Impact of Pyrolysis Temperature on Charcoal Characteristics

Oct 30, 2018 - Characterization of these charcoals was conducted using eight analytical methods. Each method describes specific changes in the tempera...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF LOUISIANA

Applied Chemistry

Impact of pyrolysis temperature on charcoal characteristics Johannes Tintner, Christoph Preimesberger, Christoph Pfeifer, Denis Soldo, Franz Ottner, Karin Wriessnig, Harald Rennhofer, Helga C. Lichtenegger, Etelvino H Novotny, and Ena Smidt Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.8b04094 • Publication Date (Web): 30 Oct 2018 Downloaded from http://pubs.acs.org on November 3, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

Impact of pyrolysis temperature on charcoal characteristics

2

Johannes Tintner1*, Christoph Preimesberger1, Christoph Pfeifer1, Denis Soldo1, Franz Ottner2,

3

Karin Wriessnig2, Harald Rennhofer1, Helga Lichtenegger1, Etelvino H. Novotny3, Ena Smidt1

4

1

5

Process Engineering, 1190 Vienna, Austria

6

2

7

Natural Hazards, 1190 Vienna, Austria

8

3

Embrapa Solos, 22460-000 Rio de Janeiro-RJ, Brazil

9

*

corresponding author: [email protected]

University of Natural Resources and Life Sciences, Department of Material Sciences and

University of Natural Resources and Life Sciences, Department of Civil Engineering and

10

Abstract

11

Charcoals were produced from spruce and beech wood under laboratory conditions at different

12

pyrolysis temperatures (300 °C to 1300 °C). Characterization of these charcoals was conducted

13

using eight analytical methods. Each method describes specific changes in the temperature range

14

till 1300 °C. Therefore the combination of these methods provides comprehensive information on

15

different pyrolysis stages. FTIR, NMR spectroscopy, and thermogravimetry display changes till

16

700 °C. A prediction model for pyrolysis temperature till 800 °C is presented based on FTIR

17

spectra with an R² of 0.98. He-pycnometry resolves the temperature range between 500 °C and

18

890 °C. Small angle X-ray scattering (SAXS) describes precisely the evolution of the porous

19

structure and completes the set of techniques by a description of the physical properties of the

20

charcoal. XRD reveals the crystallographic change of the lignocellulosic structure towards

21

precursors of graphite. The formation of calcite out of CaO and CO2 becomes evident.

1 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

22

Keywords: pyrolysis temperature; spectroscopy; X-ray diffraction; He-pycnometry;

23

thermogravimetry

24

1 Introduction

25

Charcoal production can be assumed to accompany mankind since the Iron Age. It has been

26

increased massively at the beginning of industrial revolution in the early 19th century 1. Since at

27

least the time, when biochar has become one of the main topics in the discussion about carbon

28

sequestration in soils the characterization of its properties and its stability became relevant 2,3.

29

Besides chemical characterization especially the physical properties evoke special interest in the

30

context of soil amendment. Various analytical tools have already been used (elemental analysis,

31

spectroscopic methods, thermal analysis); some of them are still rather rarely in use even if their

32

potential is proven (diffractometry, SAXS, pycnometry). Each method elucidates a specific

33

property of the material. In terms of information they provide they complement or confirm one

34

another. Up to now, the description of physical properties lags behind the description of chemical

35

changes 4. Input material, residence time, heating rate, and temperature determine char properties

36

5,6.

37

temperatures are also favorable in modern slow pyrolysis systems, whereas in fast pyrolysis

38

techniques even higher temperatures up to almost 800 °C can be used 8.

39

The purpose of this paper is the comparison of different analytical methods that describe

40

properties of charcoals produced at different pyrolysis. Besides analytical tools that focus on

41

chemical properties (elemental analyses, spectroscopic methods, thermal analysis), He-

42

pycnometry and Small Angle X-ray Scattering (SAXS) were applied to provide more insight into

43

the physical properties skeletal density and porosity. Especially SAXS is an aspiring technique to

44

gain information on the nanostructure of the sample.

Typical temperature of traditional round kilns reached levels between 400 and 600 °C 7. These

2 ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

45

2 Material and Methods

46

2.1 Material

47

Wood pieces with a diameter between 9 and 15 cm were cut from a single stem of spruce (Picea

48

abies) and one of beech (Fagus sylvatica) as the most abundant softwood and hardwood species

49

of Middle Europe. Pure graphite (purity, 99.35%) served as a reference.

50

2.2 Pyrolysis process and sample preparation

51

The stems were cut into pieces with a length between 3 and 5 cm. One piece per species was used

52

per temperature step. They were pyrolized separately in a box made of stainless steel (350 x 400

53

x 150 mm - width x depth x height = 21 L). The box was purged with nitrogen to set the

54

atmosphere around the stem pieces to oxygen free pyrolysis conditions. Right after purging the

55

box was put into a Nabertherm ® N41/13 muffle furnace to start preheating followed by the

56

pyrolysis process. Pyrolysis temperatures were 300 °C, 400 °C, 600 °C, 700 °C, 800 °C, 1000 °C,

57

1200 °C and 1300 °C (one run per temperature). Pyrolysis durations were 4 h not including the

58

cooling phase. Temperatures were recorded during the whole process (preheating, pyrolysis and

59

cooling phase) both inside the box measuring air temperature and inside the wood with a B&R ®

60

PLC (model X20CP1584 with four X20AT6402 modules). Wooden reference samples were dried

61

at 70 °C.

