Techno-economical method for the removal of alkali metals from

Publication Date (Web): August 24, 2018 ... Accordingly, we herein propose an economical method to remove the inherent ashes in the biomass using 16.6...
0 downloads 0 Views 6MB Size
Subscriber access provided by University of South Dakota

Article

Techno-economical method for the removal of alkali metals from agricultural residue and herbaceous biomass and its effect on slagging and fouling behavior Young-Joo Lee, Jong-Won Choi, Ju-Hyoung Park, Hueon Namkung, Gyu-Seob Song, SeJoon Park, Dong-Wook Lee, Joeng-Geun Kim, Chung-Hwan Jeon, and Young-Chan Choi ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b02588 • Publication Date (Web): 24 Aug 2018 Downloaded from http://pubs.acs.org on August 28, 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 36 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

ACS Sustainable Chemistry & Engineering

1

Techno-economical method for the removal of alkali metals from

2

agricultural residue and herbaceous biomass and its effect on

3

slagging and fouling behavior

4

5

Young-Joo Lee 1†, ǁ, Jong-Won Choi 1†, Ju-Hyoung Park †, Hueon Namkung †,

6

Gyu-Seob Song †, Se-Joon Park †, Dong-Wook Lee †, Joeng-Geun Kim †,

7

Chung-Hwan Jeon *ǁ, and Young-Chan Choi *†

8



Korea Institute of Energy Research (KIER), 152, Gajeong-ro, Yuseong-gu, Daejeon 34129, Republic

9

of Korea ǁ

10 11

Pusan Clean Coal Center, School of Mechanical Engineering, Pusan National University, 2, Busandaehak-ro 63beon-gil, Geumjeong-gu, Busan 46241, Republic of Korea

12 13 14 15 16 17 18 19

Corresponding Author

20

* E-mail: [email protected]. Tel. : +82-51-510-3051 (C.-H.J.).

21

* E-mail: [email protected]. Tel. : +82-42-860-3784 (Y.-C.C.).

22 23

1

These authors equally contributed to this work.

24

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

25

ABSTRACT

26

Nowadays, it is widely recognized that biomass combustion processes can contribute to the

27

mitigation of greenhouse gas emissions, and thus becomes a viable option as an alternative

28

energy source for the power industry. Among various biomasses, the herbaceous biomass is

29

regarded as abundant and relatively inexpensive fuel. However, it contains high ashes

30

(especially high levels of alkali metals), causing to operation troubles such as slagging and

31

fouling inside a heat exchanger or the efficiency deterioration. Accordingly, we herein

32

propose an economical method to remove the inherent ashes in the biomass using 16.6 M

33

acetic acid pre-treatment at 60oC for 10 min. Seven different biomasses were investigated to

34

validate the effects of method. The Kenaf shows the total mineral rejection of 93.48%. In

35

particular, the potassium and sodium elements in the Kenaf, which are major factors

36

influencing on fouling and slagging in a boiler, were removed up to 99.46 and 100%,

37

respectively. Furthermore, the proposed wet treatment was more effective for biomass with

38

higher surface areas.

39 40 41 42 43 44 45 46

Keywords: Ashless biomass, ash extraction, wet treatment, slagging, fouling

47 48

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36 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

ACS Sustainable Chemistry & Engineering

49

■ INTRODUCTION

50

Recently, global warming issues associated with the indiscriminate use of fossil fuels have

51

motivated many researchers to focus on the reduction of carbon dioxide (CO2) emissions

52

from the combustion process. The use of biomass for heat and electricity generation via

53

thermo-chemical processes such as combustion, pyrolysis, and gasification, as well as for the

54

production of various biofuels, is one means of reducing CO2 emissions.1-4 However, several

55

bottlenecks exist that affect the commercialization of biomass. Among these, the high content

56

of alkali and alkaline earth metals (AAEM) in the biomass lowers its melting temperature,

57

resulting in ash related problems such as slagging and fouling during the combustion, which

58

interrupts stable power plant operation.5-9 Therefore, to inhibit the severe ash deposition

59

phenomenon during biomass combustion, significant research effort has been expended to

60

investigate and develop operational technologies such as the blending of coal and biomass, 10

61

controlling the operation conditions,11 and the addition of additives12-13.

62

Priyanto et al.10 have investigated co-firing 4 types of woody biomasses with coal. They

63

concluded that the use of 30% of a biomass containing the ash higher than 1.0wt% for co-

64

firing produces a large number of molten ash particles with a higher calcium mineral content,

65

which is still impeding the commercial use of biomass for the power generation.

66

Szemmelveisz et al.11 reported that the softening properties of the alkaline minerals such as

67

K, Na and Cl inside three herbaceous biomasses may critically influence on operational

68

problems (slagging and fouling) during a co-combustion with coal. They suggested that co-

69

firing of coal with herbaceous biomass requires lowering the operation temperature of the

70

boiler to avoid the deposits and slagging. However, the lower thermal efficiency of boiler is

71

accompanied by lowering the combustion temperature.

72

Kassman et al.12 reported combustion of a biomass-only resulted in enhanced content of

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

73

chlorine deposit outside a super-heater surface, and thus suggested the injection of

74

ammonium sulfate to significantly lower the level of gas phase KCl. Accordingly, they found

75

almost no chlorine in the deposits. However, the effect of ammonium sulfate on other

76

environmental problems such as gaseous pollutant and PM emission has not been clear yet.

77

Wang et al.13 reported that severe melting of the ash was due to formation and fusion of

78

low temperature melting potassium silicates, thereby investigating the effects of two mineral

79

additives: zeolite 24A and kaolin to capture KCl. They concluded that addition of kaolin and

80

zeolite 24A significantly reduced the sintering tendency of the ash inside barely straw up to

81

50%.

