into this paper and are the principal leads in the work.

26

Abstract

27

Dramatic changes in molecular structure, degradation pathway and porosity of biochar are

28

observed at pyrolysis temperatures ranging from 250–550 °C, when bamboo biomass is

29

pretreated by iron-sulfate-clay slurries (iron-clay biochar), as compared to untreated

30

bamboo biochar. Electron microscopy analysis of the biochar reveals the infusion of

31

mineral species into the pores of the biochar and formation of mineral nano-structures.

32

Quantitative

Page 1 of 25

13

C Nuclear Magnetic Resonance (NMR) spectroscopy shows that the

ACS Paragon Plus Environment

Environmental Science & Technology

33

presence of the iron-clay prevents degradation of the cellulosic fraction at pyrolysis

34

temperatures of 250 °C, whereas at higher temperatures (350-550 °C), the clay promotes

35

biomass degradation, resulting in an increase in both the concentrations of condensed

36

aromatic, acidic and phenolic carbon species. The porosity of the biochar, as measured by

37

NMR cryo-porosimetry, is altered by the iron-clay pretreatment. In presence of the clay, at

38

lower pyrolysis temperatures, the biochar develops a higher pore volume while at higher

39

temperature the presence of clay causes a reduction in the biochar pore volume. The most

40

dramatic reduction in pore volume is observed in the Kaolinite infiltrated biochar at 550

41

°C, which is attributed to the blocking of the mesopores (2-50 nm pore) by the non-porous

42

metakaolinite formed from kaolinite.

43

Keywords: biochar, clays, solid state NMR, TEM, cryoporometry.

44 45

INTRODUCTION

46

Biochar production is seen as a key part of a strategy to effectively recycle biomass carbon

47

and nutrients1, improve plant growth and mycorrhizal colonization2, improve soils physical

48

and chemical properties, provide a source of renewable energy and act as a means for

49

carbon sequestration3-6. However, in China and other parts of South East Asia, an estimated

50

100 million tonnes of biomass are burnt annually, either in the field or at processing

51

factories. This results in loss of carbon and nutrients from the soil and a very large increase

52

in harmful atmospheric carbon emissions7. The vast quantities of readily available biomass,

53

which are currently treated as waste for disposal, could be readily, cost effectively and

54

widely converted into valuable biochar4,

55

matter to biomass before or after pyrolysis can result in increased crop yields and

56

improvement of soil biome at application rates similar to those used for chemical

57

fertilizers1-2, 11-15. For biochar production to be cost effective, however, its properties need

58

to be tailored for specific applications, as has been the case of activated carbons for

59

engineering applications7, 13, 16.

8-10

. The addition of clay, minerals and organic

60

Investigations into biochar residues collected from fields reveals that biochar is

61

often coated in a small, but significant quantity of soil that typically contains a range of

62

clays, organic matter, sand, salts, residual chemical fertilizers (e.g. urea) and minerals that

63

include nano-phase iron-bearing compounds17. Joseph et. al,13 and Li et. al18 have

Page 2 of 25

ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25

Environmental Science & Technology

64

suggested that these mineral-biochar composites lead to a favorable plant response due to a

65

number of factors including: the enhanced redox activity19-20 and the ensuing biotic and

66

abiotic reactions which assist in nutrient cycling and plant uptake of nutrients13, 17-18.

67

Despite the significant advantages of such enriched biochar, there has been very

68

little research on specifically how such surface coatings of soil on the biomass change the

69

properties of the biochar. For example, previous work has demonstrated that biomass

70

pyrolyzed with clay at 600 °C resulted in a biochar that had a greater surface exchange

71

capacity and adsorption capacity for contaminants than the biochar produced without such

72

amendments21. Elsewhere, it was found that addition of biochar (along with bamboo

73

vinegar) to composting piles enhances nitrogen conservation and immobilizes heavy metal

74

ions22. Recent work also shows that addition of clay and iron minerals to the biochar

75

enhances their stability within the soil23. However, the specific biochar structure that leads

76

to the observed properties is not readily understood, and although mineral-biochar

77

composites are a subject of continuing research interest, much of the work has focused on

78

their properties or applications. 21, 24-25 Within the biochar literature there is generally only

79

a qualitative understanding of the molecular structure of mineral impregnated biochar 12, 26.

80

Furthermore, standards for biochar stability and properties currently being developed by

81

the International Biochar Initiative27 are primarily based on clean feedstock in stand-alone

82

pyrolysis facilities. Biochar produced from feedstock that has been standing in fields for

83

considerable periods of time or synthesized in simple ovens and kilns may nevertheless

84

have very different properties.

85

While the structure of biochar as a function of pyrolysis conditions has been

86

investigated28-29, to the best of our knowledge there has been no systematic study of the

87

more complex changes in the molecular and structural properties of biochar produced from

88

biomass coated with mineral phases. The study herein bridges this gap in knowledge and is

89

the first in a series that examines the interaction of biomass, clay and mineral species with

90

respect to the molecular structure and the micro- and meso-porosity of the biochar

91

produced at different temperatures. Our aim is to understand the effect of iron-clay

92

pretreatment on the degree of carbon condensation, the concentration of functional groups

93

in the biochar, the porosity of the biochar and the pyrolysis degradation pathway. The

94

degree of carbon condensation is directly relevant to the stability of the biochar and its

Page 3 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

95

environmental implications relating to carbon sequestration.30-31 The concentrations of

96

different functional groups, such as carboxylic acids, and the porosity of the biochar are

97

pertinent to the agricultural applications of biochar, while the ability to control the

98

degradation pathway is crucial for efficient production of the desired biochar structure. A

99

fundamental understanding of the changes in the structure of biochar as a function of

100

pyrolysis temperature and mineral additives is therefore crucial for the implementation of

101

strategies to design biochar with superior properties, tailored to enhance performance. For

102

this purpose, a series of biochars infiltrated with iron-clay minerals were synthesized and a

103

detailed, quantitative analysis of their structure and porosity using NMR spectroscopy

104

along with electron microscopy is reported herein.

105 106

MATERIALS AND METHODS

107

Synthesis of biochars. For the comparisons of molecular structure, it was important

108

to use a biomass source which had a uniformity of structure and as such bamboo is an ideal

109

option. Bamboo is a fast growing bio-resource, which yields extensive residue from the

110

bamboo processing industry. To ensure minimal mineral and chemical variation of a

111

naturally sourced lignocellulose feedstock, the biomass for the study was procured from a

112

single 1 meter section of bamboo to synthesize all the biochars. The bamboo section was

113

cut into cubes with a dimension of approximately 10 mm. Kaolinite (a 1:1 clay) and

114

bentonite (a 2:1 clay) were chosen as clay minerals to observe the influence of different

115

clay structures. Refined kaolinite and bentonite clays were sourced from: Keens Ceramics,

116

Gindurra Rd., Sommersby, NSW, 2250, as detailed in the supplementary information under

117

Technical Specifications. Iron is also seen to influence biochar properties, and

118

FeSO4.7H2O, chosen as the iron source for mineral-biochar composites, was sourced from