62

Afterwards, all samples were powdered using a steel disc vibrating mill (Fritsch ® Pulverisette 9)

63

and stored at room temperature. Particle size was smaller 20 µm. This particle size ensures

64

reproducible analyses 9. All analyses were made with these powdered samples.

65

2.3 Elemental analyses

3 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

66

Five milligram of the samples were analysed on a HekaTech® Euro EA with column sizes of 1 m

67

length and 8 mm diameter. Helium was used as carrier gas with 8 ml min-1. For combustion

68

10 ml O2 were used additionally. Temperature of the combustion tube was 1000 °C. CO2 and

69

H2O were detected using a TCD.

70

2.4 Polycyclic Aromatic Hydrocarbons (PAH) analyses

71

The procedure applied was based on EN ISO 17993:2003. For calibration PAH mix Agilent

72

8500-6035 was used. Dry samples were extracted for 3 h and 15 min in and ultrasonic bath with a

73

mixture acetone:acetonitrile:methanol (6:3:1), the ratio was 1:10 (1 g : 10 ml). Certified

74

Reference Material BCR - 683 (Beech wood) was used as recovery standard. The HPLC Agilent

75

® consisted of the following components: Quaternary pump series 1100 G1311A, Degaser series

76

1200 G1379B, Autosampler series 1100 ALS G1313A, Column thermostat series 1100 G1316A,

77

column was a Zorbax C18 PAH, 5µ 250 mm, 3 mm. Flow rate was 0.8 ml min-1 and the injection

78

volume was 5 µl. Detection was done by means of DAD (G7117C), UV-Vis (G1314A), FLD

79

(G1321A). Sample clean-up was done by centrifugation (Mikro 20R Hettich ®, Rotor for

80

Eppendorf-cups 1195A, 3000 rpm, 10 min) and full filtration (Whatman ® syringe filter 4 mm,

81

0,45µm membrane size PTFE Membrane).

82

2.5 Fourier Transform Infrared (FTIR) spectroscopy

83

FTIR spectra were recorded in the ATR (attenuated total reflection) mode in the mid infrared

84

region (4000 – 400 cm-1) with an optical crystal of a Bruker ® Helios FTIR micro sampler

85

(Tensor 27). A total of 32 scans were collected at a spectral resolution of 4 cm-1. Five replicates

86

per sample were vector normalized (normalized by the Euclidean norm) and averaged using the

87

OPUS © (version 7.2) software.

88

2.6 Thermogravimetry (TG) 4 ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

89

Thermograms were recorded under oxidative conditions on a Netzsch ® Instrument (STA 409

90

CD Skimmer). Ten mg of the sample were combusted in an Al2O3 pan with a heating rate of

91

10 K min-1 in a temperature range from 30 °C to 950 °C, with a gas flow of 50 ml min-1

92

(80% Ar/ 20 % O2). Argon was used as flushing gas (20 ml min-1). A recalcitrance index (R50,x)

93

has been calculated as a ratio of the temperatures of 50 % mass loss in the thermogram of a

94

certain sample and of graphite 10.

95

2.7 Nuclear magnetic resonance (NMR) spectroscopy

96

Solid-State 13C NMR spectra of the samples were obtained using a Varian INOVA (11.74 T)

97

spectrometer with 13C and 1H frequencies of 125.7 and 500.0 MHz, respectively. A variable-

98

amplitude cross-polarization pulse sequence was employed. The experiments were performed

99

using magic-angle spinning (MAS) of 15 kHz, a cross-polarization time of 1 ms, an acquisition

100

time of 15 ms and a recycle delay of 500 ms. The cross-polarization time was chosen after

101

variable contact time experiments, and the recycle delays in CP (cross-polarization) experiments

102

were chosen to be five times longer than the longest 1H spin-lattice relaxation time (T1H) as

103

determined by inversion-recovery experiments. High-power TPPM (two pulse phase-modulation)

104

proton decoupling of 70 kHz was employed 11.

105

2.8 He-pycnometry

106

The measurements were performed on a micromeritics ® AccuPyc II 1340 gas pycnometer.

107

Chamber size was 100 cm³ and He pressure was 1.5 bar. 20 cycles of pressure and evacuation

108

were performed for equilibration. An additional further run of 10 measurements was averaged for

109

the final value.

110

2.9 X-ray diffractometry (XRD)

5 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

111

The samples were investigated by means of X-ray diffraction (XRD) using a Panalytical ®X´Pert

112

Pro MPD diffractometer with automatic divergent slit, Cu LFF tube (45 kV, 40 mA), with an

113

X´Celerator detector. The measuring time was 25 s, with a stepsize of 0.017°. Diffractograms

114

were recorded from 5° to 70° (2Ө). Semiquantitative mineral composition of the bulk samples

115

was estimated using Rietveld refinement with the Panalytical software © X´Pert HighScore Plus.

