Biomass Gasification and Hot Gas Upgrading in a Decoupled Dual

Jun 29, 2017 - A decoupled dual-loop gasifier (DDLG) has been developed in which biomass gasification and hot gas upgrading are separated into two ...
0 downloads 0 Views 2MB Size
Subscriber access provided by Olson Library | Northern Michigan University

Article

Biomass Gasification and Hot Gas Upgrading in a Decoupled Dual Loop Gasifier Guangyong Wang, Shaoping Xu, Chao Wang, and Junjie Zhang Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b00782 • Publication Date (Web): 29 Jun 2017 Downloaded from http://pubs.acs.org on July 2, 2017

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

Energy & Fuels 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 41

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

Biomass Gasification and Hot Gas Upgrading in a Decoupled Dual Loop Gasifier

2

Guangyong Wang, Shaoping Xu*, Chao Wang, Junjie Zhang

3

State Key Laboratory of Fine Chemicals, Institute of Coal Chemical Engineering, School of

4

Chemical Engineering, Dalian University of Technology, No.2 Linggong Road, Dalian 116024,

5

China

6

*Corresponding author, e-mail address: [email protected] (SP Xu)

7 8

Abstract: A decoupled dual loop gasifier (DDLG) has been developed in which biomass

9

gasification and hot gas upgrading are separated into two parallel loops and so that could be

10

optimized individually. In the gasification loop, the gasifier is so designed that the contact between

11

volatiles and char is restrained and therefore the steam gasification of char is enhanced. In the

12

upgrading loop, both desulfurizer and tar reforming catalyst are used for desulfurization and tar

13

destruction, respectively. As in-bed desulfurizer, an iron-bearing olivine supported ZnO

14

(Zn/olivine) was prepared and tested in a fixed bed reactor. H2S sorption over ZnO, adversely

15

affected by H2O, was accompanied by evident reduction of ZnO and vaporization of Zn at 550 oC.

16

By contrast, no obvious ZnO reduction was observed at the same condition over Zn/olivine. The

17

reduction-resistance of Zn/olivine was illustrated by temperature programmed reduction and

18

powder X-ray diffraction. In DDLG test with pine sawdust as feedstock and Zn/olivine+Ni/olivine

19

as upgrading bed materials, a synergy was found between desulfurization and tar destruction. The

20

H2O-involved reactions such as steam gasification of char and steam reforming of

21

tar/hydrocarbons were intensified at elivated gasification temperature and in presence of

22

Ni/olivine. As a result, the decrease of H2O favored H2S sorption by Zn/olivine, which in turn

23

alleviated sulfur-poisoning of Ni/olivine. Under the gasifier temperature of 850 oC, steam to

24

biomass mass ratio (S/B) of 0.3 and upgrading reactor temperature of 600 oC, H2O and tar

25

contents were effectively decreased to 8.8% and 1.5 g/Nm3, respectively. In the 2 h test, during

26

which Zn/olivine experienced about 4 cycles of sulfidation/regeneration, H2S in product gas was

27

lowered to 1.7 ppmv.

28

Keywords: Biomass gasification; Desulfurization; H2O conversion; ZnO; Olivine

29 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

30

1. Introduction

31

Synthetic natural gas (SNG) production from biomass with the advantages of sustainability and

32

CO2 neutrality drew a lot of interest recently1-3. N2-free H2-rich syngas with 10-15 vol% CH4

33

content produced by biomass gasification in dual bed gasification system could be a suitable

34

feeding gas for bio-SNG production4-7. However, there are generally trace amounts of

35

sulfur-containing compounds and high molecular weight organic compounds (tar), and a large

36

amount of steam in the raw biogenous syngas. For example, in the raw gases from biomass

37

gasification in fluidized beds up to a few hundred ppmv of sulfur (mainly as H2S) and an average

38

tar loading of about 10 g/Nm3 can be found8-12. These major impurities could deactivate the

39

downstream methanation catalyst, particularly the Ni-based catalysts (by poisoning and coking),

40

so they must be removed by gas-cleaning inside or downstream the gasifier.

41

Low temperature scrubbing downstream the gasifier is commonly adopted to remove tar,

42

moisture and sulfur compounds13, 14, in which the syngas is cooled, cleaned at relatively low

43

temperature and reheated before being sent to methanation reactor. Compared to the

44

gasification-scrubbing-methanation process, an integrated gasification-methanation process with

45

hot gas-cleaning is highly expected to significantly improve the efficiency of the SNG production

46

by avoiding cooling and reheating of the syngas10,

47

Specifically, sulfur removal could be achieved by high temperature sorption and tar could be

48

eliminated by reforming/cracking with the steam in the raw gas.

15

and the subsequent wastewater stream.

49

Incompatibility among the involved reactions is the main obstacle of the integrated

50

gasification-methanation process. The biomass/char gasification is subject to the heat and mass

51

transfer intensity between fuel particle and its surrounding heat carrier and gasification agent16, so

52

that fine particles of the fuel and solid heat carrier and high velocity of the gasification agent are

53

preferred. The raw product gas upgrading, on the other hand, desires ordinarily enough residence

54

time of the raw gas on the catalyst or sorbent surface17, and consequently a low gas velocity is

55

needed. In addition to their flow regime difference, there are remarkable difference in the

56

operating temperature range between biomass/char gasification and raw gas upgrading.

57

Biomass/char steam gasification with acceptable rate happens at about 800 oC or higher18.

58

However, sulfur compounds elimination to sufficient low levels is difficult at such high 2

ACS Paragon Plus Environment

Page 2 of 41

Page 3 of 41

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

59

temperature8, 19. The exothermic water gas shift reaction (WGS) is inhibited in this condition as

60

well.

61

Decoupling raw gas upgrading from biomass/char steam gasification seems to be necessary.

62

Biomass steam gasification on dual fluidized bed gasifier was a preferential choice for bio-SNG

63

production13, 14, during which gasification and combustion were decoupled into two reactors. The

64

endothermic biomass steam gasification was balanced by the exothermic char air combustion with

65

the help of circulating bed material, avoiding syngas consuming by the introduced O2 and the

66

product gas dilution by the CO2 generated in combustion and the N2 introduced with air5, 20.

67

However, a series of reactions, such as biomass pyrolysis, char gasification, tar steam

68

reforming/cracking and WGS, were still intertwined in the gasifier. As a result, insufficient tar

69

elimination even with in-situ cracking/reforming catalyst21, 22, suppressed char steam gasification

70

by the volatiles-char interactions23, 24, limited WGS reaction at relatively high temperature and

71

short residence time25, etc. were inevitable. Especially, more than 90% of the steam used to

72

fluidize the bed materials was not converted and the H2O content in raw gas reached up to

73

35-60%9, 12, 26, which decrease significantly the overall thermal efficiency of the process. More

74

importantly, the moisture was unfavorable for desulfurization over the commonly adopted

75

ZnO-based sorbents, especially at the relatively high temperature needed for biomass

76

gasification27,

77

pyrolysis/gasification has been tried by introducing a reforming zone in the gasifier or a

78

downstream reformer29-35. In this respect, a decoupled triple bed gasification (DTBG) system

79

characterized by separated pyrolysis/gasification, tar cracking/reforming and char combustion

80

reactions to benefit tar destruction, has been proposed in our previous studies6, 36, 37. In the DTBG,

81

however, the improvement on tar elimination was still unsatisfactory because that the gasification

82

and tar destruction were compromised in one loop and occurred over identical bed material.

83 84

28

.

To

optimize

tar

destruction,

decoupling

of

tar

destruction

from

On the other hand, hot biogenous syngas desulfurization with either in-situ or downstream desulfurizer has been paid little attention10, 15, 27.

85

As indicated by various works on desulfurization of the hot gas from coal gasification, sulfur

86

removal from the hot syngas could be achieved with reusable metal oxide, such as the oxides of

87

Zn, Fe, Ce, Ca, Cu, and Mn19, 38-40. ZnO showed good potential for H2S removal and has been 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 41

88

widely studied. However, it suffered from reduction with H2 and zinc volatilization over 550 °C8.