82

However, the problem of fouling and slagging still remains a key challenge. Furthermore,

83

such modifications or optimizations typically require extensive changes of the external

84

environment, which can be very costly to implement. Recent research efforts have been

85

shifted towards biomass pre-treatment prior to the combustion. For example, the removal of

86

the alkali metals (K and Na), and Cl with pure water,14-16 acidic solutions,17-19 and alkaline

87

solutions20 have been investigated. The hydrothermal pretreatment (HTP) process with pure

88

water have been revealed to consume long residence time, high cost and high temperature

89

even with low yield.14-16 The previous researchers have employed various strong acids such

90

as HCl, H2SO4 and HNO3 or strong base (NaOH), all of which can easily break the cellulose

91

and hemicellulose inside a biomass into glucose and xylose, respectively.17-20 Unlike the

92

cellulose and hemicellulose, glucose and xylose are soluble, leading to low solid yield during

93

the treatment process. Thereby, those reasons derived us to select the use of acetic acid as a

94

weak acid with a low solid loss.

95

It is well known that the cellulose, hemicellulose, and lignin constituents of biomass are

96

responsible for the combustion sources. Hence, a minimization of the loss of these three

97

constituents from biomass during the pre-treatment and optimization processes are necessary.

ACS Paragon Plus Environment

Page 4 of 36

Page 5 of 36 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

ACS Sustainable Chemistry & Engineering

98

We herein propose a commercially accessible and acceptable pre-treatment method for the

99

production of ashless biomass, which entails an acidic treatment of the biomass at a

100

reasonable temperature (60°C) for a substantially short time (less than 10 min) compared to

101

analogous methods already presented in the literature. The changes in the ash and carbon

102

contents of the biomass were examined with respect to the reaction temperature, residence

103

time, and pH, and were compared to other studies. In addition, we have experimentally

104

investigated the enhancement of the ash fusion temperature after the pre-treatment stage so as

105

to generate ashless biomass applying for commercial power plants.

106 107

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

108

■ MATERIALS AND METHODS

109

Preparation of raw materials. Seven biomass samples were selected for this work:

110

Miscanthus, Kenaf, Corn stalk (grown in Korea), Napiergrass, Cashew Nut Shell (CNS;

111

imported from Vietnam), Empty Fruit Bunch (EFB), and Palm Kernel Shell (PKS; imported

112

from Malaysia). All the samples were cultivated in 2017. Before leaching the alkali metals by

113

acid solution pre-treatment, all the raw biomass were dried at 105°C for 6 h in an oven to

114

reduce the water content to < 1wt%, and thereby easily chopped by the shredder (D3V-10,

115

Japan). With the exception of the PKS and CNS due to proper initial size (average 30-40 mm),

116

all the samples were crushed into small fragments and separated through a sieve, with the size

117

ranging from 10 to 20 mm.

118

Preparation of the ashless biomass. In this study, we used acetic acid (CH3COOH) to

119

remove the ash from the biomass. The mineral extraction was carried out in a 500 mL

120

autoclave reactor with a heating rate of 2°C/min up to the desired temperature. The samples

121

were stirred with a magnetic stirrer at 100 rpm. After a certain reaction time, the solid residue

122

was separated using a filter paper, with the pore size of 1 µm, washed continuously with

123

ultrapure water until it reached a neutral pH, and finally oven-dried at 105°C for 6 h. The

124

filtrate obtained from the filtration stage can be also treated to produce high-valuable mineral

125

products through ion fractionation, and the resulting fraction with the continuously depleting

126

mineral content is re-used in the ashless biomass process to extract the ash minerals. The

127

detailed ashless biomass process is shown in Figure 1.

128 129 130

In general, the mineral ions contained in the biomass can be extracted as shown by the following reactions: (Na + CH3COOH + H2O → CH3COONa + H3O)

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36 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

ACS Sustainable Chemistry & Engineering

131

(Mg + CH3COOH + H2O → [CH3COO]2Mg + H3O)

132

(K + CH3COOH + H2O → CH3COOK + H3O)

133

(Ca + CH3COOH + H2O → [CH3COO]2Ca + H3O)

134

(Mn + CH3COOH + H2O → [CH3COO]2Mn + H3O)

135

(Cl2 + CH3COOH → CH2ClCOOH + HCl)

136

There is an important concern to achieve a successful ash rejection. The ash rejection

137

efficiency was calculated from the difference in the ash weight between the initial and the

138

treated biomass (dry basis) based on the proximate analysis. The ash rejection can be defined

139

by Eq. (1).

140

141

Ash rejection (%) = =

Ash content of treated one without water Ash content of raw biomass without water M ash , treated , dry basis M ash , raw , dry basis

(1)

from Proximate analysis

142

Where, M denotes the weight.

143

The biomass sample needs to be fully immersed in the catalytic solution (i.e. acetic acid) to

144

increase the mobility of the minerals via solvation as well as the reacting surface area. The

145

biomass solid/liquid ratio is dependent on the type of biomass used, because the amount of

146

liquid absorption of each biomass differs. For example, the highly porous Miscanthus sample

147

tends to absorb more water than other samples. Therefore, a larger amount of acetic acid

148

solution is necessary to fully immerse the Miscanthus in the reaction chamber (Figure 2). To

149

evaluate the pre-treatment performance, we investigated the solid yield of the ashless biomass

150

and the amount of ash rejection in this study. The solid yield can be expressed in Eq. (2),

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 8 of 36

151

which was calculated on the basis of the amount of the organic solids between the original

152

and the treated biomass sample.

153

154

Solid yield (%) = =

Carbon and VM content of treated one without water and ash Carbon and VM content of raw biomass without water and ash ( M F .C . + MVM ) treated , dry basis ( M F .C . + MVM ) raw , dry basis

(2)

from Proximate analysis

155

Where, F.C. and VM denote the fixed carbon and volatile matter contents, respectively.

156

The mass variation was measured by a precise balance with the accuracy of 0.001 g (ME403,

157

Mettler-Toledo).

158

Experimental parameters. In the first, we observed the ash rejection efficiency with

159

different concentrations of the acetic acid solution at 40°C for 10 min. The optimization of

160

the solution concentration is a crucial concern between the solid yield of the ashless biomass

161

and ash rejection, because the highly concentrated solution dissolves more of the organic

162

solids. Therefore, it results in a lower organic solid yield even though it can guarantee a high

163

amount of ash rejection. Therefore, we carefully decided the solution concentration to

164

minimize the loss of the organic solid yield as well as to maximize the ash rejection. The

165

concentration of the acetic acid solution was 16.61 M, and then, at a given solution

166

concentration, we observed the amount of ash rejection and organic solid yield at the different

167

temperatures and residence times. The contents of pH, temperature and residence time are

168

detailed in influence of experimental parameters of the Results and discussion.