116

2.10 Small angle X-ray scattering (SAXS)

117

Measurements were carried out with a 3-pinhole SMAX-3000 SAXS camera (RIGAKU ®)

118

equipped with a MM002+ X-ray source (with copper target wavelength of λ = 0.1541 nm). Two-

119

dimensional scattering images were recorded with a TRITON ® 200 multi-wire X-ray detector

120

(200 mm diameter mapped on 1024 x 1024 pixels). The samples were put between two layers of

121

scotch tape. The scattering images were integrated azimuthally to gain information of the

122

scattered intensity I(q) in dependence of the scattering vector q, which is related to the scattering

123

angle 2θ and the wavelength λ by q=4π/ λ sin θ. The full q range was 0.1 nm-1 to 8 nm-1.The

124

obtained data were background corrected and further analyzed with respect to the pore structure,

125

following Köhnke et al. 12. Pore radii for spherical pores 13, a value related to the surface

126

roughness of structures in the nanometre regime and a value proportional to the specific inner

127

surface have been determined

128

2.11 Statistics

129

Pearson correlation coefficients were calculated by the program IBM SPSS 21®. The PLS-

130

Regression model and the PCA were calculated by the program Unscrambler X 10.1 ®. Spectral

131

regions included in both models ranged from 3728 to 2422 cm-1 and from 1896 to 400 cm-1.

132

NIPALS algorithm was applied, the model was 10-fold cross-validated. Positions of the PLS

6 ACS Paragon Plus Environment

Page 6 of 25

Page 7 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

133

validation results for all samples are given in the figure. Four factors were included in both

134

models.

135

3 Results and Discussion

136

3.1 Elemental analyses

137

Main carbon increase and corresponding oxygen and hydrogen decrease occurred between 70 °C

138

and 600 °C. At 700 °C carbon content reached values above 90% (figure 1a), hydrogen decreases

139

correspondingly (figure 1b). After that level the values did not change strongly anymore and a

140

differentiation based on elemental analyses became quite weak. Van Krevelen diagram revealed a

141

very high degree of carbonization already at about 600 °C positioning the samples in the left

142

down corner (figure 1c).

143

The highest correlation between pyrolysis temperature and carbon or hydrogen content was found

144

in the temperature range from 70 °C to 600°C resulting in a coefficient of determination R² of

145

96.3 % (carbon) and 96.4 % (hydrogen). Including higher temperature steps leads to a decrease of

146

the coefficient of determination. A better separation at least till pyrolysis temperatures of 800 °C

147

is achieved, when taking the logarithm of values for van Krevelen diagram (figure 1d). For the

148

characterization of materials charred at higher temperatures as traditional kiln samples we

149

propose to modify the diagram in that way.

150

7 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

151

(a)

(b)

(c)

(d)

152 153

Figure 1:(a) Carbon and (b) hydrogen content in relation to pyrolysis temperature; (c) van

154

Krevelen diagram; PA-Picea abies, FS-Fagus sylvatica; (d) van Krevelen diagram with log10-

155

values

156 157

3.2 Polycyclic Aromatic Hydrocarbons (PAH) analyses

158

Results of PAH revealed high concentrations, but comparable with literature 14,15. The European

159

Biochar Standard defines limit values of 12 mg/kg for a basic grade and 4 mg/kg for the premium

160

grade 16. Limit value for the International Biochar Initiative is set to 6 mg/kg 17. Spruce samples

161

contained partly far higher concentrations compared with beech, even at high temperatures. The 8 ACS Paragon Plus Environment

Page 8 of 25

Page 9 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

162

highest levels were found at pyrolysis temperatures of about 800 °C (figure 2). The ratios

163

Phenanthrene/ Anthracene and Naphthalene/ Phenanthrene are used to differentiate biochar types

164

18,19.

165

processed, unaged char. By such ratios, stability and degradation process of PAH can be

166

examined 20. Phe/Ant ratios above 10 indicate rather unstable fractions typical for combustion

167

emissions.

The ratios found in our samples are comparable to typical values from traditionally

168

(a)

(b)

169 170

Figure 2:(a) Sum of 16 EPA PAHs with limit values of EBC and IBI and (b) the two ratios of

171

Phenantrene/ Anthracene and Naphthalene/ Phenanthrene of different pyrolysis temperatures;

172

PA-Picea abies, FS-Fagus sylvatica

173 174

3.3 Fourier Transform Infrared (FTIR) spectroscopy

175

The spectral characteristics changed considerably especially in the temperature range between

176

70 °C and 800 °C. Main changes were found in the broad band of OH-stretching vibrations

177

(region between 3600 cm-1 and 3000 cm-1) and the band region of CH stretching vibrations

178

derived from methyl-, methylene, and methine groups (between 3000 cm-1 and 2800 cm-1). The 9 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

179

specific bands 21 for hemicelluloses (around 1730 cm-1), lignin (e.g. around 1510 cm-1) and

180

cellulose (peak maximum at around 1026 cm-1) display the chemical changes of pyrolysis as well.