119

Sigma Aldrich. Two different iron-clay slurries were prepared to infiltrate the biomass with

120

the minerals. The first was iron-kaolinite slurry comprising a 1:1 by weight ratio of

121

FeSO4.7H2O and refined kaolinite clay. The second slurry was iron-bentonite slurry

122

prepared from a 1:1 by weight ratio of FeSO4.7H2O and bentonite clay. Two sets of

123

samples were prepared by soaking the bamboo cubes in the two iron-clay slurries at 80 °C

124

for 3 hours, conditions optimized to yield maximum infiltration of the slurry into the

125

biomass pores without drying out the slurry over the time period. A third set of bamboo

Page 4 of 25

ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25

Environmental Science & Technology

126

cubes was left untreated to act as a control. These three sets of bamboo samples were

127

prepared and dried at 110 °C for 24 hours prior to pyrolysis. The sample for pyrolysis was

128

placed in a purpose built high nickel chromium (Sandvik 253MA stainless steel) reaction

129

vessel with 1 mm hole to allow pyrolysis gas to escape but minimize any ingress of air, and

130

this vessel then placed inside an electric kiln with an air-starved environment. An

131

electronic temperature controller fitted to the kiln allowed controlled heating of the

132

biomass. The temperature was ramped at a rate of 3 °C/minute until the desired pyrolysis

133

temperature was attained, after which point it was held at its final pyrolysis temperature for

134

30 minutes to form biochar. After this pyrolysis time was passed, cold water was sprayed

135

onto the hot metal to create a steam environment and reduce the temperature to ~200 °C.

136

The vessel was removed and quenched in iced water to lower the temperature to ~80 °C to

137

prevent any chemisorption of air by the biochar. The biochar was then stored in air tight

138

container for further characterization. This process was repeated for each sample at the

139

final pyrolysis temperatures of 250 °C, 350 °C, 450 °C and 550 °C, making twelve samples

140

in all. In the ensuing discussion, we label the biochar from untreated bamboo as BC,

141

biochar from bamboo treated with iron-kaolinite as BKF and biochar from bamboo treated

142

with iron-bentonite as BBF. Three different batches from each of the biochar samples were

143

produced and these were then crushed and combined for analysis leading to more

144

representative results. The sample codes and preparation conditions are summarized in

145

Table 1.

146

Elemental analysis. Analysis of the aggregated samples for ultimate and ash

147

analysis was carried out four times (Table 1) as per ASTM D758232. The ash constituent

148

analysis was carried out according to ASTM D432633 using x-ray fluorescence (XRF)

149

spectroscopy.

150

SEM and TEM. Scanning electron microscope (SEM) analysis was performed

151

using a Zeiss Sigma SEM fitted with a Bruker X-ray energy dispersive spectrometry (EDS)

152

detector on pieces of biochar that had been mounted on carbon tape and chromium coated.

153

For transmission electron microscopy (TEM), the biochar samples were dispersed

154

ultrasonically in ethanol and a drop of the suspension was placed on a carbon-coated

155

copper grid with a pipette. The fine particles were examined using a JEOL 2100 TEM

156

fitted with an Oxford EDS system.

Page 5 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 25

157

Solid-state NMR spectroscopy. The samples of biochar were finely ground and

158

packed into 4 mm zirconia rotors fitted with Kel-F caps. The 13C NMR experiments were

159

performed on Bruker AVANCE III 300 spectrometer with a 7 Tesla superconducting

160

magnet operating at frequencies of 300 MHz and 75 MHz for the 1H and

161

respectively. The ratios of the protonated and non-protonated carbon in the chars were

162

determined according to the method of Brewer et al28. The quantitative

163

Polarization with Magic Angle Spinning (13C DPMAS) NMR spectra of the material were

164

acquired at 12 kHz MAS with a

165

decoupling, and a Hahn-echo before signal detection to eliminate baseline distortion. 10 s

166

recycle delays were used for the measurements, where greater than 95 % of the

167

relaxation was observed as determined by

168

Selective suppression of the protonated carbon species was achieved with dipolar

169

dephasing (dephasing time: 68 µs) before acquisition, and yielded a quantitative signal of

170

the condensed and non-protonated carbon species. The typical measurement time for each

171

13

172

solid state NMR spectroscopy are provided in the supplementary information Figure S5.

13

13

C nuclei

13

C Direct

C 90° pulse length of 4 µs, 80 kHz 1H SPINAL64

13

13

C

C cross-polarization/T1 (CP/T1) filter34.

C DPMAS experiment was 11 hours. Details for the mineral analysis using 27Al and 29Si

173

NMR Cryoporometry. The pore-size distribution and pore volumes of the stable

174

carbon skeleton were assessed using a Lab-Tools Mk1 (1970s) solid-state NMR

175

spectrometer with digital-switching transmitter. The magnet used was a low-field (0.5 T)

176

Mullard permanent magnet with a 34 mm pole-gap, with a low homogeneity, often

177

deliberately degraded to 100 µT over the 12 mm long and 2.4 mm diameter sample, in

178

order to have a clear echo at 1 ms. The details of the method are given by Webber et al.35

179

Repeats were carried out on two additional samples and indicated a variation of + 0.15

180

ml/gm.

181 182

RESULTS

183

Infiltration of the mineral phase into the biochar scaffold. The pyrolysis of the iron-

184

clay slurry treated bamboo biomass results in the incorporation of mineral phases into the

185

biochar scaffold. The SEM images of the higher temperature mineral biochar composites,

186

(BKF 450, BKF 550, BBF 450 and BBF 550), which are most relevant for potential

187

applications, are shown in Figure 1 and illustrate the presence of the mineral in the

Page 6 of 25

ACS Paragon Plus Environment

Page 7 of 25

Environmental Science & Technology

188

biochars. Figure 1(a) shows the coating of clay and iron that covers the outside of all of the

189

biochars at low magnification as evidenced by the rough particulate material, while Figure

190

1(b) shows the macroporous structure of the biochar. SEM images of the lower temperature

191

mineral biochar composites are provided in supplementary information (Figures S1 and

192

S2). It is possible the surface roughening seen in Figure 1(a), may be from the texture of

193

the biomass itself, therefore images were recorded at higher magnification. Figure 1(c) and

194

(d), clearly show the mineral phases in the form of particles and agglomerates coating the

195

surface of the biochar structure. The SEM was not sufficiently resolved to see the

196

morphology of the infiltrated mineral phases, therefore TEM analysis was used to

197

investigate the mineral structure on the internal pore surface of the biochar. Figure 1(e),

198

demonstrates that the internal surface of a biochar pore is indeed coated by mineral phases

199

(visualized as darker particles). The larger pore is probably derived from the carbonization

200

of phloem or xylem36 and contains a range of amorphous and crystalline nano-phases

201

composed of ferric and alumino-silicate species. Away from this macropore, we expect that

202

minerals have impregnated existing mesopores as well. At higher magnification analysis,

203

Figure 1(f) reveals that the mineral regions have a nano-structured dendritic morphology.