89

ZnO-based multi-metal oxides with improving activity and extending operation temperature were

90

then paid much attention41-43. Typically, zinc ferrite sorbents, i.e. with a crystallite spinel structure

91

formed by ZnO and Fe2O344, 45, could effectively reduce sulfur compounds to less than 1 ppmv at

92

450 oC46. H2S level to less than 5 ppmv over zinc ferrite in Lurgi gas at temperatures as high as

93

760 °C was achieved during multiple sulfidation-regeneration cycles47. Volatilization of the metal

94

zinc happens in a way similar to the vapor-phase transport of Ostwald ripening process48, which

95

could be suppressed by using the support with higher total activation energies (determined by the

96

intrinsic interaction between the metal and support) for both the metal adatoms and metal-reactant

97

complexes49. More importantly, the interaction between active component and the support could

98

help to increase the reduction temperature of the supported active component (ZnO) and thus

99

inhibit the formation of volatilizable metal (Zn)50-53. So, intensifying the active component-support

100

interaction could be another promising approach to improve the stability of ZnO54. In addition, the

101

active components could also be utilized more efficiently by increasing its dispersion in a

102

supported form and thus decreasing the diffusional path length of H2S. As it was reported that the

103

sulfidation reaction stopped before the whole spherical ZnO-pellets were reacted at temperatures

104

below 600 °C because of the grain boundary diffusion27, 55, 56. The support plays an important role

105

in maintaining structural strength of the desulfurizer as well46.

106

The relatively cheap and abrasion-resistant iron-bearing olivine, containing (MgxFe1-x)2SiO4 as

107

the main phase and small quantities of MgSiO3 and FeOx species, was widely used in dual bed

108

gasification of biomass as active bed material or catalyst support for in-bed tar reforming4, 57.

109

Thermal induced FeOx formation over calcined olivine was detected58-60. As the catalyst support,

110

supported components (Ni, Fe and Cu) were found to be bonded to or even included inside the

111

olivine structure during calcination, suppressing their reduction54,

112

Fe-bearing olivine could be a potential support for ZnO, providing suitable iron source for zinc

113

ferrite formation and possibly improving its stability as well.

57, 61

. Thus, the calcined

114

In present work, a decoupled dual loop gasifier (DDLG) has been developed to facilitate

115

biomass gasification and in-bed hot gas upgrading for qualified biogenous syngas production. By

116

introducing an additional upgrading loop parallel with the gasification-combustion loop, raw gas 4

ACS Paragon Plus Environment

Page 5 of 41

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

117

upgrading,

i.e.

desulfurization,

could

be

optimized

independently

under

suitable

118

reaction/regeneration temperature, gas-solid contact pattern and even the atmosphere. As in-bed

119

desulfurizer, an iron-bearing olivine supported ZnO (Zn/olivine) has been prepared and tested in a

120

fixed bed reactor. The abrasion-resistant and regenerable Zn/olivine was then adopted in DDLG in

121

addition to the optimization of its reaction conditions.

122

2. Experimental

123

2.1 Desulfurization of syngas in a fixed-bed reactor

124

Several metal oxides, i.e. CeO2, ZnO, calcined limestone (CaO) and calcined olivine (Fe2O3),

125

and an olivine supported ZnO were used as H2S sorbent. The CeO2 and ZnO from Sinopharm

126

chemical reagent Co., Ltd were used after dry-pressed into tablets which were then crushed and

127

sieved into 20-40 mesh range. The limestone (20-40 mesh) came from Changxin Zhejiang and was

128

calcined at 900 oC for 4.5 h. It contains 91.5 wt.% CaO. The olivine was received from the

129

Chinese city of Yichang60 and calcined at 800 oC for 4.5 h. The olivine supported ZnO

130

desulfurizer, i.e. Zn/olivine-750, has 6 wt.% zinc loading and was prepared by incipient wetness

131

impregnation of the 800 oC calcined olivine with an aqueous solution of zinc nitrate, followed by

132

calcination at 750 oC for 4.5 h. An olivine supported 6 wt.% nickel, i.e. Ni/olivine-350, was used

133

as the methanation catalyst. It was prepared by incipient wetness impregnation of the 600 oC

134

calcined olivine with an aqueous solution of nickel nitrate and calcination at 350 oC for 4.5 h60.

135

The desulfurization of syngas with the H2S sorbent and the downstream methanation on a

136

nickel-based catalyst were performed at 550 oC and atmospheric pressure in an electric heated

137

fixed-bed quartz reactor with 8 mm inner diameter. The desulfurization performance of the

138

H2S-sorbent was evaluated based on the catalytic methanation activity of the catalyst given that

139

the H2S passed through the sorbent could definitely poison the catalyst62. For metal oxide sorbents

140

comparison, 0.5 g sorbent was loaded in the up-layer combined with a feeding gas flow of 130

141

ml/min; for sulfidation-regeneration evaluation, 1.0 g of the Zn/olivine sorbent and a feeding gas

142

flow of 52 ml/min were adopted. During metal oxide sorbents comparison, relative lower sorbent

143

weight and higher gas flow rate were adopted in order to obtain clear distinction in sulfidation

144

over different sorbents within a short reaction time. The methanation catalyst in the down-layer

145

was 1.0 g for both cases. The feeding gas was a mixture of H2 and CO2 (H2/CO2 = 4) with 500 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 41

146

ppmv H2S, named as FG-1, or a mixture of H2 and CO (H2/CO = 1) with 500 ppmv of H2S, FG-2.

147

They both were provided by Dalian Special Gases Co., Ltd. The gas was controlled by a mass

148

flow controller to flow downward through the sequentially loaded H2S sorbent and methanation

149

catalyst. To illustrate the influence of H2O on desulfurization, H2O injected by a syringe pump and

150

preheated to 250 oC was added to the feeding gas before entering into the reactor. The influence of

151

heat and mass transfer limitations in the mathanation catalyst bed has been proved to be

152

negligible60. The concentrations of CO, CO2, CH4 and other gas components in the product gas

153

were monitored using a GC7890Ⅱ gas chromatograph equipped with both TCD and FID

154

detectors. The CO or CO2 conversion and CH4 selectivity were calculated based on the product

155

gas composition. CO2 conversion decreased to 90% of the initial value was used as the threshold

156

to determine the breakthrough of sorbent sulfidation. Sulfur capacity, in (mmol of H2S)/(g of

157

sorbent), was calculated based on sorbed H2S arose from the different feeding H2S until the

158

breakthrough in the presence and absence of the H2S sorbent62. CO out

159



CO conversion (%)=(1-

160



CO2 conversion (%)=(1- CO2 out )×100

CO in

) ×100

(1)



CO



(2)

2 in

161 162 163

CH4 yield

CH4 selectivity (%)= CO

reacted

CH4 yield

×100 or CO

×100







(3)

2 reacted

-Fig. 12.2 Biomass gasification and hot gas upgrading in DDLG

164

As shown in Fig. 2, DDLG consists of two circulation loops, i.e. a gasification loop and an

165

upgrading loop with fine and coarse particles as circulating bed materials, respectively. Biomass

166

gasification happens in the bubbling fluidized bed gasifier of the gasification loop. The hot syngas

167

from the gasifier is further upgraded, typically, desulfurization and tar steam reforming, in the

168

moving bed upgrading reactor of the upgrading loop, which also acts as an efficient particulates

169

filter so that a dust-free product gas could be obtained. Both loops have an identical fast fluidized

170

bed riser followed by a separator. The fine bed material, i.e. SiO2, with the residual char from the

171

gasifier and the coarse bed material, i.e. desulfurizer and tar reforming catalyst, from the

172

upgrading reactor are together transported by air through the riser into the separator. In the

173

separator, the fine and coarse bed material particles from the riser are separated from the flue gas 6

ACS Paragon Plus Environment

Page 7 of 41

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

174

and then fluidized by the air flow introduced into the regenerator. The coarse particle drops down

175

through the bottom tube of the separator into the regenerator and the fine particle outflows the

176

separator into the combustor based on their different masses and terminal velocities. In the

177

combustor, the char is burnt out with air and the fine bed material is heated and returned to the

178

gasifier to provide heat the endothermic gasification needed. In the regenerator, the used

179

sorbent/catalyst are refreshed with air and delivered back to the upgrading reactor. The quality and

180

quantity of fine and coarse circulating bed materials in both loops are monitored. In this way, the

181

intricate reaction network is decoupled and the gasification and the raw syngas upgrading could be

182

optimized independently.