169

Analysis of the ashless biomass. The treated biomass samples were crushed and separated

170

into ~75 µm sizes for further analysis such as the proximate analysis (TGA-701, LECO Co.),

171

elemental analysis (TruSpec elemental analyzer and SC-432DR sulfur analyzer, LECO Co.),

172

and calorific value analysis (AC600 Semi-auto calorimeter, LECO Co.). In addition, we have

173

compared the surface elements between the raw and treated samples to directly identify the

ACS Paragon Plus Environment

Page 9 of 36 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

ACS Sustainable Chemistry & Engineering

174

effectiveness of the proposed treatment method using scanning electron microscope and

175

energy dispersive x-ray spectroscopy (SEM-EDX, S-4700, HITACHI). Moreover, the

176

prepared biomass samples were ashed in an electric muffle furnace at 600°C for 6 h (ASTM

177

Standard Method Number E1755-01) to identify the inorganic chemical content using X-ray

178

fluorescence (XRF, Primus II, RIGAKU Co.). The ash fusion temperature was obtained from

179

the melting test according to ASTM D1857 (5E-AF4000).

180

181

■ RESULTS AND DISCUSSION

182

Influence of experimental parameters. Figure 3 shows the variations in the biomass

183

constituents, ash rejection, and solid yield of Miscanthus at different pH (Figure 3-a),

184

temperature (Figure 3-b), and residence time (Figure 3-c) in the acetic acid solution,

185

respectively. For the acetic acid treatment, the initial ash content decreased with decreasing

186

pH at a fixed temperature and time, and the relative composition of the fixed carbon and

187

volatile matter increased. In more detail, we obtained the maximum ash rejection at pH value

188

of 1.76. The ash rejection increased from 79.25 to 84.92% when pH decreased from 2.24 to

189

1.76. However, it decreased down to 75.7% at pH value of 1.28. In contrast, the solid yield

190

continuously decreased from 99.23 to 96.62% when varying the pH from 2.24 to 1.28. This

191

was due to the increased dissolution of hemicellulose in the stronger (more concentrated) acid

192

solution. At temperatures above 170°C in the presence of acids, it is known that the xylose

193

(C5) component in biomass is converted into organic acids (e.g. glycolic acid, furfural, and

194

acetic acid) with an excessive reaction time.21 Considering the aforementioned reason, our

195

proposed pre-treatment method was performed below 100°C in order to minimize the organic

196

solid loss. In particular, at pH 1.76, we obtained an ash rejection of 84.92% and a solid yield

197

of 98.91%. In addition, above this pH, the efficiency slope became steady, because the acetate

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

198

ions in solution were sufficient to react with potassium and, hence, the excessive ions did not

199

contribute towards enhancing the leaching reaction. As a result, we set the initial solution

200

concentration as pH 1.76 for testing the influence of temperature and time.

201

According to Figure 3-b, the experimental results showed that the ash rejection increased

202

from 77.38 to 85.41% with increasing the temperature from 20oC to 60oC, however at higher

203

temperature than 60oC the ash rejection decreased down to 70.31%. The solid yield almost

204

linearly decreased from 99.91 to 95.46% when rising the temperature from 20oC to 100oC.

205

As seen in Figure 3-b, the left-side bar chart shows that volatile matter component variation

206

behaves similarly to the ash rejection. The volatile matter reduction at higher than the

207

temperature 60oC seems dominant, attributing to the relative increase of ash portion from

208

Proximate analysis. According to Eq. (1), the smaller remaining VM after a treatment can

209

increase the numerator value, and thereby showing the lower ash rejection. Thus, we decided

210

this liquid concentration (pH 1.76) and temperature (60°C) for further experiments to identify

211

the variation in ash rejection and solid yield by varying the residence time from 0 to 90 min

212

and the results are reported in Figure 3-c. Differing from pH and temperature tests results, we

213

observed the maximum ash rejection for the shortest treatment time, of which value was

214

85.41% for 10 min. Further study on the shorter residence time than 10 min will give us

215

information of more optimal residence time, which is on-going research in our laboratory. In

216

more detail, the ash rejection decreased from 85.41 to 80.52% for the treatment time from 10

217

to 90 min. The solid yield decreased from 98.75 to 94.89% for the same residence time

218

conditions.

219

The efficiency slope decreased after 10 min, which was attributed to the solution being more

220

diluted with time owing to the water generated as a by-product of the reaction. As shown in

221

Figure 4, the rejection efficiency of potassium was evaluated depending on the particle size of

222

the biomass at fixed reaction conditions (pH 1.76, 60°C and 10 min), which were already

ACS Paragon Plus Environment

Page 10 of 36

Page 11 of 36 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

ACS Sustainable Chemistry & Engineering

223

shown to be the optimal extraction conditions. As the particle size of Miscanthus decreases,

224

the available particle surface area that can be reacted and the rejection efficiency of

225

potassium both increase. Due to the low density characteristics of the biomass, the pelletizing

226

process is indispensable at the end of the ashless biomass process. For the biomass pelletizing,

227

the optimal size of the powdery biomass is about 10 mm, because the rejection efficiency of

228

potassium is more than 99wt% at that size. The more pulverization required, the more

229

uneconomical is the process due to an increase in the processing cost.

230

Basic characteristics of ashless biomass. In general, biomass contains both external ash

231

(i.e. soil) and inherent ash. While the external ash in the biomass surface is easily rinsed out

232

by water, the inherent ash is hardly extracted due to the strong chemical interactions with the

233

carbonic moieties of the cellulose, hemicellulose and lignin constituents.22,23 This is the main

234

reason for using the catalytic solution instead of just water to efficiently reduce the ash

235

contents in biomass. In this study, the acetic acid with a pH adjustment was used to remove

236

the internal ash. The acetic acid was used to mainly remove alkali-based ash such as the

237

potassium and sodium.

