Fluidized bed torrefaction of commercial wood pellets: process

Jul 30, 2018 - This paper reports an experimental study aimed at investigating the influence of fluidized bed torrefaction treatment on the quality of...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Biofuels and Biomass

Fluidized bed torrefaction of commercial wood pellets: process performance and solid product quality Paola Brachi, Riccardo Chirone, Michele MICCIO, and Giovanna Ruoppolo Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b01519 • Publication Date (Web): 30 Jul 2018 Downloaded from http://pubs.acs.org on August 12, 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 37 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

Energy & Fuels

1

Fluidized bed torrefaction of commercial wood pellets: process performance and

2

solid product quality

3 *

Paola Brachi§ , Riccardo Chirone§, Michele Miccio‡, Giovanna Ruoppolo§

4 5 6 7 8 9 10 11 12 13

§

Institute for Research on Combustion, National Research Council, P.le Tecchio 80, 80125 Napoli, Italy ‡

Department of Industrial Engineering, University of Salerno, via Giovanni Paolo II 132, 84084 Fisciano (SA), Italy

ABSTRACT

14

This paper reports an experimental study aimed at investigating the influence of fluidized bed

15

torrefaction treatment on the quality of commercial wood pellets. In particular, an experimental

16

program was performed, which allowed to investigate, at a laboratory-scale, the impact of the

17

torrefaction temperature (200, 230, 250 °C) and the reaction time (7 and 15 min) on: a) the

18

distribution and the composition of the main output products of torrefaction process (torrefied

19

solids, condensable volatiles and permanent gases); b) the quality of torrefied pellets; and c) process

20

performance in terms of mass and energy yields of the solid product. In particular, the quality of

21

pellets was characterized in terms of apparent density, bulk density, calorific values, volumetric

22

energy density, moisture uptake, swelling behavior, hardness (shore D) and nonstandard durability

23

index. Results suggest that light torrefaction (200 °C and 7-15 min) is the most suitable to ensure a

24

sustainable production (84-85% mass yield and 94-95 energy yield) of high quality torrefied wood

25

pellets (no swelling in water, about 27% decrease in the moisture uptake after 7 days of exposure in

26

an environment at a high relative humidity of 80 %, hardness and durability comparable to those of

27

untreated wood pellets, only a 2-3% decrease in the volume energy density) in a downstream

28

configuration (torrefaction after pelletization). The torrefaction treatment of wood pellets in *

Corresponding author. Tel.: +39 081 5931567; e-mail address: [email protected] 1 ACS Paragon Plus Environment

Energy & Fuels 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 2 of 37

29

fluidized bed reactor has not been investigated so far, therefore, findings of this work can be useful

30

to highlight potential advantages and drawbacks related to use of such a technology in this specific

31

application.

32 33 34

Keywords: Torrefaction, Fluidized bed reactor, Wood pellets, Shore D Hardness, Durability, Hydrophobicity.

35 36

1. INTRODUCTION

37

Biomass is considered the renewable energy source with the highest potential for replacing

38

fossil fuels in the short to medium term. 1,2 However, there are currently a number of challenges

39

related to the quality of biomass resources that still prevent them from being used on a large scale

40

for heat and power production. Biomass has, in fact, a lower energy content per unit mass compared

41

with fossil fuels, which means that a higher load of feedstock is required in the case of biomass-fed

42

plant in order get the same amount of energy when compared to fossil fuels. Moreover, biomass is a

43

relatively bulky material. Typically, biomass bulk density ranges from 80-100 kg/m3 for agricultural

44

straws and grasses to 150-200 kg/m3 for woody resources like wood chips and sawdust3, whereas

45

the bulk density of coal is about 700 kg/m3.4 As a consequence of this, the volume of feedstock to

46

be handled increases enormously, when biomass is used as a fuel instead of coal, with all the

47

consequent issues that this can bring for logistics, e.g., storage and transportation, which are factors

48

that greatly affect profit margins and thus the convenience of a biomass-fed plant. This is also the

49

reason why it is only economically feasible to transport unprocessed biomass over a distance lower

50

than about 200 km.4 Again, biomass cannot be stored outdoor without a careful protection since it is

51

prone to natural decomposition over time and breakdown with exposure to moisture and pests (e.g.,

52

flies and mosquitos), with consequent loss of quality and off-gas emissions; the high moisture

53

content of some kinds of biomass (i.e., agro-industrial residues) also accelerates the decomposition

54

process. In this regard, it is worth noting that drying biomass has a little benefit for the 2 ACS Paragon Plus Environment

Page 3 of 37 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

Energy & Fuels

55

improvement of biomass storage behavior. In fact, because of its hydrophilic nature, biomass can

56

re-absorb moisture and start to decompose again.