183

-Fig. 2-

184

Schematic of the lab-scale DDLG facility is shown in Fig. 3. The gasifier is a bubbling fluidized

185

bed reactor with a down-zone of 56 mm i.d. and 80 mm height and an up-zone 98 mm i.d. and 190

186

mm heigh. Biomass was fed into the freeboard of the gasifier rather than into the bubbling

187

fluidized bed material to diminish the contact between volatiles and char and thus to decrease the

188

inhibition of volatiles to the steam gasification of char23, 24 and improve carbon conversion. The

189

upgrading reactor is a gas-solid countercurrent moving bed reactor with an i.d. of 123 mm and a

190

height 168 mm. The riser is a fast fluidized bed reactor with 20 mm i.d. and 2600 mm height. The

191

combustor and regenerator are moving bed reactors, with 80 mm i.d. / 140 mm height and 80 mm

192

i.d. / 190 mm height, respectively. All the reactors are made of 310S stainless steel and heated by

193

independent electrical furnaces to compensate heat loss. The circulating rate of fine bed material

194

and that of coarse bed material are controlled by the rotary valve between the gasifier and the riser

195

and that between the upgrading reactor and the riser, respectively.

196

-Fig. 3-

197

Silica sand (SiO2) with the particle size of 60-100 mesh (0.15-0.25 mm) was used as the fine

198

circulating bed material in the gasification loop. Silica sand of 20-40 mesh (0.38-0.83 mm) was

199

used as the coarse circulating bed material in the upgrading loop, incorporating desulfurizer and

200

tar reforming catalyst of the same particle size for hot raw gas upgrading. The olivine supported

201

with 6 wt.% zinc, i.e. Zn/olivine-850, and the olivine supported with 6 wt.% nickel, i.e.

202

Ni/olivine-850, were used as the desulfurizer and the tar reforming catalyst, respectively. The 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

203

Zn/olivine-850 was prepared by incipient wetness impregnation of the 800 oC calcined olivine

204

with an aqueous solution of zinc nitrate followed by calcination at 850 oC for 4.5 h. The

205

Ni/olivine-850 was prepared by incipient wetness impregnation of the 1000 oC calcined olivine

206

with an aqueous solution of nickel nitrate followed by calcination at 850 oC for 4.5 h. The biomass

207

feedstock used in this work was pine sawdust from Dalian City, Liaoning Province and its

208

proximate and ultimate analyses are presented in Table 1. The pine sawdust was crushed and

209

sieved to 20-40 mesh and dried for 4 h at 105-110 °C before test.

210

-Table 1-

211

Prior to test, 3.0 kg of the fine bed material and 4.5 kg coarse bed material were added into the

212

reaction system. The bed materials were bubbling fluidized and fast fluidized/transported with air

213

in the gasifier and riser, respectively. The circulating rate of fine bed material was maintained to

214

be 3.0 kg/h and that of coarse bed material 8.0 kg/h. All of the reactors were electrically heated to

215

the desired temperature and then fluidization air of the gasifier was replaced by the preheated

216

steam. When the temperatures of all reactors reached steady, gasification began by feeding pine

217

sawdust into the gasifier at a rate of 0.3 kg/h. The product gas was extracted from the upgrading

218

reactor with the help of a vacuum pump and cooled in four sequential glycol-cooled (-12 oC)

219

condensers, in which condensable components and particulate matter were separated from the

220

permanent gas. The effluent product gas underwent further aerosol removal and was measured by

221

a wet type gas flowmeter and analyzed every 10 minutes by a gas chromatography GC-7890Ⅱ

222

equipped with a TCD and a FID. The flue gas from the separator was cooled down and the

223

entrained dust was scrubbed by a venturi gas scrubber. After about 1 hour, when the reaction

224

system reached a steady state (Fig. 10), tar and H2S were separately sampled via two sampling

225

points before the condensers. The tar sampling was performed by condensing and dissolving the

226

tar components out of the product gas with six 250 ml impinger bottles basing on the protocol

227

CEN/TS 1543963-65. The impinger bottles were located in a cooling bath cooled down to -12 °C by

228

a cryostat, with bottles 1-5 filled with toluene and bottle 1, bottle 5 and bottle 6 containing glass

229

beads. The liquid phases in the impinger bottles were unified after sampling and the aqueous

230

phase was separated from the toluene phase. The amount of water was then determined to

231

calculate H2O content in the product gas and water conversion. The tar was obtained by 8

ACS Paragon Plus Environment

Page 8 of 41

Page 9 of 41

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

232

evaporating off the toluene solvent at 50 oC under reduced pressure. The H2S was sampled by

233

separating the condensable components from the permanent gas in the impinger bottle filled with

234

glass beads, cooled down to -12 oC. H2S in the effluent permanent gas was then measured by a

235

GC9790 equipped with a packed column (GDX-303, 2 m× 3 mm) and a FPD, with the accuracy of

236

0.1 ppmv. All the experiments were kept for at least 2 h. The general operation conditions are

237

summarized in Table 2.

238 239 240

-Table 2To evaluate the performance of the process, the parameters are defined by the following equations: Mass of carbon in product gas (kg)

241

Carbon conversion (%)= Mass of carbon in biomass fed into the system (kg) ×100 (4)

242

Dry gas yield (Nm3 /kgfuel, daf)= Mass of biomass of dry ash-free basis fed into the system (kg) (5)

243 244

Volume of dry product gas (Nm3 )

Tar content of product gas (g/Nm3 )=

Mass of tar in product gas (g) Volume of dry product gas (Nm3)



Mass of H2O in product gas (kg)

Water conversion (%)=(1- Mass of H O fed as gasifying agent (kg) )×100



(6) (7)

2

245 246

Cold gas efficiency (%)=

Lower heating value of product gas (MJ/Nm3 )×Gas yield (Nm3/kg) Lower heating value of biomass fed into the system (MJ/kg)

×100 (8)

2.3 Characterization of desulfurizer and catalyst

247

The desulfurizer was characterized by powder X-ray diffraction (XRD) and temperature

248

programmed reduction (TPR). The XRD was carried out on a Rigaku D/Max-2400 diffractometer

249

with a nickel-filtered Cu Kα radiation (0.15406 nm) and with a scanning rate of 4o per minute in

250

the 2θ range of 25-65o. The TPR was conducted on a Quantachrome ChemBET 3000

251

chemisorption analyzer with a TCD detector. Prior to a TPR measurement, the as-prepared sample

252

(0.2 g) was pretreated at 250 oC in Ar for 2 h, while 5 v/v% H2/Ar with the flow rate of 40 mL/min

253

was used during the TPR experiment.

254

3. Results and discussion

255

3.1 Desulfurization of syngas in fixed-bed reactor

256

Desulfurization of syngas FG-1 and FG-2 in a fixed-bed reactor was performed over different

257

metal oxide sorbents (ZnO, calcined limestone (CaO), CeO2 and calcined olivine (Fe2O3)), with

258

Ni/olivine-350 as the methanation catalyst. The CO/CO2 conversion and CH4 selectivity were 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

259

determined after 0.5 h and are shown in Fig. 4. All the four sorbents exhibited distinct H2S

260

removal ability in FG-2 and postponed deactivation of the downstream Ni/olivine-350 by

261

sulfur-poisoning, compared with the Ni/olivine-350 exposed in FG-2 directly. In FG-1, however,

262

only ZnO showed relatively pronounced desulfurization performance and rapid deactivation of

263

Ni/olivine-350 by H2S happened with calcined limestone, CeO2 or calcined olivine as the H2S

264

sorbents.

265

-Fig. 4-

266

It has been suggested that H2O, as a product of the sulfidation reaction over metal oxides

267

sorbents, could limit H2S reaction with MeOx by two parallel routes: the shift of thermodynamic

268

equilibrium which promotes the hydrolysis of MeSx and the competitive strong adsorption of H2O

269

which blocks the diffusion path of H2S as well as the active sites of sulfidation27, 28, 66. The

270

potential H2O production by reverse water gas shift (RWGS) reaction in FG-1, as indicated in Fig.

271

5, could adversely affect sulfur removal. In FG-2, instead, much less H2O formed over the

272

sorbents (mainly by the sulfidation reaction itself), exerts limited influence on the desulfurization

273

performance. The superior desulfurization performance of ZnO compared with that of calcined

274

limestone, CeO2 and calcined olivine, especially in FG-1 atmosphere, indicates that ZnO could be

275

more suitable for deep sulfur removal from H2O-containing syngas.

