261
저작자표시-비영리-변경금지 2.0 대한민국 이용자는 아래의 조건을 따르는 경우에 한하여 자유롭게 l 이 저작물을 복제, 배포, 전송, 전시, 공연 및 방송할 수 있습니다. 다음과 같은 조건을 따라야 합니다: l 귀하는, 이 저작물의 재이용이나 배포의 경우, 이 저작물에 적용된 이용허락조건 을 명확하게 나타내어야 합니다. l 저작권자로부터 별도의 허가를 받으면 이러한 조건들은 적용되지 않습니다. 저작권법에 따른 이용자의 권리는 위의 내용에 의하여 영향을 받지 않습니다. 이것은 이용허락규약 ( Legal Code) 을 이해하기 쉽게 요약한 것입니다. Disclaimer 저작자표시. 귀하는 원저작자를 표시하여야 합니다. 비영리. 귀하는 이 저작물을 영리 목적으로 이용할 수 없습니다. 변경금지. 귀하는 이 저작물을 개작, 변형 또는 가공할 수 없습니다.

Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

저 시-비 리- 경 지 2.0 한민

는 아래 조건 르는 경 에 한하여 게

l 저 물 복제, 포, 전송, 전시, 공연 송할 수 습니다.

다 과 같 조건 라야 합니다:

l 하는, 저 물 나 포 경 , 저 물에 적 된 허락조건 명확하게 나타내어야 합니다.

l 저 터 허가를 면 러한 조건들 적 되지 않습니다.

저 에 른 리는 내 에 하여 향 지 않습니다.

것 허락규약(Legal Code) 해하 쉽게 약한 것 니다.

Disclaimer

저 시. 하는 원저 를 시하여야 합니다.

비 리. 하는 저 물 리 목적 할 수 없습니다.

경 지. 하는 저 물 개 , 형 또는 가공할 수 없습니다.

Page 2: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

A DISSERTATION FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

Improvement of hydrodeoxygenation reactivity of bio-oil

on Ni-based catalysts for application to engine fuels

니켈 촉매를 이용한 바이오오일의 수첨탈산소 개질 및

활용 가능성 평가

Advisor Professor: In-Gyu Choi

By Shinyoung Oh

PROGRAM IN ENVIRONMENTAL MATERIALS SCIENCE

DEPARTMENT OF FOREST SCIENCES

GRADUATE SCHOOL

SEOUL NATIONAL UNIVERSITY

August, 2018

Page 3: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as
Page 4: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

Improvement of hydrodeoxygenation reactivity of bio-oil

on Ni-based catalysts for application to engine fuels

니켈 촉매를 이용한 바이오오일의 수첨탈산소 개질 및

활용 가능성 평가

지도교수 최 인 규

이 논문을 농학박사 학위논문으로 제출함

2018년 5월

서울대학교 대학원

산림과학부 환경재료과학전공

오 신 영

오신영의 농학박사 학위논문을 인준함

2018년 7월

위 원 장 윤 혜 정 (인)

부위원장 최 인 규 (인)

위 원 이 학 래 (인)

위 원 최 준 원 (인)

위 원 이 수 민 (인)

Page 5: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as
Page 6: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

i

Abstract

Improvement of hydrodeoxygenation reactivity of bio-oil

on Ni-based catalysts for application to engine fuels

Shinyoung Oh

Program in Environmental Materials Science

Department of Forest Sciences

The Graduate School

Seoul National University

Due to the high energy consumption, interest in biofuel derived from biomass

developed recently. The largest carbon-based sustainable source, biomass,

could reduce carbon emission, maintain independent fuel supplies than

petroleum. Main advantage of liquid biofuel is higher energy density and

availability on demand which make easier to be stored and transported. One of

biomass to liquid (BTL) process, fast pyrolysis, is a thermochemical conversion

process that could obtain high yield of bio-oil. Generally, biofuel is mainly used

for small scale power and heat generation, however, commercial usage of bio-

oil is very limited due to its disadvantageous features (e.g. high water content,

acidity, viscosity, and oxygen content) which resulted in low heating value and

ignition or injection problems. Therefore, to solve those problems, upgrading

process is essential. The catalytic upgrading, especially hydrodeoxygenation

(HDO) upgrading originated from hydrodesulfurization (HDS) from petroleum

products, is effective, but hard to control due to the various factors. Herein, this

study focused on propose an optimal upgrading process by investigating the

effects of various reaction factors on the features and stability of biofuel and

Page 7: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

ii

blending performance with petroleum as an alternative fuel.

Firstly, the bio-oil, obtained from fast pyrolysis at 500 °C with a residence

time of <2 s, was used as a raw material of HDO process. In order to evaluate

the effect of solvent during bio-oil HDO upgrading, ethanol, acetone and ether

were used in HDO. After the HDO process, the bio-oil was converted into gas,

char, and immiscible liquids (heavy oil and light oil). Moreover, to investigate

aging performance of biofuel during storage period, 12 weeks storage test at

23 °C was conducted with bio-oil and heavy oil. Product yield and various

properties of bio-oil were influenced by the solvent used. Since the polar protic

solvent, ethanol, was able to react with bio-oil as a co-reactant during HDO,

acid-type components in the bio-oil were converted into acid, ethyl ester forms,

which were found in the light oil. Especially, the water content (0.9-4.0 wt%),

HHV (23.9-34.4 MJ/kg) and energy efficiency (59.0-65.2%) were improved the

most, when polar protic ethanol was used. Most properties of the heavy oil were

kept constant at the initial levels during the storage period, which indicates that

the oil properties were stabilized by HDO.

Secondly, to reduce the cost problems of noble metal catalysts, Ni-based

catalyst with mesoporous support (SBA-15 and Al-SBA-15) were synthesized

and their HDO effects were tested with bio-oil at 300°C and under 3MPa H2

pressure with ethanol as a solvent. The yield of heavy oil did not show

significant differences between the three catalysts (45.8-48.1 wt%). Especially,

heavy oil from Ni/Al-SBA-15 exhibited a low water content (9.3 wt%), while

that from Ni/SBA-15 revealed a high HHV (22.8 MJ/kg), energy efficiency

(62.8%), and degree of deoxygenation (54.9%) due to their high surface area

(508.7-537.7 m2/g) and pore volume (0.46-0.65 cm3/g).

To enhance the HDO activity of Ni/SBA-15, three bimetallic catalysts

(NiCu/SBA-15, NiMn/SBA-15, and NiZn/SBA-15) were prepared. High

surface area of bimetallic catalysts resulted in higher heavy oil yield (40.8-60.6

wt%) than that from monometallic catalyst. Bimetallic catalysts dramatically

decreased light oil and increased gas phase, compared with monometallic

Ni/SBA-15 catalyst. Moreover, water content (from 11.4 to 2.0-8.8 wt%),

Page 8: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

iii

acidity (from 62.2 to 49.2-59.8 mg KOH/g oil), viscosity (from 7.1 to 3.2-4.3

cSt) and oxygen content (from 32.2 to 26.2-34.6 wt%) enhanced after HDO,

compared with monometallic Ni/SBA-15. HHV was increased with bimetallic

catalysts, via demethoxylation, hydrogenation, dehydration, as well as

dehydroxylation. The heavy oil obtained from NiMn/SBA-15 revealed a high

HHV (28.4 MJ/kg) with the highest energy efficiency (90.5 %) at 350 °C.

Lastly, some of the previously used catalysts in batch type reactor were

selected and their activities were tested for bio-oil at 300 °C and under H2

pressure in continuous flow reactor. Heavy oil yielded in range of 27.9-42.3

wt%, while the char produced 2.3-11.0 wt%, which were less than those with

batch type reactor. The HDO reaction enhanced in the fuel properties of bio-oil.

Especially, higher HHV (32.8-38.0 MJ/kg), compared with bio-oil (15.7 MJ/kg)

and those from batch type reactor (24.4-34.5 MJ/kg) were measured.

Emulsification, as the blending test of heavy oil, was conducted with diesel and

gasoline. Optimal emulsification condition was investigated and bio-oil hardly

emulsified, while upgraded oil (heavy oil) emulsified with 6 -24 h, depending

on the emulsification conditions. The emulsion of heavy oil and petroleum

maintained its stabilized state for 7 days, suggesting that heavy oil has

possibility to utilize as a blend with petroleum.

Keywords: Fast pyrolysis, Hydrodeoxygenation, Bio-oil,

Ni-based catalyst, Upgrading, Continuous reactor

Student Number: 2014-30384

Page 9: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

iv

Contents

Chapter 1

Introduction ········································································· 1

1. Background ········································································· 2

1.1. Necessity of biomass to liquid (BTL) process ····························· 2

1.2. Thermochemical conversion process to produce biofuel ················· 6

1.3. Fast pyrolysis of biomass ····················································· 8

1.4. Bio-oil characteristics ························································ 10

1.5. Upgrading process for improvement bio-oil quality ····················· 13

2. Objectives ·········································································· 15

3. Literature review ·································································· 18

3.1. Fast pyrolysis process for bio-oil production ····························· 18

3.2. Limitation of bio-oil utilization as biofuel ································ 19

3.2.1. Disadvantages of bio-oil features ······································ 19

3.2.2. Aging of bio-oil during storage period ································ 22

3.3. Upgrading process to overcome the limits of bio-oil usage ············· 25

3.3.1. Hydrotreating for improving the bio-oil quality ····················· 25

3.3.2. Hydrodeoxygenative upgrading process ······························ 29

3.3.3. Various factors affected to the hydrodeoxygenation ················ 32

3.4. Biofuel application ··························································· 39

3.4.1. Blend methodology ····················································· 39

3.4.2. Emulsification to homogenize the blends ···························· 42

3.4.3. Biofuel application in commercial engine ···························· 49

Page 10: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

v

Chapter 2

Solvent effects on improvement of bio-oil features during

hydrodeoxygenative upgrading ··············································· 55

1. Introduction ········································································ 56

2. Materials and methods ···························································· 59

2.1. Fast pyrolysis for bio-oil production ······································· 59

2.2. Selection of solvent as co-reactant ········································· 62

2.3. Hydrodeoxygenation of bio-oil with supercritical fluid ················ 66

2.4. Characterization of hydrodeoxygenation products ······················· 68

2.5. Estimation of process efficiency ············································ 70

2.6. Storage stability for 12 week period ······································· 71

3. Results and discussion ···························································· 72

3.1. Effects of solvents upon the HDO products ······························· 72

3.2. Physicochemical properties of liquid products ··························· 76

3.3. Atomic H/C and O/C ratio ·················································· 82

3.4. Compositional changes of liquid products ································ 84

3.5. Gas composition ······························································ 87

3.6. Estimation of reaction efficiency ··········································· 89

3.7. Changes in the properties of heavy oil during storage ··················· 91

4. Summary ········································································· 106

Page 11: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

vi

Chapter 3

Preparation of Ni-based catalysts with various support materials

and assessment of their catalytic effects for improvement of bio-oil

properties ········································································· 107

1. Introduction ······································································ 108

2. Materials and methods ·························································· 111

2.1. Catalyst preparation by sol-gel method and wet impregnation ······· 111

2.2. Characterization of synthesized Ni-based catalysts ···················· 113

2.3. Production of bio-oil via fast pyrolysis ·································· 114

2.4. Hydrodeoxygenative upgrading of bio-oil with Ni-based catalysts ·· 115

2.5. Physicochemical properties of bio-oil and upgraded oil ··············· 116

2.6. Chemical composition of hydrodeoxygenative upgraded oil ········· 117

3. Results and discussion ·························································· 119

3.1. Synthesis of mesoporous supports and supported nickel catalysts ··· 119

3.2. Catalytic activity···························································· 130

3.3. Fuel properties of heavy oil ··············································· 132

3.4. Physicochemical features of light oil ···································· 137

3.5. Degree of deoxygenation (DOD) ········································· 139

3.6. Structural modification of low-molecular weight compounds during

HDO ·········································································· 140

3.7. Gas composition ···························································· 144

4. Summary ········································································· 146

Page 12: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

vii

Chapter 4

Enhancement of bio-oil hydrodeoxygenation over Ni-based

bimetallic catalysts on mesoporous silica support ················· 147

1. Introduction ······································································ 148

2. Materials and methods ·························································· 151

2.1. Synthesis of SBA-15 support by sol-gel method ······················· 151

2.2. Selection of transition metal and preparation of bimetallic catalysts 152

2.3. Characterization of catalysts ·············································· 153

2.4. Hydrodeoxygenative activity test of prepared catalysts ··············· 155

2.5. Measuring fuel quality of the products ·································· 156

3. Results and discussion ·························································· 158

3.1. Characterization of the bimetallic catalysts ····························· 158

3.2. Mass balance of the hydrodeoxygenation products ···················· 170

3.3. Fuel properties of the upgraded products ································ 172

3.4. Chemical composition of hydrodeoxygenatively upgraded heavy oil 178

3.5. Gas composition ···························································· 181

4. Summary ········································································· 183

Page 13: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

viii

Chapter 5

Screening of the catalysts HDO activity in continuous reactor and

biofuel blending test for engine application ····························· 185

1. Introduction ······································································ 186

2. Materials and methods ·························································· 190

2.1. Preparation and characterization of catalysts ··························· 190

2.2. Hydrodeoxygenative activity test of prepared catalyst ················ 191

2.3. Measuring fuel quality of the products ·································· 192

2.4. Gas composition ···························································· 193

2.5. Emulsification with commercial petroleum ····························· 195

3. Results and discussion ·························································· 196

3.1. Yield of the HDO products obtained from continuous flow reactor 196

3.2. Fuel properties of obtained biofuel ······································· 198

3.3. Converted chemical compounds during HDO ·························· 202

3.4. Gas composition ···························································· 205

3.5. Blends with commercial petroleum ······································ 207

4. Summary ········································································· 211

Page 14: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

ix

Chapter 6

Concluding remarks ··························································· 213

References······································································· 217

초록 ················································································ 237

Page 15: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

x

List of Tables

Table 1-1. Comparison of slow pyrolysis and fast pyrolysis ···················· 19

Table 1-2. Typical hydrotreating catalysts ········································ 36

Table 1-3. Previous studies of bio-oil blends ····································· 47

Table 1-4. List of the most significant surfactants ······························· 48

Table 1-5. Properties that affected to the properties and the requirements ···· 50

Table 1-6. Set of standard bio-oil application····································· 54

Table 2-1. Chemical compositions of Miscanthus ································ 60

Table 2-2. Critical characteristics of solvent and its classification ············· 64

Table 2-3. Major products and characteristics of hydrodeoxygenation of

solvent with hydrogen pressure ······································· 65

Table 2-4. Mass balance of essential hydrodeoxygenation products obtained

from various solvents (ethanol, acetone and ether) ················· 74

Table 2-5. Typical physicochemical features of bio-oil and heavy oil obtained

from various conditions (without solvent, with solvents such as

ethanol, acetone or ether) ·············································· 79

Table 2-6. Quantitative analysis of major micromolecules in the heavy oils by

GC/MS analysis ························································· 85

Table 2-7. Quantitative analysis of major micromolecules in the light oils by

GC/MS analysis ························································· 86

Table 2-8. Distribution of representative gas components after HDO

reaction ··································································· 88

Table 2-9. Carbon recovery rate of the products obtained from the HDO

reaction ··································································· 90

Table 2-10. Physicochemical features of heavy oil obtained from hydrodeoxy-

genative upgrading of bio-oil in the presence of Pt/C catalyst

during various storage period ·········································· 93

Page 16: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

xi

Table 2-11. Quantitative analysis of low molecular weight components in

heavy oil after 4, 8 and 12 weeks (23 °C) ·························· 100

Table 2-12. Characteristics of macromolecules extracted from bio-oil, heavy

oil, and aged heavy oil ··············································· 103

Table 3-1. Characterization of synthesized Ni-based catalysts ··············· 120

Table 3-2. Mass balance of the products obtained from HDO reaction with Ni-

catalysts ································································ 131

Table 3-3. Physicochemical properties of heavy oil···························· 134

Table 3-4. Quantitative analysis of major micromolecular components in the

bio-oil and heavy oils by GC/MS analysis ························· 142

Table 3-5. Quantitative analysis of major micromolecular components in the

bio-oil and light oils by GC/MS analysis ·························· 143

Table 3-6. Distribution of representative gas components determined by TCD-

FID after HDO reaction ·············································· 145

Table 4-1. Characterization of synthesized Ni-based catalysts ··············· 160

Table 4-2. Mass balance of the products obtained from HDO reaction with Ni-

catalysts ································································ 171

Table 4-3. Physicochemical properties of heavy oil with bimetallic catalysts

under 250, 300, and 350°C ·········································· 174

Table 4-4. Low molecular compounds in bio-oil and heavy oils ············ 180

Table 4-5. Distribution of representative gas components determined by TCD-

FID after HDO reaction ············································· 182

Table 5-1. Mass balance of the products obtained from HDO reaction with

continuous flow reactor ·············································· 197

Table 5-2. Physicochemical properties of heavy oil ··························· 200

Table 5-3. Quantitative analysis of major micromolecular components in the

bio-oil and heavy oils by GC/MS analysis ························· 204

Page 17: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

xii

List of Figures

Figure 1-1. Statistics of energy consumption and renewable energy supply; (a)

Trend of world energy consumption (source : Global energy

statistical yearbook 2017), (b) U.S. renewable energy supply

(source : Short-Term Energy Outlook, April 2018) ·················· 4

Figure 1-2. Biomass to liquid process (BTL) flow of biomass ·················· 5

Figure 1-3. Scheme of biomass to biofuel via fast pyrolysis ····················· 9

Figure 1-4. Deposits on needle surface after 3 days stirring in 75 % bio-oil in

diesel emulsions (a) and (b); and corrosion of the needle below the

deposits (c) and (d) ····················································· 11

Figure 1-5. Major upgrading technologies and requirements of HDO

process ··································································· 14

Figure 1-6. Objective of this study ················································· 17

Figure 1-7. Structure of pyrolytic lignin (major factor of bio-oil instability)

and plausible reaction during storage ································· 24

Figure 1-8. Reactivity scale of oxygenated groups under hydrotreatment

conditions ································································ 26

Figure 1-9. Reaction routes for the bio-oil hydrotreating ······················· 28

Figure 1-10. Typical reaction routes for HDO of bio-oil ························ 31

Figure 1-11. Phase diagram (pressure and temperature) for a pure compound

in a close system ························································ 38

Figure 1-12. Mechanisms of emulsion breaking: (a) stable state emulsion; (b)

creaming; (c) Ostwald ripening; (d) flocculation; (e) coalescence;

(f) breaking ······························································ 44

Figure 1-13. Scheme of emulsions blended by the HLB ························ 45

Figure 1-14. Phase diagram for the Biofuel, Petroleum, Emulsifier ternary

system ···································································· 46

Page 18: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

xiii

Figure 2-1. Schematic diagram of the experimental apparatus for the fast

pyrolysis ((1) sample hopper; (2) screw feeder; (3) fluidized bed

reactor; (4) cyclone; (5) condensers; (6) electrostatic precipitator;

(7) blower) ······························································· 61

Figure 2-2. Van Krevelen diagram of bio-oil, N2 (HDO with ethanol, without

hydrogen), and H2 (HDO with hydrogen and ethanol ·············· 63

Figure 2-3. The reaction pressure of three solvents at each HDO

temperature ······························································ 75

Figure 2-4. Water content, and acidity (TAN) of light oils obtained from HDO

with different solvents ·················································· 81

Figure 2-5. Van Krevelen diagram of bio-oil and heavy oils (No solvent: HDO

of bio-oil without solvent, only with hydrogen pressure) ·········· 83

Figure 2-6. Variation of viscosity during 12 weeks storage period ············· 95

Figure 2-7. Van Krevelen diagram of bio-oil and heavy oils after 12 weeks

storage ···································································· 98

Figure 2-8. Change of macromolecules yield during the storage period ···· 102

Figure 2-9. Degree of unsaturation for oil and macromolecules before and

after the storage period ··············································· 105

Figure 3-1. Scheme of the whole experiment ··································· 118

Figure 3-2. XRD peaks of supports and catalysts (The peaks of (100), (101),

and (110) revealed the hexagonal structure, while (100), (101),

(110), and (111) described the cubic structure) ···················· 122

Figure 3-3. XPS analysis of Ni2p spectra on (a) Ni/C, (b) Ni/SBA-15, and (c)

Ni/Al-SBA-15 (853 eV : Ni0; 856-857 eV : Ni2+ ions or NiO

particles; 860 eV : shake-up satellite peak) ························ 124

Figure 3-4. NH3-TPD peaks of supports and catalysts ························· 126

Figure 3-5. SEM images of the supports ((a): active charcoal; (b): SBA-15;

and (c): Al-SBA-15; Scale bar: 200nm) ···························· 128

Page 19: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

xiv

Figure 3-6. SEM-EDS images of Ni-based catalysts (Ni/C, Ni/SBA-15, and

Ni/Al-SBA,-15) ······················································· 129

Figure 3-7. Van Krevelen diagram of bio-oil, and heavy oils (dry basis)···· 136

Figure 3-8. Physicochemical properties of light oil ···························· 138