204

The elemental analysis by energy dispersive spectroscopy (EDS) as shown in, Figure 1(g),

205

indicates these dendritic nanostructures include an iron-bearing mineral phase. The SEM

206

and TEM analyses both substantiate that soaking biomass in an iron-clay slurry followed

207

by pyrolysis, result in the incorporation of nanostructured mineral phases within the porous

208

structure of the biochar. It is important to note that while SEM and TEM analysis provide

209

a detailed but local picture of the mineral coating to biochar, quantitative

210

spectroscopy and porosimetry were essential to establish mineral infiltration within the

211

bulk of the biochar and its influence on the molecular structure and porosity of the biochar.

13

C NMR

212 213

Elemental composition of the biochar. Table 1 summarizes the changes in the chemical

214

composition of the biochar produced at different temperatures. With increasing pyrolysis

215

temperature, the carbon content increases and the biochar yield decreases for the untreated

216

BC series of biochar. This is expected as the oxygen-bearing and hydrocarbon moieties are

217

volatilized at higher temperature. In the iron-clay biochar (BBF and BKF series), the

218

volatilization process also results in increased concentration of the clay minerals (and a

Page 7 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 25

219

relative reduction of carbon content). It is important to note that the Al, Fe and S contents

220

for the BKF and BBF samples were orders of magnitude greater than the BC samples

221

(Table 1), indicating uptake of the minerals by the biomass. In particular, the Al content in

222

Table 1 for the mineral infiltrated biochar varies from 0.01 to 0.1 wt. % (of the total

223

biochar weight). Form the chemical composition of kaolinite (Al2 Si2O5(OH)4 where Al

224

corresponds to ~20 % of the clay weight) and bentonite clay ((Na,Ca)0.3(Al,Mg)2Si4O10

225

(OH)2•n(H2O) where Al is ~10% of the clay weight), we estimate that the slurry treated

226

biochars contain between 0.1 to 1 wt% of clay mineral. Similar uptake of iron sulfate and

227

either of the clay minerals by the bamboo biomass from the iron-clay slurries was also

228

indicated by equivalent amounts of both Fe and Al for the BBF and BKF samples. The

229

total N content was observed to be lower for the BBF450 and BBF550 samples compared

230

with samples in the BC and BKF series. It is possible that the presence of the bentonite-

231

iron minerals in the biochar promotes volatilization of the nitrogenous species from the

232

biomass and further investigations are required to pinpoint the origin of this specific effect.

233

As the pyrolysis temperature increased, the concentrations of most of the mineral species

234

increased due to concomitant reduction in the biomass fraction due to loss of the oxygen

235

and hydrogen moieties during degradation and volatilization. The sulfur content for the

236

BBF and BKF samples reduced at 550 °C (after an initial increase at 250 °C due to

237

reduction of the organic content) indicating a volatilization of sulfur, consistent with the

238

degradation of iron sulfate at 550 °C37. We note that the variation in the carbon wt. % for

239

the different biochars is not monotonic and is indicative of their differing molecular

240

structures and ash content. Direct assessment of the differences in the carbon matrix of the

241

biochars, required a comprehensive analysis using quantitative

242

spectroscopy.

13

C solid-state NMR

243 244

Molecular structure of the biochar. The combination of the TEM and the elemental

245

analysis confirm the uptake and dispersion of mineral phases by the biomass from the

246

slurries. The molecular structure analysis of the biochar was conducted with quantitative

247

solid-state 13C DPMAS NMR spectra, which are shown in Figure 2. Immediately evident

248

in these spectra of the biochar are signals from cellulose (105 ppm, 90 ppm, 75 ppm and 65

249

ppm), hemicellulose (with additional peaks at 20.5 and 172 ppm) and lignin (165ppm - 110

Page 8 of 25

ACS Paragon Plus Environment

Page 9 of 25

Environmental Science & Technology

250

ppm and 55 ppm). Also evident are distinct differences in the structure of the biochar as a

251

function of the iron-clay pretreatment and the pyrolysis temperatures are observed. At the

252

lower pyrolysis temperature of 250 °C (Figure 2 a-c), the influence of the iron-clay is most

253

significant and is seen to retard the degradation of the cellulosic component. A comparison

254

of the intensity of the peak at 75 ppm (representative of the cellulose) and the -OCH3 peak

255

at 55 ppm (representing the lignin) shows that the BKF250 and BBF250 chars have a

256

significantly higher fraction of cellulose present, compared to the BC250 char.

257

Concomitantly, there is a strong signal of alkyl species in the region of 6-50 ppm for the

258

BC250 char that corresponds to volatile cellulosic degradation products (such as high

259

molecular weight alkenes). Although there is only about 1 wt. % of clay in the BKF and

260

BBF series of biochar, as found with the ultimate (elemental) analysis, the clay exerts a

261

strong effect in suppressing the degradation of the cellulose. The degradation of biomass is

262

dependent on the rate of heat / heat transfer into the biomass particle, which is driven by

263

conductive, convective and to a lesser extent by radiative mechanisms38-40. For large

264

particles, such as the bamboo sections used in the current work, and the relatively lower

265

temperature (250 °C), convective heat transfer is expected to be an important heat transfer

266

mechanism, particularly since the biomass is a poor conductor of heat. We hypothesize that

267

the coating of clay on the biomass reduces the gas permeability of the ligno-cellulosic

268

biomass thereby suppressing the convective heat transfer into the biomass and suppressing

269

degradation. The reduced gas permeability due to clays is particularly well known for the

270

case of clay-polymer nanocomposites41, where the incorporation of clay nanoparticles into

271

the polymer matrix suppresses gas diffusion through the polymer due to its very large

272

surface area (~ 1000 m2/g) of clay.

273

In contrast to the suppression of degradation at 250 °C, at elevated pyrolysis

274

temperatures (350 °C to 550 °C), the mineral species appear to have an opposite effect. At

275

350 °C (Figure 2 d-f) the structure of the untreated bamboo biochar (BC350) shows a

276

strong signal for aromatic carbon species (160 ppm to 100 ppm) as well as aliphatic carbon

277

species (50 ppm to 6 ppm). However, while the BKF350 and BBF350 biochar also show a

278

strong aromatic signal, the signal intensity of the aliphatic species is reduced compared

279

with that from the BC350 biochar. This difference is interpreted as an indicator that at the

280

higher temperatures, the presence of clays promotes the formation of condensed carbon

Page 9 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

281

structure and enhances the cross-linking of the volatile aliphatic fraction into condensed

282

aromatic species. This hypothesis is consistent with the documented ability of clay

283

minerals to catalyze condensation and polymerization reactions from a variety of substrates

284

such as aromatic ketones and aldehydydes.42-44 Investigations into low temperature (LT)

285

biochar (350 °C) has shown that they are particularly beneficial to sandy soils,16, 45 and the

286

ability to control the structure of the LT biochar may enable continued development of

287

designer biochar in this temperature regime. Our result also opens up the possibility to

288

study any possible downside46 to the presence low molecular degradation products in the

289

biochar, which may compete with its beneficial effects in the long term. At pyrolysis

290

temperatures of 450 °C - 550 °C, the

291

being centered at ~ 130 pm as see in Figure 2 (g-l). The primary species produced at these

292

temperatures have an aromatic carbon structure, in addition to some phenolic and

293

carboxylic species observed at ~150-160 ppm and 165-180 ppm, respectively.