276

-Fig. 5-

277

The influence of H2O content on desulfurization over 0.5 g ZnO at 550 oC was further studied

278

with FG-1 and 0-30% more H2O as feeding gas. The H2O feeding was stopped at 90 minutes. As

279

shown in Fig. 6, in the initial 90 minutes, CO2 conversion and CH4 selectivity decrease with the

280

increase of H2O content in the feeding gas. After that, the methanation performance of the catalyst

281

experienced 10% H2O feeding could be recovered, but those experienced more H2O feedings

282

could not. The more the steam feeding, the worse the recovered catalytic performance of the

283

catalyst. The results indicate that deep sulfur removal from feedgas containing up to 10% H2O

284

could be achieved over ZnO, while H2O contents as high as 30% could remarkably deteriorate the

285

desulfurization performance at 550 oC and result in a rapid deactivation of the downstream

286

Ni/olivine-350. Meanwhile, some white powder material was observed to deposit on the

287

downstream inside wall of reactor tube after the experiments, suggesting that ZnO was partially 10

ACS Paragon Plus Environment

Page 10 of 41

Page 11 of 41

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

288

reduced to metallic zinc by H2 and the produced Zn underwent vaporization and condensation in

289

the reactor67.

290

-Fig. 6-

291

As mentioned above, ZnO showed relatively good potential for H2S removal, nevertheless, it

292

worked unfavorably in H2O-rich atmosphere and was prone to reduction and loss in hot syngas as

293

well. Such defects limit its application for deep sulfur removal from the conventional hot

294

biogenous syngas with H2O contents upto 35-60%, especially as the in-bed desulfurizer.

295

3.2 Characterization of Zn/olivine and its desulfurization performance in fixed-bed reactor

296

To suppress ZnO reduction, it was supported on an iron-bearing olivine calcined at 800 °C for

297

4.5 h. The dehydration and extraction of iron oxides from the olivine structure onto surface during

298

calcination58 develops the porous structure of the olivine. Specifically for the olivine used here,

299

800 °C was shown to be the optimal calcination temperature to improve the iron oxides

300

formation60. The thermal induced FeOx shells could improve the dispersion of the supported ZnO

301

and provide opportunities for the interactions between iron oxides and ZnO as well. The XRD spectra of the 800 oC calcined olivine supported ZnO after further calcination at 750

302 303

o

304

Besides the main olivine phase58 maintained, the new peaks appeared at 31.7° and 34.4° (2θ)

305

could be attributed to the main reflection of ZnO45. The relative intensity of the peaks at 29.8° and

306

35.6° (2θ) increase a lot, indicating the formation of ZnFe2O4 whose main diffraction peaks are

307

somewhat overlapped with that of the olivine phase45, 68.

308

C for 4.5 h (Zn/olivine-750) compared with that of the calcined olivine is shown in Fig. 7.

-Fig. 7-

309

The TPR profiles of the calcined olivine and the Zn/olivine-750 are presented in Fig. 8. The

310

calcined olivine exhibits a broad reduction peak above 350 °C, which is mainly attributed to the

311

reduction of the free iron oxides associated with the olivine phase. It corresponds to about 3 wt.%

312

Fe0 in the reduced olivine calculated from the hydrogen consumption during TPR experiments, as

313

1 mole H2 could reduce 1/3 mole Fe2O3 which is the main form of reducible iron in the calcined

314

olivine60. A similar reduction peak starting at 350 °C and ending at 850 °C appears for the

315

Zn/olivine-750, with an additional reduction peak at higher temperature. The new peak (>850 °C)

316

is in good accordance with the reported ones during reduction and modification of ZnFe2O469. 11

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

317

Notably, the potential reduction peak of ZnO is not so apparent67, 70. The slight decrease of H2

318

consumption below 850 oC of the Zn/olivine-750, compared with that of the calcined olivine,

319

could attribute to the strong interaction between ZnO and the support and the formation of

320

reduction-resistant structure (ZnFe2O4, as mentioned above), which may suppress the reduction of

321

the supported ZnO47, 54, 57, 61.

322

-Fig. 8-

323

Sulfidation and regeneration behaviors of the Zn/olivine-750 were evaluated in FG-1 at 550 oC

324

with the spent sorbent regenerated in air flow at 750 or 850 oC (Fig. 9). Meanwhile, sulfur

325

capacity of the Zn/olivine-750 was determined. The breakthrough point for the fresh

326

Zn/olivine-750 was 2.25 h, corresponding to a H2S capacity of 0.16 mmol/gZn/olivine. After

327

regeneration at 750 oC for 1 h, the sulfidation time till breakthrough of the refreshed

328

Zn/olivine-750 extended to 3.25 h, indicating that the sulfur removal ability was maintained and

329

even developed. To investigate the effect of regeneration condition, the second and the third

330

regeneration were conducted at 750 oC for 0.5 h and 850 oC for 0.5 h, respectively. As shown in

331

Fig. 9, after the second regeneration, the working sulfidation time was shorted to 2.75 h, indicating

332

that the regeneration at 750 oC for 0.5 h was insufficient. After the third regeneration, possibly

333

because of the improved regeneration rate at elevated temperature, the sulfidation time was

334

extended to 4 h, corresponding to a H2S capacity of 0.28 mmol/gZn/olivine. Obviously, the

335

Zn/olivine-750, underwent repeated sulfidation-regeneration cycles, was well recovered after

336

regeneration in air at 850 oC for 0.5 h.

337

Despite of the varied regeneration conditions, the reactivity of the Zn/olivine-750 appears to be

338

improved with the sulfidation-regeneration cycling, suggesting that the sulfidation of this sorbent

339

can be improved by activation treatments38. The enhanced desulfurization performance of the

340

refreshed Zn/olivine could mainly result from the thermal induced structural modification during

341

regeneration, i.e. iron thermal transformation in the olivine grain60 and possibly an improved

342

interaction between ZnO and olivine. Furthermore, the white material evolved during the

343

sulfidation of the ZnO pellets at 550 oC was not observed after desulfurization over the

344

Zn/olivine-750, implying an improved reduction-resistance of the supported ZnO as indicated by

345

the TPR test. 12

ACS Paragon Plus Environment

Page 12 of 41

Page 13 of 41

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

346 347

-Fig. 93.3 Biomass gasification and hot gas upgrading in DDLG

348

Low H2O conversion could greatly decrease thermal efficiency of the overall gasification

349

process and increase the cost, remaining a weakness of biomass steam gasification in dual bed

350

gasifier12. Furthermore, high H2O content is detrimental to the downstream utilization of the

351

as-produced hot gas, typically, deteriorating the hot gas desulfurization behavior as indicated

352

above. Therefore, less H2O feeding for bed material fluidization and intensifying H2O conversion

353

were attempted in DDLG to decrease the H2O content in the biogenous syngas.

354

The pressure drop across the gasifier at different H2O feeding rates, as one of the most

355

important gas-solid flow features, could help to determine the flow regime71. Variation of pressure

356

drop with increasing H2O feeding at 800 oC indicates that the bed material (SiO2 of 60-100 mesh)

357

starts to fluidize with H2O feeding rate of about 60 g/h. H2O feeding rates with the interval of 90

358

g/h to 300 g/h, corresponding to S/B of 0.3-1.0, were used in this work. Preliminary tests have

359

proved that the segregation of char in the gasifier was not apparent in this interval. The relatively

360

low superficial gas velocity in the gasifier, approaching to the minimum fluidization velocity of

361

the bed material, benefits the contact between gasification agent and biomass/char, as the mean

362

bubble diameter in the fluidized bed increases with the superficial gas velocity which would

363

deteriorate gas-solid contact25. In addition, excessive unreacted H2O effluent can be alleviated, as

364

it was reported that H2O conversions in the gasifier were generally maintained to be less than 10%

365

with S/B of 0.4-2.6 and the feeding H2O was significantly beyond the consumed12. In a typical

366

gasification test, the pressure drop and the product gas compositions as a function of time on

367

stream were shown in Fig. 10. The stably pulsing pressure drop indicates that the bed material was

368

sufficiently fluidized and bubbled vigorously. The gas composition reached a steady state after

369

about 1 hour.

370

-Fig. 10-

371

The optimization of biomass gasification over DDLG was then performed at varied S/B and

372

gasifier temperature. In these tests, SiO2 was used as bed materials in both loops and the

373

upgrading reactor temperature was maintained to be 350 ºC to avoid volatiles cracking and

374

condensation. 13

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

375

Effect of S/B on biomass steam gasification is shown in Fig. 11. Carbon conversion and gas

376

yield increase with the S/B rising up to 0.7 and then decrease. Tar content decreases slightly and

377

product gas composition remains almost constant with the increase of S/B. The increasing steam

378

addition could promote char steam gasification at low S/B, which is responsible for the

379

improvement of both carbon conversion and gas yield. Increasing S/B to 1.0 from 0.7, however,

380

deteriorated the gasification performances, due to the unfavorable steam-char contact in the

381

gasifier at higher superficial gas velocity and heat removal by the increasing H2O effluent. H2O

382

conversion decreases monotonously with the increase of S/B, in accordance with the published

383

work72. H2O conversion remains below 18%, suggesting that the feeding steam was still evidently

384

exceeded the needed even with S/B as low as about 0.4, i.e. the lower limit of the recommended

385

S/B12.