57

Recent developments in mechanical densification technologies, including pelletization and

58

briquetting, have substantially improved the economics of moving biomass around the globe4,5.

59

Typically, these technologies increase the biomass energy density (MJ/m3) through the increase of

60

its bulk density (kg/m3), but have a very little benefit for the improvement of properties such as the

61

low calorific values (MJ/kg) and the remarkable hydrophobic behavior. The lower heating value

62

(LHV) of currently marketed wood pellets is, in fact, approximately 15-19 MJ/kg, which still limits

63

the mixing ratio that can be used in co-firing with coal typically having a lower heating value

64

(LHV) of about 23-28 MJ/kg.6

65

Fuel pellets with a very high bulk energy density in the range from 15 to 18 GJ/m3 and a

66

lower heating value as high as 20-24 MJ/kg can be obtained when torrefaction, a relatively new

67

thermal pretreatment of biomass, is combined with pelletization.6 It is worth mentioning that wood

68

pellets, which are known to be a very energy-dense biomass fuel, have a bulk energy density that

69

typically range from 7.5 to 10 GJ/m3 whereas that of coal is about 18.4-23.8 GJ/m3.

70

Basically, torrefaction is a thermo-chemical treatment where a solid biomass is heated in an

71

inert environment up to a temperature ranging between 200 and 300 °C. It is traditionally

72

characterized by low particle heating rate (typically less than 50 °C/min) and a relatively long

73

reactor residence time that typically ranges from 30 to 120 minutes depending on the specific

74

feedstock, technology and temperature. The benefits accomplished by torrefaction include the

75

possibility to convert any lignocellulosic raw materials into a hydrophobic solid, which can be

76

stored outdoors, can be co-fired efficiently with coal and have improved properties even for other

77

applications such as pyrolysis and gasification. The bulk density of the torrefied material, however,

78

is generally lower than that of the raw biomass, making transport and storage economically

79

challenging. Therefore, combining torrefaction and pelletization has great potential to upgrade raw

80

biomass into a standard commodity fuel. 3 ACS Paragon Plus Environment

Energy & Fuels 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 4 of 37

81

So far, two potential pathways, either an upstream (torrefaction before pelletization) or

82

downstream (torrefaction after pelletization) configuration, have been assessed for retrofitting

83

torrefaction within pellet production facilities. In particular, since a successful upstream integration

84

certainly produces highly dense and durable torrefied pellets, whereas a downstream integration

85

could potentially results in a deterioration of pellet quality (e.g., a loss in strength and density), most

86

of the research efforts performed so far, have been mainly focused on the upstream integration.7,8

87

However, many challenges still remain with such configuration, which includes: (a) difficulty in

88

densification with a consequent need for binders along with a conditioning step in order to obtain

89

reasonable pelletization efficiency;9,10 (b) frequent maintenance requirements due to the abrasive

90

nature of torrefied biomass that lower useful life of pellet;11 (c) safety concerns from fine generation

91

during grinding and pelletizing torrefied biomass, in particular possibility of dust explosion11.

92

Moreover, it is worth noting that, most of the currently used binders are hydrophilic and, hence, in

93

addition to impart additional costs and lower the pellet heating value, they may also compromise

94

pellet hydrophobicity.11 On the other hand, torrefaction as a downstream operation, where white

95

pellets are torrefied to produce black pellets, has the benefit of being a simple bolt-on integration. In

96

addition, downstream integration eliminates the need for additional grinding and pelletizing units,

97

while minimizing plant contamination due to dust generated from processing torrefied biomass.

98

Again, CAPEX can be reduced, since a typical torrefier has significantly higher throughput when

99

processing pellets compared to wood chips.11

100

In this context, taking into account the operational benefits of downstream integration

101

compared to the upstream one, an experimental study on the torrefaction treatment of commercial

102

wood pellets, which has been poorly investigated so far12-14, has been focused in this work.