181

Figure 3a displays spectra of Fagus samples at different pyrolysis temperature compared with

182

pure graphite. Picea revealed similar characteristics (not shown). From 70 °C to 300°C cellulosic

183

compounds decrease and lignin is relatively enriched; as the bands stay more or less the same we

184

conclude that the compounds do not change substantially, however. This effect of cellulosic

185

decomposition can be monitored also in TG, NMR, XRD, and SAXS. Main changes take place

186

between 300 °C and 400 °C. It demonstrates a structural breakdown between these two

187

temperature steps. At 400 °C and 600 °C typical bands of aromatic CH bands (1596 cm-1,

188

877 cm-1, 820 cm-1, 759 cm-1) occur. Above that temperature all band signals decrease and spare

189

just a smooth spectrum due to carbon concentration. Results correspond well to literature 22.

190

Based on the spectral pattern PLS1-regression models were established to predict pyrolysis

191

temperature. The spectral pattern allows conclusions to be drawn about the production

192

temperature of charcoals up to 800 °C. Within the temperature range from 70 °C to 800 °C the

193

highest R² with 98 % was reached (figures 3b and 3c). Higher temperatures led to similar spectra

194

and their inclusion into the model did not lead to any further increase of the quality criterion.

195

According to the loadings plot (not shown) the first component is dominated by the loss of wood

196

characteristics especially the decrease of the cellulose peak around 1026 cm-1 till 600 °C, whereas

197

the second component reveals aromatic vibrations of an intermediate stage with a maximum at

198

about 400 °C indicated by a peak around 1596 cm-1. The increase of pyrolysis temperature is

199

paralleled by disappearance of distinct bands in the infrared spectrum. The loss of functional

200

groups in the spectral pattern at 400 °C and beyond is caused by the strong relative increase of

201

carbon that is equally reflected by elemental analyses 14,23.

202

10 ACS Paragon Plus Environment

Page 10 of 25

Page 11 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

(a)

203 (b)

(c)

204 205

Figure 3:(a) Spectral characteristics of Fagus samples heated up to temperatures from 70 °C till

206

800 °C. (b) Scores Plot from a PCA with FTIR spectra of Fagus and Picea samples, samples

207

pyrolyzed at 1000 °C, 1200 °C, and 1300 °C are marked by “x” symbols; (c) PLS1-regression

208

line including confidence bands (based on a significance level of α =0.05, PLS1 model with 4

209

factors)

210 211

The progress of carbonization leads to higher hydrophobicity of charcoals. Oxidation due to

212

aging reverses the process and results again in hydrophilic properties. They are favorable for soil

213

amendment with regard to water and nutrient storage 24. In FTIR spectra these oxidation 11 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

214

processes due to aging can be separated, however, from lower stages of pyrolysis with spectral

215

characteristics described in this section 25.

216

3.4 Thermogravimetry (TG)

217

As long as functional groups are present in spectra of charcoals produced at pyrolysis

218

temperatures up to 400 °C a slight mass loss in the thermograms between 80 °C and 130 °C is

219

observed. It is caused by water evaporation. With increasing pyrolysis temperature and the

220

resulting hydrophobicity this effect disappears. Thermal data of oxidative combustion stress

221

again the strong carbonization of our materials. The thermal behavior confirms the results of

222

FTIR spectroscopy. Shape and temperature range of the thermograms up to pyrolysis

223

temperatures of 400 °C indicate the variety of chemical components. Thermograms of charcoals

224

produced at 600 °C and above shift to higher temperatures and feature a more uniform thermal

225

behavior according to their chemical composition. Figure 4a displays this pattern for Fagus

226

samples. Similar results are obtained for Picea (not shown).

227 (a)

228

12 ACS Paragon Plus Environment

Page 12 of 25

Page 13 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

(b)

229 230

Figure 4: (a) Thermograms of beech (Fagus sylvatica) samples pyrolyzed at different

231

temperatures (70 °C – 1300 °C) during oxidative combustion, (b) the recalcitrance index R50,x

232 233

Harvey et al. 10 proposed a recalcitrance index R50,x. They distinguished three groups of

234

recalcitrance with the limit values of 0.5 and 0.7. Below 0.5 materials are comparably recalcitrant

235

as uncharred plant biomass and above 0.7 recalcitrance draws near to soot. According to this

236

classification our materials would reach the highest class of carbon sequestration at a pyrolysis

237

temperature around 800 °C. In terms of carbon sequestration these pyrolysis products indicate a

238

high stability.

239

3.5 NMR spectroscopy

240

The low pyrolysis temperatures up to 300 °C demonstrate decreasing hemicelluloses and relative

241

increases of cellulose and lignin (figure 5a) 26. With higher temperatures progressive

242

decomposition of aliphatic remains (alkyl groups) takes place. At 400 °C some O-aryl groups

243

remain, but they decrease with increasing pyrolysis temperature (figure 5b). Similar to FTIR

244

spectra NMR spectra reveal the chemical change with increasing pyrolysis temperatures and the

245

loss of distinct peaks assigned to molecules and molecule groups, respectively. Compared to 13 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

246

FTIR spectra NMR spectra provide clear signals of the aryl-group in charcoals produced at

247

pyrolysis temperatures above 400 °C. The progressive upfield shift of aryl signal is due to the

248

aromatic ring polycondensation 11,27–29. The results are in accordance with elemental analyses.