238

Table 1 summarizes the results of the fuel characterization. As shown, we observed a

239

significant ash rejection between the initial and the treated sample, while the difference in the

240

carbon content was relatively small. In addition, the calorific values of most samples

241

increased up to 100-600 kcal/kg, which was attributed to increase in portion of the fixed

242

carbon and volatile matter due to the higher rejected ash content per fuel. Miscanthus and

243

Corn stalk showed a relatively higher difference in their calorific values after the treatment,

244

implying that the proposed treatment method is more effective for the naturally low calorific

245

value biomasses in terms of the fuel upgrade. In addition, when the five organic components

246

such as carbon (C), hydrogen (H), nitrogen (N), oxygen (O) and sulfur (S) were investigated

247

by ultimate analysis, the precursors of the NOx and SOx also decreased by up to 80%, while

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

248

the C, H and O contents remained constant. In other words, our proposed treatment method

249

can be evaluated to improve the fuel quality in terms of environmental benefits. Since the

250

pre-treatment temperature (60°C) of the ashless process for extracting the ash is very low, the

251

biomass studied in this experiment has a very high solid yield, above 97%. The overall mass

252

balance for the production process of the ashless biomass using Miscanthus is illustrated in

253

Figure 5. The constituents of the biomasses were analyzed according to the NREL Laboratory

254

Analytical Procedures (NREL / TP-510-42618, structural carbohydrates and lignin, NREL /

255

TP-510-42623, sugars and by-products in the liquid fraction). In case of Miscanthus, the solid

256

to aqueous acetic acid solution ratio is 1:8, meaning that 1 kg of many pieces of the

257

Miscanthus and 8 kg of solution are mixed for the mineral extraction. As a next step, the

258

separation with a filter produces 0.876 kg of ashless biomass fuel and 8.124 kg of solution

259

waste, respectively. And then, after extracting 0.1932 kg of the minerals, the remaining

260

7.9308 kg acid liquid is further re-used in the extraction process. Therefore, in this study, the

261

extraction process was designed for minimizing the wastewater by reusing the acidic solution

262

while increasing the yield of the ashless biomass to be used as fuel.

263

According to the previous literatures, an ash rejection of 10-50%, with water as a pre-

264

treatment solution,15,24,25 and an ash rejection of 20-70% with either an acidic or alkaline

265

catalyst solution17,18,20 have been reported. Furthermore, the ratio of the solid (biomass) to

266

liquid (solution) and residence time has been reported as 1:20~50 and 4-24 h, respectively.

267

However, in this study, the solid-liquid ratio was reduced to 1: 8 or less, and the residence

268

time was set to 10 minutes, which was different from the previous study. In addition, we

269

prepared the high concentration solution with 16.6 M, which resulted in high ash rejection

270

while the previous researchers employed the molar concentration of acetic acid as below 3 M.

271

Considering the relatively high ash rejection, solid yield, low solid-liquid ratio, and short

272

residence time, we believe that the proposed method is fairly competitive for practical

ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36 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

273

ACS Sustainable Chemistry & Engineering

applications.

274

Comparison of mineral composition. Ash emission is directly related to the initial amount

275

of minerals present inside the biomass prior to combustion. As summarized in Table 2, the

276

ash content of the raw Miscanthus is 30 times higher than that of ashless-treated biomass

277

after combustion. Although the initial compositions of the minerals are dependent on the

278

different types of biomass, we found that all the biomass mainly contained a large portion of

279

potassium, calcium and silicon. Therefore, the development of the ash rejection method for

280

ashless biomass production is important. For instance, the acetic acid solution may be more

281

effective on biomass stalk, rather than the shell of the biomass such as PKS and CNS,

282

because the higher surface area of the stalk can accelerate the reaction between the solid and

283

the liquid agent, which has been confirmed by our comparisons of the ash rejection values

284

(See Table 1). Particularly, almost all the groups of biomass used in this experiment showed a

285

high rejection value of up to 98-99% for potassium oxide, 70-93% for calcium oxide, and 60-

286

95% for silica, which was significantly higher compared to the previous works in the

287

literature24-27. Karnowo et al.26 have suggested the potassium extraction method, from which

288

the potassium present inside of the rice husk was removed up to 94.0% via the use of bio-oil

289

containing phenolic and dehydrated sugars with the ratio of oil to rice husk of 20 for 24 h.

290

Liaw et al.27 reported that the hydrothermal process with the operating conditions of 10Mpa

291

and 150°C removed more than 90% of the calcium contained inside the wood chips. On the

292

other hand, our proposed extraction process uses relatively lower temperature, lower pressure

293

and a shorter residence time than those found in the previous works24-27. We believe that our

294

process can maximize the solid yield of the hemicellulose and cellulose.

295

As depicted in Figure 6 and 7, the raw and ashless Miscanthus were produced as a char at

296

700-900°C, and the characteristics of the char surface were examined at various

297

magnifications. The difference between two samples was apparent above 1,000 magnification

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

298

condition. The observed raw Miscanthus char surface was covered with many mineral grains

299

whereas the cylindrical shape was not clearly visible. However, the char surface of the treated

300

Miscanthus was clearer than the raw sample, allowing for a smoother appearance. During the

301

surface elemental analysis via the EDX, a few peaks of inorganic matter such as K, Fe, Si, Al,

302

and Ca were observed for the raw biomass. All of these materials result in fouling, slagging,

303

and clinker during biomass combustion. C and O were the main elements that remained in the

304

ashless biomass (Figure 8).

305

Influence of the ash melting behavior by the mineral extraction. The melting

306

characteristics after ash extraction in biomass are shown using a ternary phase diagram.