294

The solid state

13

13

C NMR signal is very similar in all the materials

C DPMAS NMR spectra of the biochar samples enable the

295

quantitative assessment of the changes in the biochar structure due to the presence of the

296

mineral phases. The normalized

297

temperatures (Supplementary information Figure S3), illustrating the quantities of specific

298

carbon functionalities, viz: alkyl, HCOH, aromatic and acidic functionalities, present in the

299

biochar. The data indicate that the BKF and BBF biochars generally have a higher fraction

300

of non-protonated aromatic carbon as compared to the total carbon, with the BBF550

301

biochar having the highest concentration. This implies that the BBF and BKF series of

302

biochar have a greater degree of aromatic condensation that is correlated to greater stability

303

of the biochar within the soil.30-31 Additionally the BBF and BKF series also have a higher

304

concentration of phenolic species, particularly at 450 °C and 550 °C.

305

13

C NMR intensities are plotted against the pyrolysis

For the amount of detailed information

13

C NMR spectroscopy provides, a proper

306

quantification of carbon species is important to avoid errors in estimating the influence of

307

mineral additives on biochar structure. As a case in point, we offer a specific study

308

(Supplementary Information Figure S6), in which the mistaken application and evaluation

309

of the NMR data has led to incorrect conclusions regarding the influence of the mineral

310

additives on biochar structure.

311

The overall effect of the pyrolysis temperature and the presence of clay on the

Page 10 of 25

ACS Paragon Plus Environment

Page 10 of 25

Page 11 of 25

Environmental Science & Technology

312

biochar is better visualized by plotting the Van Kervelen diagram (the O:C vs. H:C atomic

313

ratios along the x and y axes respectively), for each pyrolysis temperature, as shown in

314

Figure 3. The data for the Van Kervelen diagram is extracted directly from the 13C NMR

315

data (Supplementary Information Figure S3) which provides the ratios of the different C-H,

316

C-O and quaternary C species in each biochar sample. For a given biochar, we can

317

determine precisely which functionality is responsible for the observed hydrogen or oxygen

318

content. Figure 3 indicates that the evolution of the biochar structure is very dependent on

319

the presence of the iron-clay mineral during pyrolysis and that this effect is more

320

pronounced at the lower temperatures. At lower temperatures, (250 °C), the BKF and BBF

321

biochars have an increased amount of oxygen and a reduced amount of hydrogen as

322

compared to the biochar without clay. At the intermediate temperatures, (350 °C to 450 °C)

323

the compositions of the BKF and BBF biochars deviate the most from BC biochar.

324

Specifically, the presence of mineral species suppresses the removal of the oxygen

325

moieties, while enhancing the removal of the hydrogen moieties. The differences in the

326

Van Kervelen diagram at 350 °C and 450 °C can be deduced from the data in Figure S3,

327

and assigned to the reduced amounts of alkyl CH, CH2, CH3 species and increased amount

328

of acidic and phenolic species in the BBF and BKF biochars as compared to the BC

329

biochar. At the highest temperature (550 °C), the different biochars yield similar molecular

330

structures. This similarity can be seen as a strong indicator that the clay-mineral exerts a

331

catalytic influence on the structure of the evolving biochar by lowering the free energy for

332

the condensation reactions (particularly at the lower temperatures). While individual

333

reaction mechanisms during pyrolysis of biomass are exceedingly complex38,

334

hypothesize that since reaction rates follow an exponential function of temperature, at the

335

increased temperature (550 °C), there is sufficient energy to overcome the barriers to

336

promoting carbon condensation, resulting in the similar biochar structures observed.

337

Further kinetic studies need to be undertaken to confirm this hypothesis. With respect to

338

the stability of the biochar, Figure 3 indicates that the BBF550 biochar yields the most

339

condensed and therefore most stable material,12,

340

biochar having comparatively less condensed aromatic carbon, would be expected to have

341

lower stability in the soil. The differences in the BBF550 and BKF550 biochar can be

342

related to the dramatic changes in the structure of the clay (vide infra) which occurs at the

Page 11 of 25

30-31, 48

47

, we

while the BC550 and BKF550

ACS Paragon Plus Environment

Environmental Science & Technology

343

higher pyrolysis temperature49.

344 345

Inclusion and nature of the iron-bearing phases in the biochar. The intensity of the 13C

346

spinning sidebands (ssb) gives an insight into the structure of the iron-bearing phases

347

(Figure 2 and Supplementary Information, Figure S4). Generally, in organic solids, the

348

ssb’s arise as a result of the Magic Angle Spinning (MAS) averaging out the chemical shift

349

anisotropy (CSA), and occur at multiples of the spinning frequency from the center-band.

350

A faster MAS speed results in smaller ssb’s which are further away from the isotropic

351

peaks. The full spinning sideband envelope for the

352

supplementary information Figure S4. At the experimental conditions of 7 Tesla magnetic

353

field strength and 12 kHz MAS, we can expect the

354

intensity first order ssb’s and no higher order ssb’s. This is indeed the case for the all the

355

biochar produced at 250 °C as well as the BC series pyrolyzed at higher temperatures

356

(Supplementary information Figures S4 a-d, g and j respectively). In contrast, the BBF

357

biochars pyrolyzed at 450 °C and 550 °C have particularly high intensity ssb’s, with fourth

358

order ssb’s being visible (Supplementary information Figure S4, i and l). Such unusually

359

high intensity of the ssb’s can be attributed to a paramagnetic induced anisotropy arising

360

from the presence of highly paramagnetic phases formed from the adsorded iron salts, in

361

close (nanometer scale) proximity to the condensed carbon structure of the biochar. It is

362

well known that changes in the structure of iron oxides, such as Fe2O3, from the bulk scale

363

(micron-sized or larger) to nano-scale results in their transition from a ferromagnetic to a

364

super-paramagnetic state.50 Indeed in Figure 1, the TEM images corroborate that there are

365

nano-scale iron-bearing phases within the porous structure of the biochar. The intensity of

366

the ssb’s in the 13C solid-state NMR spectra, particularly of the BBF at 450 °C and 550 °C

367

therefore indicate that the nano-sized iron phases are distributed fairly uniformly

368

throughout the biochar. It is important to note that while both the bentonite and kaolinite

369

clay slurries were prepared with iron sulfate, the strongest influence from the paramagnetic

370

phases is observed with the biochar derived from the iron-bentonite slurry treated biomass.