249

Due to the strong carbonization the highest temperature samples do not produce useful spectra.

250

As discussed in Novotny et al. 30 this is a result of the high heterogeneity in local magnetic

251

susceptibility and/ or chemical shift anisotropy, which is not completely averaged out at the usual

252

rates of magic angle sample spinning.

253

(a)

(b)

254 255

Figure 5:(a) VACP 13C NMR spectra of heated biomass (until 300 °C). FS and PA are Fagus

256

sylvatica and Picea abies respectively with the heating temperature. C: cellulose; H: 14 ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

257

hemicellulose; L: lignin (including the whole grey box); A: C6 and C4 from amorphous

258

cellulose; Cr: C6 and C4 from crystalline cellulose. The dashed ellipses indicate the useful region

259

for cellulose crystallinity evaluation. (b) Fagus sylvatica spectra for pyrolysis temperatures above

260

300 °C, the dot line indicates the upfield shift of aryl C signal due to the increase of aromatic ring

261

poly-condensation.

262 263

3.6 X-ray diffractometry (XRD)

264

XRD displays well the transformation of lignocellulosic wood structure into charred organic

265

matter between 300 and 400 °C (figure 6). The shoulder at around 16° and the peaks at 21.9° and

266

34.7° can be assigned to cellulosic crystallinity 31. At pyrolysis temperatures of 400 °C the

267

cellulose peaks disappeared corresponding to the loss of the band in FTIR (1026 cm-1 in figure

268

3a) and NMR (indicated with “C” in figure 5a). With increasing pyrolysis temperature the broad

269

two peaks of 24° and 43° dominate the diffractograms. Their shift towards the 002 and 011 peaks

270

of graphite 32,33, the increase of intensity and sharpness of the peaks indicate the transition

271

towards graphite, although it cannot be called “graphite” yet, Finally several sharp peaks can be

272

assigned to calcite: 23.0°, the main peak at 29.4°, furthermore 35.9°, 39.3°, 43.0°, 47.6°, 48.4°,

273

57.6° 34. Best representation of all peaks is found in the Fagus samples pyrolyzed at 1200 and

274

1300 °C. But even at lower temperatures at least the main peak at 29.4° is present. Other peaks

275

can be assigned to CaO. Presence of calcite in biochar is well described in literature 35. Our

276

results, however, give new insights in its formation. The amount of calcite increases with

277

pyrolysis temperature. This can be explained by a relative enrichment. Interestingly, calcite is

278

present in materials that reached temperatures > 650 °C at which this mineral is destroyed

279

thermally 36. Therefore CaO reacts with CO2 trapped in the pores for carbonation 37 after cooling

280

down. CaO might be a relic of Ca-oxalate (in the form of whewellite visible by a small peak at 15 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

281

14.9° till 400 °C), a crystal commonly formed in different parts of plants 38. Calcite in biochar

282

can influence its behavior as a soil amendment especially in acidic soils 39.

283 (a)

(b)

284 285

Figure 6: X-ray diffractograms of (a) Picea and (b) Fagus samples; cellulose (cell.), whewellite

286

(w.) and calcite (C) peaks are indicated with vertical guiding lines. The shift of graphitic peaks is

287

described in the text.

288 289

Main results of this method are: XRD displays the transition of celluloses into precursor phases

290

of graphite above 300 °C. Main strength of the method is the elucidation of relative enrichment

291

of minerals especially calcite, generated by a de-novo synthesis from CaO and CO2 trapped in the

292

porous char structure after cooling down from the pyrolysis process. 16 ACS Paragon Plus Environment

Page 16 of 25

Page 17 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

293

3.7 He-pycnometry

294

Results of pycnometry can be described asymptotically by three lines. Skeletal density

295

maintained a level of about 1.4 g/cm³. Then the values increased quite linearly up to a level of

296

about 2 g/cm³. The asymptotic constant lines and the inclined one cross each other shortly above

297

500 °C at a temperature about 890 °C (figure 7). It can be assumed that polycondesation of

298

aromatic rings goes along with prominent physical changes above 400 °C. The density of

299

charcoal approximates graphite density of 2.2 and 2.3 g/cm³. This parameter provides

300

information in a temperature range that is not resolved well by other methods. It covers perfectly

301

the common temperature range of charcoal kilns. The relation between pyrolysis temperature and

302

skeletal density was also reported by Brewer et al. 40 and Wiedemeier et al. 41. Wiedemeier et al.