386

-Fig. 11-

387

The gasifier temperature could greatly affect the biomass steam gasification, as indicated in Fig.

388

12. Carbon conversion, H2O conversion and gas yield increase with raising gasifier temperature,

389

while tar content shows reverse tendency, contributed by the enhanced char steam gasification and

390

steam reforming/cracking of tar at higher temperature. H2 concentration increases with the

391

increase of temperature and reaches to 36.8% at 800 oC, and then remains more or less unchanged.

392

Lower heating value (LHV) of the product gas increases slightly. From the point of view of

393

improving H2O conversion as well as product gas yield, higher gasifier temperature is preferred.

394

Nevertheless, it is very difficult to work in the gasifier above 900°C for biomass gasification with

395

pure steam12.

396

-Fig. 12-

397

Notably, the tar contents in the raw biogenous syngas from the gasifier of DDLG are of 100

398

g/Nm3 magnitude, remarkably exceeded the reported values from fluidized bed gasifiers11, 12 and

399

close to those from the updraft (counterflow) gasifiers11. This could be due to the specific design

400

of the gasifier. As indicated in Fig. 3, biomass was fed into the freeboard rather than the

401

commonly way into the bubbling fluidized bed material. Biomass underwent rapid pyrolysis upon

402

touching the hot bed material in surface layer of the bubbling fluidized bed. The as-produced

403

pyrolysis volatiles outflowed with the effluent steam and syngas from char steam gasification, 14

ACS Paragon Plus Environment

Page 14 of 41

Page 15 of 41

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

404

rather than flowed into the deeper zone of the fluidized bed. The nascent char from biomass

405

pyrolysis, instead, sank into the fluidized bed and mixed with the bed material sufficiently

406

(confirmed by the insignificant segregation of char in bed material). Consequently, in-situ tar

407

elimination was suppressed by the short residence time of the volatiles and the limited contact

408

between volatiles and the hot bed material (char and SiO2), resulting in tar-rich product gas. In

409

contrast, steam gasification of the nascent char dispersed in the fluidized bed and the derived H2O

410

consumption benefited from the restrained volatiles and char interactions, as the volatiles were

411

suggested to inhibit steam gasification of char23, 24. As expected, carbon conversion reached to

412

81.6% at the gasifier temperature of 850 oC and S/B of 0.4, approaching to or even exceeding the

413

results of the ordinary dual fluidized bed gasifiers7. Meanwhile, H2O conversion of 20.6% was

414

achieved over DDLG, corresponding to 27.0% H2O content in the raw gas which is less than the

415

published data from biomass steam gasification in fluidized bed10, 12.

416

In the configuration of DDLG, the as-produced volatiles is delivered to and upgraded in the

417

upgrading reactor under optimized condition. The improvements of the steam involving reactions

418

during volatiles upgrading, such as tar/hydrocarbon steam reforming and WGS, could promote

419

H2O conversion further and thus facilitate desulfurization of the raw syngas.

420

Thus, Ni/olivine-850 was included in the upgrading loop, acting as tar reforming catalyst, along

421

with the Zn/olivine-850 desulfurizer, for in-bed hot gas upgrading. Meanwhile, both the spent

422

Zn/olivine-850 and the deactivated Ni/olivine-850 underwent regeneration in the regenerator at

423

850 oC, as the sulfurized Zn/olivine-850 could be well recovered at 750-850 oC in air flow and the

424

potential carbon and even sulfur deposits over Ni/olivine-850 were suggested to be removed at

425

about 850 oC14.

426

As shown in Table 3, tar in the raw syngas was partially reformed over SiO2+Ni/olivine at the

427

upgrading reactor temperature of 600 oC. The catalytic activity was not good enough, mainly due

428

to sulfur-poisoning of nickel. Meanwhile, H2O conversion decreases a lot compared with that over

429

the inert bed material, because Ni/olivine-850 as an oxygen carrier is prone to be reduced by H2 to

430

produce H2O. In comparison, with the introduction of Zn/olivine-850 additionally, tar and gaseous

431

C2+ hydrocarbons were mostly eliminated. CH4 decreases and H2/CO ratio increases distinctly.

432

63.8% H2O conversion and H2O content as low as 8.8% were achieved. The results confirm that 15

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

433

tar/hydrocarbon

434

SiO2+Ni/olivine+Zn/olivine and consequently H2O conversion was promoted as expected. In the 2

435

h test, during which Zn/olivine-850 experienced about 4 cycles of sulfidation/regeneration, the

436

H2S in product gas was lowered to 1.7 ppmv.

437 438

steam

reforming

and

WGS

were

Page 16 of 41

remarkably

intensified

over

-Table 3Conclusions

439

Biomass gasification and hot gas upgrading are decoupled in the DDLG system basing on their

440

different requirements on flow regime, operating temperature, etc. Both desulfurizer and tar

441

reforming catalyst are needed for desulfurization and tar destruction during in-bed hot gas

442

upgrading.

443

As in-bed desulfurizer, the Zn/olivine was prepared and tested in a fixed bed reactor. H2S

444

sorption over ZnO, adversely affected by H2O, was accompanied by evident reduction of ZnO and

445

vaporization of Zn at 550 oC in H2/CO2/H2O/H2S. By contrast, no obvious ZnO reduction was

446

observed at the same condition over Zn/olivine-750. The result confirmed the improved

447

reduction-resistance of the olivine supported ZnO as indicated by the TPR test, due to the

448

interaction between ZnO and the support and additionally the formation of ZnFe2O4 between the

449

supported ZnO and the thermal induced iron oxides on olivine. Meanwhile, H2S was efficiently

450

removed over Zn/olivine-750 at 550 °C and the spent Zn/olivine-750 could be well recovered at

451

750-850 °C during repeated sulfidation-regeneration cycles.

452

In DDLG test with pine sawdust as feedstock, optimization of the biomass gasification itself

453

showed that low S/B and high gasifier temperature benefited H2O conversion and thus diminished

454

the adverse effects of H2O on H2S sorption. With the introduction of Zn/olivine-850 and

455

Ni/olivine-850 as upgrading bed materials, a synergy was found between desulfurization and tar

456

destruction during hot gas upgrading. The H2O-involved reactions such as steam reforming of

457

tar/hydrocarbons and WGS were intensified in presence of Ni/olivine-850. As a result, the

458

decrease of H2O favored H2S sorption by Zn/olivine-850, which in turn alleviated sulfur-poisoning

459

of Ni/olivine-850. Under the gasifier temperature of 850 oC, S/B of 0.3 and upgrading reactor

460

temperature of 600 oC, H2O content and tar were effectively decreased to 8.8% and 1.5 g/Nm3,

461

respectively. In the 2 h test, during which Zn/olivine experienced about 4 cycles of 16

ACS Paragon Plus Environment

Page 17 of 41

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

462

sulfidation/regeneration, H2S in product gas was lowered to 1.7 ppmv.

463

Acknowledgments

464

This work is supported by the Natural Science Foundation of China (No. 50776013) and the

465

National High Technology Research and Development Program of China (No. 2008AA05Z407).

466

References

467

1.

468

catalytic hydrothermal gasification. Industrial & engineering chemistry research 2005, 44 (13),

469

4543-4551.

470

2.

471

wood-derived producer gas for the production of synthetic natural gas. Industrial & engineering

472

chemistry research 2010, 49 (15), 7034-7038.

473

3.

474

in gasification of biomass at intermediate temperature and pressure. II. Process performance

475

analysis. Energy & Fuels 2011, 25 (9), 4085-4094.

476

4.

477

Gasifier: Possible Interactions of Fuels. Energy & Fuels 2013, 27 (6), 3261-3273.

478

5.

479

chemical engineering fundamentals for fuel pyrolysis and gasification in dual fluidized bed.

480

Industrial & Engineering Chemistry Research 2013, 52 (19), 6283-6302.

481

6.

482

countercurrent moving bed gasifier. Fuel 2013, 112, 635-640.

483

7.

484

University, 2007. URL< http://www.chemeng.lth.se/exjobb/E450.pdf >.

485

8.

486

sulfide from biomass gasification gas. Catalysis Reviews 2007, 49 (4), 407-456.

487

9.