371

This implies that the specific presence of the bentonite clay acts as a catalyst for the

372

formation of super-paramagnetic nano-Fe2O3 phases at temperatures above 450 °C. Kaolin

373

undergoes a structural transformation to metakaolinite in this temperature range, as

Page 12 of 25

13

C NMR spectra is plotted in the

13

C NMR spectrum to have very low

ACS Paragon Plus Environment

Page 12 of 25

Page 13 of 25

Environmental Science & Technology

374

revealed by the solid state 27Al MAS NMR spectra (Supplementary Figure S5 (e) and (f))

375

and may not be able to exert a similar catalytic effect

376 377

Porosity of the biochar. The SEM and TEM (Figure 1) images show a complex porous

378

structure of the biochar with multiple length scales. NMR cryoporometry provides the total

379

pore volume of the biochar samples as shown in Table 2, as well as the distribution of pore

380

sizes as shown in Figure 4. NMR cryoporometry has the additional and significant

381

advantage of being applicable even when there are liquids and volatile components already

382

in the pores. The untreated bamboo biochar, BC (Table 2) follows the normal pattern of

383

increased pore volume with pyrolysis temperature from 250 °C to 350 °C, while above 350

384

°C, the pore volume is nearly constant.35 The lower pore volume at 250 °C in BC biochar is

385

attributable to hydrophobic volatiles formed from the biomass degradation, condensing into

386

the pores of the forming biochar.51-52 In comparison to the BC250 biochar, the BKF250

387

and BBF250 biochars have nearly double the pore volume which is attributable to the

388

reduced concentration of hydrophobic volatiles trapped in these chars as seen from the 13C

389

NMR (Figure 2 b and c). Within the error bar of the measurement the BKF and BBF

390

samples show pore-volume and pore-size distribution similar to the corresponding BC

391

biochar at 350 °C and 450 °C. The increase in pyrolysis temperature accelerates the

392

reaction between the iron, the clay and the volatiles that are liberated and the data indicate

393

that the continued aromatic condensation reactions at higher temperatures may lead to a

394

closing of the sub-micron pores or at least to an occlusion of the access to the pore space.

395

We also note that to some degree, the reduction may also be attributable to the enhanced

396

1

397

due to the increased aromaticity in the BBF and BKF series, as compared to the BC series,

398

which may yield lower than expected values for the porosity. However, in the particular

399

case of the BKF550 biochar, there is a large reduction in the pore volume as seen in Figure

400

4 (b) which cannot be assigned to any effect other that the collapse of the mesopore

401

volume. Considering that at 550 °C the molecular structure of the carbon matrix for both

402

BBF and BKF is quite similar (as seen in Figure 2), the change in porosity for the BKF550

403

must be due to another factor. Examination of the thermally treated neat iron-sulfate-clay

404

mixtures by solid state 29Si and 27Al NMR (Supplementary information Figure S5) shows

H-T2 relaxation of the water caused by the iron phases and reduced adsorption of water

Page 13 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

405

that while the iron-bentonite is stable to temperatures above 550 °C, the iron-kaolinite

406

undergoes a degradation-transformation into metakaolinite at 550 °C. Thermal

407

decomposition of kaolinite in this way, can explain the low porosity of the BKF550

408

biochar, as well as the sudden change in the Van Kervelen diagram (Figure 3). Below the

409

thermal decomposition temperature of the kaolinite (i.e. at 450 °C) both bentonite and

410

kaolinite have high surface areas that we deduce results in producing biochars with similar

411

structures. However, the metakaolinite formed at 550 °C, has a low surface area and lacks

412

the layered structure of kaolinite. This reduction in porosity, specifically in the mesoporous

413

regime, can be directly attributed to the metakaolinite blocking the mesoporous structure in

414

the biochar. Concomitantly, the reduced surface area of the metakaolinite, and, hence, the

415

reduced mineral-biochar interactions, significantly reduces the influence of the clay on the

416

biochar structure. Therefore, the BKF biochar shifts to a structure very similar to that of the

417

BC biochar. In contrast to the kaolinite, the bentonite clay does not undergo a structural

418

collapse at 550 °C as suggested by the

419

influence through the high interfacial contact with the biochar. This insight is crucial since

420

it can explain results observed in literature, such as the reduced surface adsorption capacity

421

of kaolinite-bamboo biochar (600 °C) seen by Yao et al.21

422 423

Implications for the synthesis of enhanced biochar. Previous work has demonstrated

424

that properties of the biochar, such as the concentration of acidic functionalities, the degree

425

of aromatic condensation and porosity of the biochar play key roles in determining its

426

stability within the soil, its ability to enhance soil fertility and its capability for pollutant

427

sequestration6, 17. Here, we provide quantitative

428

which demonstrate that the concentration of acidic functionalities (particularly the phenolic

429

species), the degree of aromatic condensation and the pore volume can be tailored to a high

430

degree by infiltrating the biomass with high surface area clay minerals and iron salts prior

431

to pyrolysis. Such control on the structure of the biochar will allow for selecting biochar

432

for specific applications. For example, as a means for carbon sequestration, the BBF550

433

biochar that has the highest degree of aromatic condensation, would be the most suitable,

434

as a higher degree of aromatic condensation is linked to enhanced biochar stability within

435

the soil3, 31. On the other hand, to enhance the soil fertility or immobilize soil contaminants,

Page 14 of 25

29

Si and

27

Al NMR, and continues to exert its

13

C NMR and porosimetry measurements

ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25

Environmental Science & Technology

436

the BKF450 and BBF450 biochar with their higher concentration of phenolic and

437

carboxylic species as well as higher porosity would be well suited. Furthermore, infiltration

438

and formation of clay and metal oxide nanostructures within the pores of the biochar will

439

promote the redox reactions that can have a significant impact on nutrient availability and

440

uptake in plants, as well as increase beneficial micro-organisms which improve the health

441

of the rhizosphere18,

442

structure is enabled by use of materials which are inexpensive and readily available across

443

a wide geographical distribution. Detailed investigations of the composition, nanostructure

444

and magnetic properties of the mineral phases will soon be forthcoming elsewhere. The

445

next step will be to demonstrate the differences in the activity of these biochar materials

446

under field conditions and correlate them to the molecular structure measured herein.

53-58

. Finally, it is important to note that the control of the biochar

447 448

ASSOCIATED CONTENT

449

Supporting Information. Text figures and tables addressing the transformation of minera

450

impregnated biochar under changing pyrolysis conditions. SEM images of mineral on

451

biochar pores. Variation of biochar functionality as function of temperature. 13C NMR

452

detection of paramagnetic mineral infiltration into biochar pores. Effect of heating on 29Si

453

and

454

kaolinite from supplier. This information is available free of charge via the Internet at

455

http://pubs.acs.org/.

27

Al NMR of bentonite and kaolinite. Technical data/specifications of bentonite and

456 457

AUTHOR INFORMATION

458

Corresponding Author

459

*Phone:+61 02 93854616; e-,mail [email protected]

460

ACKNOWLEDGMENTS

461

We acknowledge the assistance of the Electron Microscope and X-ray unit of University of

462

Newcastle and the EMU and NMR Facility of the Mark Wainwright Analytical Centre,

463

University of NSW. This work was supported by ARC grant LP120200418, Renewed

464

Carbon Pty Ltd. and the DAFF Carbon Farming Futures Filling the Research Gap

465

(RG134978). We acknowledge extensive discussions with Dr. B. Njegic for insightful

466

comments regarding the manuscript.