307

Figure 9 shows the seven biomasses plotted in the diagram with the three points of SiO2 +

308

Al2O3 + Fe2O3 + Na2O + TiO2, K2O + P2O5 + SO3 + Cl2O and CaO + MgO + MnO. According

309

to the composition of the mineral oxides, the upper part is composed of the silicon type “S

310

type”, the lower left is the calcium type “C type”, the right side is the potassium type “K

311

type”, and the middle type is “CK type”, which is an intermediate between calcium and

312

potassium.28 The melting temperature of the ash is the highest (1,300°C) near silicon and

313

calcium, while the lowest temperature is 1,100°C for the K type, and the remaining area is

314

1,100-1,300°C. In addition, the upper part of the plot shows the characteristics of high acidity

315

due to the increase in SiO2, Al2O3, and TiO2 belonging to the acid series. To avoid fouling and

316

slagging problems, the biomass should be desirably placed as far as possible from the right

317

corner, in other words, low potassium content. For instance, raw Kenaf and Napiergrass are

318

initially located in the right corner (K type) due to the high K2O contents. After the treatment,

319

they migrate to the upper S type and C type, which means the alleviation of the low melting

320

risk during combustion. Likewise, the other biomass are also shifted toward the top or left

321

corner of the triangle by means of this proposed process.

322

Furthermore, the ash fusion temperature and deposition tendency with the base to acid (B/A)

ACS Paragon Plus Environment

Page 14 of 36

Page 15 of 36 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

ACS Sustainable Chemistry & Engineering

323

ratio, comparison of silica and alkali, slagging, fouling, and the chlorine indices are evaluated

324

and listed in Table 3. The ash fusion temperature of the biomass is very important, because it

325

can be used to determine the melting behavior of the ash. Most biomass causes slagging and

326

fouling to occur in the boiler, since the initial deformation temperature (IDT) is lower than

327

1,150°C. However, when the alkali metals are extracted from the biomass, the IDT of the

328

ashless biomass is increased to a value higher than 30% compared to that of the raw biomass.

329

When all the biomass are treated under the ashless biomass process, the IDT is higher than

330

1,200°C. Among them, in particular, the IDT of the Kenaf is higher than 1,550°C after the

331

pre-treatment. There is a significant level of ash composition in the biomass;29-34 (a) B/A ratio:

332

>1.0, (b) silica percentage: 4.0, (f) chlorine content: >0.5 (Table 4). The raw biomass possesses a severe

334

slagging/fouling potential as determined based on the B/A ratio and the total alkali indicators,

335

thus presenting a dangerous impact in the boilers. However, the ashless biomass showed a

336

near zero value. Similar to the data presented in Table 2, it can be seen that the chlorine,

337

which causes high temperature corrosion, can be removed by the ashless biomass process by

338

more than 90%. Thus, one can reduce the generation of chlorine compounds such as KCl and

339

HCl, which may corrode the inner part of the boiler or the super-heater and the economizer

340

tube. Therefore, the proposed ash extraction process in this study can contribute to the clean

341

biomass combustion technology in the near future.

342 343

■ CONCLUSIONS

344

In this study, we proposed a acetic acid based pretreatment method to extract the alkali

345

minerals inside herbaceous biomass and investigated the effect of treatment parameters such

346

as concentration, temperature and residence time. The optimal pretreatment conditions for the

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

347

maximum ash rejection were found to be 1.76 in pH, 60°C in temperature and 10 min in

348

residence time, at which the hemicellulose, cellulose, and lignin losses were minimized. The

349

further investigation of shorter treatment time than 10 min will be the next step in our group.

350

Among seven biomass samples, Kenaf showed the maximum ash rejection of 93.48%. In

351

particular, the potassium and sodium elements in Kenaf, which are the major chemical factors

352

influencing the fouling and slagging in a boiler, were removed up to 99.46 and 100%,

353

respectively. Comparing the wet treatment effect between stalk-based and shell based

354

biomasses, the proposed treatment seems more effective for the stalk-based biomass such as

355

Miscanthus, Kenaf, Corn stalk, Napiergrass, and EFB than the shell-based biomass such as

356

CNS and PKS due to their higher surface areas, of which result was validated by BET

357

measurement. From the experimental results, we opine that the proposed pretreatment

358

method may contribute to the significant reduction of chloride-induced corrosion and the

359

slagging/fouling by ash deposition in a boiler during combustion. Finally, we believe that the

360

proposed method will pave a way to the fuel switching from coals to agricultural residues and

361

herbaceous biomass, which have a shorter harvest interval than the lignocellulosic biomass.

362 363 364 365

ACS Paragon Plus Environment

Page 16 of 36

Page 17 of 36 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

ACS Sustainable Chemistry & Engineering

366

■ AUTHOR INFORMATION

367

Corresponding Authors

368

*C.-H. JEON. E-mail: [email protected]. Tel: +82-51-510 3051

369

*Y.-C. CHOI. E-mail: [email protected]. Tel: +82-42-860 3784

370

Notes

371

The authors declare no competing financial interest.

372 373

■ ACKNOWLEDGMENTS

374

This research was supported by the Clean Power Core Technology Program of the Korea

375

Institute of Energy Technology Evaluation and Planning (KETEP) granted financial resource

376

from the Ministry of Trade, Industry & Energy, and Republic of Korea (No.

377

20151120100180).

378

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

379

■ REFERENCES

380

(1) Carpenter, D.; Westover, T.; Czernik, S.; Jablonski, W. Biomass feedstocks for renewable

381

fuel production: a review of the impacts of feedstock and pretreatment on the yield and

382

product distribution of fast pyrolysis bio-oils and vapors. Green Chem. 2014, 16(2), 384-406,

383

DOI 10.1039/C3GC41631C.

384

(2) Ranzi, E.; Debiagi, P.; Frassoldati, A. Mathematical modeling of fast biomass pyrolysis

385

and bio-oil formation. Note I: Kinetic mechanism of biomass pyrolysis. ACS Sustainable

386

Chem. Eng. 2017, 5(4), 2867-2881, DOI 10.1021/acssuschemeng.6b03096.

387

(3) Cheah, S.; Jablonski, W.; Olstad, J.; Carpenter, D.; Barthelemy, K.; Robichaud, D.;

388

Westover, T. Effects of thermal pretreatment and catalyst on biomass gasification efficiency

389

and syngas composition. Green Chem. 2016, 18(23), 6291-6304, DOI 10.1039/C6GC01661H.