303

41

304

linear increase from 500 °C till 890 °C is documented by all authors.

found stable values till 400 °C but a further increase even till 1000 °C. At least the comparable

305

306 307

Figure 7: Density of the solid matter measured by means of He-pycnometry. PA-Picea abies,

308

FS-Fagus sylvatica

309 310

3.8 Small Angle X-ray Scattering (SAXS) 17 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

311

The small-angle x-ray scattering curves of Fagus and Picea both show characteristic features

312

evolving during the temperature treatment. For temperatures up to 300 °C the wood structure is

313

rather intact, featuring a low q signal arising from small voids and cracks in the wood and a

314

shoulder around 1.5 nm-1 attributed to the wood structure, which is destroyed at higher pyrolysis

315

temperatures (figure 8a, 8b and 8d). Degradation of the cellulose around 250°C is reported in

316

Popovski et al. 42. For higher temperatures a clear development of a shoulder attributed to pores

317

in the range of nanometers is visible around 2 nm-1. Starting with a quite wide distribution, a

318

broad shoulder shifting to the left indicates increasing pore size. The pores for both wood types

319

start around 0.3 nm and show a strong increase in pore size at 600 °C and slow further increase in

320

pore size with higher temperatures (figure 8d).

321

Both wood types show changing nanostructure with increasing treatment temperature. The

322

change of the power law exponent in the low q regime shows a change in surface roughness of

323

the higher order structure, i.e. cracks and voids (n = 4 smooth surface, 3 < n < 4 surface fractal

324

and 1 < n < 3 volume fractal). While the wood treated with 70 °C features slightly rough surfaces

325

(n = 3.8 and n = 3.9 for Picea and Fagus, respectively). The surfaces are smooth after the 300 °C

326

temperature treatment. For temperatures higher than 400 °C the absolute value of the power law

327

exponent decreases again i.e. from n = 4 to n = 2.7 for Fagus and n = 2.3 for Picea, respectively,

328

indicating the generation of a fractal pore structure (figure 8c).

329

18 ACS Paragon Plus Environment

Page 18 of 25

Page 19 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

(a)

(b)

330

19 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(d)

(c)

331

(e)

332 333

Figure 8:(a) scattering curves for Fagus sylvatica (FS) samples and (b) for Picea abies (PA)

334

samples, intensities I (q) shifted on the vertical axis, arrows in (a) and (b) indicating the peak shift

335

were included to guide the eye; (c) roughness of the pore structure; (d) pore size given as pore

336

radius; (e) specific inner surface

337 338

The change of the nanostructure is also clearly seen in the specific inner surface. From the SAXS curves a

339

value P/Q can be determined, which is proportional to the specific inner surface 12. The value is increasing

340

over the full temperature rang (Figure 8e) for both wood types.

341

The results can be compiled in the following way: SAXS indicates a structural development due

342

to heat treatment by the pyrolysis process: With increasing temperature first the wood structure 20 ACS Paragon Plus Environment

Page 20 of 25

Page 21 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

343

decomposed at temperatures above 250 °C. Originally slightly rough surfaces attributed to cracks

344

and voids get smoother. At temperatures higher than 300 °C the surfaces increases in roughness

345

again, featuring also the development of nanopores in the size range of around 0.5 nm, which are

346

increasing in size towards 1 nm for temperatures higher than 400 °C and which onwards only

347

slightly increase. This process is accompanied by an increase of the specific inner surface.

348

4 Conclusions

349

Three ranges of pyrolysis temperatures can be roughly distinguished. Chemical changes mainly

350

take place up to temperatures between 400 °C and 600 °C. They are well reflected by elemental

351

analyses, FTIR and NMR spectroscopy, and thermal analyses. With increasing pyrolysis

352

temperatures physical and structural processes dominate till a temperature between 800 °C and

353

1000 °C, paralleled by slight further chemical rearrangement. Physical characteristics are

354

described by means of pycnometry and SAXS. With a further increase of pyrolysis temperature

355

chemical and physical properties appear constant according to the results obtained except the

356

specific inner surface. Altogether a concise picture of changing charcoal properties depending on

357

pyrolysis temperature can be drawn. This overview is an important pre-requisite for the

358

investigation of biochar functions in soils. Porosity and inner surface are relevant properties in

359

soils, as well as the chemical reactivity that corresponds to degree of carbonation.

360

5 Acknowledgments

361

We thank Reinhold Ottner for thermal analysis of graphite, Axel Mentler for PAH analyses and

362

Marion Sumetzberger-Hasinger for elemental analyses.

363

6 Competing interests, Funding, Data availability

21 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

364

The authors declare that there are no competing interests. This research did not receive any

365

specific grant from funding agencies in the public, commercial, or not-for-profit sectors. Data are

366

available upon request.