103

In detail, batch experimental runs were performed at three different temperatures for two

104

values of the reaction time in order to investigate the effect of such process variables on both the

105

quality of torrefied wood pellets (i.e., elemental composition, bulk and energy densities, sizes and

106

shape, hydrophobic behavior, mechanical durability and fine particles content) and process 4 ACS Paragon Plus Environment

Page 5 of 37 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

Energy & Fuels

107

performances (i.e., mass and energy yields). To the best of the Authors’ knowledge, the torrefaction

108

treatment of wood pellets in fluidized bed reactor has not been investigated so far. Therefore,

109

findings of this work can also useful to highlight potential advantages and drawbacks related to use

110

of the fluidization technology in this specific application.

111 112

2. EXPERIMENTAL SECTION

113 114 115

2.1 Materials sampling and chemical characterization Commercial fir pellets (6 mm diameter and 3.15-4 mm length) were used in this work as biomass

116

feedstock. Ticino sand in size range of 150-300 µm (Particle density = 2651 kg/m3; packed bulk

117

density = 1537 kg/m3; Sauter mean diameter = 210 µm; minimum fluidization velocity = 4.14·10-2

118

m/s at 25 °C) was used for the operation of the fluidized bed reactor.

119

The determination of moisture content (M), volatile matter (VM), fixed carbon (FC), and ashes

120

(ASH) in raw and torrefied biomass samples was performed by using a TGA 701 LECO

121

thermogravimetric analyzer by following the ASTM D5142. The elemental composition (CHN) of

122

samples was determined by using a CHN 2000 LECO analyzer according to the ASTM D5373

123

standard method. The oxygen content was estimated by subtracting the sum of the percentages (dry

124

basis) of C, H, N and ash from 100%. All the analyses were performed in triplicate at least.

125

Liquid products were analyzed by gas chromatography coupled to mass spectrometry (GC‐MS

126

Agilent HP6890/HP5975, equipped with a capillary column DB-35MS, 30 m ×0.25 mm ID) using

127

helium as carrier gas.

128 129

pH measurements of liquid products were measured by using a Thomas Scientific 675 pH/ISE meter at ambient temperature.

130

The analysis of permanent gases evolved during the torrefaction treatment of wood pellets was

131

performed in real time. In particular, the Testo 350 advanced portable emission analyzer was used

132

for the real-time monitoring of evolved carbon monoxide (CO) and carbon dioxide (CO2), which

133

are the main components of permanent gases evolved during torrefaction processes11. 5 ACS Paragon Plus Environment

Energy & Fuels 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 6 of 37

134 135 136

2.2. Calorific values, mass density and energy density of raw and torrefied pellets The higher heating value (HHV, MJ/kg, dry basis) was measured by using a Parr 6200 Isoperibol

137

Calorimeter. The conversion of higher to lower heating values (LHV, MJ/kg, dry basis) was

138

performed according to the following Eq. (1):

139 140

LHV (MJ/kg) = HHV (MJ/kg) - 2.442(8.936·H/100)

(1)

141 142

where the constant 2.442 is the enthalpy difference in MJ/kg between the gaseous and liquid

143

water at 25 °C, the constant 8.936 is the ratio between the molar masses of water (H2O) and

144

hydrogen (H2)15 and H is the weight fraction (%) of hydrogen in the samples on a dry basis, All the

145

analyses were performed in duplicate at least.

146

The apparent density of wood pellets was calculated as the ratio of the pellet mass to its volume.

147

In particular, in this work, the ends of the tested wood pellets were flattened by using sandpaper in

148

order to make them perfect cylinders. The length (L) and diameter (D) of cylinders were then

149

measured by using a Vernier caliper (Mod. Metrica, 0.05mm and 1/128 inch Accuracy; 0-150 mm

150

and 0-6 inch measuring Ranges), which were used to calculate the volume of pellets and hence their

151

apparent density. The bulk density was calculated as the ratio of the mass to the volume (including

152

the contribution of the interparticle void volume) of a batch of pellets, poured into a 25 ml

153

graduated cylinder. The packed bulk density was obtained by mechanically tapping the cylinder

154

containing the sample until no further volume change was observed. Density measurements were

155

performed after the samples were oven dried at 105 ± 5 °C for 24 h. The density values reported

156

later in Fig. 5A are the average of 5 measurements for each sample.