390

(4) Miedema, J.; Benders, R.; Moll, H.; Pierie, F. Renew, reduce or become more efficient?

391

The climate contribution of biomass co-combustion in a coal-fired power plant. Appl.

392

Energy 2017, 187, 873-885, DOI 10.1016/j.apenergy.2016.11.033.

393

(5) Sikarwar, V.; Zhao, M.; Clough, P.; Yao, J.; Zhong, X.; Memon, M.; Fennell, P. An

394

overview of advances in biomass gasification. Energy Environ. Sci. 2016, 9(10), 2939-2977,

395

DOI 10.1039/C6EE00935B.

396

(6) Wang, L.; Skreiberg, Ø.; Becidan, M.; Li, H. Investigation of rye straw ash sintering

397

characteristics and the effect of additives. Appl. Energy 2016, 162, 1195-1204, DOI

398

10.1016/j.apenergy.2015.05.027.

399

(7) Niu, Y.; Wang, Z.; Zhu, Y.; Zhang, X.; Tan, H.; Hui, S. Experimental evaluation of

400

additives and K2O–SiO2–Al2O3 diagrams on high-temperature silicate melt-induced slagging

401

during biomass combustion. Fuel 2016, 179, 52-59, DOI 10.1016/j.fuel.2016.03.077.

402

(8) Zhang, X. Essential scientific mapping of the value chain of thermochemically converted

ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36 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

ACS Sustainable Chemistry & Engineering

403

second-generation

404

10.1039/C6GC02335E.

405

(9) Zhao, X.; Zhou, H.; Sikarwar, V.; Zhao, M.; Park, A.; Fennell, P.; Fan, L. Biomass-based

406

chemical looping technologies: the good, the bad and the future. Energy Environ. Sci. 2017,

407

10(9), 1885-1910, DOI 10.1039/C6EE03718F.

408

(10) Kassman, H.; Pettersson, J.; Steenari, B; Åmand, L. Two strategies to reduce gaseous

409

KCl and chlorine in deposits during biomass combustion-injection of ammonium sulphate

410

and co-combustion with peat. Fuel Process. Technol. 2013, 105, 170-180, DOI

411

10.1016/j.fuproc.2011.06.025.

412

(11) Szemmelveisz, K.; Szűcs, I.; Palotas, A.; Winkler, L.; Eddings, E. Examination of the

413

combustion conditions of herbaceous biomass. Fuel Process. Technol. 2009, 90(6), 839-847,

414

DOI 10.1016/j.fuproc.2009.03.001.

415

(12) Priyanto, D.; Ueno, S.; Sato, N.; Kasai, H.; Tanoue, T.; Fukushima, H. Ash

416

transformation by co-firing of coal with high ratios of woody biomass and effect on slagging

417

propensity. Fuel 2016, 174, 172-179, DOI 10.1016/j.fuel.2016.01.072.

418

(13) Wang, L.; Skreiberg, Ø.; Becidan, M. Investigation of additives for preventing ash

419

fouling and sintering during barley straw combustion. Appl. Therm. Eng. 2014, 70(2), 1262-

420

1269, DOI 10.1016/j.applthermaleng.2014.05.075.

421

(14) Chen, W.; Ye, S.; Sheen, H. Hydrothermal carbonization of sugarcane bagasse via wet

422

torrefaction in association with microwave heating. Bioresour. Technol. 2012, 118, 195-203,

423

DOI 10.1016/j.biortech.2012.04.101.

424

(15) Yu, C.; Thy, P.; Wang, L.; Anderson, S.; VanderGheynst, J.; Upadhyaya, S.; Jenkins, B.

425

Influence of leaching pretreatment on fuel properties of biomass. Fuel Process. Technol. 2014,

426

128, 43-53, DOI 10.1016/j.fuproc.2014.06.030.

427

(16) Bach, Q.; Tran, K.; Skreiberg, Ø.; Khalil, R.; Phan, A. Effects of wet torrefaction on

bio-fuels.

Green

Chem.

2016,

ACS Paragon Plus Environment

18(19),

5086-5117,

DOI

ACS Sustainable Chemistry & Engineering 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

Page 20 of 36

428

reactivity and kinetics of wood under air combustion conditions. Fuel 2014, 137, 375-383,

429

DOI 10.1016/j.fuel.2014.08.011.

430

(17) Vamvuka, D.; Troulinos, S.; Kastanaki, E. The effect of mineral matter on the physical

431

and chemical activation of low rank coal and biomass materials. Fuel 2006, 85(12), 1763-

432

1771, DOI 10.1016/j.fuel.2006.03.005.

433

(18) Liu, X.; Bi, X. Removal of inorganic constituents from pine barks and switchgrass. Fuel

434

Process. Technol. 2011, 92(7), 1273-1279, DOI 10.1016/j.fuproc.2011.01.016.

435

(19) Asadieraghi, M.; Daud, W. Characterization of lignocellulosic biomass thermal

436

degradation and physiochemical structure: effects of demineralization by diverse acid

437

solutions. Energy Convers. Manage. 2014, 82, 71-82, DOI 10.1016/j.enconman.2014.03.007.

438

(20) Park, D.; Yun, Y.; Park, J. Studies on hexavalent chromium biosorption by chemically-

439

treated

440

10.1016/j.chemosphere.2005.02.020.

441

(21) Rackemann, D.; Doherty, W. The conversion of lignocellulosics to levulinic

442

acid. Biofuels Bioprod. Bioref. 2011, 5(2), 198-214, DOI 10.1002/bbb.267.

443

(22) Liu, Q.; Chmely, S.; Abdoulmoumine, N. Biomass treatment strategies for

444

thermochemical

445

10.1021/acs.energyfuels.7b00258.

446

(23) Bhatnagar, A.; Sillanpää, M.; Witek-Krowiak, A. Agricultural waste peels as versatile

447

biomass for water purification–a review. Chem. Eng. J. 2015, 270, 244-271, DOI

448

10.1016/j.cej.2015.01.135.

449

(24) Davidsson, K.; Korsgren, J.; Pettersson, J.; Jaglid U. The effects of fuel washing

450

techniques on alkali release from biomass. Fuel 2002, 81(2), 137-142, DOI 10.1016/S0016-

451

2361(01)00132-6.