Page 15 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

467 468

References

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

1. Kammann, C. I.; Schmidt, H.-P.; Messerschmidt, N.; Linsel, S.; Steffens, D.; Mueller, C.; Koyro, H.-W.; Conte, P.; Stephen, J., Plant growth improvement mediated by nitrate capture in co-composted biochar. Scientific Reports 2015, 5. 2. Blackwell, P.; Joseph, S.; Munroe, P.; Anawar, H. M.; Storer, P.; Gilkes, R. J.; Solaiman, Z. M., Influences of Biochar and Biochar-Mineral Complex on Mycorrhizal Colonisation and Nutrition of Wheat and Sorghum. Pedosphere 2015, 25 (5), 686-695. 3. Singh, B. P.; Cowie, A. L.; Smernik, R. J., Biochar carbon stability in a clayey soil as a aunction of feedstock and pyrolysis temperature. Environ. Sci. Technol. 2012, 46 (21), 11770-11778. 4. Woolf, D.; Amonette, J. E.; Street-Perrott, F. A.; Lehmann, J.; Joseph, S., Sustainable biochar to mitigate global climate change. Nat. Commun. 2010, 1 (56). 5. Harvey, O. R.; Kuo, L. J.; Zimmerman, A. R.; Louchouarn, P.; Amonette, J. E.; Herbert, B. E., An Index-Based Approach to Assessing Recalcitrance and Soil Carbon Sequestration Potential of Engineered Black Carbons (Biochars). Environ. Sci. Technol. 2012, 46 (3), 1415-1421. 6. Atkinson, C. J.; Fitzgerald, J. D.; Hipps, N. A., Potential mechanisms for achieving agricultural benefits from biochar application to temperate soils: a review. Plant Soil 2010, 337 (1-2), 1-18. 7. Clare, A.; Barnes, A.; McDonagh, J.; Shackley, S., From rhetoric to reality: farmer perspectives on the economic potential of biochar in China. Int. J. Agr. Sustain. 2014, 12 (4), 440-458. 8. Clare, A.; Shackley, S.; Joseph, S.; Hammond, J.; Pan, G.; Bloom, A., Competing uses for China's straw: the economic and carbon abatement potential of biochar. Glob. Change Biol. Bioenergy 2015, 7 (6), 1272-1282. 9. Roberts, K. G.; Gloy, B. A.; Joseph, S.; Scott, N. R.; Lehmann, J., Life Cycle Assessment of Biochar Systems: Estimating the Energetic, Economic, and Climate Change Potential. Environ. Sci. Technol. 2010, 44 (2), 827-833. 10. Shackley, S.; Clare, A.; Joseph, S.; McCarl, B.; Schmidt, H.-P., Economic evaluation of biochar systems: : current evidence and challenges In Biochar for Environmental Management : Science, Technology and Implementation, 2 ed.; Lehmann, J.; Joseph, S., Eds. Taylor & Francis: London and New York, 2015; Vol. 1, pp 813-851. 11. Nielsen, S.; Minchin, T.; Kimber, S.; van Zwieten, L.; Gilbert, J.; Munroe, P.; Joseph, S.; Thomas, T., Comparative analysis of the microbial communities in agricultural soil amended with enhanced biochars or traditional fertilisers. Agric. Ecosyst. Environ. 2014, 191, 73-82. 12. Chia, C. H.; Singh, B. P.; Joseph, S.; Graber, E. R.; Munroe, P., Characterization of an enriched biochar. J. Anal. Appl. Pyrolysis 2014, 108, 26-34. 13. Joseph, S.; van Zwieten, L.; Chia, C. H., "Designing" of Biochar for Specific Applications to Soils; A Technology in Its Infancy. In Biochar and Soil Biota, Ladygina, N.; Rineau, F., Eds. CRC Press: Boca Raton, 2013. 14. Yao, C.; Joseph, S.; Li, L.; Pan, G.; Lin, Y.; Munroe, P.; Pace, B.; Taherymoosavi, S.; Van Zwieten, L.; Thomas, T.; Nielsen, S.; Ye, J.; Donne, S., Developing More Effective Enhanced Biochar Fertilisers for Improvement of Pepper Yield and Quality.

Page 16 of 25

ACS Paragon Plus Environment

Page 16 of 25

Page 17 of 25

512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557

Environmental Science & Technology

Pedosphere 2015, 25 (5), 703-712. 15. Joseph, S.; Anawar, H. M.; Storer, P.; Blackwell, P.; Chia, C.; Lin, Y.; Munroe, P.; Donne, S.; Horvat, J.; Wang, J.; Solaiman, Z. M., Effects of Enriched Biochars Containing Magnetic Iron Nanoparticles on Mycorrhizal Colonisation, Plant Growth, Nutrient Uptake and Soil Quality Improvement. Pedosphere 2015, 25 (5), 749-760. 16. Novak, J. M.; Lima, I. M.; Xing, B.; Gaskin, J. W.; Steiner, C.; Das, K. C.; Ahmedna, M.; Rehrah, D.; Watts, D. W.; Busscher, W. J.; Schomberg, H., Characterization of designer biochar produced at different temperatures and their effects on a loamy sand. Ann. Environ. Sci. 2009, 3, 195-206. 17. Joseph, S.; Graber, E. R.; Chia, C.; Munroe, P.; Donne, S.; Thomas, T.; Nielsen, S.; Marjo, C.; Rutlidge, H.; Pan, G. X.; Li, L.; Taylor, P.; Rawal, A.; Hook, J., Shifting paradigms: development of high-efficiency biochar fertilizers based on nano-structures and soluble components. Carbon Manag. 2013, 4 (3), 323-343. 18. Li, Y.; Yu, S.; Strong, J.; Wang, H., Are the biogeochemical cycles of carbon, nitrogen, sulfur, and phosphorus driven by the "Fe-III-Fe-II redox wheel" in dynamic redox environments? J. Soils Sediments 2012, 12 (5), 683-693. 19. Graber, E. R.; Tsechansky, L.; Lew, B.; Cohen, E., Reducing capacity of water extracts of biochars and their solubilization of soil Mn and Fe. Eur. J. Soil. Sci. 2014, 65 (1), 162-172. 20. Kluepfel, L.; Keiluweit, M.; Kleber, M.; Sander, M., Redox Properties of Plant Biomass-Derived Black Carbon (Biochar). Environ. Sci. Technol. 2014, 48 (10), 56015611. 21. Yao, Y.; Gao, B.; Fang, J.; Zhang, M.; Chen, H.; Zhou, Y.; Creamer, A. E.; Sun, Y.; Yang, L., Characterization and environmental applications of clay-biochar composites. Chem. Eng. J. 2014, 242, 136-143. 22. Chen, Y.-X.; Huang, X.-D.; Han, Z.-Y.; Huang, X.; Hu, B.; Shi, D.-Z.; Wu, W.-X., Effects of bamboo charcoal and bamboo vinegar on nitrogen conservation and heavy metals immobility during pig manure composting. Chemosphere 2010, 78 (9), 1177-1181. 23. Yang, F.; Zhao, L.; Gao, B.; Xu, X.; Cao, X., The Interfacial Behavior between Biochar and Soil Minerals and Its Effect on Biochar Stability. Environ. Sci. Technol. 2016, 50 (5), 2264-2271. 24. Wang, S.; Gao, B.; Zimmerman, A. R.; Li, Y.; Ma, L.; Harris, W. G.; Migliaccio, K. W., Removal of arsenic by magnetic biochar prepared from pinewood and natural hematite. Bioresour. Technol. 2015, 175, 391-395. 25. Wang, S.; Gao, B.; Li, Y.; Mosa, A.; Zimmerman, A. R.; Ma, L. Q.; Harris, W. G.; Migliaccio, K. W., Manganese oxide-modified biochars: Preparation, characterization, and sorption of arsenate and lead. Bioresour. Technol. 2015, 181, 13-17. 26. Li, F.; Cao, X.; Zhao, L.; Wang, J.; Ding, Z., Effects of Mineral Additives on Biochar Formation: Carbon Retention, Stability, and Properties. Environ. Sci. Technol. 2014, 48 (19), 11211-11217. 27. http://www.biocharinternational.org/sites/default/files/IBI_Biochar_Standards_V2%200_final_2014.pdf. 28. Brewer, C. E.; Schmidt-Rohr, K.; Satrio, J. A.; Brown, R. C., Characterization of biochar from fast pyrolysis and gasification systems. Environ. Prog. Sustain. 2009, 28 (3), 386-396. 29. Keiluweit, M.; Nico, P. S.; Johnson, M. G.; Kleber, M., Dynamic molecular