488

allothermal biomass gasifier milena // Proceedings of 15th European Conference on Biomass for

489

Energy Industry and Climate Protection. Florence: ETA, 2007.

490

URL.

Waldner, M. H.; Vogel, F. Renewable production of methane from woody biomass by

Seemann, M. C.; Schildhauer, T. J.; Biollaz, S. M. Fluidized bed methanation of

Nanou, P.; van Swaaij, W. P.; Kersten, S. R.; van Rossum, G. Evaluation of catalytic effects

Wilk, V.; Hofbauer, H. Co-gasification of plastics and biomass in a dual Fluidized-bed Steam

Zhang, J.; Wu, R.; Zhang, G.; Yao, C.; Zhang, Y.; Wang, Y.; Xu, G. Recent studies on

Zou, W.; Song, C.; Xu, S.; Lu, C.; Tursun, Y. Biomass gasification in an external circulating

Bengtsson, K. Twin-bed gasification concepts for Bio-SNG Production. Lund: Lund

Torres, W.; Pansare, S. S.; Goodwin Jr, J. G. Hot gas removal of tars, ammonia, and hydrogen

van der Meijden, C. M.; van der Drift, A.; Vreugdenhil, B. J. Experimental results from the

17

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

491

10. Kienberger, T.; Zuber, C.; Novosel, K.; Baumhakl, C.; Karl, J. Desulfurization and in situ tar

492

reduction within catalytic methanation of biogenous synthesis gas. Fuel 2013, 107, 102-112.

493

11. Milne, T. A.; Evans, R. J.; Abatzoglou, N. Biomass gasifier" tars": Their nature, formation,

494

and conversion. National Renewable Energy Laboratory Golden, CO, 1998, 570.

495

URL.

496

12. Corella, J.; Toledo, J. M.; Molina, G. Biomass gasification with pure steam in fluidised bed:

497

12 variables that affect the effectiveness of the biomass gasifier. International Journal of Oil, Gas

498

and Coal Technology 2008, 1 (1-2), 194-207.

499

13. van der Meijden, C. M. Development of the MILENA gasification technology for the

500

production of Bio-SNG. Technische Universieit Eindhoven 2010.

501

URL.

502

14. Seiffert, M.; Rönsch, S.; Schmersahl, R.; Zeymer, M.; Majer, S.; Pätz, C.; Kaltschmitt, M.;

503

Rauch, R.; Rehling, B.; Tremmel, H.; Ulrich, D.; Biollaz, S.; Schildhauer, T.

504

Bio-SNG-Demonstration of the production and utilization of synthetic natural gas (SNG) from

505

solid biofuels. Final project report. TREN/05/FP6EN/S07.56632/19895. 2009.

506

15. Rhyner, U.; Edinger, P.; Schildhauer, T. J.; Biollaz, S. M. Experimental study on high

507

temperature catalytic conversion of tars and organic sulfur compounds. International Journal of

508

Hydrogen Energy 2014, 39 (10), 4926-4937.

509

16. Di Blasi, C. Combustion and gasification rates of lignocellulosic chars. Progress in Energy &

510

Combustion Science 2009, 35 (2), 121-140.

511

17. Xu, G.; Murakami, T.; Suda, T.; Yoshiaki Matsuzawa, A.; Tani, H. The superior technical

512

choice for dual fluidized bed gasification. Industrial & Engineering Chemistry Research 2006, 45

513

(45), 2281-2286.

514

18. Corella, J.; Toledo, J. M.; Molina, G. A review on dual fluidized-bed biomass gasifiers.

515

Industrial & Engineering Chemistry Research 2007, 46 (21), 6831-6839.

516

19. Swisher, J.; Schwerdtfeger, K. Review of metals and binary oxides as sorbents for removing

517

sulfur from coal-derived gases. Journal of Materials Engineering and Performance 1992, 1 (3),

518

399-407.

519

20. Kern, S. J.; Pfeifer, C.; Hofbauer, H. Cogasification of polyethylene and lignite in a dual 18

ACS Paragon Plus Environment

Page 18 of 41

Page 19 of 41

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

520

fluidized bed gasifier. Industrial & Engineering Chemistry Research 2013, 52 (11), 4360-4371.

521

21. Aznar, M. P.; Caballero, M. A.; Gil, J.; Martín, J. A.; Corella, J. Commercial steam reforming

522

catalysts to improve biomass gasification with steam-oxygen mixtures. 2. Catalytic tar removal.

523

Industrial & engineering chemistry research 1998, 37 (7), 2668-2680.

524

22. Rapagna, S.; Jand, N.; Kiennemann, A.; Foscolo, P. Steam-gasification of biomass in a

525

fluidised-bed of olivine particles. Biomass and Bioenergy 2000, 19 (3), 187-197.

526

23. Bayarsaikhan, B.; Sonoyama, N.; Hosokai, S.; Shimada, T.; Hayashi, J. I.; Li, C. Z.; Chiba, T.

527

Inhibition of steam gasification of char by volatiles in a fluidized bed under continuous feeding of

528

a brown coal. Fuel 2006, 85 (3), 340-349.

529

24. Li, C. Z. Importance of volatile-char interactions during the pyrolysis and gasification of

530

low-rank fuels-a review. Fuel 2013, 112, 609-623.

531

25. Corella, J.; Aznar, M. P.; Delgado, J.; Aldea, E. Steam gasification of cellulosic wastes in a

532

fluidized bed with downstream vessels. Industrial & engineering chemistry research 1991, 30 (10),

533

2252-2262.

534

26. Pfeifer, C.; Rauch, R.; Hofbauer, H. In-bed catalytic tar reduction in a dual fluidized bed

535

biomass steam gasifier. Industrial & engineering chemistry research 2004, 43 (7), 1634-1640.

536

27. Zuber, C.; Husmann, M.; Schroettner, H.; Hochenauer, C.; Kienberger, T. Investigation of

537

sulfidation and regeneration of a ZnO-adsorbent used in a biomass tar removal process based on

538

catalytic steam reforming. Fuel 2015, 153, 143-153.

539

28. Kim, K.; Jeon, S.; Vo, C.; Park, C. S.; Norbeck, J. M. Removal of hydrogen sulfide from a

540

steam-hydrogasifier product gas by zinc oxide sorbent. Industrial & engineering chemistry

541

research 2007, 46 (18), 5848-5854.

542

29. Pinto, F.; Lopes, H.; Cabrita, I.; Rui, N. A.; Gulyurtlu, I. Effect of catalysts in the quality of

543

syngas and by-products obtained by co-gasification of coal and wastes. 2: Heavy metals, sulphur

544

and halogen compounds abatement. Fuel 2008, 87 (7), 1050-1062.

545

30. García, G.; Campos, E.; Fonts, I.; Sánchez, J. L.; Herguido, J. Gas catalytic upgrading in a

546

two-zone fluidized bed reactor coupled to a cogasification plant. Energy & Fuels 2013, 27 (5),

547

2835-2845.

548

31. Li, D.; Asadullah, M.; Zhang, S.; Wang, X. S.; Wu, H.; Li, C. Z. An advanced biomass 19

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

549

gasification technology with integrated catalytic hot gas cleaning: Part I. Technology and initial

550

experimental results in a lab-scale facility. Fuel 2013, 108 (11), 409–416.

551

32. Sutton, D.; Kelleher, B.; Doyle, A.; Ross, J. R. H. Investigation of nickel supported catalysts

552

for the upgrading of brown peat derived gasification products. Bioresource Technology 2001, 80

553

(2), 111-6.

554

33. Wang, J. B.; Xiao, B.; Liu, S. M.; Hu, Z. Q.; He, P. W.; Guo, D. B.; Hu, M.; Qi, F. J.; Luo, S.

555

Y. Catalytic steam gasification of pig compost for hydrogen-rich gas production in a fixed bed

556

reactor. Bioresource Technology 2013, 133, 127-133.

557

34. Li, D.; Tamura, M.; Nakagawa, Y.; Tomishige, K. Metal catalysts for steam reforming of tar

558

derived from the gasification of lignocellulosic biomass. Bioresource Technology 2015, 178 (2),

559

53-64.

560

35. Abu El-Rub, Z.; Bramer, E. A.; Brem, G. Review of catalysts for tar elimination in biomass

561

gasification processes. Industrial & Engineering Chemistry Research 2004, 43 (22), 6911-6919.

562

36. Wei, L.; Xu, S.; Liu, J.; Lu, C.; Liu, S.; Liu, C. A novel process of biomass gasification for

563

hydrogen-rich gas with solid heat carrier: preliminary experimental results. Energy & Fuels 2006,

564

20 (5), 2266-2273.