Figure 4-1. N2 adsorption-desorption isotherms and pore size distribution

curves of bimetallic catalysts ········································ 159

Figure 4-2. NH3-TPD peaks of Ni- and bimetallic catalysts on SBA-15 ···· 162

Figure 4-3. Small angle XRD pattern of SBA-15 support and XRD peaks of

Ni- and bimetallic catalysts on SBA-15 ···························· 164

Figure 4-4. XPS of (a)Ni/SBA-15; (b) NiCu/SBA-15; (c) NiMn/SBA-15; (d)

NiZn/SBA-15 (853 eV : Ni0; 856-857 eV : Ni2+ ions or NiO

particles; 860 eV : shake-up satellite peak). The shake-up satellite

peak only observed in Ni/SBA-15 catalyst. ························ 166

Figure 4-5. SEM micrographs of (a) SBA-15 support; (b) Ni/SBA-15; and Ni-

base bimetallic catalysts ((c) NiCu/SBA-15; (d) NiMn/SBA-15; (e)

NiZn/SBA-15) ························································· 168

Figure 4-6. SEM-EDS images of (a) SBA-15 support; (b) Ni/SBA-15; and Ni-

base bimetallic catalysts ((c) NiCu/SBA-15; (d) NiMn/SBA-15; (e)

NiZn/SBA-15) ························································· 169

Figure 4-7. Van Krevelen diagram of bio-oil, and heavy oils ················· 177

Figure 5-1. Van Krevelen diagram of bio-oil and heavy oils, dry basis ····· 201

Figure 5-2. Gas composition determined by TCD-FID after HDO reaction 206

Figure 5-3. Ternary diagram of emulsification test of Heavy oil (Upgraded oil)

with petroleum; (a) gasoline and (b) diesel (Yellow dotted region

meant the conditions of forming stable emulsion) ················ 209

Figure 5-4. Images of before and after emulsification, and stable state

emulsion (W/O emulsification with Span60 as emulsifier at

60 °C) ·································································· 210

Page 20: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

xv

List of Abbreviations

HDO Hydrodeoxygenation

CR Carbon recovery

DOD Degree of deoxygenation

TAN Total acid number

HHV Higher heating value

RF Response factors

GC/MS Gas chromatography/mass spectroscopy

FID Flame ionization detector

TCD Thermal conductivity detector

BET Brunauer–Emmett–Teller

BJH Barrett-Joyner-Halenda

XRD X-ray diffraction

TPD Temperature programmed desorption

SAXS Small angle X-ray scattering

XPS X-ray photoelectron spectroscopy

SEM Scanning electron microscope

EDS Energy Dispersive X-Ray Spectroscopy

ICP-ES Inductively coupled plasma emission spectroscopy

GPC Gel permeation chromatography

Mw Weight-average molecular weight

Mn Number-average molecular weight

PDI Polydispersity index

PP Phenol polymer

PL Pyrolytic lignin

APP Aged phenol polymer

APL Aged pyrolytic lignin

Page 21: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

xvi

CPU Condensed phase upgrading

VPU Vapor phase upgrading

LEFR Laminar entrained flow reactor

NDIR Nondispersive infrared sensor

Pluronic P-123 Poly(ethylene glycol)-block-poly(propylene glycol)-

block-poly(ethylene glycol)

TEOS Tetraethyl orthosilicate

AIP Aluminum isopropoxide

IC engine Internal combustion engine

Page 22: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

1

Chapter 1

Introduction

Page 23: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

2

1. Background

1.1. Necessity of biomass to liquid (BTL) process

Demand for energy and its resources, has never been higher worldwide than

it is today due to the rapid population and urbanization growth. As the world

population has grown and transportation has become more readily available, the

increase in fuel demand has become natural. Transportation sector, constituting

about one fifth of the total energy consumption, is one of the main energy

consumption fields. Present sources of energy are not sufficient to overcome the

increasing needs. Since 1973 the energy resources have been doubled in

developed countries but the demand is still higher. It is unavoidable that World

energy demand will increase approximately two and half times the present level

(Goyal et al. 2008; Mortensen et al. 2011).

The concerns of fossil fuel related to the environment or the economy, such

as the depletion of fossil fuels, variation of oil price, growth of carbon dioxide

emission and global warming, resulted in high demands for renewable energy.

Future Bioenergy availability in 2050 estimated from 20 to 400 EJ (20 to 400

1018 J), while use of current global primary energy is 465 EJ (Lysen 2000;

Gielen et al. 2002). Therefore, the alternative energy provided from biomass,

which is carbon dioxide neutral energy source and has low sulfur, nitrogen and

ash, has been studied. The liquid fuel, such as biocrude oil, biodiesel, bioalcohol,

biochar, biogas and syngas, that can be used as transportation fuel in high

demand with low-cost (Mohan et al. 2006; Grahn et al. 2007; Baliban et al.

2013), produced by thermo-chemical and bio-chemical conversion processes

(Bridgwater 2012) is considered as promising renewable liquid fuel. Among

electricity, industrial and transportation fuel, one third of whole energy is used

for transportation, typically 63% of nonrenewable petroleum used for

Page 24: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

3

transportation energy in the US (Mohan et al. 2006). Therefore, the conversion

of biomass to liquid fuel might result in important advantages at environmental

and economic strategies (Hussain et al. 2013).

Biomass can be converted to liquid fuel via biochemical and thermochemical

system (Huber et al. 2006). Biochemical conversion (sugar platform), involves

depolymerization of polysaccharides and fermentation of the resulting sugars.

Compared with sugar-based biomass, biochemical conversion of lignocellulosic

biomass is harder due to the complexity of cross-linked lignin-hemicellulose

matrix structure (Brown and Brown 2013). Thus, pretreatment process is added

to the biochemical conversion of lignocellulosic biomass to improve the

accessibility of cellulase on cellulose followed by hydrolysis and fermentation

to ethanol, while thermochemical conversion does not require such a

complicated process (Lee 1997).

Page 25: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

4

Figure 1-1. Statistics of energy consumption and renewable energy supply; (a)

Trend of world energy consumption (source : Global energy

statistical yearbook 2017), (b) U.S. renewable energy supply

(source : Short-Term Energy Outlook, April 2018)

*: Liquid biofuels include ethanol and biodiesel

**: Other biomass includes municipal waste from biogenic sources, landfill gas, and

other non-wood waste

Page 26: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

5

Figure 1-2. Biomass to liquid process (BTL) flow of biomass

Page 27: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

6

1.2. Thermochemical conversion process to produce

biofuel

Biomass can be converted into useful forms of energy, like power/heat

generation and transportation fuels, or chemicals through a number of different

processes. Biomass to energy conversion divided into two main process

technologies: thermo-chemical and bio-chemical. Thermochemical conversion

have three basic approaches: combustion, gasification and liquefy. Biomass

liquefy included high-temperature pyrolysis, high-pressure liquefaction, or

supercritical extraction.

Combustion meant the burning of biomass in air, and is used over a wide

range of outputs to convert the chemical energy stored in biomass into heat,

mechanical power, or electricity using various items of process equipment, e.g.

stoves, furnaces, boilers, steam turbines, turbo-generators, etc. However, net

bioenergy conversion efficiency for combustion range from 20-40%

(McKendry 2002b).

Gasification is the conversion of biomass into a gas mixture by the partial

oxidation of biomass at high temperature ranged 800-900 °C with oxygen, air,

steam, or mixtures of these as gasifying agent (Damartzis and Zabaniotou 2011).

Biomass gasification performed through the following reactions typically

(Demirbaş 2001; McKendry 2002a; White and Plaskett 1981):

C + O2 → CO2

C + 1

2O2 → CO

CO + 1

2O2 → CO2

CO2 + C ↔ 2CO

CO2 + 4H2 → CH4 + 2H2O

Page 28: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

7

Produced gas is a mixture of H2, CO, CO2, CH4, C2H4, and other impurities

such as nitrogen, sulfur, alkali compounds and tar, and had low calorific value

(4-6 MJ/N m3). Moreover, the produced gas can be used as direct for gas engines

and turbine as well as chemical (e.g. methanol) production (McKendry 2002c).

Pyrolysis is converted biomass into liquid fraction by heating around 500 °C

in the absence of air. It enables the conversion of biomass to bio-oil with an

efficiency of up to 80 % (McKendry 2002b; Bridgewater 2004; Bridgwater

2012). Different conditions (temperature, residence time, particle size, and

feedstock) leads to varied proportions of products (Goyalet al. 2008).

Moreover, hydrothermal upgrading and liquefaction could produce bio-oil

from biomass. Hydrothermal upgrading converted biomass in a wet

environment at high pressure to partly oxygenated hydrocarbons. Stable liquid

hydrocarbon was yielded by liquefaction with low temperature and high

hydrogen pressure. But the process is more complex and expensive then

pyrolysis (McKendry 2002b).

Recently, some processes are integrated to produce second generation

biofuels. For instance, integrated gasification combined cycle could produce

electrical energy by combining gasification unit with gas and steam turbines.

Pyrolysis/gasification, syngas cleaning and condition, and Fischer-Tropsch

synthesis were integrated for second generation liquid fuel production

(Damartzis and Zabaniotou 2011).

Page 29: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

8

1.3. Fast pyrolysis for bio-oil production

During the thermochemical conversion, biomass degraded to release the

inherent energy (combustion) or to produce fuel intermediates (energy carriers)

such as synthesis gas (syngas) and pyrolysis liquids under medium or high

temperature in an oxidized or reduced atmosphere (Boateng et al. 2008).

Pyrolysis, which is one of the thermochemical conversion technologies,

thermally decomposed organic materials in the absence of oxygen, producing

of liquid, gases and a highly reactive carbonaceous char. Fast pyrolysis is a

thermal decomposition process that occurs at moderate temperatures with a high

heat transfer rate to the biomass particles and a short hot vapor residence time

in the reaction zone (Czernik and Bridgwater 2004). Since fast pyrolysis system

is simple and the yields of liquid products (bio-oil) are high (50-75%), bio-oil

from fast pyrolysis becomes one of the most promising methods to convert

renewable resources such as lignocellulosic biomass to liquid fuels and

chemicals (Bridgewater 2004).

Page 30: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

9

Figure 1-3. Scheme of biomass to biofuel via fast pyrolysis

Page 31: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

10

1.4. Bio-oil characteristics

During the fast pyrolysis, bio-oil is formed by rapidly and simultaneously

depolymerizing and fragmenting of cellulose, hemicelluloses and lignin with a

rapid temperature increasing (Mohan et al. 2006). Therefore, the product quality

depends on reaction temperature, residence time, heating rate, reactor type as

well as feedstock. Multi-component dark brown homogeneous mobile liquid,

bio-oil, can be used in a burner, furnace, boiler system, diesel and gasoline

engines and turbines for electricity generation, and also converted into the

useful chemicals (Czernik and Bridgwater 2004). Bio-oil is consisted of four

distinct fractions 40% of medium-polar monomers detectable by GC, 12% of

polar monomers detectable directly by HPLC or GC after derivatization, 28%

of water derived from reaction water and feedstock moisture, 20% of oligomeric

material, known as pyrolytic lignin. Generally the 40% of oxygen is present in

amounts of compounds that have been identified in the bio-oil (Czernik and

Bridgwater 2004). Water in bio-oil, varying a wide range of 15-30%, might be

originated from the moisture in the feedstock and dehydration occurring during

pyrolysis (Oasmaa and Czernik 1999). From those, typical heating value of bio-

oil is about 17 MJ/kg, which is only 40-45% of that of hydrocarbon fuels (40

MJ/kg). The viscosity, which related to the flowability, ranged from 18.5-24.0

cSt and at 40 °C (Wildschut et al. 2009) depending on the feedstock,

experimental conditions and especially on the efficiency of collection of low

boiling components (Czernik and Bridgwater 2004). Also, bio-oil obtained from

biomass is quite acidic in the range of pH 2-3. When the bio-oil blends with

petroleum used in diesel engine, the needle was corroded due to its acidity.

Stirring in the emulsion formed very hard, cracked and adhering deposits on the

needles, corrosive damages were also found (Chiaramonti et al. 2003b) (Figure

1-4)

Page 32: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

11

Figure 1-4. Deposits on needle surface after 3 days stirring in 75 % bio-oil in

diesel emulsions (a) and (b); and corrosion of the needle below the

deposits (c) and (d) (Chiaramonti et al.,2003b).

Page 33: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

12

Those features such as high viscosity, high water contents, and high acidity

as well as a high amount of unsaturated oxygen-containing obtained from

thermal decomposition (Torri et al. 2010) resulted in lower energy density, less

than 50% of that for conventional fuels and immiscibility with hydrocarbon

fuels.

The pyrolytic lignin, consisted of small amounts of phenol, eugenol, cresols,

and large quantities of alkylated phenols, is one of the major factors of

determining viscosity, reactivity, and stability of bio-oil (Bayerbach and Meier

2009). Homogeneous and stable properties are required during storage or at the

power generation facility for liquid fuel (Boucher et al. 2000; Chiaramonti et al.

2007; Van de Beld et al. 2013). However, the large amounts of volatiles, many

reactive oxygenated compounds, and pyrolytic lignin in bio-oil produce a glue-

like material when stored for extended periods at ambient temperature, which

leads to instability (Czernik and Bridgwater 2004; Diebold and Czernik 1997;

Boucher et al. 2000; Mante and Agblevor 2012). Moreover, polymerization,

esterification, and etherification of these compounds can increase the viscosity

of the bio-oil during storage, resulting in injection problems when used as a

biofuel (Boucher et al. 2000). Those factors cause bio-oil instable and preclude

direct using of bio-oil as transportation fuel, due to the possibility of corrosion

of engine, low mobility, low flowability, multiphase flow, clogging, high-

pressure drops, and stoppage during provision bio-oil to vehicle engine (Czernik

and Bridgwater 2004). Therefore, maintenance and improvement of the

properties and stability of biofuels are the main factors regarding their

introduction into the market (Diebold 2000).

Page 34: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

13

1.5. Upgrading process for improvement bio-oil quality

Some disadvantageous features of bio-oil mentioned above make it not

suitable for direct application in biofuel. Therefore, bio-oil needs to reform its

undesirable features by reducing the oxygen content. There are physical and

catalytic upgrading technologies. Simplify or intensify processes is strong point

of hot-gas filtration process, solvent addition is used to homogenize the bio-oil

to decrease viscosity. For diesel, emulsified with surfactant is performed for its

utilization. But these physical upgrading technologies cannot improve bio-oil to

suitable properties regardless of its simple and inexpensive process. Thus, to

obtain high quality biofuel, multi-stage catalytic hydroprocessing, e.g. hydrode-

oxygenation (HDO), catalytic cracking of pyrolysis vapors and esterification is

studied recently. In detail, catalytic cracking which decomposes pyrolytic

vapors to hydrocarbons by the oxygen removal with a porous zeolite catalyst

such as MCM-41, ZSM-5 an HZSM-5. In company with alcohol and catalyst,

acids in bio-oil, which is a major problem of engine corrosion and most reactive

compounds, are esterified. The HDO upgrading process has been widely

researched and is well known as an effective way to improve bio-oil quality and

stability. One of the disadvantages of bio-oil, high viscosity, which can cause

nozzle plugging or injection problems, is improved by removing oxygen in

oxygenated compounds as water or other compounds under high temperature

and high pressure with the catalyst such as NiMo, CoMo, solid acid catalyst or

noble metal catalyst with various supports (Czernik and Bridgwater 2004; Qin

et al. 2013; Ardiyanti et al. 2011).

Page 35: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

14

Figure 1-5. Major upgrading technologies and requirements of HDO process

Page 36: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

15

2. Objectives

Biomass has great potential as a promising resource to produce a liquid

biofuel, that could alter the conventional petroleum based fuel. Although, fast

pyrolysis is an encouraging thermochemical conversion technology due to its

simple process and production of high yield of liquid bio-oil, the product quality

is not that suitable. To overcome the disadvantageous features, hydrotreating is

required. But the complexity of various influence factors makes it hard to

understand mechanism and expect the product features. Therefore, investigation

of the bio-oil hydrodeoxygenative (HDO) upgrading mechanism and

production of high quality and stable biofuel for application to conventional

engines provides positive effects on sufficiency of rapidly increasing energy

demand as well as reducing the concerns related to fossil fuel usage.

In this study, the effects of various reaction parameters (solvent, catalyst,

temperature, and reaction type) on fuel quality and stability via HDO were

investigated. Especially, the Ni, widely applied due to activity, selectivity and

possible reaction pathways for the hydrodeoxygenation, was selected to

improve economic efficiency. In order to better comprehensive understanding

of catalytic effect on product quality, both catalyst characterization and product

analysis were conducted. Catalyst surface features, morphologies and structural

properties of support were analyzed by BET, ICP-ES, XRD, XPS, NH3-TPD,

and SEM-EDS. Fuel properties, (e.g. water content, acidity, viscosity, heating

value, and stability), and process efficiency were also estimated. From this

result, correlations between process parameters and product features were

suggested. Additionally, storage stability of produced biofuel was carried out,

and its improved stability was compared to bio-oil aging performance.

Furthermore, the blending test of upgraded bio-oil (heavy oil) with commercial

petroleum was performed and also compared with bio-oil. Three emulsifiers

Page 37: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

16

with three types of emulsification were applied to heavy oil blending with

petroleum. After stable form emulsion was obtained, the stability of it was

investigated by observing the period of maintaining the emulsion. Based on

these results, the possibility of utilization to conventional engines was discussed.

Therefore, the objectives of this study were:

(1) To evaluate the effects of solvent type on the improved properties of

hydrodeoxygenative upgraded bio-oil (heavy oil),

(2) Using the Ni-based catalysts, to enhance the economic efficiency

compared with noble metal catalysts, the effects of catalyst supports,

bimetallic active sites and reactor type on the properties of heavy oil

during the HDO process were assessed.

(3) Then, the storage stability of hydrodeoxygenative upgraded bio-oil

(heavy oil) was observed.

(4) Based on the above results, to determine the possibility of heavy oil

application to the conventional engines by emulsification with petroleum.

Page 38: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

17

Figure 1-6. Objective of this study

Page 39: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

18

3. Literature review

3.1. Fast pyrolysis process for bio-oil production

Initially, pyrolysis of biomass was conducted for the production of charcoal

(biochar). After that, fast pyrolysis is defined as the thermal decomposition of

the biomass organic matrix in non-oxidizing atmospheres resulting in liquid

bio-oil, since oil crisis in the 1970s (Kan et al. 2016; Czernik and Bridgwater

2004). In the 1990s, fast pyrolysis technologies have been developed several

reactors and reached near-commercial status by Red Arrow Products Co., Inc.,

in Wisconsin, Dynamotive in Vancouver and BTG in Netherlands (Czernik and

Bridgwater 2004). Fast pyrolysis could obtain high yield (ca. 50-75%) of liquid

bio-oil with 15-25% of solid char and 10~20% of non-condensable gas without

pretreatment. Bio-oil is produced by thermal decomposition under moderate

temperature (450-550 °C), ambient pressure and very short reaction time (1-2 s)

in the absence of oxygen. Therefore, chemical reaction kinetics, heat and mass

transfer processes, as well as a phase transition phenomena, play important roles

in fast pyrolysis (Bridgewater 2004). Lower process temperature and longer

vapor residence times favor the production of char, whereas high temperature

and long vapor residence time increase the gas yields. Therefore, moderate

temperature and short vapor residence time are optimized for higher liquid

yields.

Page 40: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

19

Table 1-1. Comparison of slow pyrolysis and fast pyrolysis (Brownsort 2009)

Slow pyrolysis Fast pyrolysis

Temperature (°C) Range 250-750 400-750

Typical 350-400 450-550

Residence time Unit mins-days secs

Range 2-30 min <2 s

Yield (wt%)

Char Range 2-60 0-50

Typical 25-35 10-25

Bio-oil Range 0-60 10-80

Typical 20-50 50-70

Gas Range 0-60 5-60

Typical 20-50 10-30

Page 41: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

20

3.2. Limitation of bio-oil utilization as biofuel

3.2.1. Disadvantages of bio-oil features

Despite the benefits of the fast pyrolysis system, potential of direct

substitution of bio-oil for petroleum fuels and chemical feedstocks is limited

due to their high viscosity, high water and ash contents, low heating value,

instability and high corrosiveness. Ensyn has produced 100t of constant quality

pyrolysis liquid, which has been successively combusted in 8MW boilers at

Oilon (Finland) and in 10 MW boilers at Fortum Varme (Sweden) (Oasmaa et

al. 2001; Oasmaa et al. 2005). However, problems with the combustion of bio-

oil in diesel engines have been reported by in large-scale (> 80 kg/h) units in

Ormrod. The bio-oil was separated into a heavy fraction of pyrolysis liquid and

water-soluble phase, due to the characteristics of feedstock (Bridgwater,1999).