367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401

7 References (1) Straka, T. J. Historic Charcoal Production in the US and Forest Depletion: Development of Production Parameters. AHS 2014, 03, 104–114. (2) Smith, P. Soil carbon sequestration and biochar as negative emission technologies. GCB 2016, 22, 1315–1324. (3) Sohi, S.; Lopez-Capel, E.; Krull, E.; Bol, R. Biochar, climate change and soil: A review to guide future research. CSIRO Land Water Sci. Report 2009, 5, 17–31. (4) Aller, M. F. Biochar properties: Transport, fate, and impact. Crit. Rev. Environ. Sci. Technol. 2016, 46, 1183–1296. (5) Keiluweit, M.; Kleber, M.; Sparrow, M. A.; Simoneit, B. R. T.; Prahl, F. G. Solvent-extractable polycyclic aromatic hydrocarbons in biochar: Influence of pyrolysis temperature and feedstock. Environ. Sci. Technol. 2012, 46, 9333–9341. (6) Singh, B. P.; Cowie, A. L.; Smernik, R. J. Biochar carbon stability in a clayey soil as a function of feedstock and pyrolysis temperature. Environ. Sci. Technol. 2012, 46, 11770–11778. (7) Powell, A. J.; Wheeler, J.; Batt, C. M. Identifying archaeological wood stack charcoal production sites using geophysical prospection: Magnetic characteristics from a modern wood stack charcoal burn site. J. Archaeol. Sci. 2012, 39, 1197–1204. (8) Garcia-Nunez, J. A.; Pelaez-Samaniego, M. R.; Garcia-Perez, M. E.; Fonts, I.; Abrego, J.; Westerhof, R. J. M.; Garcia-Perez, M. Historical Developments of Pyrolysis Reactors: A Review. Energy Fuels 2017, 31, 5751–5775. (9) Smidt, E.; Meissl, K.; Tintner, J. The influence of waste sample preparation on reproducibility of thermal data. Thermochim. Acta 2008, 468, 55–60. (10) Harvey, O. R.; Kuo, L.-J.; Zimmerman, A. R.; Louchouarn, P.; Amonette, J. E.; Herbert, B. E. An indexbased approach to assessing recalcitrance and soil carbon sequestration potential of engineered black carbons (biochars). Environ. Sci. Technol. 2012, 46, 1415–1421. (11) Novotny, E. H.; deAzevedo, E. R.; Bonagamba, T. J.; Cunha, T. J. F.; Madari, B. E.; Benites, V. d. M.; Hayes. Studies of the Compositions of Humic Acids from Amazonian Dark Earth Soils. Environ. Sci. Technol. 2007, 41, 400–405. (12) Small Angle X-ray Scattering; Glatter, O.; Kratky, O., Eds.; Academic Press: London, 1982. (13) Köhnke, J.; Fürst, C.; Unterweger, C.; Rennhofer, H.; Lichtenegger, H. C.; Keckes, J.; Emsenhuber, G.; Mahendran, A. R.; Liebner, F.; Gindl-Altmutter, W. Carbon microparticles from organosolv lignin as filler for conducting Poly(lactic acid). Polym. 2016, 8, DOI: 10.3390/polym8060205. (14) Kloss, S.; Zehetner, F.; Dellantonio, A.; Hamid, R.; Ottner, F.; Liedtke, V.; Schwanninger, M.; Gerzabek, M. H.; Soja, G. Characterization of slow pyrolysis biochars: Eff ects of feedstocks and pyrolysis temperature on biochar properties. J. Environ. Qual. 2012, 41, 990–1000. 22 ACS Paragon Plus Environment