157

Bulk and apparent energy densities (GJ/m3, dry basis) of raw and torrefied pellets were

158

calculated as the ratio of the lower heating value (LHV) to the bulk and apparent density of the

159

pellets, respectively. 6 ACS Paragon Plus Environment

Page 7 of 37 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

Energy & Fuels

160 161 162

2.3 Hygroscopic behavior of raw and torrefied pellets Two types of tests were used to investigate the hygroscopic behavior of both raw and torrefied

163

wood pellets. The first one, coded below as exposure test, involved pellets that were placed in a

164

climate-controlled chamber at specified humidity and temperature conditions. The second one,

165

coded below as immersion test, involved the immersion of pellets in water. In both cases, pellets

166

were analyzed to determine the resulting moisture uptake after a prefixed time. In more details,

167

exposure tests were performed by using conventional static desiccator technique.16,17 Specifically, a

168

glass desiccator containing a supersaturated solution of either KBr or K2CO3 in water was used as a

169

humidity control chamber, which allowed exposing samples to 80 ±2 or 43 ±1 % RH at room

170

temperature (i.e., 25 ± 2 °C), respectively. In order to minimize daily temperature changes the

171

desiccator was partially submerged in a water bath. Approximately 5 g of oven-dried pellets were

172

put into an open weighing bottle and then placed into the desiccator for testing. The humidity and

173

temperature in the glass desiccator were checked by using a digital thermo-hygrometer (30.5005

174

TFA Dostmann). The moisture content (MC, %wt) of pellets samples after an exposure time of 7

175

days was measured by using a Kern DBS Halogen Moisture analyzer.

176

Immersion tests were performed putting pellets of similar size into a 50 ml volumetric vials filled

177

with about 30 mL of distilled water. At the end of the test, the solid samples were separated from

178

water by using a sieve with 1 mm diameter and letting the sample drip water for about 30 minutes.

179

After the dripping time, the moisture content (MC, %wt.) of wet pellets was measured by using the

180

above mentioned Kern DBS analyzer.

181 182

In both cases, the water uptake (mass gained), Wg (%wt.), with respect the initial mass of dry samples, was calculated by using the following formula:

183 184



 = ( ) · 100

(2)

7 ACS Paragon Plus Environment

Energy & Fuels 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 37

185 186

2.4 Hardness and durability of raw and torrefied pellets

187

A non-standard method was purposely set up in this work in order to compare at a small scale

188

the quality of raw and torrefied wood pellets in terms of their durability, which represents the ability

189

to hand pellets without experiencing unacceptable breakage or generating a significant amount of

190

fines. This newly developed method allows for the determination of a pellet durability index

191

(PDI)18, which is defined as the weight fraction (%) of whole pellets remaining intact in a sample

192

tumbled according the method described below.

193

About 5 g of dust-free pellets are placed on a 1 mm mesh sieve. The sieve is tightly covered at

194

the top to prevent the loss of the material sample during sifting or mechanical agitation. A pan is

195

also used to collect fines passing through the sieve. Four nitrile butadiene rubber (NBR) balls (30

196

mm diameters, 70 Shore A hardness) and 7 stainless steel ball bearings (4.5 mm diameter, grade

197

100 AISI 316 ) are added to the sieve to make the test more stressing and somehow simulating the

198

bulk handling during actual manufacturing and delivery steps. The durability test set (i.e., sieve, pan

199

and cover) is placed on an orbital shaker moving at 210 rpm for 30 min. Afterward, fines are

200

removed from the tumbled samples by screening and the remaining pellets are weighed against their

201

initial mass to calculate the PDI as follows:

202 203

    !"#$

 (%. ) =    "

 !"#$

∙ 100

(3)

204 205

Analysis of pellets hardness was carried out by using a Shore D durometer (Sauter HBD 100-0)

206

in accordance with ASTM D2240. This test method is based on the penetration of a specific type of

207

indentor when forced into the material under specified conditions. The indentation hardness is

208

inversely related to the penetration and is dependent on the elastic modulus and viscoelastic

209

behavior of the material. Shore durometer hardness measurements were repeated at least 10 times 8 ACS Paragon Plus Environment

Page 9 of 37 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

Energy & Fuels

210

for each sample by using both different pellets and different points on each pellet. The average

211

values and the standard deviation of measurements were accurately taken into account (see Fig. 8

212

below).

213 214

2.5. Apparatus and procedures for pellets torrefaction

215

A schematic representation of the laboratory-scale fluidized-bed apparatus is shown in Figure 1.