The water and solid contents are the major factors of affecting the bio-oil quality,

since the water content has good correlations with density and heating value.

The factors are influenced by feedstock moisture, process parameter, or oxygen

leak (water content), and low gas velocities (<15 m/s) in the cyclones, or it can

be due to a large portion of “fines” <10 μm in size (solid contents) (Oasmaa et

al. 2005). Bio-oil has over 30 wt% of water content is defined as poor-quality,

while solids content less than 0.1 wt% belongs to good quality bio-oil (Oasmaa

et al. 2004).

Moreover, bio-oil is a complex mixture of organic compounds which

produced from thermal degradation of cellulose, hemicellulose and lignin. High

amount of unsaturated oxygen-containing compounds like acid, aldehydes,

Page 42: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

21

alcohol and sugars were obtained from thermal decomposition (Torri et al.

2010). It also results in a low energy density, less than 50% of that for

conventional fuels and immiscibility with hydrocarbon fuels. These make the

bio-oil instable and preclude direct using of bio-oil as transportation fuel. Also,

those disadvantages resulted in corrosion of engine, low mobility, flowability,

multiphase flow, clogging, high-pressure drops, and stoppage during provision

bio-oil to vehicle engine (Czernik and Bridgwater 2004). Consequently,

upgrading of bio-oil is necessary to give a liquid product that can be used as a

liquid fuel or chemical feedstocks in various applications.

Page 43: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

22

3.2.2. Aging of bio-oil during storage period

Due to the concerns of fluctuation of oil price, elasticities of gasoline or

crude oil demand, the strategic solution to decline or prevent the adverse

economic consequences of a serious oil supply outage is needed. The Strategic

Petroleum Reserve came into effect in 1975, the period steadily increase from

38 to 71 days in 2013 (Difiglio 2014). The largest emergency supply with the

capacity to hold up to 727 million barrels is in Louisiana and Texas by the

United States Department of Energy. Thus, the storage stability of bio-oil is

important for utilization as biofuel. Diebold (Diebold 2000) presented diverse

chemical reactions were expected to occur in bio-oil via the catalytic reaction,

but these reactions also occurred in the absence of catalysts when bio-oil was

stored at room temperature for a long time. The large amounts of volatiles and

many reactive oxygenated compounds which lead to instability, including acids,

aldehydes, sugars, and alcohols in bio-oil, could produce a glue-like material

when stored for extended periods at ambient temperature (Czernik and

Bridgwater 2004; Diebold and Czernik 1997; Boucher et al. 2000; Mante and

Agblevor 2012). Especially, esterification of acids with alcohols, aldehyde

polymerization, hydration of aldehydes or ketones with water, hemiacetal or

acetal formation between aldehydes and alcohols, resin formation from

aldehydes and phenols, olefin condensation, oxidation of alcohols and

aldehydes to form carboxylic acids, and oxidative esterification of aldehydes

were suggested to occur in bio-oil during storage (Mante and Agblevor 2012;

Diebold 2000). These reactions are involved in property changes such as

viscosity (Boucher et al. 2000). Maintaining a lower viscosity is important for

Page 44: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

23

atomization and injection. Therefore, studies have been performed to

experimentally demonstrate the occurrence of the above reactions (Mante and

Agblevor 2012; Kim et al. 2012).

Page 45: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

24

Figure 1-7. Structure of pyrolytic lignin (major factor of bio-oil instability) and plausible reaction during storage (Bayerbach

and Meier 2009)

Page 46: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

25

3.3. Upgrading process to overcome the limits of bio-oil

usage

3.3.1. Types of upgrading process for improving the bio-oil quality

Since the drawbacks of bio-oil mentioned above, the process that improves

the bio-oil quality, and maintains the stability is the main factor regarding their

introduction into the market (Diebold 2000). The bio-oil reforming process can

be roughly categorized into two types, physical and catalytic type process.

Representative physical upgrading technology, emulsification, is the simplest

way to use bio-oil as a transport fuel, or combine with diesel fuel. Although bio-

oil are not miscible with hydrocarbons, they can be emulsified with diesel fuel

by the aid of surfactants (Czernik and Bridgwater 2004). The emulsion is more

stable and has lower viscosity than bio-oil. The physical upgrading technologies,

such as hot gas filtration, solvent addition, or emulsification, could simply

upgrade the fuel properties, while the quality is not enough, compared with

commercial petroleum. While, catalytic treatments, including catalytic cracking,

esterification and hydrodeoxygenation (HDO), have been pursued to improve

the composition of bio-oil, compared with physical methods (Bridgwater 2012).

Many previous studies have investigated HDO significantly improved the

stability of bio-oil (Venderbosch et al. 2010; Wildschut et al. 2009; de Wild et

al. 2009; De Miguel Mercader et al. 2011; Zacher et al. 2014; Elliott and Hart

2008; Elliott 2013). Hydrotreatment could effectively convert oxygenated

compounds with temperature. The reactivity of oxygenated groups under

hydrotreatment was shown in Figure 1-8 (Elliott 2007).

Page 47: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

26

Figure 1-8. Reactivity scale of oxygenated groups under hydrotreatment

conditions (Elliott 2007; Wang et al. 2013)

Page 48: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

27

Catalytic cracking is usually accomplished deoxygenation through

simultaneous dehydration, decarboxylation, and decarbonylation reactions

occurring in the presence of zeolite catalysts. In the late 1970s, synthetic zeolites

(e.g. ZSM-5) were successfully used to convert oxygenated organic compounds

into hydrocarbons resulting in commercialization of the methanol-to-gasoline

process. This discovery also stimulated research focused on the production of

hydrocarbons from biomass-derived pyrolysis oil and pyrolysis vapors (Qin et

al. 2013). Esterification is a potential route to remove the organic acids in bio-

oil by reacting them with alcohol present in bio-oil or with added alcohol,

resulting in improvement of stability (Miao and Shanks,2009). The overall bio-

oil hydrotreating route is roughly presented in Figure 1-9 (Wang et al. 2013).

Page 49: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

28

Figure 1-9. Reaction routes for the bio-oil hydrotreating (Wang et al. 2013)

Page 50: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

29

3.3.2. Hydrodeoxygenative upgrading process

HDO is one of the most effective hydrotreatment process using hydrogen

to reduce the high oxygen content in bio-oil in the presence of suitable catalysts

such as sulfided CoMo or NiMo (Ohta et al. 2011). Via HDO, bio-oil is

effectively reformed, not only removal of the oxygen in bio-oil, but also

improvement of bio-oil stability by converting acid, aldehydes and alcohols and

unsaturated compounds into some more stable ones (Bridgwater 2012). HDO is

a low-hydrogen consumption process compared with other hydroprocesses as

well as a method for manufacturing a more useful product from bio-oil (Elliott

2007). Furthermore, the process is effective in stabilizing pyrolytic lignin,

which makes up approximately 20% of bio-oil, and in enhancing bio-oil

stability with supercritical ethanol (Zhang et al. 2011; Bayerbach and Meier

2009). Thus, modification of pyrolytic lignin and oxygenated compound during

storage might be accomplished for assessing bio-oil stability. Non-volatility and

thermal instability of pyrolytic lignin make bio-oil reforming difficult

(Bayerbach and Meier 2009). Among lots of reforming technologies, HDO with

supercritical ethanol with the proper catalyst can convert pyrolytic lignin into

stable and noncorrosive organic compounds consisted of liquid phase (Tang et

al. 2010). Generally, hydrogenation/HDO reactions were suggested to be in

competition with the repolymerization reactions during hydrotreating of bio-

oils. At the temperature up to 250 °C, parallel reactions like repolymerization,

decarboxylation, and hydrotreating took place. Since the polymerization

reaction rate was faster than that of hydrotreating reactions, the high molecular

weight intermediate fractions were formed char or coke during the reaction.

Page 51: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

30

However, bio-oil components and the high molecular weight fraction could be

hydrotreated or deoxygenated to produce stable bio-oil when H2 and a proper

catalyst were available (Al-Sabawi and Chen 2012; Wang et al. 2013). However,

the coke formation during the reaction could deactivate the catalysts, and make

it hard to convert bio-oil into hydrocarbon, directly. Therefore, the alternative

processes, such as two-step HDO are needed for the production of hydrocarbon

fuel from bio-oil. According to Elliott et al. (Elliott and Hart 2008), two-step

HDO was effective to obtain suitable hydrocarbon fuels. During the two-step

HDO, the partial oxygenated acids, aldehydes, or furans can be removed under

mild condition. Then, the stabilized bio-oil is performed deep HDO at severe

condition, to improve the fuel quality and convert the compounds into

hydrocarbon. Elliott and Baker (Elliott and Baker 1989) initially accomplished

the two-step hydrotreating with CoMo/Al2O3 catalysts. However, the product

had ca. 30 wt% of oxygen level, and the coke formation problem was also

remained. After that, Samolada et al. (Samolada et al. 1998), Wildschut et al.,

Venderbosch et al.(Venderbosch et al. 2010), and Xu et al.(Xu et al. 2013), etc.

performed two-step HDO of the bio-oil. However, the coke formation, 6-11 wt%

of oxygen level, as well as a partial conversion into hydrocarbon had remained

as a problem. Therefore, the second step, which rendered phenolic compounds

into the hydrocarbon fuels, is still a challenge (Crossley et al. 2010; Yan et al.

2010).

Page 52: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

31

Figure 1-10. Typical reaction routes for HDO of bio-oil (Wang et al. 2013; Diebold 2000)

Page 53: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

32

3.3.3. Various factors affected to the hydrodeoxygenation

Hydrotreating become a promising modern refinery process, for example,

the process that involves fixed beds of Ru-based catalysts and conventional

petroleum hydrotreating catalysts is currently developed (Snowden-Swan et al.

2016). However, it is difficult to control, due to the various factors affected in

the process. There are numerous data available in the literature concerning

hydrotreating of model compounds, but most of them are often given for the

catalysts that differ in their composition. Moreover, the operating conditions as

well as reactor types are also differed. There are the factors that concern to the

hydrodeoxygenative upgrading of bio-oil; temperature, pressure, catalyst,

solvent, or reactor type.

Various catalysts with different active phases, promoters, and supports

have been studied in HDO of bio-oils and model compounds. First of all, the

homogeneous catalyst is more effective than heterogeneous one. Moreover, due

to catalyst recovery, heterogeneous catalyst is preferred. Metals are present as

zerovalent metal, sulfide, oxide, and others and mono- or bimetallic form.

Supports are generally referred carbon, silica, alumina, zirconia, titania,

amorphous silica-alumina, and zeolites. Typical hydrotreating catalysts were

listed in Table 1-2. Generally and widely used sulfided metal catalysts such as

NiMo and CoMo show good performance in the enhancement of bio-oil.

However, they have problems that addition of sulfur into bio-oil, fast

deactivation of catalyst by poisoning and contamination of water while washing

the reactor. Noble metal catalysts, generally dispersed on zeolites such as ZSM-

5 or active charcoal supports, are efficiently applied to the HDO of bio-oil.

Page 54: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

33

However, the noble metal phase is expensive, and its utilization affects the

economy of the HDO reaction (Narani et al. 2015; Yan et al. 2010). Also, liquid

acid catalyst, e.g. H3PO4, catalyzed dehydration of cycloalkanol to cycloalkene.

However, the recovery of the acid from the reaction mixture is difficult (Ohta

et al. 2011). Acids are also considered as major reactive compounds, addition

problems such as increasing instability and acidity. Thus, non-sulfided and acid-

free catalysts, e.g. novel metal catalysts, bimetallic catalysts with activated

carbon or aluminum oxide supported catalysts, are now on researching and

show effective results (Wildschut et al. 2010). Therefore, noble metal catalyst

is advantageous to HDO because it has highly active and stable toward HDO

with high hydrogenation properties. Also, those noble metal catalysts were less

deactivated than generally used the metal sulfides. Supported non-noble metal

catalysts, such as Fe/SiO2, Co/SiO2, and Ni-based materials, have been recently

used in HDO studies. Although Ni-based catalysts show greater activity than

other catalysts, coke deposition is still a problem for bio-oil upgrading (Moulijn

et al. 2001; Huynh et al. 2014). According to Wang (Wang and Rinaldi 2012),

Raney nickel and other nickel catalysts can catalyze the transfer hydrogenation

of ketones, phenols, aldehydes, olefins, and aromatic alcohols with 2-PrOH as

an H-donor. Ni2P was also known as the effective HDO catalyst among metal

phosphides (Ni2P, Co2P, WP, MoP, and NiMoP) (Li et al. 2011; Zhao et al. 2011).

But they might be oxidized by water to form phosphate, which cover the active

sites and deactivate the catalyst active site (Li et al. 2011). Therefore, recently,

some researches related to bifunctional catalysts (metal and acid function) were

conducted and showed improved HDO activity compared with metal catalyst

(Zhao et al. 2011; Zhao and Lercher 2012). However, similar in Al2O3, acidic

Page 55: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

34

features could easily form the coke and become unstable under presence of

water (Centeno et al. 1995; Furimsky and Massoth 1999). To avoid these,

neutral form support (activated carbon), TiO2, ZrO2, mesoporous materials

(MCM-41, SBA-15), or zeolites (Hβ, HY, and HZSM-5) have been used (Elliott

et al. 2009; Wildschut et al. 2009; Wildschut et al. 2010; Wildschut et al. 2010;

Elliott and Hart 2008; Bui et al. 2011; Chiranjeevi et al. 2008; Zhu et al. 2011;

Hong et al. 2010). Mesoporous silica materials have noted as catalyst supports

due to their high specific surface areas, narrow range of pore size distributions,

and large pore volumes. Therefore, mesoporous silica supports are desirable

alternative silica supports based on their control of catalyst particle size and

influence in product selectivity through transport effects (Ghampson et al. 2012).

Alumina is occasionally added to produce structures with silica mesoporous

supports in order to assign acidic sites that can act as active sites. To create

catalytic active sites for mesoporous silica supports, the supports are doped with

metal. Among the mesoporous silicate supports (M41S, SBA, MSU, HMS, etc.),

SBA-15 has the largest pore, with hexagonally regio-regular arranged

mesopores, adjustable pore size, and high thermal stability (Lu et al. 2010;

Lestari et al. 2010; Choudhary and Phillips 2011). Moreover, SBA-15 was

found to be significantly more effective in deoxygenation than the HMS-

supported catalyst. Supporting these active phases in meso-structured supports

might contribute to the complete utilization of their catalytic properties by

promoting dispersion and increasing the number of active sites. Based on the

catalytic researches, Pacific Northwest National Laboratory (PNNL) and other

entities have led to the development of a general multi-stage catalytic

hydrotreating scheme for the successful upgrading of bio-oil into transportation

Page 56: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

35

fuels (Zacher et al. 2014). Including conventional petroleum hydrotreating

catalysts such as NiMoS and CoMoS, and newer catalysts such as Ru/C, several

catalysts have been studied for estimating the suitability for process (Zacher et

al. 2014; Elliott 2007). This research at PNNL has focused on improving reactor

stability and testing and developing improved catalysts to achieve optimal

technical and cost performance (Snowden-Swan et al. 2016).

Page 57: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

36

Table 1-2. Typical hydrotreating catalysts

Catalog Catalysts

Mo-based sulfides bulk or supported MoS2, NiMoS2, and CoMoS2

Nobel metals supported Ru, Rh, Pd, Pt, Re, PtRh, PdRh, PdCu,

PdFe, PtRe, and RuNo; Ru, Pt, Rh nanoparticles

Base metals supported Cu, Ni, NiCu, and NiFe and Raney

nickel

Metal phosphides supported Ni2P, MoP. NiMoP, CoMoP, Fe2P, WP,

and RuP

Other metal catalysts bulk NiMo-B; supported nitrides (Mo2N) and

carbides (Mo2C); supported Mo-based oxide

(MoO2, MoO3)

Bifunctional catalysts noble metal or base metal catalysts with aqueous

acid (CH3COOH, H3PO4), Nafion, solid acid

(HZSM-5, Hβ, HY, sulfated zirconia), or bulk

acidic salt

Page 58: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

37

The organic solvent is added to the bio-oil during the HDO reaction, to

reducing the viscosity of the bio-oil, the presence of the solvent reduces the

formation of coke, which disturbs the reaction and can deactivate the catalyst.

Generally, hydrogen-donating hydrocarbons such as tetralin have been used as

HDO solvents (Xu et al. 2014). However, owing to the hydrophobicity of these

hydrocarbons, they scarcely blend with the bio-oil and improve its

characteristics only somewhat. Supercritical fluid has also been used as an HDO

solvent; its gas-like characteristics (high diffusion rate, low viscosity) and fluid-

like characteristics (high density) allow enhancements in heat and mass transfer

as well as increased concentrations of solvent around a solute (Mazaheri et al.

2010; Antal et al. 2000). Supercritical alcohols such as methanol and ethanol

and supercritical acetone have been reported to be effective in biomass

liquefaction and diesel production by means of esterification (Chen et al. 2015);

these solvents are also known as hydrogen donors, and thus they assist the

hydrogenation of unsaturated compounds in bio-oil during HDO (Huang et al.

2014). Solvents can be categorized according to their polarity (polar protic,

polar aprotic, and nonpolar), and these different solvents undergo different

reactions during the processes (Mazaheri et al. 2010). Additionally, it has been

confirmed that the hydrogen donor during the reaction was hydrogen gas, not

ethanol (known as hydrogen donor), according to the preliminary HDO of bio-

oil and ethanol mixtures with and without added hydrogen gas.

Page 59: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

38

Figure 1-11. Phase diagram (pressure and temperature) for a pure compound in

a close system

Page 60: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

39

3.4. Biofuel application

3.4.1. Blend methodology

The United States set biofuels blending volumes to 49.0 billion liters for 2010,

increasing to 136 billion liters by 2022, based on the Renewable Fuels Standard-

2. Started with the United States, similar guidelines that the need of easily and

accurately measuring biofuel blends comes from other countries, and becomes

more important (Zhang 2012).

Bi-fueling or blending is the simplest technique for admitting low cetane

fuels in high compression engines (Patel et al. 2012). Blending with other

biofuels/additives could accommodate the engine requirements with allowing

the utilization of low-cost large-scale mass production diesel engines with only

minor modifications as the dual injection system even with the bio-oil and

petroleum diesel blends (Czernik and Bridgwater 2004; Chiaramonti et al.

2003a, 2003b; Ikura et al. 2003). The optimization of injection timing, injection

pressure, compression ratio and load are also needed after blending to utilization

of blend into the engines (Patel et al. 2012). Generally, the biodiesel was

produced from raw jatropha and karanj oil, and its blends with diesel were tested

for power generation in a 7.5 kVA diesel engine generator set (Kalbande et al.

2008). With the splash blend method, often used for blending biodiesel, the

petrodiesel fuel and B100 (i.e., 100% biodiesel) are pumped separately into a

transport or storage tank, and it is assumed the blend will have been adequately

mixed in the tank during the delivery period. For better blend consistency, in-

line (injection) blending is offered for ethanol blending at pipeline racks and

terminals (Zhang 2012). ASTM D975 method currently allows up to 5%

biodiesel in diesel without the requirement for biodiesel content labeling. As a

high amount of biofuel enters the power generator, knowing the ratio of

Page 61: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

40

biodiesel/diesel or ethanol/gasoline blend becomes more important for

operating. Most of the studies were focused on the biodiesel/diesel blends.

Currently, utilization of bio-oil in engines required preheating prior to

combustion and flushing by diesel or methanol after using to prevent corrosion

and scale formation. These requirements resulted in a complex start-up and shut

down process, as well as modifications of the engine. The ignition process of

the fuel spray in diesel engines depends on a various parameters, such as droplet

size or dispersion properties (Sazhina et al. 2000). Those factors as well as

viscosity were also important in utilization of bio-oil, including diesel engines,

CHP generator, even when it transported through pipelines. To improve its low

mobility and flowability, blend or emulsion with other fuel is required (Aburto

et al. 2009). CANMET developed method for microemulsion production from

bio-oil and diesel oil, and concluded 5-40% bio-oil produced stable emulsion.

Previous studies (Garcia-Perez et al. 2007; Garcia-Perez et al. 2010; Doll et al.

2008; Yang et al. 2013) presented blends of bio-oil with biodiesel. Moreover,

blend of bio-oil with diesel, pyrolytic aqueous phase or methanol were also

studied (Xu et al. 2010; Boucher et al. 2000). The results of bio-oil blends are

important in the utilization of bio-oils as a transportation fuel has been an

elusive goal. The increase viscosity rate of the blends was considerably less than

that of the bio-oil under the same aging conditions (Boucher et al. 2000). While,

Diebold and Czernik (Diebold and Czernik 1997), and Zhang and Wu (Zhang

and Wu 2014) reported that 10-20 wt% of additives increased stability and

reduced initial viscosity during storage. In particular, Ikura et al. (Ikura et al.