565

37. Tursun, Y.; Xu, S.; Wang, G.; Wang, C.; Xiao, Y. Tar formation during co-gasification of

566

biomass and coal under different gasification condition. Journal of Analytical & Applied Pyrolysis

567

2015, 111, 191-199.

568

38. Slimane, R. B.; Abbasian, J. Utilization of metal oxide-containing waste materials for hot

569

coal gas desulfurization. Fuel Processing Technology 2001, 70 (2), 97-113.

570

39. Zeng, Y.; Zhang, S.; Groves, F.; Harrison, D. High temperature gas desulfurization with

571

elemental sulfur production. Chemical Engineering Science 1999, 54 (15), 3007-3017.

572

40. Westmoreland, P. R.; Harrison, D. P. Evaluation of candidate solids for high-temperature

573

desulfurization of low-Btu gases. Environmental Science & Technology 1976, 10 (7), 659-661.

574

41. Focht, G.; Ranade, P.; Harrison, D. High-temperature desulfurization using zinc ferrite:

575

reduction and sulfidation kinetics. Chemical Engineering Science 1988, 43 (11), 3005-3013.

576

42. Park, N. K.; Lee, D. H.; Jun, J. H.; Lee, J. D.; Ryu, S. O.; Lee, T. J.; Kim, J. C.; Chang, C. H.

577

Two-stage desulfurization process for hot gas ultra cleanup in IGCC. Fuel 2006, 85 (2), 227-234. 20

ACS Paragon Plus Environment

Page 20 of 41

Page 21 of 41

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

578

43. Si, O. R.; Park, N. K.; Chang, C. H.; Kim, J. C.; Lee, T. J. Multicyclic study on improved

579

Zn/Ti-based desulfurization sorbents in mid-temperature conditions. Industrial & Engineering

580

Chemistry Research 2004, 43, 1466-1471.

581

44. Ayala, R. E.; Marsh, D. W. Characterization and long-range reactivity of zinc ferrite in

582

high-temperature desulfurization processes. Industrial & Engineering Chemistry Research 1991,

583

30 (1), 55-60.

584

45. Chomiak, M.; Trawczyński, J. Effect of titania on the properties of Zn-Fe-O sorbents of

585

hydrogen sulfide. Fuel Processing Technology 2015, 134, 92-97.

586

46. Kobayashi, M.; Shirai, H.; Nunokawa, M. Investigation on desulfurization performance and

587

pore structure of sorbents containing zinc ferrite. Energy & Fuels 1997, 11, (4), 887-896.

588

47. Grindley, T.; Steinfeld, G. Development and testing of regenerable hot coal gas

589

desulfurization sorbents. Final Report DoE/MC/16545-1125. 1981.

590

URL.

591

48. Lai, X.; Goodman, D. W. Structure-reactivity correlations for oxide-supported metal catalysts:

592

new perspectives from STM. Journal of Molecular Catalysis A Chemical 2000, 162 (1), 33-50.

593

49. Ouyang, R.; Liu, J. X.; Li, W. X. Atomistic theory of Ostwald ripening and disintegration of

594

supported metal particles under reaction conditions. Journal of the American Chemical Society

595

2013, 135 (5), 1760-1771.

596

50. Pfeifer, C.; Rauch, R.; Hofbauer, H.; Świerczyński, D.; Courson, C.; Kiennemann, A.

597

Hydrogen-rich gas production with a Ni-catalyst in a dual fluidized bed biomass gasifier. Sci

598

Therm Chem Biomass Convers. 2004, 677-690.

599

URL

600

51. Swierczynski, D.; Courson, C.; Bedel, L.; Kiennemann, A.; Guille, J. Characterization of

601

Ni-Fe/MgO/olivine catalyst for fluidized bed steam gasification of biomass. Chemistry of

602

materials 2006, 18 (17), 4025-4032.

603

52. Zhao, Z.; Lakshminarayanan, N.; Swartz, S. L.; Arkenberg, G. B.; Felix, L. G.; Slimane, R.

604

B.; Choi, C. C.; Ozkan, U. S. Characterization of olivine-supported nickel silicate as potential

605

catalysts for tar removal from biomass gasification. Applied Catalysis A: General 2015, 489,

606

42-50. 21

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

607

53. Cui, D.; Liu, J.; Yu, J.; Yue, J.; Su, F.; Xu, G. Necessity of moderate metal-support interaction

608

in Ni/Al2O3 for syngas methanation at high temperatures. RSC Advances 2015, 5 (14),

609

10187-10196.

610

54. Hachimi, A.; Vilcocq, L.; Courson, C.; Kiennemann, A. Study of olivine supported copper

611

sorbents performances in the desulfurization process in link with biomass gasification. Fuel

612

Processing Technology 2014, 118, 254-263.

613

55. Ahmed, M.; Alonso, L.; Palacios, J. M.; Cilleruelo, C.; Abanades, J. Structural changes in

614

zinc ferrites as regenerable sorbents for hot coal gas desulfurization. Solid State Ionics 2000, 138

615

(1), 51-62.

616

56. Gibson, J. B.; Harrison, D. P. The reaction between hydrogen sulfide and spherical pellets of

617

zinc oxide. Industrial & Engineering Chemistry Process Design and Development 1980, 19 (2),

618

231-237.

619

57. Virginie, M.; Adánez, J.; Courson, C.; Diego, L. F. D.; García-Labiano, F.; Niznansky, D.;

620

Kiennemann, A.; Gayán, P.; Abad, A. Effect of Fe-olivine on the tar content during biomass

621

gasification in a dual fluidized bed. Applied Catalysis B Environmental 2012, 121, 214-222.

622

58. Swierczynski, D.; Courson, C.; Bedel, L.; Kiennemann, A.; Vilminot, S. Oxidation reduction

623

behavior of iron-bearing olivines (FexMg1-x)2SiO4 used as catalysts for biomass gasification.

624

Chemistry of materials 2006, 18 (4), 897-905.

625

59. Yang, X.; Xu, S.; Xu, H.; Liu, X.; Liu, C. Nickel supported on modified olivine catalysts for

626

steam reforming of biomass gasification tar. Catalysis Communications 2010, 11 (5), 383-386.

627

60. Wang, G.; Xu, S.; Jiang, L.; Wang, C. Nickel supported on iron-bearing olivine for CO2

628

methanation. International Journal of Hydrogen Energy 2016, 41 (30), 12910-12919.

629

61. Courson, C.; Udron, L.; Petit, C.; Kiennemann, A. Grafted NiO on natural olivine for dry

630

reforming of methane. Science & Technology of Advanced Materials 2002, 3 (3), 271-282.

631

62. Köchermann, J.; Schneider, J.; Matthischke, S.; Rönsch, S. Sorptive H2S removal by

632

impregnated activated carbons for the production of SNG. Fuel Processing Technology 2015, 138

633

(94), 37-41.

634

63. Kern, S.; Pfeifer, C.; Hofbauer, H. Gasification of lignite in a dual fluidized bed

635

gasifier-Influence of bed material particle size and the amount of steam. Fuel Processing 22

ACS Paragon Plus Environment

Page 22 of 41

Page 23 of 41

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

636

Technology 2013, 111, 1-13.

637

64. Neeft, J.; Knoef, H.; Zielke, U.; Sjöström, K.; Hasler, P.; Simell, P.; Dorrington, M.; Thomas,

638

L.; Abatzoglou, N.; Deutch, S.; Greil, C.; Buffinga, G.J.; Brage, C.; Suomalainen, M. Guideline

639

for sampling and analysis of tar and particles in biomass producer gases, version 3.3. Energy

640

project ERK6-CT1999-20002 (Tar Protocol).

641

URL.

642

65. van Paasen, S. V. B.; Kiel, J. H. A.; Neeft, J. P. A.; Knoef, H. A. M.; Buffinga, G. J.; Zielke,

643

U.; Sjöström, K.; Brage, C.; Hasler, P.; Simell, P.A.; Suomalainen, M.; Dorrington, M.A.; Thomas,

644

L. Guideline for sampling and analysis of tar and particles in biomass producer gases.

645

ECN-C-02-090. URL.

646

66. Novochinskii, II; Song, C. S.; Ma, X. L.; Liu, X. S.; Shore, L.; Lampert, J.; Farrauto, R. J.

647

Low-temperature H2S removal from steam-containing gas mixtures with ZnO for fuel cell

648

application. 1. ZnO particles and extrudates. Energy & Fuels 2004, 18 (2), 576-583.