Page 17 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603

structure of plant biomass-derived black carbon (biochar). Environ. Sci. Technol. 2010, 44 (4), 1247-1253. 30. McBeath, A. V.; Smernik, R. J.; Schneider, M. P. W.; Schmidt, M. W. I.; Plant, E. L., Determination of the aromaticity and the degree of aromatic condensation of a thermosequence of wood charcoal using NMR. Org. Geochem. 2011, 42 (10), 1194-1202. 31. Zimmerman, A. R.; Gao, B., The Stability of Biochar in the Environment: Introduction. CRC Press: Boca Raton, Fl, 2013; p 1-40. 32. ASTM D7582. http://www.astm.org/Standards/D7582.htm. 33. D4326, A. http://www.astm.org/Standards/D4326.htm. 34. Torchia, D. A., Measurement of proton-enhanced C-13 T1 values by a method which suppresses artifacts J. Magn. Reson. 1978, 30 (3), 613-616. 35. Webber, J. B. W.; Corbett, P.; Semple, K. T.; Ogbonnaya, U.; Teel, W. S.; Masiello, C. A.; Fisher, Q. J.; Valenza, J. J.; Song, Y. Q.; Hu, Q. H., An NMR study of porous rock and biochar containing organic material. Microporous Mesoporous Mater. 2013, 178, 94-98. 36. Ronsse, F.; Nachenius, R. W.; Prins, W., Carbonization of Biomass. In Recent Advances in Thermochemical Conversion of Biomass, I ed.; Pandey, A.; Bhaskar, T.; Stocker, M.; Sukumaran, R., Eds. Elsevier: Amsterdam, 2015; pp 293-324. 37. Coombs, P. G.; Munir, Z. A., The decomposition of iron(III) sulfate in air. J. Therm. Anal. 1989, 35 (3), 967-976. 38. Ko, K. H.; Rawal, A.; Sahajwalla, V., Analysis of thermal degradation kinetics and carbon structure changes of co-pyrolysis between macadamia nut shell and PET using thermogravimetric analysis and C-13 solid state nuclear magnetic resonance. Energy Convers. Manage. 2014, 86, 154-164. 39. Dhungana, A.; Basu, P.; Dutta, A., Effects of Reactor Design on the Torrefaction of Biomass. J. Energ. Resour. -ASME 2012, 134 (4). 40. Papadikis, K.; Gu, S.; Bridgwater, A. V., Computational modelling of the impact of particle size to the heat transfer coefficient between biomass particles and a fluidised bed. Fuel Process. Technol. 2010, 91 (1), 68-79. 41. Choudalakis, G.; Gotsis, A. D., Permeability of polymer/clay nanocomposites: A review. Eur. Polym. J. 2009, 45 (4), 967-984. 42. Rocchi, D.; Gonzalez, J. F.; Menendez, J. C., Montmorillonite Clay-Promoted, Solvent-Free Cross-Aldol Condensations under Focused Microwave Irradiation. Molecules 2014, 19 (6), 7317-7326. 43. Varma, R. S., Clay and clay-supported reagents in organic synthesis. Tetrahedron 2002, 58 (7), 1235-1255. 44. Nagendrappa, G., Organic synthesis using clay and clay-supported catalysts. App. Clay Sci. 2011, 53 (2), 106-138. 45. Novak, J. M.; Busscher, W. J.; Watts, D. W.; Amonette, J. E.; Ippolito, J. A.; Lima, I. M.; Gaskin, J.; Das, K. C.; Steiner, C.; Ahmedna, M.; Rehrah, D.; Schomberg, H., Biochars Impact on Soil-Moisture Storage in an Ultisol and Two Aridisols. Soil Science 2012, 177 (5), 310-320. 46. Anjum, R.; Krakat, N.; Reza, M. T.; Klocke, M., Assessment of mutagenic potential of pyrolysis biochars by Ames Salmonella/mammalian-microsomal mutagenicity test. Ecotox. Environ. Safe. 2014, 107, 306-312. 47. Sinha, S.; Jhalani, A.; Ravi, M. R.; Ray, A., Modelling of pyrolysis in wood: A