2003) has obtained stable blend of diesel and 5-40% bio-oil. The corrosiveness,

heating value, cetane number and viscosity of the obtained emulsion fuel was

improved obviously compared to the bio-oil. Qi et al. (Qi et al. 2008) prepared

the blends of bio-oil and diesel by rapid stirring, but blend remained stable

shortly. By analyzing biodiesel rich fraction, improvement in oxidation stability,

viscosity compared with bio-oil and reduced CO and NOx emissions and smoke

Page 62: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

41

opacity compared with petroleum diesel were obtained. And Dhanajay et al.

reported that 5% bio-oil blends with diesel showed high mechanical efficiency

(Joshi and Patel 2012).

Page 63: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

42

3.4.2. Emulsification to homogenize the blends

Since bio-oil has high polarity and hydrophilic nature, it is immiscible with

non-polar hydrocarbon diesel fuel. This could result the fluctuation of

coalescence, which is destabilized (Figure 1-12). Therefore, when the bio-oil

blends with petroleum diesel, the aid of co-solvents (or surfactant, additives) is

required. Weerachanchai at el. tested alcohol, ethanol and n-butanol, and large

amount of n-butanol proved to be a better co-solvent (Weerachanchai et al.

2009). More studies of bio-oil blends were listed in Table 1-3 (Krutof and

Hawboldt 2016). Bio-oil performed only in low and medium speed diesel

engine, but polar solvent which increase the stability and reduce the viscosity,

reduced in flash point (Czernik and Bridgwater 2004; Diebold 2000; Diebold

and Czernik 1997). Therefore, previous study added few amount of cetane

enhancer and success to ignition in high speed engines almost similar to that of

diesel oil (Maher et al. 2007). Bio-oil can also be emulsified in petroleum based

diesel with non-polar hydrocarbons. Resulted single-phase fuel could ignite in

a compression engine (Chiaramonti et al. 2003a, 2003b; Huber et al. 2006; Ikura

et al. 2003). Chiaramonti et al. reported three types of emulsions by the base

phase of the blends : (1) water-in-oil emulsion (W/O) based on the diesel phase,

(2) oil-in-water emulsion (O/W) based on the bio-oil phase, and (3) bi-

continuous emulsion with 50% of each phase (Figure 1-13). Those emulsions

have different hydrophilic lipophilic balance (HLB), which is important in

decision of emulsifying agent (stabilization the interfacial properties)

(Chiaramonti et al. 2003a; Kunitake et al. 2016). Emulsion, obtained from the

blends of biofuel, petroleum and emulsifier, could be presented as a phase

diagram, which is exampled in Figure 1-14. The left to dashed line of phase

diagram presented those three emulsion types. Generally less than 5 wt% of

emulsifier was added due to the viscosity control, than the type of it was

depended on the biofuel quality, which is representative to HLB. The

Page 64: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

43

emulsifiers can be derived from fatty acids and polyoxyethylene glycol, sorbitol,

or polyethoxylated alcohols wit long aliphatic chains, and representative

emulsifiers were presented in Table 1-3 (Diebold 2000; Chiaramonti et al.

2003a).

But according to Ikura et al., none of the emulsions were remaining

homogeneous phase after two year, due to the emulsifier reacted with the

compounds in bio-oil during storage (Ikura et al. 2003). However, the blends of

bio-oil with diesel still have problems like requirement of large percent of diesel,

addition of surfactant, as well as occurring corrosion after used. Moreover, high

water content and aging performance of bio-oil resulted in poor stability of

emulsion and necessity of fresh bio-oil production (Krutof and Hawboldt 2016;

Chiaramonti et al. 2003b).

Page 65: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

44

Figure 1-12. Mechanisms of emulsion breaking: (a) stable state emulsion; (b) creaming; (c) Ostwald ripening; (d) flocculation;

(e) coalescence; (f) breaking (Chiaramonti et al. 2003a)

Page 66: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

45

Figure 1-13. Scheme of emulsions blended by the HLB (Kunitake et al.,2016; Chiaramonti et al. 2003a)

Page 67: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

46

Figure 1-14. Phase diagram for the Biofuel, Petroleum, Emulsifier ternary

system (Chiaramonti et al. 2003a)

Page 68: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

47

Table 1-3. Previous studies of bio-oil blends

Bio-oil type Blend with Blending process Additives Ref.

Bio-oil from mallee

eucalyptus and pine

wood

Biodiesel

derived from

canola oil Heated at 60 °C for 30 min then

cooled to room temperature

Ethyl acetate/biodiesel solution to enhance

solubility of pyrolysis oils in biodiesel

(Garcia-Perez et

al. 2007)

Bio-oil from pine wood

Biodiesel

derived from

poultry fat

5 wt% methanol added to each blend to

improve thermal stability

(Garcia-Perez et

al. 2007; Garcia-

Perez et al. 2010)

Wood bio-oil Jatropha methyl

ester biodiesel

Addition of surfactant,

mechanical blending at 1500 rpm

for 30 min

2 vol% surfactant Span20 sorbitan monolaurate

(Prakash et al.

2013)

Bio-oil derived from

softwood and hardwood Diesel

Emulsification at 60-65 °C then

mechanically treated

Non-ionic block polymers (or homopolymers)

surfactants with a HLB from 4 to 18

(Chiaramonti et al.

2003a, 2003b)

Hardwood bio-oil Diesel

10-30 wt% of bio-oil in diesel

(W/O emulsification)

Hypermer B246SF and Hypermer 2234

surfactants CANMET’s CETC surfactant

(Ikura et al. 2003)

Slow pyrolysis bio-oil Diesel Alcohols Ethanol and n-butanol (Weerachanchai et

al. 2009)

Page 69: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

48

Table 1-4. List of the most significant surfactants. (Chiaramonti et al. 2003a)

Emulsifier Producer Stab* Emulsifier Producer Stab Emulsifier Producer Stab

Ampholak Akzo-Nobel 0 Carbopol BFGoodrich 0 Marasperse Reed Lignin Inc. 0

Amphoteen Berol-Nobel 1 Carbonwax Carlo Erba 0 Morwet PC

Petrochemicals

1

Armotan Akzo-Nobel 1 Chimipal Cesalpina 0 PEMULEN BF Goodrich 0

Armoteric Akzo-Nobel 0 CTAB Fluka 0 OMA-4 Akzo-Nobel 0

Atlas Uniqema 7 Dapral Akzo-Nobel 2 Reodal Dalton 1

Atlox Uniqema 7 Dowfax Dow 1 Rolfor Cesalpina 0

Atsurf Uniqema 3 Ethomeen Akzo-Nobel 1 SCS Uniqema 2

Bevaloid Rhodia 2 Hypermer Uniqema 5 Span Fluka 1

Bretax CTP 0 Igepal Aldrich 1 Syn-Fac MilliKen Chem. 0

Brij Aldrich 1 Kelzan Kelco 0 Synperonic Uniqema 2

*: Stab meant emulsion stability tested at room temperature, and presented the phase separated days

Page 70: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

49

3.4.3. Biofuel application in commercial engine

The development of the biorefinery framework for bio-oil utilization into

biofuels is an effective way to lower petroleum dependency when the biofuels

required the properties listed in Table 1-5 (Smith and Consultancy 2007; Ng and

Sadhukhan 2011; Marketing 2007). In recent years, roadmaps for exploiting

biomass resources and deploying bioenergy technologies have been developed

by various countries (Gonzalez-Salazar et al. 2016; Soriano and Mulatero 2011;

Amer and Daim 2010; Ou et al. 2010; Bauen et al. 2013); (1) global technology

roadmaps on biofuels for transport, (2) bioenergy for heat and power, (3)

European Union roadmaps on biomass technology , (4) biofuels for transport,

(5) biogas, (6) United States roadmaps on bioenergy and bio based products, (7)

algal biofuels technology, (8) a roadmap for sustainable aviation biofuels for

Brazil, (9) China roadmaps on biomass energy technologies, (10) rural biomass

energy, and (11) a roadmap for biorefineries in Germany.

The blends of biodiesel with diesel were tested for power generation in a 7.5

kVA diesel engine generator set. The overall efficiency of the generator for

4,500 W loading condition of fuel blends was recorded in the range of 21-27%,

while 6,000 W loading conditions was improved for fuel blends were found in

the range of 31-39% (Kalbande et al. 2008). Moreover, the blends were used to

run electricity generator for power generation. According to previous study

(Kalbande et al. 2008), biodiesel blends and petroleum diesel at 4500 and 6000

watt electricity generator resulted in similar overall efficiency ranged 21.5-23.2,

and 31.1-33.8%, respectively. The overall efficiency of electricity generation

was compared between biodiesel blends and petroleum diesel, by the following

equation:

Overall efficiency = [Energy output (kWh)/Energy input (kWh)] × 100

Page 71: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

50

Table 1-5. Properties that affected to the properties and the requirements (Marketing 2007)

Property Effect on performances Time frame of effect Test method

Standard

for

petroleum

Standard

for biofuel

(Biodiesel)

Flash point (°C) Safety in handling and use - D 93 52 93

Water and Sediment (%) Fuel filters and injectors Long-term D2709 0.05 0.05

Volatility (90% volume

recovered °C) Ease of starting and smoke Immediate D86 300-356 360

Viscosity (mm2/s) Fuel spray atomization and fuel

system lubrication or leakage Long-term and Immediate D 445 1.9-4.1 1.9-6.0

Ash (%) Damage fuel injection system and

cause combustion chamber deposits Long-term D 482 0.01 0.02

Sulfur (ppm) Particulate emissions, cylinder wear,

and deposits

Particulates: Immediate

Wear : Long-term D5453 15 15-500

Cetane number

Measure of ignition quality – affects

cold starting, smoke, combustion,

and emissions

Immediate D613 40 47

Carbon residue (%) Measures coking tendency of fuel;

may relate to engine deposits Long-term D524 0.35 0.05

Heating value (MJ/kg) Affects fuel economy Immediate DIN 51900 45 35

Density (at 15°C, kg/m3) Affects heating value Immediate EN ISO 3675 820-845 860-900

Stability (43°C, weeks)

Indicates potential to form insoluble

particles and gum/residues in the

fuel during use and/or in storage

Immediate and Long-term D4625 24 -

Lubricity (at 60°C,

micron)

Affects fuel injection system (pump

and injector) wear

Moderate: Long-term

Severe: Short-term D 6079 520-560 314

Page 72: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

51

Thus, biodiesel blends has disponibility toward all types of compression

ignition engines for electricity generation.

Bio-oil, obtained from fast pyrolysis, has lower fuel quality and stability than

those of petroleum, but has a capacity to use as a fuel in boilers, diesel engines,

or gas turbines for the production of heat, power or combined heat and power

(CHP). Differed from biodiesel, the use of bio-oil in diesel engine required

modification of the bio-oil quality, or the engine system (Chiaramonti et al.

2003b). For example, gas turbines (Andrews et al. 1997; Overend and Chornet

1997), and diesel engines (Leech 1997), have been tested with bio-oil (Iwamoto

et al. 1997). The utilization method is modified to use a high compression ratio,

or design the dual injection into the combustion chamber. The availability of

bio-oil for the generation of transportation fuels and CHP was reported (Ng and

Sadhukhan 2011). Previous studies performed various pyrolysis oil combustion

demonstrations, such as boilers, diesel engines, and gas turbines (Czernik and

Bridgwater 2004; Shaddix and Hardesty 1999). Through various technologies,

the bio-oil quality is improved, and achieves the goal of using it as

transportation fuel, presented high engine performance characteristics and

energy efficiencies (Sipilä et al. 1998; Chiaramonti et al. 2003b; Williams and

Horne 1994). The lubricity of the blend may lead to wear and damage of diesel

engines, primarily fuel injection systems (Goodrum and Geller 2005; Sulek et

al. 2010; Anastopoulos et al. 2002). Suarez et al. (Suarez et al. 2009) compared

the lubricity of bio-fuels from pyrolysis and alcoholysis of soybean oil and their

blends with petroleum diesel. It was observed that the lubricity of the blend was

enhanced when either biodiesel or pyrodiesel were added. Putun et al. (Pütün et

al. 2004) studied the lubrication and tribological performances of the blend of

bio-oil and diesel. The lubricity of the blend was of the better friction reduction

performances, worse wear resistance than the commercial number zero diesel

fuel (Xu et al. 2010). The phenolic compounds in bio-oil were extracted to

biodiesel could utilize bio-oils as additives for transportation fuels, while also

Page 73: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

52

improving the oxidation stability and cold flow properties of biodiesel. Manuel

et al. (Garcia-Perez et al. 2007) studied how much and what fractions of bio-oil

are soluble in biodiesel, as well as using bio-oil as a diesel extender or

replacement in only slightly modified diesel engines. Soluble fractions of bio-

oils in biodiesel blends showed only small increases in density and viscosity

and a small change in the calorific value and solid residues. The oily bottom

phases of the blends were considerably more soluble in biodiesel than the polar

oils obtained from the aqueous phases.

The weaknesses of electrochemical energy storage relative to conventional

petroleum-based fuels include low specific energy, low energy density and low

refueling/recharging rate (Ronning 1997). Bradly et al. (Bradley and Frank 2009)

studied about plug-in hybrid electric vehicles (PHEVs) that used both

electrochemical energy storage and a conventional fuel to overcome these

drawbacks and to provide additional benefits to the consumer and society.

Therefore, the interest in PHEVs is growing (Graham 2001), and allowing a

PHEV to use both electrical energy and petroleum as a transportation fuel, with

the benefits of high energy efficiency, low carbon emissions, cost, and

improvement of transportation energy sector sustainability (Bradley and Frank

2009).

Moreover, EC-ALTENER project analyzed the possibility of bio-oil

utilization in European heat and power generation with the data from Aug. 2002,

to evaluate the opportunities for bio-oil in the heat and power markets of Europe.

As a result, they suggest the complete set of standard bio-oil application, and

shown in Table 1-6 (Brammer et al. 2006). The project is related to bio-oil

systems for heat and power. The evaluation of opportunities is carried out by a

quantitative assessment of the economic competitiveness of a set of standard

applications in 14 European countries which are members of the PyNe biomass

pyrolysis network (Brammer et al. 2006)

Considering with bio-oil production cost, investment cost, operation and

Page 74: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

53

maintenance cost as well as energy input and output, heat-only applications

were the most economically competitive, followed by CHP applications, with

vary rarely competitive in electricity-only applications. In detail, the boiler for

heat applications showed on average the best economic competitiveness across

Europe, followed by the internal combustion (IC) engines for CHP and the

Rankine cycle for CHP (Brammer et al. 2006).

Page 75: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

54

Table 1-6. Set of standard bio-oil application (Brammer et al. 2006)

Technology Product Rated outputa

Bio-oil boiler Heat 0.5, 1.0, 2.0 MWth

Bio-oil IC engine (dual fuel diesel)

Electricity 1.0, 5.0 MWe

CHP 1.0, 5.0 MWe

2.0, 10.0 MWth

Bio-oil gas turbine

Electricity 5.0, 15.0 MWe

CHP 5.0, 15.0 MWe

10.0, 25.0 MWth

Bio-oil gas turbine combined cycle

Electricity 15.0 MWe

CHP 15.0 MWe

25.0 MWth

Bio-oil boiler (Rankine cycle) CHP 15.0 MWe

25.0 MWth

a : Rated heat output is given as MWth, rated electrical (or power) output as MWe.

CHP assumed that the heat is supplied as hot water, typically for space heating

purposes.

* Standard applications were evaluated for the Austria, Belgium, Denmark, France,

Finland, Germany, Greece, Ireland, Italy, Netherlands, Norway, Portugal, Spain and

UK

Page 76: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

55

Chapter 2

Solvent effects on improvement of

bio-oil features during hydrodeoxygenative

upgrading

Papers published in Fuel 182 (2016) 154-160

and Energy 113 (2016) 116-123

Page 77: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

56

1. Introduction

Demand for fossil fuels has been consistently increasing. Particularly, the

consumption of petroleum in 2030 might increase by up to 1.5 times relative to

the consumption in 2004 (Williams 2005). This trend is causing problems such

as the depletion of fossil fuels, hikes in oil prices, and global warming.

Consequently, alternative energy sources are required, particularly liquid-phase

bioenergy sources. Through the biomass to liquid (BTL) process, biomass can

be converted into liquid fuel. Among thermochemical BTL processes, fast

pyrolysis has advantages including high yield of bio-oil, easy processing, and

short reactions that do not require pretreatment (Bridgwater 2012). However,

the bio-oil has disadvantages (high water content, high oxygen content, high

viscosity, and low heating value) that result in aging, corrosion, and injection

problems. To overcome those limitations of bio-oils and make them useful in

existing engines, bio-oil upgrading is essential.

The hydrodeoxygenation (HDO) upgrading process has been widely

researched and is well known as an effective way to improve bio-oil quality and

stability. One of the disadvantages of bio-oil is its high viscosity, which can

cause nozzle plugging or injection problems when it is used as fuel. Furthermore,

the viscosity can increase with storage time, making these problems quite severe.

In previous research, the strategy has been developed of adding organic solvent

to the bio-oil during the HDO reaction; in addition to reducing the viscosity of

the bio-oil, the presence of the solvent reduces the formation of coke, which

disturbs the reaction and can deactivate the catalyst. The solvent can react as a

co-reactant, and prominent composition of light oil at some conditions.

Generally, hydrogen-donating hydrocarbons such as tetralin have been used

as HDO solvents (Xu et al. 2014). However, owing to the hydrophobicity of

these hydrocarbons, they scarcely blend with the bio-oil and improve its

Page 78: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

57

characteristics only somewhat. Supercritical fluid has also been used as an HDO

solvent; its gas-like characteristics (high diffusion rate, low viscosity) and fluid-

like characteristics (high density) allow enhancements in heat and mass transfer

as well as increased concentrations of solvent around a solute (Mazaheri et al.

2010; Antal et al. 2000). Supercritical alcohols such as methanol and ethanol

and supercritical acetone have been reported to be effective in biomass

liquefaction and diesel production by means of esterification (Chen et al. 2015);

these solvents are also known as hydrogen donors, and thus they assist the

hydrogenation of unsaturated compounds in bio-oil during HDO (Huang et al.

2014). Solvents can be categorized according to their polarity (polar protic,

polar aprotic, and nonpolar), and these different solvents undergo different

reactions during the processes (Mazaheri et al. 2010). Additionally, it has been

confirmed that the hydrogen donor during the reaction was hydrogen gas, not

ethanol (known as hydrogen donor), according to the preliminary HDO of bio-

oil and ethanol mixtures with and without added hydrogen gas.

The large amounts of volatiles and many reactive oxygenated compounds

including acids, aldehydes, sugars, and alcohols in bio-oil produce a glue-like

material when stored for extended periods at ambient temperature, which leads

to instability (Czernik and Bridgwater 2004; Diebold and Czernik 1997;

Boucher et al. 2000; Mante and Agblevor 2012). Moreover, polymerization,

esterification, and etherification of these compounds can increase the viscosity

of the bio-oil during storage, resulting in injection problems when used as a

biofuel (Boucher et al. 2000). Therefore, maintenance and improvement of the

properties and stability of biofuels are the main factors regarding their

introduction into the market (Diebold 2000). Catalytic treatments have been

pursued to improve the composition of bio-oil. Many previous studies have

investigated various upgrading processes, including HDO, to significantly

improve the stability of bio-oil (Venderbosch et al. 2010; Wildschut et al. 2009;

de Wild et al. 2009; De Miguel Mercader et al. 2011; Zacher et al. 2014; Elliott

Page 79: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

58

and Hart 2008; Elliott 2013). In general, HDO is a low-hydrogen consumption

process compared with other hydroprocesses as well as a method for

manufacturing a more useful product from bio-oil (Elliott 2007). Furthermore,

the process is effective in stabilizing pyrolytic lignin, which makes up

approximately 20% of bio-oil, and in enhancing bio-oil stability with

supercritical ethanol (Zhang et al. 2011; Bayerbach and Meier 2009). Thus,

modification of pyrolytic lignin and oxygenated compound during storage

might be accomplished for assessing bio-oil stability.

Diebold presented diverse chemical reactions expected to occur in bio-oil via

the catalytic reaction, but these reactions also occurred in the absence of

catalysts when bio-oil was stored at room temperature for a long time (Diebold

2000). Chemical reactions such as esterification of acids with alcohols,

aldehyde polymerization, hydration of aldehydes or ketones with water,

hemiacetal or acetal formation between aldehydes and alcohols, resin formation

from aldehydes and phenols, olefin condensation, oxidation of alcohols and

aldehydes to form carboxylic acids, and oxidative esterification of aldehydes

were suggested to occur in bio-oil during storage (Mante and Agblevor 2012).