216

The torrefaction unit of the apparatus consists of a steel tubular column (38 mm inner diameter and

217

350 mm height) surrounded by an electrical heating tape (FGR-100/240V V-ROPE HEATER

218

500W by Omegalux). The thermal insulation of the reactor is ensured by a mineral wool heat-

219

insulating cylinder. The temperature of the reactor is regulated by a proportional integral derivative

220

(PID) controller (Gefran 600 PID), which reads the bed temperature trough a K-type thermocouple

221

inserted in the reactor at 7 cm above the gas distributor. A disk-shaped ceramic fiber felt (6 mm

222

thick) set at the bottom of the column and topped by a 15 mm height bed of steel spheres

223

(4.5 mm diameter) acts as the gas distributor. A flowmeter (Asameter Model E, by ASA) with a

224

100-1000 NL/h flow range supplies nitrogen as a fluidization gas during the torrefaction tests. A

225

glass tubular trap followed by a water-cooled Liebig condenser and an air-cooled Liebig condenser

226

are used as gas-liquid separators in order to recover the condensable fraction of the torgas evolved

227

during the torrefaction treatment of wood pellets.

228

Batch torrefaction tests were performed at three different temperatures (i.e., 200, 230 and 250

229

°C) and by fixing the reaction time equal to either 7 or 15 min. Torrefaction tests and the resulting

230

solid products will be coded below as “TWP-T-t” where TWP stands for torrefied wood pellet

231

(TWP), T indicates the torrefaction temperature and t the reaction time: for example,TWP-200-15

232

denotes the torrefied pellets obtained from a test performed at 200 °C by fixing the reaction time at

233

15 min. In a typical experimental run, about 20 g of raw wood pellets (RWP) having a moisture

234

content of approximately 6-7 % wt. was charged into the reactor. The aspect ratio of the granular

235

solid bed, which is defined as the ratio of height to diameter of the bed, was set as high as 4. At 9 ACS Paragon Plus Environment

Energy & Fuels 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 10 of 37

236

startup, the bed of granular solid was heated to the selected torrefaction temperature while using air

237

as a fluidizing gas. When the fluidized bed reached its prefixed steady-state temperature, the

238

fluidizing gas was switched to nitrogen and the biomass was dropped from the top into the reactor.

239

During torrefaction tests, the nitrogen flow rate was set equal to 200 Nl/h corresponding to a

240

fluidization velocity in the range from 7.8·10-2 to 9.7·10-2 m/s at the operating temperatures. This

241

choice, which was suggested by preliminary trial and error fluidization tests, allowed achieving a

242

good mixing of biomass and inert particles in the bed during torrefaction tests while preventing

243

their elutriation and segregation. The resulting gas−solid contact time, calculated as the ratio of the

244

static bed height to the superficial gas velocity, slightly increased with the rise in the torrefaction

245

temperature. Once the prefixed test time was passed, the bed was cooled down to less than 100 °C

246

as fast as possible (approximately 2-3 min) by turning the electrical heater off, removing the reactor

247

from the insulating cylinder and blowing cold compressed air onto the reactor surface. Finally, the

248

solid and liquid products were recovered and weighted. In more details, the condensable volatiles

249

evolved during torrefaction tests were quantitatively recovered from tail end of the tube reactor and

250

the three adopted condensation traps (see Fig. 1) by washing these latter with acetone. Afterward,

251

they were collected as trap-1 liquids (T1-L), when recovered from the tail end of the reactor

252

column, trap2 liquids (T2-L), when recovered from the glass tubular trap, trap-3 liquids (T3-L),

253

when recovered from the water-cooled Liebig condenser and trap-4 liquids (T4-L), when recovered

254

from the air-cooled Liebig condenser. The solid torrefied product was separated from the granular

255

solid by manual sieving. Mass yields of solid (MYS), liquid (MYL), and gaseous (MYG) products

256

were evaluated on an as-received basis (ar) through the following Eqs. 4−6. It is worth noting that

257

the mass of condensable compounds used in Eq. 2 does not include the mass of the T1-L fraction

258

recovered from the tail end of the reactor column, which was of difficult determination. The energy

259

densification index (IED) and the energy yield (EYS) of torrefied solids were also evaluated on a dry

260

basis by means of Eqs. 7 and 8.

261 10 ACS Paragon Plus Environment

Page 11 of 37 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

Energy & Fuels

-.//01203 4.523

262

&'( (%, *+) = ,

263

&': (%, *+) = ,

264

&'? (%, *+) = ,

265

DE (−, GH) =

/67 7..3 80550-4

9 ∙ 100

;.