452

(25) Banks, S.; Nowakowski, D.; Bridgwater, A. Fast pyrolysis processing of surfactant

biomass

of

Ecklonia

conversion. Energy

sp. Chem. 2005,

Fuels 2017,

ACS Paragon Plus Environment

60(10),

31(4),

1356-1364,

3525-3536,

DOI

DOI

Page 21 of 36 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

ACS Sustainable Chemistry & Engineering

453

washed

454

10.1016/j.fuproc.2014.07.005.

455

(26) Zahara, Z.; Kudo, S.; Norinaga, K.; Hayashi, J. Leaching of alkali and alkaline earth

456

metallic species from rice husk with bio-oil from its pyrolysis. Energy Fuels 2014, 28(10),

457

6459-6466, DOI 10.1021/ef501748h.

458

(27) Liaw, S.; Yu, Y.; Wu, H. Association of inorganic species release with sugar recovery

459

during

460

10.1016/j.fuel.2015.11.044.

461

(28) Zeng, T.; Pollex, A.; Weller, N.; Lenz, V.; Nelles, M. Blended biomass pellets as fuel for

462

small scale combustion appliances: Effect of blending on slag formation in the bottom ash

463

and pre-evaluation options. Fuel 2018, 212, 108-116, DOI 10.1016/j.fuel.2017.10.036.

464

(29) Pronobis, M. Evaluation of the influence of biomass co-combustion on boiler furnace

465

slagging by means of fusibility correlations. Biomass Bioenergy 2005, 28(4), 375-383, DOI

466

10.1016/j.biombioe.2004.11.003.

467

(30) Vamvuka, D.; Pitharoulis, M.; Alevizos, G.; Repouskou, E.; Pentari, D. Ash effects

468

during combustion of lignite/biomass blends in fluidized bed. Renewable Energy 2009,

469

34(12), 2662-2671, DOI 10.1016/j.renene.2009.05.005.

470

(31) Xiao, R.; Chen, X.; Wang, F.; Yu, G. The physicochemical properties of different

471

biomass ashes at different ashing temperature. Renewable Energy 2011, 36(1), 244-249, DOI

472

10.1016/j.renene.2010.06.027.

473

(32) Teixeira, P.; Lopes, H.; Gulyurtlu, I.; Lapa, N.; Abelha, P. Evaluation of slagging and

474

fouling tendency during biomass co-firing with coal in a fluidized bed. Biomass Bioenergy

475

2012, 39, 192-203, DOI 10.1016/j.biombioe.2012.01.010.

476

(33) Yadav, V.; Baruah, B.; Khare, P. Comparative study of thermal properties of bio-coal

477

from aromatic spent with low rank sub-bituminous coals. Bioresour. Technol. 2013, 137, 376-

Miscanthus. Fuel

wood

Process.

hydrothermal

Technol.

processing. Fuel

2014,

2016,

ACS Paragon Plus Environment

128,

166,

94-103,

581-584,

DOI

DOI

ACS Sustainable Chemistry & Engineering 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

478

385, DOI 10.1016/j.biortech.2013.03.131.

479

(34) García, R.; Pizarro, C.; Álvarez, A.; Lavín, A.; Bueno, J. Study of biomass combustion

480

wastes. Fuel 2015, 148, 152-159, DOI 10.1016/j.fuel.2015.01.079.

481

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36 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

ACS Sustainable Chemistry & Engineering

Table 1 Fuel characteristics of raw and ashless biomass

Contents

Proximate analysis (dry basis, wt %)

Ultimate analysis (daf, wt %)

Solid yield (daf*, wt %)

V. M.

Ash

F.C.

C

H

N

O

S

O/C

H/C

-

71.26

13.25

15.49

49.30

6.67

1.28

42.72

0.03

0.65

98.75

81.14

1.93

16.93

52.30

5.59

0.36

41.72

0.03

-

75.45

5.13

19.42

49.97

5.84

1.77

42.33

97.98

84.54

0.33

15.13

51.67

6.01

0.07

-

71.57

11.70

16.73

50.77

6.24

97.84

81.67

1.98

16.35

51.43

-

74.05

6.10

19.85

98.52

80.87

2.16

-

79.22

95.25

Raw Miscanthus Ashless miscanthus Raw Kenaf Ashless kenaf Raw Corn stalk Ashless Corn stalk Raw Napierglass Ashless Napierglass Raw EFB Ashless EFB Raw PKS Ashless PKS Raw CNS Ashless CNS

Atomic ratio

HHV [kcal/kg]

LHV [kcal/kg]

1.62

4,020

3,650

0.59

1.28

4,340

4,040

0.08

0.85

0.12

4,390

4,050

42.24

0.02

0.82

0.12

4,290

3,940

1.02

41.82

0.15

0.82

0.12

3,980

3,650

5.91

0.05

42.57

0.03

0.83

0.11

4,590

4,260

48.83

6.41

1.24

43.44

0.08

0.89

0.13

4,270

3,890

16.97

53.56

5.97

1.56

38.85

0.07

0.73

0.11

4,490

4,120

5.01

15.78

55.59

6.30

1.82

36.23

0.07

0.65

0.11

4,870

4,240

81.15

1.40

17.46

53.19

6.30

1.68

38.75

0.08

0.73

0.12

4,630

4,290

-

72.73

3.64

23.63

57.04

5.67

1.51

35.75

0.03

0.63

0.10

4,860

4,510

98.46

75.92

1.78

22.30

56.52

6.00

0.07

37.38

0.03

0.66

0.11

4,880

4,540

-

80.27

1.47

18.26

58.88

6.64

1.49

33.00

0.00

0.56

0.11

5,430

5,010

98.58

78.56

0.62

20.82

54.28

6.01

0.11

39.56

0.04

0.73

0.11

4,970

4,640

*daf : dry and ash free

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 24 of 36

Table 2 Mineral oxide contents of raw and ashless biomass Mineral oxide (mg/kg) Contents Na2O