These reactions are involved in property changes such as viscosity. Maintaining

a lower viscosity is important for atomization and injection.

Therefore, in this study, the effects of the different types of solvent upon the

HDO of bio-oil; namely, ethanol (polar protic), acetone (polar aprotic), and

diethyl ether (ether, non-polar) solvents as well as the improved stability of bio-

oil after hydrodeoxygenative upgrading were studied. Especially, the physico-

chemical features of viscosity, water content, acidity, and degree of deoxy-

genation were compared. Furthermore, the economic factors of energy

efficiency and carbon recovery were also compared. Then, heavy oil, bio-oil

upgraded via the hydrodeoxgygenative process, was stored at 23 °C for 2, 4, 8,

or 12 weeks. After storage, characterization of the stored heavy oil properties

and the amounts of pyrolytic lignin were analyzed.

Page 80: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

59

2. Materials and methods

2.1. Fast pyrolysis for bio-oil production

Miscanthus, which belongs to the genus of C4 perennial grasses, is

considered a promising bio-energy crop due to high growth rate and low

nutritional requirements. Moreover, it has additional advantages like high

biomass yield, high energy density, low soil erosion, low establishment costs

and low water content compared to other biomass sources (Lewandowski et al.

2003). Therefore, Miscanthus sinensis was selected as a raw material in this

study. Holocellulose (the sum of cellulose and hemicellulose), lignin and ash

contents were determined by NREL procedures (Sluiter et al. 2008). Elemental

analysis was performed using a Flash EA 1112 instrument (Thermo Electron

Corporation, USA). Compositional analytical data are presented in Table 2-1.

Miscanthus sinensis was ground to pass a 2 mm sieve and then subjected to

fast pyrolysis. Fast pyrolysis was carried out at 500 °C; the residence time was

maintained at 1.63 s by means of a flow of inert nitrogen gas. 2 kg of biomass

was fed into the fluidized-bed fast pyrolysis reactor (Figure 2-1), for 33.3 g/min.

The flow rate of the nitrogen gas in the reactor was 50 L/min. The volatile

pyrolytic products were cooled and condensed to a liquid bio-oil in the yield of

48.3 wt% (Bok et al. 2013).

Page 81: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

60

Table 2-1. Chemical compositions of Miscanthus

Sample

Elemental composition (%) Extractives

(%)

Holocellulose

(%)b

Lignin

(%)b

Ash

(%) C H N Oa S

Miscanthus 58.9 5.4 0.7 35.1 0.0 4.4 (0.6)c 70.4 (0.4) 25.8 (0.1) 4.1 (0.1)

a: By difference b: Based on weight of extractive free Miscanthus c: Standard deviation

Page 82: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

61

Figure 2-1. Schematic diagram of the experimental apparatus for the fast pyrolysis ((1) sample hopper; (2) screw feeder; (3)

fluidized bed reactor; (4) cyclone; (5) condensers; (6) electrostatic precipitator; (7) blower) (Bok et al.,2013)

Page 83: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

62

2.2. Selection of solvent as co-reactant

Generally, alcohol solvents are known to serve as hydrogen donors during

HDO. However, the H/C atomic ratio of heavy oil varied from 1.6 (without H2)

to 2.0 (with H2) (Figure 2-2). Since the alcohols were typical polar protic solvent,

the effect of the other categories; polar aprotic, and nonpolar on HDO should

be investigated. Thus, three solvents were selected, one each having polar protic,

polar aroptic, and nonpolar character. Moreover, supercritical fluid has both of

gas, and liquid properties. Supercritical fluid has high diffusion rate, low

viscosity, low surface tension, and high density. Therefore, it has advantages

that the increase the solvent concentration around the solute. Previous study

mentioned that supercritical fluid is effective to improve the bio-oil stability by

removal of pyrolytic lignin (Tang et al. 2010). Each of the solvents was chosen,

which existed in the supercritical phase in the mildest running conditions

(250 °C, 3 MPa) used in this study (Table 2-2). The solvents acted as co-reactant

during the HDO and also affected the quality of the bio-oil. Therefore, HDO of

the solvent itself was accomplished, the features of which are described in Table

2-3.

Page 84: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

63

Figure 2-2. Van Krevelen diagram of bio-oil, N2 (HDO with ethanol, without

hydrogen), and H2 (HDO with hydrogen and ethanol)

Page 85: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

64

Table 2-2. Critical characteristics of solvent and its classification

Solvent Classification Molecular

weight (g/mol)

Critical

temperature (°C)

Critical

pressure (MPa)

Critical density

(g/cm3)

Acetone Polar aprotic 58.08 235.1 4.7 0.278

Ethanol Polar protic 46.07 240.9 6.1 0.276

Ether Non-polar 74.12 194.0 3.6 0.277

Page 86: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

65

Table 2-3. Major products and characteristics of hydrodeoxygenation of solvent with hydrogen pressure

HDO Liquid Yield

(%)

Major liquid

compound

Atomic H/C ratio

of liquid phase

Atomic O/C ratio

of liquid phase

Ethanol 17.6 Ethyl esters 4.1 1.3

Acetone 17.1 Ethane, isopropanol 4.1 2.7

Ether 0.0 - - -

Page 87: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

66

2.3. Hydrodeoxygenation of bio-oil with supercritical fluid

The bio-oil described above was subjected to hydrodeoxygenation under the

temperature of 250, 300, or 350 °C for 45min under 3 MPa of H2. One of three

different solvents (ethanol, acetone, and diethyl ether) was added to each

experimental sample, and Pt/C catalyst was added to each sample; this catalyst

has been previously judged to be effective for deoxygenation and improving

bio-oil quality.

In detail, the bio-oil was subjected to HDO in a batch-type autoclave reactor

composed of a mechanically driven stirrer, an electric heating mantle, and a

cooling coil. The reactor was filled with bio-oil and solvent (ethanol, acetone,

or diethyl ether) in the ratio of 4:1 (w/w) and catalyst of 4 wt% relative to the

bio-oil was also added; the ratio of reactants and catalysts were based on the

optimal ratio of the previous study (Kim et al. 2014). The Pt/C used contained

5 wt% of active metals on active charcoal (Sigma Aldrich, catalog number

205931). The air inside the reactor was first flushed thoroughly with nitrogen

gas, and then hydrogen gas was injected to 3 MPa. According to Elliott et al.,

two-step HDO is effective to obtain suitable hydrocarbon fuels (Elliott and Hart

2008). However, we performed only mild HDO to improve the bio-oil quality

and to investigate the features of the HDO products obtained by means of using

different solvents. Because HDO activated over 250 °C (Gutierrez et al. 2009),

the reactor was heated to the reaction temperature of either 250, 300, or 350 °C

at the average heating rate of 25 °C/min, and was maintained for 45 min at the

reaction temperature. After the reaction, the reactor was then cooled to 30 °C at

the average cooling rate of 10 °C/min. The reaction products obtained included

three phases (liquid, char, and gas), with the liquid phase comprising an

immiscible water-rich solvent-based phase (light oil) and an organic phase of

improved quality (heavy oil). Each product was separated by means of filtering

Page 88: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

67

or use of a separatory funnel and then subjected to further analysis. Experiments

were conducted in duplicate for each condition. The yield of char, light oil, and

heavy oil were calculated by means of the following equations:

Yproduct (%) = Mproduct (g)/ Mbio-oil+solvent (g) 100

Ygas (%) = 100 − [Ylight oil + Yheavyoil + Ychar],

where Yproduct and Mproduct are respectively the yield and mass of the products

(light oil, heavy oil, and char).

Page 89: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

68

2.4. Characterization of hydrodeoxygenation products

2.4.1. Physicochemical properties

Karl-Fischer titration (Titro Line KF, Schott Instruments, Germany) was

carried out to measure the water content of the bio-oil and heavy oils. To

investigate the acidity of the bio-oil and heavy oil, the total acid number (TAN)

was measured according to ASTM D664. The viscosity at 40 °C was determined

using a capillary-type viscometer and a ViscoClock measuring instrument

(Schott Instruments). Elemental analysis of carbon, hydrogen, and nitrogen was

carried out using a LECO CHNS-932 analyzer. The higher heating value (HHV)

was calculated by means of following equation (Friedl et al. 2005). The degree

of deoxygenation (DOD), an important parameter of the HDO reaction, was

calculated by means of the following equation:

HHV (J/kg) = 3.55 [C]2 - 232 [C] - 2230 [H] + 51.2 [H] [C] + 131

[N] + 20,600

Degree of deoxygenation (%) = (MObio-oil MOheavy oil) / MObio-oil 100

where [C], [H], and [N] are the mass% of carbon, hydrogen, and nitrogen on

dry basis, and MOheavy oil and MObio-oil are the molar oxygen / carbon ratio of

heavy oil and bio-oil, respectively.

2.4.2. GC/MS analysis of low molecular compounds

For the qualitative and quantitative analysis of low-molecular-weight

compounds in bio-oil and heavy oils, GC/MS analysis was carried out with

fluoranthene as an internal standard (I.S.). This analysis was carried out using

Page 90: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

69

an Agilent 7890B GC instrument equipped with a DB-5 capillary column (60

m × 0.25 mm × 0.25 um) and coupled with a 5975C mass selective detector

(MSD) and a flame ionization detector (FID). The oven was maintained at 50 °C

for 5 min, heated at the rate of 3 °C/min to 280 °C, and then held at 280 °C for

20 min. The injector and FID detector temperatures were 250 and 300 °C,

respectively. The amount of low-molecular-weight compounds was calculated

by means of the following equation:

Amount of compounds (mg/mL oil) = RF relative area AI.S Voil Yheavy oil / 100,

where the response factor RF depends on the type of compounds, the relative

area is the peak area relative to the I.S. (area of the compounds / area of the I.S.),

AI.S is the mass of the I.S. (mg), Voil is the volume of oil (mL), and Yheavy oil /

100 is the yield of heavy oil. In the case of bio-oil, the value of Yheavy oil / 100

was supposed to be 1.

Each peak present in the chromatograph was identified by reference to the

National Institute of Standards and Technology (NIST) mass spectral library.

Page 91: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

70

2.5. Estimation of process efficiency

The goal of carrying out the HDO reaction was to improve the quality of the

bio-oil to increase the possibility of using it as an alternative energy source in

current engines without requiring design modifications. Consequently, the

energy efficiency of each procedure using the various solvents was calculated

by means of the following equation.

Energy efficiency = HHVheavy oil (Yheavy oil / 100) / HHVBio-oil

Along with energy efficiency, the carbon flow rate is also important, because

condensed organics could be deposited on the catalyst surface, or dealkylated

methane could change to the gas phase. Because the carbon and hydrogen

content determine the HHV, the quantity of carbon remaining in the heavy oil

is important. Therefore, the carbon transformed during the HDO reaction, by

means of dealkylation, decarbonylation, fragmentation, degradation, etc., were

calculated by means of the following equation:

Carbon recoveryproduct = Yproduct Cproduct,

where Yproduct indicates the yield of each product (gas, light oil, heavy oil, and

char) and Cproduct is the carbon content of that product. To compare the efficiency

of each process, the yield of the product was also considered. Thus, the carbon

flow from the bio-oil to the HDO products could be compared, as well as the

carbon recovery of heavy oil under each set of experimental conditions.

Page 92: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

71

2.6. Storage stability for 12 week period

The bio-oil has a low thermal stability, and an increase in temperature showed

a similar effect on bio-oil viscosity as storage. However, thermal aging of oil

might involve other properties or depend on the composition. Therefore, for

stability testing, heavy oil was stored under fixed conditions (23 °C) for 2, 4, 8,

or 12 weeks. Heavy oils (obtained with HDO catalysts of Pt/C) were weighed

(25 g), placed in glass vials for each storage period, and tightly sealed with a

plastic cap. Since bio-oil obtained from pine and oak wood showed stable

viscosity for 60 days (Mante and Agblevor 2012), heavy oils were stored for 3

months (84 days) to assess the role of aging on the stability of the oil. The

weight of heavy oil was measured every two weeks for the first month and then

monthly for the remainder of the storage period in order to determine the

amount of lost volatile compounds. The physicochemical properties and

chemical composition of aged bio-oil and aged heavy oil was analyzed in the

same way of above.

Page 93: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

72

3. Results and Discussion

3.1. Effects of solvents upon the HDO products

The yields of the HDO products based on the weight of the bio-oil are listed

in Table 2-4; these yields varied considerably for the different solvents used.

When ethanol or acetone was used as the HDO solvent, the product yields

showed considerable differences in composition with different reaction

temperatures. As shown in the Figure 2-3, the liquid solvents converted into

supercritical fluids with increasing temperature. Generally, when the initial

liquid phase was heated, thermal expansion occurred and it reduced the density

(Tucker, S.C. 1999). However, in the batch type reactor, this could increase the

pressure and the solvent converted into supercritical phase. Since the high

density accelerated the solubility, this conversion trend supposed that the HDO

of bio-oil effectively performed at the targeted temperature. For these solvents,

the heavy oil yield relative to the weight of the bio-oil decreased with increasing

temperature, from 51.5 to 16.5% for ethanol solvent, and from 62.2 to 26.8%

for acetone solvent. Contrastingly, when ether was used as a solvent, the yield

of heavy oil increased from 38.7 to 47.2% with increasing temperature to 300 °C,

and decreased to 38.3% with further increases in temperature. These results

indicate that the HDO ability of the ether was maximized at a sustained

temperature of about 300 °C, with the slight effect of accelerating further

decomposition as well as of preventing char formation over 300 °C.

With increasing reaction temperature from 250 to 350 °C, the yield of the gas

obtained from HDO when the polar protic solvent ethanol was used increased

steadily from 7.4 to 35.6%. When acetone was used, the yield increased

significantly over 300 °C (from 17.2 to 41.7%). Therefore, it was determined

that the polar solvents (ethanol and acetone) accelerated further decomposition

Page 94: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

73

with increasing reaction temperature. This might be explained as a result of the

reduction in polarity and increase in oil affinity of polar organic compounds

(water-soluble compounds) caused by the HDO.

In the case of ether as the solvent, the yields of the four main products varied

least with the different reaction temperatures used. In particular, the range of

gas yield was 28.6–31.6%, whereas those of the polar solvents were 7.4–35.6%

(for ethanol) and 17.1–41.7% (for acetone). This occurred because the non-

polar ether was entirely hydrodeoxygenated to gas (Table 2-3) whereas slight

amounts of hydrodeoxygenated polar solvent (ca. 17%) remained in the liquid

phase. Thus, the non-polar solvent ether prevented char formation uniformly

with increasing temperature.

Among the three types of solvent used in this experiment, only the polar

protic solvent (ethanol) could undergo hydrogen bonding with hydroxyl groups

in the bio-oil. In the case of the polar aprotic solvent (acetone), hydrogen

bonding might occur by means of hydrogenation of acetone and reduction,

which competes with HDO. As shown in Table 2-4, the yield of light oil, the

portion based on the aqueous solvent, was higher when ethanol was used in the

HDO than when acetone or ether was used. Hydrogen bonding between the

solvent and the bio-oil could accelerate the further decomposition of organic

compounds into the light oil (Joshi and Lawal 2012).

Of the three solvents used in these experiments, the non-polar ether was

found to be the most stable with increasing reaction temperature. However, the

polar aprotic acetone was more effective in preventing char formation. The

unstable changes of product yield obtained from ethanol might have resulted

from hydrogen bonding between the solvent and the bio-oil.

Page 95: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

74

Table 2-4. Mass balance of essential HDO products obtained from various solvents (ethanol, acetone and ether)

Yield

(%, based

on the

bio-oil)

Without

solvent

Ethanol Acetone Ether

250 °C 300 °C 350 °C 250 °C 300 °C 350 °C 250 °C 300 °C 350 °C

Char 25.8 9.4(0.9) * 5.5(1.1) 16.1(0.8) 6.2(2.1) 3.8(0.4) 7.1(0.7) 12.6(1.3) 11.2(0.6) 9.5(0.2)

Light

oil 32.4 31.8(1.5) 41.1(0.3) 31.8(0.5) 14.5(1.0) 27.0(1.6) 24.5(1.5) 19.5(0.1) 13.1(1.6) 20.6(0.0)

Heavy

oil 21.7 51.5(1.5) 41.6(1.4) 16.5(4.4) 62.2(2.6) 52.0(1.4) 26.8(0.7) 38.7(1.5) 47.2(0.5) 38.3(0.1)

Heavy

oil (based

on the

biomass)

10.5 24.9 20.9 8.0 30.0 25.1 12.9 18.7 22.8 18.5

Gas 20.1 7.4(0.8) 11.7(0.0) 35.6(4.1) 17.1(0.4) 17.2(0.6) 41.7(2.8) 29.2(0.3) 28.6(0.6) 31.6(0.2)

a: Standard deviation

Page 96: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

75

Figure 2-3. The reaction pressure of three solvents at each HDO temperature

Page 97: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

76

3.2. Physicochemical properties of liquid products

Properties of the bio-oil such as acidity, heating value, and thermal stability

were improved by hydrocracking, hydrogenolysis, and demoisturization, as

well as by hydrogenation and deoxygenation during HDO (Joshi and Lawal

2012). However, the solvent used in the reaction also acted as a co-reactant, the

improvement in the bio-oil might differ depending on the HDO solvent used. In

this work, to elucidate how the solvent (co-reactant) influences bio-oil quality,

we investigated the differences in physical properties arising from the use of

different solvents, including the water content, acidity (total acid number, TAN),

viscosity, elemental composition, and chemical composition. Bio-oil becomes

phase separated into an aqueous phase (light oil) and an organic phase (heavy

oil) after the HDO. Thus, herein we discuss the properties of both phases.

3.2.1. Heavy oil properties

Several representative physicochemical features of bio-oil and heavy oils

were measured and are displayed in Table 2-5. Even when no solvent was added

to the reactant, the HDO reaction improved the oil properties. However, HDO

did not sufficiently improve the features of viscosity and acidity, which can lead

to major engine problems. Moreover, in the experiment in which no solvent was

added, tar-like sticky char was produced. Therefore, solvent should indeed be

used. Compared with the bio-oil, the heavy oil had reduced water content (from

17.7 to 0.5–13.7 wt%), acidity (from 164.8 to 63.2–126.3 mg KOH/g oil) and

viscosity (from 29.6 to 2.2–6.7 cSt), as well as improved HHV (from 17.3 to

19.2–27.8 MJ/kg).

The water content, which is related to the heating value and affects the

combustion properties when a fuel is used in an engine, decreased from 17.7%

Page 98: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

77

in the bio-oil to 0.5–4.0% in the heavy oil fraction when ethanol was used, to

4.1–13.5% when acetone was used, and to 10.1–13.7% when ether was used,

owing to demoisturization or dehydration of the organic phase.

Demoisturization or dehydration was carried out most effectively in ethanol,

followed by acetone and then ether, but the uniformity of these reactions with

increasing temperature followed the reverse sequence of these solvents, being

greatest in ether. This was due to hydrogen bonding between water and the polar

ethanol or acetone. The non-polar ether does not undergo hydrogen bonding

with water that has been demoisturized from the bio-oil, so the water was not

entirely separated to the light oil when ether was used as the solvent; also, this

was unaffected by the HDO temperature. Demoisturization or dehydration

showed a steady rising trend with increasing temperature. The TAN of the heavy

oil was lowered (to 63.2–126.3 mg KOH/g oil) compared to that of the bio-oil

(164.8 mg KOH/g oil). This followed a similar trend as the water content,

because the phenolic hydroxyl group as well as acids could undergo hydrogen

bonding or reactions with polar solvents. Compared to the polar protic ethanol,

acetone and ether showed less activity in removing acids from the bio-oil (Table

2-6). This is likely because ethanol reacted with acids derived from the

decomposition of carbohydrates, converting them into desirable compounds

such as esters and ketones. However, ether that underwent HDO was converted

into the gas phase, and thus did not react with the compounds in the bio-oil.

Reducing the viscosity is an important point of reforming bio-oil into fuel,

because the high viscosity of bio-oil produces many disadvantages during its

injection into an engine. Aldehydes, and sugars such as levoglucosan, and

nitrogen containing compounds contribute especially to the viscosity of bio-oil

(Oasmaa and Kuoppala,2003; Mante and Agblevor,2012). However, the HDO

reaction converts levoglucosan into esters by means of ring opening,

hydrogenation, decarbonylation, and dehydration. Therefore, the viscosity

decreased from 29.6 cSt for bio-oil to 2.2–6.7 cSt for heavy oil. All heavy oil

Page 99: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

78

fractions had reduced viscosity of over 77.4% compared to that of the bio-oil,

but the trend of changing viscosity also similar with other properties.