Page 22 of 25

Page 23 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442

Industrial & Engineering Chemistry Research

(15) Bucheli, T. D.; Hilber, I.; Schmidt, H.-P. Polycyclic aromatic hydrocarbons and polychlorinated aromatic compounds in biochar. In Biochar for Environmental Management: Science, Technology and Implementation; Joseph, S., Lehmann, J., Eds.; pp 595–624. (16) Schmidt, H.-P. European Biochar Certificate (EBC) - guidelines version 6.1, 2015. (17) International Biochar Initiative. Standardized Product Definition and Product Testing Guidelines for Biochar That Is Used in Soil: Version 2.1, 2015. (18) Schimmelpfennig, S.; Glaser, B. One step forward toward characterization: Some important material properties to distinguish biochars. J. Environ. Qual. 2012, 41, 1001–1013. (19) La Rosa, J. M. de; Paneque, M.; Hilber, I.; Blum, F.; Knicker, H. E.; Bucheli, T. D. Assessment of polycyclic aromatic hydrocarbons in biochar and biochar-amended agricultural soil from Southern Spain. J Soils Sediments 2016, 16, 557–565. (20) Yunker, M. B.; Macdonald, R. W.; Vingarzan, R.; Mitchell, R. H.; Goyette, D.; Sylvestre, S. PAHs in the Fraser River basin: A critical appraisal of PAH ratios as indicators of PAH source and composition. Org. Geochem. 2002, 33, 489–515. (21) Schwanninger, M.; Rodrigues, J. C.; Pereira, H.; Hinterstoisser, B. Effects of short-time vibratory ball milling on the shape of FT-IR spectra of wood and cellulose. Vib. Spectrosc. 2004, 36, 23–40. (22) Kwon, S.-M.; Jang, J.-H.; Lee, S.-H.; Park, S.-B.; Kim, N.-H. Change of Heating Value, pH and FT-IR Spectra of Charcoal at Different Carbonization Temperatures. J. Korean Wood Sci. Technol. 2013, 41, 440–446. (23) Hao, F.; Zhao, X.; Ouyang, W.; Lin, C.; Chen, S.; Shan, Y.; Lai, X. Molecular structure of corncobderived Biochars and the mechanism of Atrazine sorption. Agron. J. 2013, 105, 773–782. (24) Wiedner, K.; Fischer, D.; Walther, S.; Criscuoli, I.; Favilli, F.; Nelle, O.; Glaser, B. Acceleration of Biochar Surface Oxidation during Composting? J. Agric. Food Chem. 2015, 63, 3830–3837. (25) Smidt, E.; Tintner, J.; Klemm, S.; Scholz, U. FT-IR spectral and thermal characterization of ancient charcoals - A tool to support archeological and historical data interpretation. Quat. Int. 2017, 457, 43–49. (26) Bernandelli, O. D.; Novotny, E. H.; Azevêdo, E. R. d.; Colnago, L. A. Analyses of Biomass Products by Nuclear Magnetic Resonance Spectroscopy. In Analytical Techniques and Methods for Biomass; Vaz, S., Ed.; Springer International Publishing: Cham, Switzerland, 2016; pp 143–172. (27) Freitas, J. C. C.; Bonagamba, T. J.; Emmerich, F. G. Investigation of biomass- and polymer-based carbon materials using 13C high-resolution solid-state NMR. Carbon 2001, 39, 535–545. (28) Knicker, H.; Totsche, K. U.; Almendros, G.; González-Vila, F. J. Condensation degree of burnt peat and plant residues and the reliability of solid-state VACP MAS 13C NMR spectra obtained from pyrogenic humic material. Org. Geochem. 2005, 36, 1359–1377. (29) Smernik, R. J.; Kookana, R. S.; Skjemstad, J. O. NMR characterization of 13C-benzene sorbed to natural and prepared charcoals. Environ. Sci. Technol. 2006, 40, 1764–1769. (30) Novotny, E. H.; Maia, C. M. B. d. F.; Carvalho, M. T. d. M.; Madari, B. E. BIOCHAR: PYROGENIC CARBON FOR AGRICULTURAL USE - A CRITICAL REVIEW. Rev. Bras. Ciênc. Solo 2015, 39, 321–344. (31) Chen, X.; Yu, J.; Zhang, Z.; Lu, C. Study on structure and thermal stability properties of cellulose fibers from rice straw. Carbohydr. Polym. 2011, 85, 245–250. (32) Katsumi, N.; Yonebayashi, K.; Okazaki, M. Evaluation of stacking nanostructure in soil humic acids by analysis of the 002 band of their X-ray diffraction profiles. Soil Sci. Plant Nutr. 2015, 61, 603–612.

23 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468

(33) Hansora, D. P.; Shimpi, N. G.; Mishra, S. Graphite to Graphene via Graphene Oxide: An Overview on Synthesis, Properties, and Applications. JOM 2015, 67, 2855–2868. (34) Xu, B.; Poduska, K. M. Linking crystal structure with temperature-sensitive vibrational modes in calcium carbonate minerals. Phys. Chem. Chem. Phys. 2014, 16, 17634–17639. (35) Yuan, J.-H.; Xu, R.-K.; Zhang, H. The forms of alkalis in the biochar produced from crop residues at different temperatures. Bioresour. Technol. 2011, 102, 3488–3497. (36) Siewert, C. Rapid screening of soil properties using thermogravimetry. Soil Sci. Soc. Am. J. 2004, 68, 1656–1661. (37) Smidt, E.; Meissl, K.; Tintner, J.; Ottner, F. Interferences of carbonate quantification in municipal solid waste incinerator bottom ash: Evaluation of different methods. Environ. Chem. Lett. 2010, 8, 217– 222. (38) Franceschi, V. R.; Nakata, P. A. Calcium oxalate in plants: Formation and function. Ann. Rev. Plant Biol. 2005, 56, 41–71. (39) Rechberger, M. V.; Kloss, S.; Rennhofer, H.; Tintner, J.; Watzinger, A.; Soja, G.; Lichtenegger, H.; Zehetner, F. Changes in biochar physical and chemical properties: Accelerated biochar aging in an acidic soil. Carbon 2017, 115, 209–219. (40) Brewer, C. E.; Chuang, V. J.; Masiello, C. A.; Gonnermann, H.; Gao, X.; Dugan, B.; Driver, L. E.; Panzacchi, P.; Zygourakis, K.; Davies, C. A. New approaches to measuring biochar density and porosity. Biom. Bioenergy 2014, 66, 176–185. (41) Wiedemeier, D. B.; Abiven, S.; Hockaday, W. C.; Keiluweit, M.; Kleber, M.; Masiello, C. A.; McBeath, A. V.; Nico, P. S.; Pyle, L. A.; Schneider, M. P. W.. Aromaticity and degree of aromatic condensation of char. Org. Geochem. 2015, 78, 135–143. (42) Fritz-Popovski, G.; van Opdenbosch, D.; Zollfrank, C.; Aichmayer, B.; Paris, O. Development of the fibrillar and microfibrillar structure during biomimetic mineralization of wood. Adv. Funct. Mater. 2013, 23, 1265–1272.

24 ACS Paragon Plus Environment

Page 24 of 25

Page 25 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Industrial & Engineering Chemistry Research

PAH C, H, O NMR

STA

FTIR

porosity pycnometry

ACS Paragon Plus Environment