MgO

Al2O3

SiO2

P2O5

K2O

CaO

TiO2

MnO

Fe2O3

Raw Miscanthus

398

2,346

7,793

98,187

2,479

14,699

3,685

570

398

2,001

Ashless Miscanthus

21

46

1,004

16,988

166

188

248

104

12

559

2,368

4,142

630

1,579

8,109

28,627

5,372

82

92

267

Ashless Kenaf

0

51

99

499

632

155

1,495

31

11

369

Raw Corn stalk

105

7,485

3,158

50,841

20,818

24,760

7,883

363

129

1,427

0

39

250

16,117

1,212

219

1,269

61

13

575

122

878

1,098

9,103

4,841

38,242

1,768

177

79

4,658

Ashless Napierglass

0

192

2,162

11,490

2,285

214

2,631

296

22

2,329

Raw EFB

60

2,263

1,842

18,467

2,353

21,145

2,723

235

65

906

Ashless EFB

0

0

0

1,844

0

380

380

243

151

10,964

Raw PKS

73

1,862

2,121

16,669

1,137

2,357

10,428

168

44

1,581

Ashless PKS

2

81

5421

6,436

387

795

2,421

280

15

1,967

141

1,068

181

347

1,545

9,913

1,002

100

118

300

0

404

267

542

518

989

1,843

180

188

1,263

Raw Kenaf

Ashless Corn stalk Raw Napierglass

Raw CNS Ashless CNS

ACS Paragon Plus Environment

Page 25 of 36 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

ACS Sustainable Chemistry & Engineering

Table 3 Ash fusion temperature and deposition tendency of raw and ashless biomass Ash fusion temperature (oC)

Slagging/Fouling indices

Contents IDT

ST

HT

FT

B/A ratio

Silica Percentage

Slagging index

Fouling index

Total Alkali

Chlorine contain

Raw Miscanthus

1,174

1,261

1,297

1,329

0.22

0.92

1,198

0.07

11.39

1.22

Ashless Miscanthus

1,363

1,411

1,444

1,464

0.06

0.95

1,379

0.01

1.08

0.01

Raw Kenaf

1,178

1,220

1,263

1,291

17.80

0.14

1,195

82.22

60.47

0.55

Ashless Kenaf

>1,550

>1,550

>1,550

>1,550

3.29

0.21

1,550

0.00

4.63

0.02

Raw Corn stalk

1,085

1,139

1,161

1,178

0.77

0.75

1,100

0.07

21.26

0.65

Ashless Corn stalk

1,344

>1,500

>1,500

>1,500

0.13

0.90

1,375

0.00

1.11

0.01

Raw Napierglass

1,040

1,128

1,189

1,213

4.40

0.55

1,069

0.88

62.92

0.22

Ashless Napierglass

1,258

1,282

1,290

1,301

0.38

0.69

1,264

0.00

0.99

0.01

918

1,124

1,151

1,185

1.32

0.76

964

0.16

42.36

0.31

Ashless EFB

1,393

1,483

1,492

1,498

5.62

0.14

1,412

0.00

2.72

0.01

Raw PKS

1,205

1,220

1,224

1,229

0.86

0.55

1,208

0.17

6.67

0.01

Ashless PKS

1,318

1,452

>1,500

>1,500

0.43

0.59

1,354

0.01

4.47

0.01

Raw CNS

1,084

1,399

1,461

>1,500

19.78

0.13

1,159

18.98

68.33

0.10

Ashless CNS

1,210

1,258

1,261

1,266

4.55

0.13

1,220

0.00

15.96

0.01

Raw EFB

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 26 of 36

Table 4 Summary of deposition tendency Indices

Low

Medium

High

Severe

(a) Base/acid ratio : (Fe2O3 + CaO + MgO + Na2O + K2O)/(SiO2 + Al2O3 + TiO2)

< 0.5

0.5-0.7

0.7-1.0

> 1.0

(b) Silica percentage : (SiO2)/( SiO2 + Fe2O3 + CaO + MgO) × 100

> 50

50-30

30-5

1,340

1,340-1,250

1,250-1,150

< 1,150

(d) Fouling index : (Fe2O3 + CaO + MgO + Na2O + K2O)/(SiO2 + Al2O3 + TiO2) × Na2O

< 0.2

0.2-0.5

0.5-1.0

> 1.0

(e) Total alkali : Na2O + K2O

< 2.0

2.0-3.0

3.0-4.0

> 4.0

(f) Chlorine contain

< 0.2

0.2-0.3

0.3-0.5

> 0.5

(c) Slagging index : (Rs) = 4 IT + HT/5

ACS Paragon Plus Environment

Page 27 of 36 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

ACS Sustainable Chemistry & Engineering

Figure 1 Flow chart of the proposed ashless biomass process

27

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Figure 2 S/L ratio during ash rejection experiments

28

ACS Paragon Plus Environment

Page 28 of 36

Page 29 of 36 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

ACS Sustainable Chemistry & Engineering

Figure 3 Comparison of constituents, ash rejection and solid yield of Miscanthus at different pH ((a), fixed in 40oC and 10 min), temperature ((b), fixed in 16.61M CH3COOH and 10 min) and residence time ((c), fixed in 16.61M CH3COOH and 60oC)

29

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Figure 4 The rejection efficiency of potassium from Miscanthus depending on particle size

30

ACS Paragon Plus Environment

Page 30 of 36

Page 31 of 36 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

ACS Sustainable Chemistry & Engineering

Figure 5 The overall mass balance for the production of ashless biomass (Miscanthus)

31

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Figure 6 SEM on the surface of raw Miscanthus char

32

ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36 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

ACS Sustainable Chemistry & Engineering

Figure 7 SEM on the surface of ashless Miscanthus char

33

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Figure 8 EDX on the surface of raw and ashless Miscanthus char at 900oC

34

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36 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

ACS Sustainable Chemistry & Engineering

Figure 9 Ternary phase diagram using new chemical classification system

35

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

■ For Table of Contents Use Only

Synopsis: We suggest the biomass pretreatment method, including the successive acidic solutions based treatment at below 60oC for 10 min.

36

ACS Paragon Plus Environment

Page 36 of 36