The calorific value was increased from 17.7 MJ/kg in the bio-oil to 21.5–27.8

MJ/kg in the heavy oil. The calorific values of all heavy oils were similar

regardless of the reaction conditions. Especially, in the polar solvents, it was

found that certain temperature levels accelerated the deoxygenation of organic

compounds. For instance, ethanol catalyzed deoxygenation even at 250 °C, but

was more effective in accelerating the deoxygenation over 300 °C. In case of

polar aprotic acetone, it presented no significant deoxygenation efficiency.

However, polar aprotic acetone displayed the deoxygenation performance over

that temperature (degree of deoxygenation increased from 14.7 to 65.7–66.5%).

Page 100: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

79

Table 2-5. Typical physicochemical features of bio-oil and heavy oil obtained from various conditions (without solvent,

with solvents such as ethanol, acetone or ether)

Bio-oil No

solvent**

Ethanol Acetone Ether

250 °C 300 °C 350 °C 250 °C 300 °C 350 °C 250 °C 300 °C 350 °C

Water content

(wt%) 17.7 10.3 4.0 3.9 0.5 13.5 8.6 4.1 13.7 12.2 10.1

TAN

(mg KOH/g oil) 164.8 151.7 75.2 63.8 63.2 126.3 113.5 112.6 109.3 107.2 93.5

Viscosity (cSt) 29.6 26.7 2.5 2.2 4.4 3.0 3.5 4.2 6.7 6.5 6.7

Elemental analysis (wt%)

C 41.3 55.1 55.3 57.8 70.7 43.4 64.3 65.2 59.2 59.6 63.2

H 7.4 9.0 9.1 8.6 9.0 3.6 5.9 6.1 4.9 5.1 5.6

N 0.9 0.3 0.0 0.0 0.2 0.2 0.3 0.3 0.3 0.4 0.4

S 0.1 0.1 0.1 0.0 0.0 0.0 0.0 0.0 0.1 0.0 0.1

O* 50.3 35.4 35.5 33.5 20.1 44.5 25.8 26.8 29.1 28.5 24.7

HHV (MJ/kg) 16.5 23.4 23.9 25.1 34.4 17.2 26.1 27.1 22.7 23.1 25.2

Degree of

deoxygenation

(%)

- 44.4 23.0 30.3 78.2 14.7 66.5 65.7 59.1 60.2 67.5

Energy efficiency - 64.3 59.0 65.2 56.1 61.9 76.0 39.8 51.3 63.3 54.6

*: Calculated by difference **: HDO without solvent, with hydrogen pressure

Page 101: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

80

3.2.2. Light oil properties

After the HDO reaction, a solvent-based aqueous phase termed light oil was

obtained. Light oil could also be obtained when solvent was not used in the

reaction, but in this case its color was dark brown, similar to that of the heavy

oil. Also, the organic compounds in the light oil were of similar composition as

those in the bio-oil. Contrastingly, when solvent was used in the reaction, the

light oil was opaque and light yellow, and consisted of some monomers and

water. Figure 2-4 described representative features of the light oils yielded, such

as their water content and acidity. Remarkable features of the light oils included

high water content and acidity. The heavy oils obtained from reactions including

ethanol or acetone had lower water contents compared with that of the reaction

including ether, owing to the effect of hydrogen bonding discussed above.

However, the water content of the light oil was also lower when the polar

ethanol and acetone were used (63.3–65.0 wt%) than when the nonpolar ether

was used (76.6 wt%). This result suggests that dehydration and hydrogenolysis

reaction might be carried out simultaneously in the polar atmosphere (Mane and

Rode 2012). Furthermore, the fact that most ether was converted into the gas

phase also affected the yield and features of the light oil when ether was used

(Table 2-3). It could also be supported by the water content of light oil obtained

without solvent (73.9 wt%). Acidity of light oil also increased (159.2–264.7 mg

KOH/g oil), and exhibited a similar tendency of water content. The high acidity

resulted from the high water content or from the presence of monomeric

phenolic compounds (Oasmaa et al. 2010).

Page 102: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

81

Figure 2-4. Water content, and acidity (TAN) of light oils obtained from HDO

with different solvents

0

50

100

150

200

250

300

0

10

20

30

40

50

60

70

80

90

No solvent Ethanol Acetone Ether

TA

N (

mg

KO

H/g

oil

)

Wa

ter

con

ten

t (w

t%)

Water content (wt%) TAN (mg KOH/g oil)

Page 103: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

82

3.3. Atomic H/C and O/C ratio

The atomic H/C and O/C ratios in the heavy oil resulting from the use of

different HDO solvents are shown in Figure 2-5. Conversion and removal of

unstable acids, aldehydes, sugars, and water might reduce the O/C ratio. The

atomic ratio was calculated by the value of water excluded heavy oil, because

the H/C ratio could also decrease despite hydrogenation. The atomic O/C ratio

was effectively reduced by the HDO reaction, decreasing from 0.9 in bio-oil to

0.2–0.8 in the heavy oil. Similarly, the H/C ratio decreased from 2.1 in the bio-

oil to 1.0–2.0 in the heavy oil owing to the deoxygenation or dehydration. In the

case of using polar protic ethanol, the O/C ratio steadily decreased from 0.6 to

0.3 with increasing temperature, whereas the H/C ratio maintained the value of

0.1. However, the use of the polar aprotic acetone and the nonpolar ether

displayed decreasing O/C and increasing H/C ratios with temperature.

Page 104: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

83

Figure 2-5. Van Krevelen diagram of bio-oil and heavy oils (No solvent: HDO

of bio-oil without solvent, only with hydrogen pressure)

0.0

0.4

0.8

1.2

1.6

2.0

2.4

0.2 0.4 0.6 0.8 1.0

Ato

mic

H/C

ra

tio

Atomic O/C ratio

Ethanol 250 Ethanol 300 Ethanol 350Acetone 250 Acetone 300 Acetone 350Ether 250 Ether 300 Ether 350Bio-oil No solvent

Page 105: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

84

3.4. Compositional changes of liquid products

Thirteen major compounds were identified in the GC/MS analysis and

integrated; Tables 2-6 and 2-7 respectively present the relative amounts of these

compounds in heavy oil and in light oil.

In the heavy oil, most of the compounds derived from lignin fragments,

including phenol, guaiacol, and syringol; acetic acid and cyclopentanone were

also detected. Acetic acid, which was a major factor in increased acidity,

instability, and decreased HHV, was effectively converted into ethyl acetate or

moved into the light oil phase when ethanol was used as the solvent. When

ethanol was hydrodeoxygenated over noble metal catalyst, 72.4% of the ethanol

was converted into gas and only 27.6% remained. The heavy oil fraction

included ethyl esters (such as ethyl acetate and ethyl propanoate). Therefore,

ethanol was considered to act as a co-reactant in the HDO reaction, converting

acids in the bio-oil into esters and thereby reducing the bio-oil’s acidity and

enhancing its stability.

However, trace amounts of acetic acid might not undergo hydrogen bonding

with polar aprotic acetone (Kratzer 1953). Therefore, acetic acid remained in

the heavy oil when acetone was used as the solvent. This trend presented same

in light oil phase. However, the nonpolar ether also converted acetic acid into

ethyl acetate. Furthermore, the amounts of esters increased significantly. This

may have occurred owing to the hydration of ether to form ethanol (Varisli et al.

2007). According to this result, it seemed that the nonpolar ether could undergo

the same reactions as the polar protic ethanol. Most of the phenolic compounds

were retained in the heavy oil when polar aprotic acetone was used. Especially,

guaiacol and syringol contents in the heavy oil were higher when acetone was

used. Therefore, acetone might not accelerate demethoxylation as well as

dealkylation.

Page 106: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

85

Table 2-6. Quantitative analysis of major micromolecules in the heavy oils by GC/MS analysis

RT Compound

Relative area (Compound area/I.S. area)

W/O solvent Ethanol Acetone Ether

250 300 350 250 300 350 250 300 350

3.3 Acetic acid, ethyl ester - 3.9 6.6 0.3 - - - 12.9 14.1 10.2

3.8 Acetic acid 30.2 - - - 6.0 0.7 12.6 18.5 24.5 17.8

6.8 Cyclopentanone 4.8 0.4 0.9 1.0 1.7 1.2 0.5 1.1 0.9 0.7

8.7 2-Methyl-2-cyclopentenone 7.5 1.0 0.3 1.6 1.6 1.2 1.2 1.8 1.9 2.3

15.4 Phenol 16.3 1.2 7.3 5.4 3.3 3.1 4.0 3.9 5.6 8.3

20.2 p-Cresol 10.0 0.7 2.6 6.0 12.6 16.5 25.2 3.4 4.4 5.7

20.7 Guaiacol 10.3 1.7 4.5 4.1 9.9 9.8 20.2 4.9 7.0 3.8

23.8 Creosol 3.0 0.9 0.8 4.8 9.4 1.2 12.1 - 0.6 2.4

26.1 Catechol 7.9 0.3 1.2 1.1 1.4 1.9 5.1 2.5 3.2 2.7

29.9 4-Ethylguaiacol 7.7 1.2 1.8 4.8 4.5 8.8 5.2 2.5 2.8 2.8

33.2 Syringol 7.0 1.0 3.9 2.0 5.3 5.8 7.9 2.7 2.4 1.2

33.8 p-Propyl-guaiacol 4.7 0.9 2.0 3.6 4.3 7.4 7.3 2.2 2.9 2.1

43.5 4 – Propyl-syringol - 0.4 2.0 1.5 2.8 3.6 0.9 1.5 1.8 0.7

Total 109.6 19.7 34.7 49.0 63.4 65.8 95.0 57.8 72.0 60.8

Page 107: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

86

Table 2-7. Quantitative analysis of major micromolecules in the light oils by GC/MS analysis

RT Compound Relative area (Compound area/I.S. area)

W/O solvent Acetone Ethanol Ether

3.3 Acetic acid, ethyl ester - - 21.7 2.8

3.5 formic acid, methyl ester - - - -

3.8 Acetic acid 89.9 41.4 52.3 88.4

5.6 Propanoic acid, ethyl ester 9.3 - 2.5 -

6.8 Cyclopentanone 1.6 2.9 0.1 2.1

8.7 2-Methyl-2-cyclopentenone 1.7 2.9 1.6 1.5

15.4 Phenol - 4.8 0.3 2.3

20.2 p-Cresol 5.8 1.8 2.3 -

20.7 Guaiacol 1.9 2.4 1.9 1.0

23.8 Creosol 2.1 - - -

26.1 Catechol 5.7 4.9 2.5 1.6

29.9 4-Ethylguaiacol - - - -

33.2 Syringol 1.6 2.1 1.9 -

33.8 p-Propyl-guaiacol - - - -

35.7 4-(1-Hydroxyethyl)-guaiacol - - - -

43.5 4 -Propyl-syringol - - - -

Total 119.6 63.2 87.1 99.7

Page 108: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

87

3.5. Gas composition

Table 2-8 shows the distributions of HDO gas products resulting when

different solvents were used. The gas phase was mainly composed of CH4, CO2,

C2H4, C2H6, and C3H8. CH4 concentration exhibited upward trends with added

solvent during the HDO, whereas CO2 concentrations declined when solvent

was added. Nevertheless, CO2 was the main component in the HDO reaction.

While some unsaturated C2H4 was observed, the solvent could prevent carbon

loss into the gas phase. Up to 81.2% CO2 was present in the gas phase, which

could be attributed to its release by means of decarboxylation (Şenol et al. 2005).

The 8.7–24.1% of CH4 observed generally originated from demethylation and

slight hydrogenation (Wu et al. 2014). The highest CH4 concentration (24.1%)

was observed when polar protic ethanol was used, whereas the use of the

nonpolar ether allowed the release of much more CO2 and the use of polar

aprotic acetone allowed the release of more C2H6, and C3H8, which likely

originated from dealkylation of phenolic side chains. However, the remaining

amounts of 4-ethylguaiacol, 4-propylguaiacol and 4-propylsyringol in the

heavy oil after HDO were highest when polar aprotic acetone was used.

Therefore, the polar ethanol stabilized the alkyl-substituted phenolic

compounds by means of dealkylation, and the nonpolar ether stabilized the

oxygenated compounds by means of decarboxylation.

Page 109: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

88

Table 2-8. Distribution of representative gas components after HDO reaction

Concentration

(%)

Without

solvent Ethanol Acetone Ether

CH4 8.7 24.1 14.5 14.3

CO2 81.2 70.9 67.6 73.9

C2H4 0.2 0.3 0.0 0.1

C2H6 6.8 3.2 9.4 8.7

C3H8 3.2 1.6 8.5 3.1

Page 110: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

89

3.6. Estimation of reaction efficiency

Table 2-9 lists the carbon flows from the bio-oil to the HDO products. The

carbon recovery to the heavy oil was 55.3%, 72.8%, and 64.7% in ethanol,

acetone, and ether, respectively. Acetone yielded the highest carbon recovery

(72.8%) in the heavy oil, whereas ether yielded a transfer of 18.3% of the carbon

into the char. Acetone’s high carbon recovery may have been due to its

promotion of esterification and its suppression of demethoxylation or

dealkylation. Esterification, which was effective in ethanol solvent, can reduce

the carbon recovery owing to the modification rather than removal of oxygen-

containing compounds (de Miguel Mercader et al. 2010). Therefore, the heavy

oil obtained from HDO with acetone presented the highest amount of

micromolecule, consisted of guaiacols, and syringols which has a high carbon

number. While, the bio-oil has lost an amount of carbon to light oil phase when

HDO with ethanol. As shown in Table 2-7, the micromolecules in light oil

obtained from HDO, with ethanol were mostly consisted of phenols, while that

from ether was consisted of acetic acid, which has a low carbon number. HDO

with ether showed that a lot of carbon in bio-oil and ether flowed into the gas,

and char (18.7, and 18.3%, respectively). However, light oil obtained from ether

contained only 4.1% of carbon in reactant. It might be due to the low

homogeneity of non-polar ether, and polar water or acetic acid; light oil obtained

from ether showed high water content (77.1%), and had a lot of acetic acid (88.7%

of total micromolecules).

After the HDO reaction including solvent, the product quality, stability, and

economics (carbon recovery, CR) were respectively improved when using

ethanol, ether, and acetone. Therefore, the choice of solvent should be made

considering the goals of carrying out the reaction.

Page 111: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

90

Table 2-9. Carbon recovery rate of the products obtained from the HDO

reaction

Carbon yield

(%) Ethanol Acetone Ether

Char 8.7 6.0 18.3

Light oil 28.4 10.3 4.1

Heavy oil 55.3 72.8 64.7

Gas 7.6 10.9 18.7

Page 112: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

91

3.7. Changes in the properties of heavy oil during storage

The physical properties of bio-oil changed undesirably during storage period.

However, the HDO reaction improved the bio-oil quality by lowering its

viscosity, acidity, and water content. Moreover, the process might also enhance

the stability of bio-oil via removal of reactive compounds (levoglucosan,

furfural, acetic acid, and unsaturated compounds) as well as promote

decomposition of phenol polymers. Especially, Pt/C with ethanol was effective

to enhancing stability of oil with reducing of phenol polymers.

According to a previous study, bio-oil separates into a water phase and a

viscous sludge phase (Garcia-Perez et al. 2006). In contrast with bio-oil, there

were no visual changes in the heavy oil during the storage period. To test these

changes, the heavy oils were weighed before and after storage. None of the

samples lost more than 0.4 wt% of volatiles, and no substantial volatile loss was

noted over the 12 week storage period.

3.7.1. Water content

Changes in the physical properties of the heavy oil during the storage period

are shown in Table 2-10. Significant changes were observed in the properties of

heavy oils, including water content and viscosity, which was the most

problematic feature during bio-oil storage. For example, an excessive amount

of water generally led to corrosiveness and phase separation during aging.

However, the HDO process decreased the water content in bio-oil from 17.7 to

0.3-2.1 wt% in heavy oil via dehydration. The water content in heavy oil

increased an average of 0.2 wt%/week (Pt/C) over the 12 week period. This

increase was faster within the first 2 weeks for Pt/C and the water content after

storage (3.4 wt%) was still lower than that of bio-oil (23.6 wt% water content

after 10 weeks). The increase in water content suggests the occurrence of

Page 113: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

92

reactions generating water, such as esterification, acetalization, condensation,

and alcohol addition (Oasmaa et al. 2010). Water formation reactions were

predominant in the initial stage.

3.7.2. Acidity

The acidity of heavy oil was strikingly reduced after the HDO reaction, due

to the conversion of acids, like acetic acid, in bio-oil into other compounds

during HDO. TAN of heavy oil changed slightly during the 12 weeks. The

acidity of heavy oil decreased from 73.2 to 70.9 mg KOH/g oil during the 12

weeks of storage. The level of TAN observed in heavy oil resulted from

phenolic hydroxyl groups in lignin derivatives (Meng et al. 2014). Moreover,

the decrease in acidity after storage might reflect re-polymerization of

monomeric compounds, which was ascertainable from the average molecular

weights of heavy oil and phenol polymers (Tables 2-10 and 2-12).

Page 114: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

93

Table 2-10. Physicochemical features of heavy oil obtained from hydrodeoxgygenative upgrading of bio-oil in the presence

of Pt/C catalyst during various storage period

Storage period (week) Bio-oil Heavy oil - Pt/C (storage period, week)

0 2 4 8 12

Water content (wt%) 17.7 1.3(0.4) 2.9(0.2) 3.0(0.2) 3.4(0.2) 3.4(0.0)

TAN (mg KOH/g oil) 164.8 73.2(0.1) 73.0(0.3) 73.8(0.1) 72.8(0.0) 70.9(0.2)

Average molecular weight (Mw) 419 762 749 771 830 867

Elemental analysis (wt%)

C 40.2 57.8 58.0 58.7 59.4 59.4

H 7.5 8.6 9.0 9.0 9.1 9.3

N 0.2 0.0 0.0 0.1 0.1 0.2

O b 52.2 33.5 33.0 32.2 31.4 31.2

HHV (MJ/kg) c 18.2 23.9 24.2 24.4 24.6 24.8

a: Standard deviation b: calculated by difference c: high heating value was calculated following formula,

HHV(MJ/kg) = -1.3675 + (0.3137 X C) + (0.7009 X H) + (0.0318 X O)

Page 115: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

94

3.7.3. Viscosity

Figure 2-6 shows the change in viscosity of heavy oil after various storage

periods. The HDO reaction reduced the viscosity of the bio-oil from 30 cSt to 3

cSt. Additionally, the rate of viscosity increase in heavy oil ranged from 0.0

cSt/week during the storage periods, while that of bio-oil was 1.1 cSt/week. The

increase in bio-oil viscosity during storage was due to olefin condensation,

polymerization, phase separation, coke formation, evaporation of volatiles, and

oxidation. These reactions were accelerated by acids, aldehydes, and sugars,

such as levoglucosan, and nitrogen-containing compounds in bio-oil that

catalyze these reactions and increase the viscosity (Mante and Agblevor 2012;

Garcia-Perez et al. 2006). However, the HDO process converted acetic acid,

vanillin, furfural, and levoglucosan into esters via ring opening, hydrogenation,

decarbonylation, and dehydration. The process also removed acetic acid and

converted the unstable pyrolytic lignin structure into stable phenol polymers via

depolymerization, hydrogenation, dehydroxylation, and demethoxylation.

Additionally, both bio-oil and heavy oil contained low levels of nitrogen,

demonstrating the minor role of this element in the viscosity increase. The stable

viscosity of heavy oil during storage is thought to be due to the lack of reactants

for the above reactions.

Page 116: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

95

Figure 2-6. Variation of viscosity during 12 weeks storage period

Page 117: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

96

3.7.4. Molecular weight

The average molecular weight (Mw) of aged heavy oil is presented in Table

2-10. Heavy oil had an immaterial increase in average molecular weight upon

aging. Before storage, the Mw of heavy oil was 762 Da, which increased slightly

after storage and stabilized after 2 weeks. For 12 weeks, the Mw of heavy oil

increased 105 Da (from 762 to 867 Da). As mentioned in previous research,

Pt/C was effective in enhancing the stability of pyrolytic lignin (PL) in bio-oil

via conversion to phenol polymer (PP) by demethoxylation, dehydroxylation,

and decomposition. These reactions decreased the reactivity of PP; thus, the

dimerization and trimerization reactions scarcely occurred during aging. Free

radicals are associated with polymerization, which is known as the

Trommsdorff effect. Moreover, the GPC chromatogram of heavy oil presented

smooth peak, which was different from that of free radical polymerization

(Cioffi et al. 2001; Long et al. 2015). As a result, HDO reduced the incidence

of radicals and stabilized lignin fragments in bio-oil, preventing such reactions

during storage.