649

67. Sasaoka, E.; Hirano, S.; Kasaoka, S.; Sakata, Y. Stability of zinc oxide high-temperature

650

desulfurization sorbents for reduction. Energy & Fuels 1994, 8 (3), 763-769.

651

68. Šepelák, V.; Steinike, U.; Uecker, D.-C.; Trettin, R.; Wiβmann, S.; Becker, K.

652

High-temperature reactivity of mechanosynthesized zinc ferrite. Solid State Ionics 1997, 101,

653

1343-1349.

654

69. Valenzuela, M. A.; Bosch, P.; Jiménez-Becerrill, J.; Quiroz, O.; Páez, A. I. Preparation,

655

characterization and photocatalytic activity of ZnO, Fe2O3 and ZnFe2O4. Journal of

656

Photochemistry and photobiology A: Chemistry 2002, 148 (1), 177-182.

657

70. Park, S. W.; Joo, O. S.; Jung, K. D.; Kim, H.; Han, S. H. Development of ZnO/Al2O3

658

catalyst for reverse-water-gas-shift reaction of CAMERE (carbon dioxide hydrogenation to form

659

methanol via a reverse-water-gas-shift reaction) process. Applied Catalysis A: General 2001, 211

660

(1), 81-90.

661

71. Park, H. C.; Choi, H. S. The segregation characteristics of char in a fluidized bed with

662

varying column shapes. Powder technology 2013, 246, 561-571.

663

72. Hofbauer, H.; Rauch, R. Stoichiometric water consumption of steam gasification by the

664

FICFB-gasification process // Progress in Thermochemical Biomass Conversion. 2008, 199-208. 23

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

665

URL.

666

24

ACS Paragon Plus Environment

Page 24 of 41

Page 25 of 41

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

667

Figure captions

668

Fig. 1. Schematic of the fixed-bed test.

669

Fig. 2. Concept of the DDLG system.

670

Fig. 3. Schematic of the lab-scale DDLG facility.

671

Fig. 4. Methanation activity of Ni/olivine downstream different H2S sorbents at 550 oC in (a) FG-2

672

and (b) FG-1.

673

Fig. 5. Product gas distribution for RWGS reaction of FG-1 over ZnO at 550 oC.

674

Fig. 6. Methanation activity of Ni/olivine downstream ZnO in FG-1 with varied H2O addition: (a)

675

CO2 conversion and (b) CH4 selectivity.

676

Fig. 7. XRD spectra of the 800 oC calcined olivine and Zn/olivine-750.

677

Fig. 8. TPR profiles of the 800 oC calcined olivine and Zn/olivine-750.

678

Fig. 9. Breakthrough curves of the fresh and regenerated Zn/olivine-750 in FG-1 with 10 vol.%

679

H2O addition: (a) CO2 conversion and (b) CH4 selectivity.

680

Fig. 10. Product gas compositions and the pressure drop of the gasifier as a function of time,

681

biomass feeding rate 0.3 kgfuel,db/h, gasifier temperature 800 oC and S/B 0.7.

682

Fig. 11. Effect of S/B on biomass steam gasification with gasifier temperature at 800 oC.

683

Fig. 12. Effect of gasifier temperature on biomass steam gasification with S/B of 0.4.

25

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

684

Table captions

685

Table 1. Proximate and ultimate analyses of the pine sawdust.

686

Table 2. General operation parameters of the DDLG system.

687

Table 3. Key parameters and results for biomass gasification with raw syngas upgrading over

688

varied bed materials.

689 690

26

ACS Paragon Plus Environment

Page 26 of 41

Page 27 of 41

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

691 692

Fig. 1. Schematic of the fixed-bed test.

693

27

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

694 695

Fig. 2. Concept of the DDLG system.

696

28

ACS Paragon Plus Environment

Page 28 of 41

Page 29 of 41

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

697 698

Fig. 3. Schematic of the lab-scale DDLG facility.

699

29

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

700 701

Fig. 4. Methanation activity of Ni/olivine downstream different H2S sorbents at 550 oC in (a) FG-2

702

and (b) FG-1.

703

30

ACS Paragon Plus Environment

Page 30 of 41

Page 31 of 41

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

704 705

Fig. 5. Product gas distribution for RWGS reaction of FG-1 over ZnO at 550 oC.

706

31

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

707 708

Fig. 6. Methanation activity of Ni/olivine downstream ZnO in FG-1 with varied H2O addition: (a)

709

CO2 conversion and (b) CH4 selectivity.

710

32

ACS Paragon Plus Environment

Page 32 of 41

Page 33 of 41

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

711 712

Fig. 7. XRD spectra of the 800 oC calcined olivine and Zn/olivine-750.

713

33

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

714 715

Fig. 8. TPR profiles of the 800 oC calcined olivine and Zn/olivine-750.

716

34

ACS Paragon Plus Environment

Page 34 of 41

Page 35 of 41

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

717 718

Fig. 9. Breakthrough curves of the fresh and regenerated Zn/olivine-750 in FG-1 with 10 vol.%

719

H2O addition: (a) CO2 conversion and (b) CH4 selectivity.

720

35

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

721 722

Fig. 10. Product gas compositions and the pressure drop of the gasifier as a function of time,

723

biomass feeding rate 0.3 kgfuel,db/h, gasifier temperature 800 oC and S/B 0.7.

724

36

ACS Paragon Plus Environment

Page 36 of 41

Page 37 of 41

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

725 726

Fig. 11. Effect of S/B on biomass steam gasification with gasifier temperature at 800 oC.

727

37

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

728 729

Fig. 12. Effect of gasifier temperature on biomass steam gasification with S/B of 0.4.

730 731

38

ACS Paragon Plus Environment

Page 38 of 41

Page 39 of 41

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

732

Table 1. Proximate and ultimate analyses of the pine sawdust. Proximate analysis (wt.%, ad)

733

Moisture

Ash

Volatile

8.26

0.61

78.40

Fixed carbon 12.73

Ultimate analysis (wt.%, daf) Carbon

Hydrogen

Oxygen1)

Nitrogen

Sulfur

47.75

6.98

44.84

0.07

0.36

1) by difference

734 735

39

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

736

Page 40 of 41

Table 2. General operation conditions of the DDLG system. Fine bed material

60-100 mesh SiO2

Fine bed material inventory (kg)

3.0

Fine particles circulating rate to biomass feeding rate (kg/kg)

10 20-40 mesh Zn/olivine,

Coarse bed material

Ni/olivine and SiO2

Coarse bed material inventory (kg)

4.5

Coarse particles circulating rate to biomass feeding rate (kg/kg)

27

Biomass feeding rate (kgfuel,db/h)

0.3

Steam to biomass mass ratio (S/B) (kgH2O/kgfuel,db)

0.3-1.0

o

Gasifier temperature ( C)

700-850 o

Upgrading reactor temperature ( C)

350-600

o

Combustor temperature ( C)

850

o

Regenerator temperature ( C)

850

o

Riser temperature ( C)

850 o

Steam generator temperature ( C)

800

Operating pressure

atmospheric

737 738 739

40

ACS Paragon Plus Environment

Page 41 of 41

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

740

Table 3. Key parameters and results for biomass gasification with raw syngas upgrading over

741

varied bed materials. Feeding rate (kg/h) S/B o

Gasifier temperature ( C) Fine bed material o

Upgrading reactor temperature ( C) Coarse bed material composition (wt.%) 3

Gas yield (Nm /kgfuel, daf) 3

LHV (MJ/Nm ) Cold gas efficiency (%)

0.30

0.30

0.30

0.4

0.3

0.3

850

850

850

SiO2

SiO2

SiO2

350

600

600

SiO2+Ni/olivine

SiO2+Ni/olivine+

(88/12)

Zn/olivine (76/12/12)

1.03

1.14

1.51

14.6

12.8

11.4

SiO2 (100)

81.2

78.7

92.9

3

Tar content (g/Nm )

61.2

23.2

1.5

H2O conversion /%

20.6

10.5

63.8

H2O content /%

27.0

24.0

8.8

H2S concerntration /ppmv

146.1

147.1

1.7

H2

34.9

41.0

47.3

CO

33.3

28.5

28.5

CH4

10.1

8.4

6.3

CO2

17.2

19.4

17.1

C2H4

2.7

1.6

0.3

C2H6

0.6

0.5

0.3

C3H6

1.0

0.5

0.1

C3H8

0.2

0.1

0.1

Dry gas composition /vol.%

742 743

41

ACS Paragon Plus Environment