Page 18 of 25

ACS Paragon Plus Environment

Page 18 of 25

Page 19 of 25

604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649

Environmental Science & Technology

review. J Solar Energy Soc India (SESI) 2000, 10, 41-62. 48. Mao, J. D.; Johnson, R. L.; Lehmann, J.; Olk, D. C.; Neves, E. G.; Thompson, M. L.; Schmidt-Rohr, K., Abundant and Stable Char Residues in Soils: Implications for Soil Fertility and Carbon Sequestration. Environ. Sci. Technol. 2012, 46 (17), 9571-9576. 49. Meinhold, R. H.; Mackenzie, K. J. D.; Brown, I. W. M., Thermal-reactions of kaolinite studied by solid-state 2-Al and 29-Si NMR. J. Mater. Sci. Lett. 1985, 4 (2), 163166. 50. Kucheryavy, P.; He, J.; John, V. T.; Maharjan, P.; Spinu, L.; Goloverda, G. Z.; Kolesnichenko, V. L., Superparamagnetic iron oxide nanoparticles with variable size and an iron oxidation state as prospective imaging agents. Langmuir 2013, 29 (2), 710-716. 51. Park, J.; Hung, I.; Gan, Z.; Rojas, O. J.; Lim, K. H.; Park, S., Activated carbon from biochar: Influence of its physicochemical properties on the sorption characteristics of phenanthrene. Bioresour. Technol. 2013, 149, 383-389. 52. Zhang, J. N.; Lu, F.; Zhang, H.; Shao, L. M.; Chen, D. Z.; He, P. J., Multiscale visualization of the structural and characteristic changes of sewage sludge biochar oriented towards potential agronomic and environmental implication. Scientific Reports 2015, 5. 53. Joseph, S. D.; Camps-Arbestain, M.; Lin, Y.; Munroe, P.; Chia, C. H.; Hook, J.; van Zwieten, L.; Kimber, S.; Cowie, A.; Singh, B. P.; Lehmann, J.; Foidl, N.; Smernik, R. J.; Amonette, J. E., An investigation into the reactions of biochar in soil. Aust. J. Soil Res. 2010, 48 (6-7), 501-515. 54. Bartlett, R. J.; James, B. R., Redox Chemistry of Soils. Advan. Agron. 1993, 50, 151-208. 55. Chen, S.; Rotaru, A.-E.; Shrestha, P. M.; Malvankar, N. S.; Liu, F.; Fan, W.; Nevin, K. P.; Lovley, D. R., Promoting Interspecies Electron Transfer with Biochar. Scientific Reports 2014, 4. 56. Husson, O., Redox potential (Eh) and pH as drivers of soil/plant/microorganism systems: a transdisciplinary overview pointing to integrative opportunities for agronomy. Plant Soil 2013, 362 (1-2), 389-417. 57. Joseph, S.; Husson, O.; Graber, E. R.; van Zwieten, L.; Taherymoosavi, S.; Thomas, T.; Nielsen, S.; Ye, J.; Pan, G.; Chia, C.; Munroe, P.; Allen, J.; Lin, Y.; Fan, X.; Donne, S., The Electrochemical Properties of Biochars and How They Affect Soil Redox Properties and Processes. Agronomy-Basel 2015, 5 (3), 322-340. 58. Lin, Y.; Munroe, P.; Joseph, S.; Kimber, S.; Van Zwieten, L., Nanoscale organomineral reactions of biochars in ferrosol: an investigation using microscopy. Plant Soil 2012, 357 (1-2), 369-380.

Page 19 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

650 651 652 653 654

Page 20 of 25

Table 1. Composition (weight %) of the biochar materials by ASTM D758232 and ASTM D432633. Kao refers to kaolinite and Ben refers to bentonite. Sample

Pyrolysis temp.

Minerals

BC250

250 °C

-

78

BC350

350 °C

-

BC450

450 °C

BC550

550 °C

Yield % C %

H%

N%

S % Fe %

Al %

Ca % K % Ash %

51.9

5.45

0.87

0.08 0.014 0.0016 0.050 0.83 14.38

60

71.5

4.02

1.11

0.07 0.011 0.0022 0.075

1.4

28.08

-

35

75

3.42

1.38

0.12 0.023 0.0051 0.098

1.4

19.10

-

33

79.2

2.72

1.28

0.06 0.030 0.0050 0.078 0.85 26.34

BKF250 250 °C FeSO4 +Kao.

86

48.7

5.15

0.5

1.0

2.0

0.026 0.046 0.74 15.43

BKF350 350 °C FeSO4 +Kao.

59

67.6

3.3

0.87

1.1

2.1

0.042 0.054

1.0

11.40

BKF450 450 °C FeSO4 +Kao.

48

62.9

2.82

1.45

1.1

2.9

0.093 0.069

1.2

35.60

BKF550 550 °C FeSO4 +Kao.

38

63.6

1.72

1.39

0.52

2.3

0.16

BBF250 250 °C FeSO4 + Ben.

76

48.7

5.15

0.5

0.73

1.2

0.010 0.053 0.35 15.91

BBF350 350 °C FeSO4 + Ben.

63

57.2

3.51

0.8

1.4

2.5

0.029

0.12

0.82 24.14

BBF450 450 °C FeSO4 + Ben.

42

61

2.79

0.84

1.3

2.5

0.034

0.12

0.63 56.87

BBF550 550 °C FeSO4 + Ben.

38

70.2

2.49

0.79

0.63

1.7

0.10

0.10

0.37 31.64

0.049 0.64 14.61

655 656 657 658

Table 2. Pore Volume as a function of temperature and clay treatment from the NMR cryo-

659

porosimetry. Error bar for the measurements is + 0.15 ml/g

660 661 Pyrolysis BC series of BKF series of BBF series of Temperature biochar biochar biochar o -1 -1 C (Pore vol ml.g ) (Pore vol ml.g ) (Pore vol ml.g-1) 250

0.36

0.73

0.88

350

1.05

0.86

0.77

450

0.77

0.56

0.73

550

1.01

0.16

0.42

662 Page 20 of 25

ACS Paragon Plus Environment

Page 21 of 25

Environmental Science & Technology

663 664 665

666 667 668

Figure 1. SEM images of BKF550 biochar (a), BKF450 biochar (b), BBC550 biochar (c)

669

and BBC450 biochar (d). The TEM image of a macropore of BKF550 (e) shows there are a

670

range of minerals (dark phases) both in the mesopores of the amorphous carbon lattice and

671

on the surface. TEM image of the nanostructured mineral decorating in the inner pore

672

surface (f). TEM-EDS spectra of the nanostructured mineral identifying it as a Fe-rich

Page 21 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

673

mineral (g).

674 675

Figure 2. Quantitative 13C DP MAS NMR spectra of the BC biochar series (a, d, g, and j),

676

the BKF biochar series (b, e, h, and k) and the BBF biochar series (c, f, i, and l). The

677

spectra of the chars are quantitatively measured by 13C DPMAS NMR. The spectra plotted

678

in bold lines represent the total 13C NMR signal while the spectra plotted in thin line were

679

acquired after 67 µs of recoupled dipolar dephasing and represent the signal of the non-

680

protonated carbon species. The difference between the bold and narrow spectra represents

681

the amount of protonated carbon species. * represents the spinning sidebands

682

Page 22 of 25

ACS Paragon Plus Environment

Page 22 of 25

Page 23 of 25

Environmental Science & Technology

683 684 685

Figure 3. Van Kervelen diagram (H:C vs. O:C atomic ratios) of the BC biochar series

686

(squares), the BKF biochar series (circles) and the BBF biochar series (triangles). The

687

atomic ratios are derived from the quantitative analysis of the functional groups in the

688

biochars plotted in Supplementary information Figure S2. The dashed lines are guides to

689

the eye.

690 691 692 693 694 695 696 697 698 699

Page 23 of 25

ACS Paragon Plus Environment

Environmental Science & Technology

700 701 702

Figure 4. Porosity as a function of pore diameter for BC biochar (a), BKF biochar (b) and

703

BBF biochar (c) as measured by NMR cryo-porosimetry. The curves are for biochars

704

pyrolyzed at 250 °C (thin line), 350 °C thin dashed line, 450 °C bold line and 550 °C (bold

705

dashed line).

706 707

Page 24 of 25

ACS Paragon Plus Environment

Page 24 of 25

Page 25 of 25

708 709 710

Environmental Science & Technology

For Table of Contents Only

Page 25 of 25

ACS Paragon Plus Environment