3.7.5. Atomic H/C and O/C ratios

The atomic H/C and O/C ratios for the original and aged heavy oils obtained

from different HDO catalysts are presented in Figure 2-7. Conversion and

removal of unstable acid, aldehyde, sugar, and water reduce the O/C ratio of

heavy oil. Since the elemental analysis and calculated atomic ratios include the

water content of heavy oil, the H/C ratio of heavy oil was lower than that of bio-

oil, despite the hydrogenation conditions. Compared with bio-oil (1.0), the

atomic O/C ratio was effectively reduced via the HDO reaction (0.4), The H/C

ratio was reduced by 1.8-1.9 (compared with that of bio-oil; 2.2) due to de-

moisturization and dehydration. Heavy oil obtained from Pd/C and Pt/C

Page 118: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

97

demonstrated an increased atomic O/C ratio of heavy oil of Ru/C (from 0.4 to

0.5), while neither the O/C nor H/C ratio of the other heavy oils of Pd/C and

Pt/C changed during storage.

Page 119: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

98

Figure 2-7. Van Krevelen diagram of bio-oil and heavy oils after 12 weeks

storage

1.7

1.8

1.9

2.0

2.1

2.2

2.3

0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1

Ato

mic

H/C

ra

tio

Atomic O/C ratio

Page 120: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

99

3.7.6. Changes in micro-molecule composition during 12 weeks of aging

Low-molecular weight compounds in the original and aged heavy oil were

analyzed using GC/MS and were identified by comparison with the NIST Mass

Spectral Library (Table 2-11). The major peaks of the gas chromatogram after

storage were qualitatively almost identical with those of the original heavy oil.

Some of the peak areas varied with storage time. Generally, the amount of

compounds slightly decreased, but the variation was not significant for heavy

oil storage. For instance, heavy oil obtained from Ru/C treatment showed very

small changes in micro-molecules. The total amount of micro-molecules

decreased from 20.14 to 17.74 mg/ml for Pd heavy oil, from 12.93 to 11.54

mg/ml for Ru heavy oil, and from 15.86 to 10.61 mg/mL for Pt heavy oil. As

mentioned above, aldehydes and sugars in bio-oil, which affect viscosity during

storage, were converted into esters or ketones during the HDO process.

Compared with the notable reduction in untreated bio-oil stability (Kim et al.

2012), the results with heavy oil indicated that HDO improved bio-oil stability.

Related to the average molecular weight of heavy oil, a slight decrease in the

amount of low molecular weight compounds might be due to dimerization or

trimerization reactions.

Page 121: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

100

Table 2-11. Quantitative analysis of low molecular weight components in

heavy oil after 4, 8 and 12 weeks (23 °C)

Compound

Amount (mg/ml heavy oil)

Storage period (weeks)

0 4 8 12

Phenol

Phenol 0.6 0.5 0.6 0.3

2-Methyl-phenol 0.4 0.5 0.5 -

4-Methyl-phenol 0.2 0.4 0.4 -

Guaiacol 2.6 2.1 2.1 1.7

2,4-Dimethyl-phenol 0.2 0.1 0.1 -

4-Ethyl-phenol 0.0 0.1 0.2 1.7

4-Methyl-guaiacol 2.7 2.0 2.2 1.1

2-Propyl-phenol - - - -

4-Ethyl-guaiacol 3.1 2.6 2.6 2.0

Syringol 0.7 0.7 0.7 0.5

4-Propyl-guaiacol 1.7 1.5 1.7 1.3

Sum of Phenols 12.2 10.5 11.1 8.6

Ketone

2-Methyl-cyclopentanone 0.9 0.4 0.4 1.1

2-Ethyl-cyclopentanone - 1.5 1.7 1.1

2,5-Hexanedione 0.7 0.4 0.4 -

2-Ethyl-2-cyclopentenone 1.7 1.5 - -

2,3-Dimethyl-2-cyclopentenone 0.4 0.4 0.4 -

Sum of Ketones 3.7 4.2 2.9 2.2

Total 15.9 14.7 14.0 10.6

Page 122: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

101

3.7.7. Changes in phenol polymer after storage

The yields of lignin fragments in heavy oil (PP: phenol polymer extracted

from heavy oil before storage, APP: phenol polymer extracted from heavy oil

after storage) before and after storage are described in Figure 2-8. PP yielded

almost the same amounts of APP (6.0 wt%), while PL (pyrolytic lignin; 13.3

wt%) increased approximately 1.8-fold (24.3 wt%) after 10 weeks of storage

(Kim et al. 2012).

The difference between Mw values of PP before and after storage ranged from

129 to 396 Da (Table 2-12). A similar trend was observed with the Mw of heavy

oil (Table 2-10). Therefore, polymerization (actually dimerization and

trimerization) was slightly pronounced between PP and lignin-derived

monomers (micro-molecules). Essential functional groups of phenolic hydroxy

and methoxy are known as two of the more reactive sites in the lignin structure.

In the case of phenolic hydroxyl groups, which are involved in hemiacetal and

acetal formation, the amount decreased from 1.9 to 0.8 wt%. However, the

oxygen level did not change after storage. This result suggests that condensation

or esterification instead of acetal and hemiacetal formation might have occurred.

The content variation of methoxy groups differed with HDO catalyst. The

amount of methoxy group stayed the same (1.0 wt%). The properties were

differed from PP and APP, their structural discrepancies (yield, molecular

weight and functional groups) were less than for PL and APL (aged pyrolytic

lignin). Therefore, it appears that the stability of biofuel can be enhanced via

HDO treatment.

Page 123: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

102

Figure 2-8. Change of macromolecules yield during the storage period

0.0

5.0

10.0

15.0

20.0

25.0

30.0

0 week 2 weeks 4 weeks 8 weeks 12weeks

Yie

ld (

%)

PL (Bio-oil) PP (Heavy oil)

Page 124: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

103

Table 2-12. Characteristics of macromolecules extracted from bio-oil, heavy oil, and aged heavy oil

Lignin

fragment

Elemental analysis

(wt%)

Average molecular Weight

(Da) Functional groups (%)

C H N Oa Mw Mn PDIb Phen-OH OCH3

PL* 65.0 6.4 0.6 27.8 1,065 544 2.0 6.8(0.6) c 11.6(1.1)

APL** 63.5 5.7 0.6 29.8 1,310 788 1.7 1.4(0.1) 5.5(0.1)

Pt PP*** 75.6 7.9 0.6 15.9 818 576 1.4 1.9(0.5) 1.0(0.2)

Pt APP**** 75.9 8.0 0.7 15.4 947 647 1.5 0.8(0.0) 1.1(0.0) a: Calculated by difference b: Polydispersity index(PDI) means the value of Mw/Mn c: Standard deviation *: Pyrolytic lignin (macromolecules in bio-oil) **: Aged pyrolytic lignin (macromolecules in aged bio-oil) ***: Phenol polymer (macromolecules in heavy oil) ****: Aged phenol polymer (macromolecules in aged heavy oil)

Page 125: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

104

3.7.8. Degree of unsaturation

The degree of unsaturation was obtained by carbon, hydrogen, and nitrogen

analysis. Figure 2-9 shows that the degree of unsaturation of heavy oil presented

no notable changes (from 1.1-1.4 to 1.3-1.9) during storage; that of PP also

remained unchanged (Pd-PP: 3.8; Ru-PP: 4.0; Pt-PP: 3.4). Compared with the

degree of unsaturation of lignin (5.4), that of PP was lower (3.4-4.0), indicating

a more stable state via HDO treatment with supercritical ethanol. This finding

accounts for the hydrogenation and dealkylation of side chains and the stability

improvement achieved through the use of HDO. These reactions reduce double

bonds, enhance stability, and decrease the degree of unsaturation. Although

there were structural modifications during storage (Table 2-12), they appeared

not to affect the degree of unsaturation.

Page 126: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

105

Figure 2-9. Degree of unsaturation for oil and macromolecules before and after

the storage period

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

Bio-oil Heavy oil

Deg

ree

of

un

satu

rati

on

0 week (Oil) 12 weeks (Oil)0 week (Macromolecule) 12 weeks (Macromolecule)

Page 127: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

106

4. Summary

In this study, three solvents - polar protic (ethanol), polar aprotic (acetone),

and nonpolar (ether) - were chosen and their effects on the HDO reaction of bio-

oil and the resulting quality as well as storage stability of heavy oil were

evaluated. Bio-oil was converted into gas, char, light oil and heavy oil by the

HDO reaction. The polar protic ethanol effectively improved the typical bio-oil

features: water content increased from 17.7 to 0.5%, acidity decreased from

164.8 to 63.2 mg KOH/g oil, viscosity decreased from 29.6 to 2.2cSt, and HHV

increased from 17.3 to 27.8 MJ/kg). The nonpolar ether did not affect the

organics in the bio-oil, but improved its thermal stability (the degree of

deoxygenation varied 12.4% with increasing reaction temperature). When the

polar aprotic acetone was used, 72.8% of the carbon in the bio-oil remained in

the heavy oil, which was presented that the highest carbon recovery rate. The

energy efficiency calculated from the yield and the HHV of heavy oil was

almost the same for all solvents used. During the HDO reaction, the polar

solvents (ethanol and acetone) acted as co-reactants. Therefore, ethyl esters

were observed in the liquid phase when ethanol was used (and also in trace

amounts when acetone was used), and alcohols were observed when acetone

was used, owing to the possibility of hydrogen bonding between the polar

solvents and organic compounds. Heavy oil was stored at 23 °C for 12 weeks,

and the essential physicochemical properties (water content, viscosity, and

acidity) were almost unchanged during storage. In particular, the principal and

problematic property of viscosity was maintained due to the lack of sugars and

aldehydes in heavy oil. The yield and properties of PP were modified to a lesser

extent than those of PL. Based on the degree of unsaturation, heavy oil and PP

are more stable than bio-oil and PL.

Page 128: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

107

Chapter 3

Preparation of Ni-based catalysts

with various support materials

and assessment of their catalytic effects

for improvement of bio-oil properties

A paper published in

RSC Advances 7(67) (2017) 15116-15126

Page 129: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

108

1. Introduction

Improvements in bio-oil quality and stability can be achieved by the

upgrading process hydrodeoxygenation, an interesting alternative to well-

enhanced biofuel production. This process occurs simultaneously with

hydrodesulphrization, hydrodenitrogenation, hydrodemetallization, and

removes the oxygens of oxygenated compounds and converts them to esters, or

ketones via hydrogenolysis, hydrogenation, dealkylation, dehydration,

decarbonylation, or dehydroxylation. Furthermore, depolymerization,

transalkylation, hydrogenation, and further cracking of compounds in bio-oil

can also be significant depending on the characteristics of the catalysts, and

these reactions can contribute to improve the bio-oil quality (Furimsky 2000;

Rezaei et al. 2014). Hydrodeoxygenation (HDO) is attracting a great deal of

attention because it can enhance the bio-oil quality to overcome disadvantages

such as high water content, acidity, viscosity, oxygen level, and low heating

value under high pressure (Zhang et al. 2013). Some of previous studies

conducted HDO with lignin model compounds. Guaiacol converted into o-

cresol, phenol or benzene (González-Borja and Resasco 2011). Anisole and

catechol also converted into phenol, benzene or cyclohexane via

hydrodeoxygenation (Zhu et al. 2011; Jongerius et al. 2012). Generally, most of

phenolic compounds were methylated, hydrogenated, demethoxylated of

dehydroxylated during the HDO process. Coumaran alkyl ether bond and b-O-

4 bond in dimeric lignin model compounds were cleaved during HDO, while 5-

5 linkages were not (Jongerius et al. 2012).

General HDO catalysts, which are based on sulfided CoMo, NiMo or noble

metal catalysts, show good performance in the enhancement of bio-oil.

However, the need for a supply of sulfur-containing compounds and potential

of sulfur contamination in bio-oil represent significant difficulties. Noble metals,

Page 130: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

109

generally dispersed on zeolites such as ZSM-5 or active charcoal supports, are

efficiently applied to the HDO of bio-oil. However, the noble metal phase is

expensive, and its utilization affects the economy of the HDO reaction (Narani

et al. 2015; Yan et al. 2010).

Since the high price and limited amount of noble metal, supported non-noble

metal catalysts, such as Fe/SiO2, Co/SiO2, and Ni-based materials, have been

recently used in HDO studies. Transition metal catalysts, especially Ni or Cu,

were promoted hydrogenation, decarboxylation and decarbonylation (Wang et

al. 2016). Moreover, its electronic properties can perform similar reactions as

Pd or Pt such as cleavage C-C or C-H for hydrocarbon or alcohol reforming (De

et al. 2016). Due to its high activity on hydrogenation as well as cheaper price

than noble metals, Ni-based catalysts could widely applied in the field.

According to Wang (Wang et al. 2016), RANEY nickel and other nickel

catalysts can catalyze the transfer hydrogenation of ketones, phenols, aldehydes,

olefins, and aromatic alcohols with 2-PrOH as an H-donor. Zhang et al.

investigated that Ni could cleave the C-O bonds in lignin under severe condition.

Especially Ni/C catalyzed lignin hydrogenolysis under mild condition with

polar solvent (Zhang et al. 2014). Although Ni-based catalysts show greater

hydrogenation activity than other catalysts, coke deposition is still a problem

for bio-oil upgrading (Huynh et al. 2014).

Mesoporous silica materials have noted as catalyst supports due to their high

specific surface areas, narrow range of pore size distributions, and large pore

volumes. Therefore, mesoporous silica supports are desirable alternative silica

supports based on their control of catalyst particle size and influence in product

selectivity through transport effects (Ghampson et al. 2012). Alumina is

occasionally added to produce structures with silica mesoporous supports in

order to assign acidic sites that can act as active sites. To create catalytic active

sites for mesoporous silica supports, the supports are doped with metal. Among

the mesoporous silicate supports (M41S, SBA, MSU, HMS, etc.), SBA-15 has

Page 131: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

110

the largest pore, with hexagonally regio-regular arranged mesopores, adjustable

pore size, and high thermal stability (Choudhary and Phillips 2011; Lu et al.

2010; Lestari et al. 2010). Moreover, SBA-15 was found to be significantly

more effective in deoxygenation than the HMS-supported catalyst (Zhao et al.

1998). Supporting these active phases in meso-structured supports might

contribute to complete utilization of their catalytic properties by promoting

dispersion and increasing the number of active sites.

Therefore, in the present work, Ni particles were distributed uniformly inside

active charcoal and meso-structured supports (SBA-15 and Al-SBA-15) to

investigate the effect of Ni based catalysts on the HDO reaction of bio-oil.

Page 132: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

111

2. Materials and methods

2.1. Catalyst preparation by sol-gel method and wet

impregnation

Mesoporous SBA-15 or Al-SBA-15 was synthesized by sol-gel method

(Mohan et al. 2012), using poly(ethylene glycol)-block-poly(propylene glycol)-

block-poly(ethylene glycol) (Pluronic P-123, average Mn ~5,800, Aldrich,

catalog number : 435465) as the structure-directing agent, tetraethyl ortho

silicate (TEOS, Aldrich, catalog number : 131903) as the silicate source, and

aluminum isopropoxide (AIP, Aldrich, catalog number : 220418) in order to

obtain the desired Si/Al ratio and acid sites on the supports under acidic

conditions. All commercial chemicals were used without further purifications.

SBA-15 and Al-SBA-15 were obtained as follows:

8 g of Pluronic P-123 was dissolved in 250 ml of 1.9 M HCl, heated to 40 °C,

and TEOS was added. After stirring for 20 h at 40 °C, the mixture was aged at

110 °C for 24 h. For Al-SBA-15, 8.5 g of TEOS and 0.125 g of AIP were added

to 10 ml of 1.9M HCl. After stirring for 3 h, the homogenous mixture was added

to a solution of 4 g of Pluronic P-123 dissolved in 150 ml of 1.9M HCl. The

solution was stirred for 20 h at 40°C and then aged at 110 °C for 24 h. The solid

product was recovered by filtration, dried overnight and calcined at 550 °C for

5 h under static air.

Synthesized SBA-15, Al-SBA-15, and commercial activated charcoal (~100

mesh particle size, Sigma-Aldrich, catalog number : 242276) were used as Ni-

based catalyst supports. The catalytic active site, nickel, was loaded on SBA-

15, Al-SBA-15, or activated charcoal support by a wet impregnation method

(Mohan et al.,2012; Galarneau et al.,2003), as follows.

An equal amount of Ni(NO3)2·6H2O solution and supports (active charcoal,

Page 133: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

112

SBA-15, and Al-SBA-15) were stirred for 1 h. The mixture was then dried at

80 °C and calcined in air at 550 °C for 3 h to obtain Ni/C, Ni/SBA-15, and

Ni/Al-SBA-15. The catalysts were reduced at 800 °C for 3 h under an H2

atmosphere before use.

Page 134: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

113

2.2. Characterization of synthesized Ni-based catalysts

The morphology of each of the three catalysts (Ni/C, Ni/SBA-15, and Ni/Al-

SBA-15) was investigated by FE-SEM and SEM-EDS (field-emission scanning

electron microscopy, SIGMA). The specific surface areas of the support and

catalysts were measured by the Brunauer-Emmett-Teller (BET) test, and the

pore distribution and cumulative volumes of the pores were obtained by the

Barret–Joyner–Halenda (BJH) method from the desorption branches of the N2

isotherms measured at 77.3K. Samples were degassed at 100 °C for 4 h before

the tests. The acidity of the catalysts was measured by the NH3-TPD (BELCAT

II). Before the adsorption, the 0.05 g of the catalyst was dried in a 50 ml/min of

pre-dried (500 °C) He for 110 min. Then adsorption took place with 5% NH3 in

He at 100 °C for 30 min. X-ray powder diffraction patterns (XRD) were

analyzed by a Bruker D8 Advance with DAVINCI using Cu Kα radiation (λ=

1.5418 Å ), operated at 40 kV and 40mA with a scan speed of 0.5 sec/min in a

range of 10-70° (2θ). The Si/Al ratio of Al-SBA-15 and the metal content of the

catalysts were measured by inductively coupled plasma optical emission

spectroscopy (ICP-ES, MultiwavePRO, Anton Parr). The samples were

digested in acids (HF, HNO3, H2O2) for ICP-ES analysis. The surface properties

of the catalysts were analyzed by AXIS-HIS (Kractos Inc.). X-ray photoelectron

spectroscopy (XPS) was performed using Mg in an Mg/Al dual anode (15 mA,

12kV), and the spectra were analyzed using Casa XPS software.

Page 135: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

114

2.3. Production of bio-oil via fast pyrolysis

Ground to pass a 0.5 mm sieve yellow poplar (Liriodendron tulipifera) was

used as a raw material for fast pyrolysis. The sample was fed into the reactor

with the feeding rate of 1 kg/h. Fast pyrolysis was accomplished at 500 °C and

the residence time was regulated at 1.63 s by inert nitrogen gas flow. The

volatile pyrolytic products were cooled and condensed to a liquid bio-oil.

Page 136: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

115

2.4. Hydrodeoxygenative upgrading of bio-oil with Ni-

based catalysts

The prepared catalysts were tested in HDO of bio-oil in a 200 mL stainless

steel (SUS316) autoclave reactor. The reactor was equipped with a magnetic

stirrer, a thermocouple, a pressure sensor, and a system for controlling the

stirring rate, temperature and pressure. Before the reaction, the catalysts were

activated directly in the reactor by reduction under a H2 atmosphere at 550 °C

for 3 h. After activation, same as the method mentioned in Chapter 2, 2 g of the

catalysts or supports were added to a mixture of 40 g bio-oil and 10 g ethanol,

then placed into the sealed reactor and flushed with N2. The reactor was then

pressurized to 3 MPa H2, and the temperature was increased to 300 °C. After

the reaction, there were three phases of products: liquid, char, and gas. The

liquid product, consisting of an upper water-rich phase (light oil) and lower

organic phase (heavy oil), was recovered from the reactor and separated using

a separating funnel. The yields of char, light oil and heavy oil were calculated

according to the following equations.

Yield of char (%) = [Solid (g) Catalyst (g)] / [Bio-oil (g) + Ethanol (g)] 100

Yield of light oil (%) = [Light oil (g)] / [Bio-oil (g) + Ethanol (g)] 100

Yield of heavy oil (%) = [Heavy oil (g)] / [Bio-oil (g) + Ethanol (g)] 100

Yield of gas (%) = 100 [Yield of light oil + Yield of heavy oil + Yield of char]

Page 137: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

116

2.5. Physicochemical properties of bio-oil and upgraded oil

Karl-Fischer titration (TitroLine KF, Schott Instruments, Germany) was used

to measure the water content of the bio-oil and heavy oils. In order to investigate

the acidity of the bio-oil and heavy oil, the total acid number (TAN) was

measured using ASTM D664. The viscosity was determined using a capillary-

type viscometer and ViscoClock (Schott Instruments) at 40 °C. Elemental

analysis of carbon, hydrogen, and nitrogen was performed out with a LECO

CHNS-932 analyzer. The higher heating value (HHV) was calculated using the

following equation (Friedl et al. 2005). The degree of deoxygenation (DOD)

was estimated as follows:

HHV (J/kg) = 3.55 [C]2 - 232 [C] - 2230 [H] + 51.2 [H] [C] + 131

[N] + 20,600

Degree of deoxygenation (%) = (MObio-oil MOheavy oil) / MObio-oil 100

where [C], [H], and [N] are the mass% of carbon, hydrogen, and nitrogen on

dry basis, and MOheavy oil and MObio-oil are the molar oxygen / carbon ratio of

heavy oil and bio-oil, respectively.

Page 138: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

117

2.6. Chemical composition of hydrodeoxygenative

upgraded oil

GC/MS analysis with internal standard (I.S.; fluoranthene) was performed for

qualitative and quantitative analysis of the low-molecular weight compounds in

bio-oil and heavy oils. The identification of low-molecular weight compounds

in the bio-oil and heavy oil was conducted with an Agilent 7890A coupled with

a 5975C mass selective detector (MSD) and a flame ionization detector (FID)

equipped with a DB-5 capillary column (60m × 0.25 mm × 0.25 um). The oven

temperature was maintained at 50 °C for 5 min, followed by heating at a rate of

3 °C/min to 280 °C and holding for 20 min. The injector and FID detector

temperatures were 250 °C and 300 °C, respectively. Figure 3-1 describes the

overall experimental scheme of this study.

Page 139: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

118

Figure 3-1. Scheme of the whole experiment

Page 140: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

119

3. Results and discussion

3.1. Synthesis of mesoporous supports and supported

nickel catalysts

3.1.1. Physical properties of the supports and Ni catalysts

Table 3-1 presents the physical properties of the catalysts used in this study.

The specific surface areas of Ni/C, Ni/SBA-15, and Ni/Al-SBA-15 were 436.5,

508.7, and 537.7 m2/g, while the pores volumes were 0.46, 0.35, and 0.46 cc/g,

respectively. These values were lower than those of Pt/C (5 wt%, Sigma-

Aldrich), which might be due to the high metal loading (14.6-17.1 wt%). As

SBA-15 and Al-SBA-15 were mesoporous-type supports, the micropore areas

of Ni/SBA-15 and Ni/Al-SBA-15 (57.0, and 39.1 m2/g, respectively) were

lower than that of Ni/C (169.3 m2/g), while the surface area had the opposite

tendency. Therefore, Ni/Al-SBA-15, with a high surface area and large pore size,

was expected to be effective for good dispersion of the reactant during the

reaction.

Page 141: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

120

Table 3-1.Characterization of synthesized Ni-based catalysts

Sample

Metal

content

(%)

Si/Al

ratio

BET surface

area

(m2/g)

Micropore

area

(m2/g)

Pore

size

(Å )

V pore

(cm3/g)

Reduction

yield (%)

Active carbon - - 896.4 449.8 55.2 0.62 -

SBA-15 - - 800.2 54.3 55.0 1.08 -

Al-SBA-15 - 2.5 889.2 36.2 32.0 0.73 -

Ni/C 17.1 - 436.5 169.3 59.3 0.43 92.8

Ni/SBA-15 15.4 - 508.7 57.0 44.2 0.65 94.9

Ni/Al-SBA-15 14.6 2.5 537.7 39.1 32.4 0.46 89.5

Pt/C

(commercial) 5.0 - 1317.4 243.0 37.6 0.99 -

Page 142: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

121

3.1.2. XRD patterns of the Ni catalysts

Figure 3-2 presents the XRD peaks of Ni/C, Ni/SBA-15, and Ni/Al-SBA-15.

All samples presented three typical peaks at 2θ=32°, 38° and 58°, marked as

(100), (101), and (110), while Ni/C showed more peaks, such as (111) (2θ=47°).

These exhibited that the crystalline properties of SBA-15 and Al-SBA-15 were

almost identical. The peaks illustrate that Ni/SBA-15 and Ni/Al-SBA-15 have

a hexagonal structure, while Ni/C has a cubic structure (Wang et al. 2013; Zhang

et al. 2012; Yu et al. 2011; Kruk et al. 2000). Hydroxide (Ni(OH)2) or oxide

form (NiO) was investigated in the patterns of Ni/C (2θ=22, 39°, respectively).

It was also found in the XPS results (section 3.1.3.).

According to the peaks at XRD patterns, Ni particles on Ni/C retained higher

dispersion than Ni/SBA-15 or Ni/Al-SBA-15. It was due to the mesoporous

structure of SBA-15 and Al-SBA-15, which resulted multi-layered dispersion

inside the structure.

Page 143: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

122

Figure 3-2. XRD peaks of supports and catalysts (The peaks of (100), (101), and (110) revealed the hexagonal structure,

while (100), (101), (110), and (111) described the cubic structure)

Page 144: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

123

3.1.3. Surface binding properties of the catalysts

X-ray photoelectron spectroscopy (XPS) of the catalysts was performed to

identify the surface binding properties. Figure 3-3 presents the Ni2p XPS spectra,

which exhibit a large band centered at 855 eV that can be de-convoluted into

two components. The first component at 853 eV was assigned to Ni0; the low

intensity of this band reveals incomplete reduction of nickel (Yu et al. 2011).

However, the apparent low intensity of this band could be due to the subsequent

reabsorption of photoelectrons emerging from the core level of nickel by other

nickel atoms of the metallic particles. The peak of the more energetic signal at

856-857 eV reveals either Ni2+ ions or very small NiO particles on the walls of

the mesopores (de Miguel Mercader et al. 2011). These peaks indicated that

NiO was partially present, even after the reduction. This band indicated that the

reduction is not completely accomplished or should be assigned to the

phenomenon that the photoelectrons emerged from the core level of nickel are

subsequently reabsorbed by other nickel atoms of the metallic particles.

Additionally, the peak at 860 eV is known as the shake-up satellite peak (Joshi

and Lawal 2012).

Page 145: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

124

Figure 3-3. XPS analysis of Ni2p spectra on (a) Ni/C, (b) Ni/SBA-15, and (c)

Ni/Al-SBA-15 (853 eV : Ni0; 856-857 eV : Ni2+ ions or NiO

particles; 860 eV : shake-up satellite peak)

Page 146: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

125

3.1.4. Surface acidity of the catalysts

The type of surface acidity of the three catalysts was examined by NH3-TPD

measurements (Figure 3-4). Since the acid strength of the solid catalyst was

defined as its proton-donating ability, the acidity of the synthesized catalyst

might have affected the HDO of bio-oil. As shown in Figure 3-4, a broad peak

near 160-360 °C and a sharp peak at 600 °C were observed for Ni/SBA-15 and

Ni/Al-SBA-15, while Ni/C showed no acidic sites. According to previous

studies (Aburto et al. 2009), the peaks produced at temperatures above 400 °C

are generally attributed to the desorption of NH3 from strong Brønsted and

Lewis sites, while the low-temperature peaks can be attributed (a) desorption of

weakly adsorbed NH3 at weak acid sites or non-acidic sites, (b) weak acid

Brønsted and Lewis sites, and (c) formation of NH4+(NH3)n groups (Aburto et

al. 2009). During calcination, Ni/SBA-15 and Ni/Al-SBA-15 were quickly

increase (Aponte and de Lasa 2017), and this led to higher strong acid sites and

a shift of NH3-TPD peaks toward higher temperatures.

Page 147: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

126

Figure 3-4. NH3-TPD peaks of supports and catalysts

Page 148: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

127

3.1.5. The morphologies of the supports and Ni catalysts

The morphologies of the supports and catalysts were investigated by SEM

(Figure 3-5), and SEM-EDS (Figure 3-6), respectively. Active charcoal had a

rough surface with a spherical shape and small pores (Figure 3-5 (a)). However,

SBA-15 and Al-SBA-15 showed larger structures with smooth surfaces, and

round hexagonal prisms. Figures 3-6 (a)-(c) show that Ni particles in 0.2-0.5

μm were well dispersed within the (a) active charcoal, (b) silica, and (c)

alumina-silica matrix. Ni particles were dispersed as multi-layers on the SBA-

15, and Al-SBA-15 supports, while Ni on the active charcoal was uniformly

dispersed on the active charcoal agglomerate, which might be due to the large

hexagonal structures of SBA-15 and Al-SBA-15. The impregnation of Ni over

the supports led to the formation of Ni/C, Ni/SBA-15 and Ni/Al-SBA-15

catalysts as well as led to the formation of crystalline Ni aggregation on the

surface contained partial NiO as evidenced from the phase formation in the high

angle XRD.

Page 149: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

128

Figure 3-5. SEM images of the supports ((a): active charcoal; (b): SBA-15; and

(c): Al-SBA-15; Scale bar: 200nm)

Page 150: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

129

Figure 3-6. SEM-EDS images of Ni-based catalysts ((a): Ni/C; (b): Ni/SBA-15;

and (c): Ni/Al-SBA-15; Scale bar: 2μm)

Page 151: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

130

3.2. Catalytic activity

The yields of the HDO products obtained with supports, Ni/C, Ni/SBA-15,

and Ni/Al-SBA-15 are shown in Table 3-2 based on the weight of the bio-oil.

The yield of each product revealed that the composition of the HDO products

was considerably affected by the support. The products obtained from the HDO

with supports showed high yield of tar-like char (71.7-80.4 wt%). The oil phase

was not separated into the light and heavy oil. It might be due to the lack of

active sites (metal particles) which accelerated the hydrogenation, or

deoxygenation. The yield of each product was differed with the type of support.

The silica-base supports (SBA-15, and Al-SBA-15) produced less char, more

oil and gas phase than active charcoal. The yield of heavy oil showed no

significant changes (45.8-48.1 %) with Ni-based catalysts. However, the yield

was greater than that of Pt/C due to the high metal loading (ca. 14-17 %) and

hexagonal mesoporous structures of the supports. Alternatively, the yields of

char, gas, and light oil presented different tendencies with respect to the

supports. The active charcoal support produced 9.1 % char, 27.1 % gas, and

18.0 % light oil. When silica-based supports (SBA-15, and Al-SBA-15) were

used, 16.8-18.6 % char, 6.6-7.1 % gas, and 26.7-28.3 % light oil were produced.

Therefore, active charcoal supports leads to both HDO and further

decomposition, while SBA-15 and Al-SBA-15 enhance depolymerization and

condensation (charring). The high water content of the light oil might be a result

of the dehydration of organic compounds in the oil phase (Zhao et al. 2009).

This might indicate that the structure of the support plays an important role in

upgrading bio-oil to heavy oil, as well as producing by-products such as light

oil, char, and gas. This process is due to further decomposition of organic

compounds in the liquid phase (Furimsky et al. 1986).

Page 152: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

131

Table 3-2. Mass balance of the products obtained from HDO reaction with Ni-catalysts

Yield (%) Char Oil phase

Gas Light oil Heavy oil

Active charcoal 80.4 (0.3) 6.8 (0.0) 12.7 (0.3)

SBA-15 73.8 (0.9) 10.2 (0.3) 16.0 (1.2)

Al/SBA-15 71.7 (0.2) 11.8 (0.2) 16.5 (0.4)

Pt/C a 7.7 (0.0) b 23.7 (0.4) 35.7 (5.2) 32.8 (2.5)

Ni/C 9.1 (0.7) 18.0 (0.1) 45.8 (2.4) 27.1 (1.4)

Ni/SBA-15 16.8 (0.4) 28.3 (0.5) 47.7 (1.3) 7.1 (0.7)

Ni/Al-SBA-15 18.6 (0.5) 26.7 (0.1) 48.1 (2.0) 6.6 (0.9)

a: Previous study (Chapter 2)

b: standard deviation

Page 153: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

132

3.3. Fuel properties of heavy oil

3.3.1. Physicochemical features of heavy oil

Typical fuel properties of bio-oil, such as acidity, heating value, and thermal

stability, are improved by hydrocracking, hydrogenolysis, dehydration,

hydrogenation and deoxygenation during HDO (Furimsky et al. 1986).

Therefore, several improved physicochemical properties of heavy oil were

measured and are displayed in Table 3-3. The water content, related to the

heating value and combustion properties in the engine, decreased from 17.7 %

in the bio-oil to 13.0 wt% (Ni/C), 11.4 wt% (Ni/SBA-15), and 9.3 wt% (Ni/Al-

SBA-15) in the heavy oils, due to dehydration of the organic phase. Despite the

high metal loading of Ni/C, the dehydration and de-moisturization capacity

were still lower than those of Pt/C. However, the mesoporous hexagonal

structure (SBA-15, and Al-SBA-15) effectively catalyzed the de-moisturization,

especially the acidic support (Al-SBA-15). Previous study (Shemfe et al. 2017)

showed that zeolite is effective to dehydration. Since alumina-silica Al-SBA-

15 support is consisted of same as zeolite, it is also effective to dehydration.

The TAN of the heavy oil was significantly reduced (50.2-62.2 mg KOH/g

oil) in comparison to that of bio-oil (164.8 mg KOH/g oil) and heavy oil

obtained from Pt/C (93.2 mg KOH/g oil). Despite acid removal, phenolic

compounds should be measured as a weak acid by KOH. Therefore, the acidity

level was 50.2-62.2 mg KOH/g in the heavy oils. The compounds derived from

carbohydrates, which enhanced the acidity, can be converted into esters and

ketones. Moreover, phenolic compounds might also be converted and play an

important role in reducing the acidity. Reducing the viscosity is an important

point of reforming bio-oil into fuel because the high viscosity of bio-oil results

in many disadvantages during injection into an engine (Ochoa-Hernández et al.

Page 154: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

133

2013). In this experiment, the viscosity decreased from 29.6 cSt for bio-oil to

6.8-7.1cSt for heavy oil. The viscosity is greater than that of the heavy oil

obtained from Pt/C, however, the heavy oils also had fluidity and did not stick

to the vials.

Page 155: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

134

Table 3-3. Physicochemical properties of heavy oil

Bio-oil a Pt/Ca Ni/C Ni/SBA-15 Ni/Al-SBA-15

Water content (wt%) 17.7 (0.2)d 0.5 (0.0) 13.0 (0.1) 11.4 (0.2) 9.3 (0.4)

TAN

(mg KOH/g oil)

164.8

(0.8) 93.2 (0.5) 50.2 (0.8) 62.2 (1.0) 55.6 (0.1)

Viscosity(cSt)b 29.6 (0.2) 4.4 (0.1) 6.8 (0.3) 7.1 (0.0) 6.8 (0.2)

C (wt%) 40.0 (1.0) 70.7 (1.4) 54.9 (1.1) 58.9 (0.7) 57.3 (1.3)

H (wt%) 7.5 (0.2) 9.0 (0.7) 9.9 (1.5) 8.8 (0.6) 8.0 (0.7)

N (wt%) 0.2 (0.0) 0.2 (0.0) 0.7 (0.1) 0.1 (0.0) 0.1 (0.0)

Oc (wt%) 52.2 (1.4) 20.1 (1.5) 34.4 (0.8) 32.2 (1.4) 34.5 (1.2)

S (wt%) 0.2 (0.0) 0.0 (0.0) 0.0 (0.0) 0.0 (0.0) 0.0 (0.0)

HHV(MJ/kg) 15.9 34.4 23.7 25.4 24.2

Energy efficiency (%) - 60.3 58.1 62.8 61.7

Degree of deoxygenation - 73.7 49.2 54.9 49.2

a: Previous study (Chapter 2)

b: measured at 40°C c:calculated by difference d:Standard deviation

Page 156: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

135

3.3.2. Elemental composition and calorific value Physicochemical features

of heavy oil

The elemental compositions of the bio-oil and heavy oil were determined and

are presented in Table 3-3. Compared to the carbon (40.0 wt%), hydrogen (7.5

wt%), and oxygen (52.2 wt%) contents of the bio-oil, the carbon and hydrogen

levels increased (54.9-58.9 wt%, and 8.0-9.9 wt%, respectively), while the

oxygen level decreased (32.2-34.5 wt%) in heavy oil. The Van Krevelen

diagram of the bio-oil, and heavy oils (Figure 3-7, dry basis) shows the

significant tendency for active charcoal to enhance the hydrogenation, due to

the effective dispersion of metal particles on high surface area, and micropore

area. Mesoporous silica supports (SBA-15, and Al-SBA-15) exhibited good

dehydration due to their acidic sites on the surface. Both supports reduced the

atomic O/C ratio from 0.8 (bio-oil) to 0.4 (heavy oils) during the HDO reaction

due to the removal of oxygenated compounds in the bio-oil to the gas (Ni/C) or

light oil phase (Ni/SBA-15, and Ni/Al-SBA-15) via deoxygenation, and the

dehydration. Since hydrogenation and deoxygenation was accelerated by the

metal sites, the well-dispersed Ni/C (Figures 3-2 and 3-6) is effective to

hydrogenation. However, deoxygenation is generally occurred in lignin-derived

phenolic compounds, which has large size (Yu et al. 2012). Therefore,

deoxygenation process might be effectively occurred in mesoporous SBA-15

(Table 3-1). The reaction also resulted in a greater HHV of the heavy oils (21.9-

22.8 MJ/kg) compared to that of the bio-oil (17.3 MJ/kg).

Page 157: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

136

Figure 3-7. Van Krevelen diagram of bio-oil, and heavy oils (dry basis)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

0.0 0.2 0.4 0.6 0.8 1.0

Ato

mic

H/C

ra

tio

Atomic O/C ratio

Bio-oil

Pt/C

Ni/C

Ni/SBA-15

Page 158: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

137

3.4. Physicochemical features of light oil

After the HDO reaction, a solvent-based aqueous phase termed light oil was

produced as by-product. The physicochemical properties, and organic

compounds in the light oil were differed from those of heavy oil. Figure 3-8

described representative features of the light oils yielded, such as their water

content and acidity. High water content and acidity is remarkable in light oil

properties. The water content of the light oil was lower when the Ni-based

catalysts were used (43.8–49.7 wt%) than when the Pt/C was used (65.0 wt%).

Acidity of light oil presented no significant differences between Pt/C (159.2 mg

KOH/g oil), and Ni-based catalysts (160.8-167.4 mg KOH/g oil).

Page 159: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

138

Figure 3-8. Physicochemical properties of light oil

Page 160: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

139

3.5. Degree of deoxygenation (DOD)

Since high oxygen levels result in high acidity, viscosity, and low heating

values, deoxygenation is vital to the value-added utilization of bio-oil. HDO

leads to decreased oxygen levels in bio-oil through deoxygenation, such as

demethoxylation, dehydroxylation, or dehydration. The DOD was calculated

based on the atomic O/C ratio. After the HDO, the O/C ratio of the bio-oil (0.8)

decreased to ca. 0.4 due to deoxygenation. There were no significant differences

among the various supports. The DOD varied between 49.2 and 54.9% with Ni-

based catalysts. Ni/SBA-15 showed the highest DOD level (54.9%), which

suggests the high pore volume of the catalyst (0.65 cc/g), and large pore size

(44.2 Å ) reacted with the oxygenated compounds (Furimsky et al. 1986). The

decrease in H/C ratio in the heavy oil obtained from Ni/SBA-15 and Ni/Al-

SBA-15 (1.4) compared to that of bio-oil (1.5) could have resulted from

dehydration and deoxygenation. In conclusion, dehydration can occur on both

Ni particles and mesoporous silica supports (SBA-15 and Al-SBA-15), while

active charcoal enables metal dispersion (Wu et al. 2014).

Page 161: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

140

3.6. Structural modification of low-molecular weight

compounds during HDO

Twenty-two typical compounds in bio-oil, heavy oil, and light oil were

identified from the GC/MS analysis and classified based on chemical structure

and functional groups. Their relative amounts are presented in Tables 3-4 and

3-5 (heavy oil and light oil, respectively). The Ni-based catalysts mainly

generated compounds in heavy oil categorized as guaiacols, such as guaiacol,

4-ethyl-guaiacol, and 4-propyl-guaiacol. The minor compounds, such as esters,

ketones, and other phenolic compounds, were also present in heavy oil after the

HDO. The heavy oils showed a 65.7-76.4% decrease in the total compounds

compared to bio-oil. The decrease typically resulted from decarboxylation and

further decomposition of acetic acid, furfural, vanillin, and levoglucosan or

separation into the light oil phase. Significantly, the light oil phase consisted

mostly of acetic acid and acetic acid ethyl ester (Table 3-5).

Generation of catechol and p-cresol and removal of 4-ethyl-phenol and 3,4-

dihydroxyacetophenone suggest that Ni-based catalyst can accelerate

dealkylation or transalkylation. However, the remaining 4-propyl-guaiacol and

production of 4-propyl-syringol suggest that Ni-based catalysts might not

effectively reduce the activation energy of dealkylation, or decomposition.

Additionally, less CH4 was observed than in Pt/C (Table 3-6). Guaiacol and

syringol remained abundant in the heavy oil, despite demethoxylation to phenol.

This result was due to the dealkylation, or demethoxylation via hydrogenation

(Furimsky et al.,1986) and was confirmed by the elimination of 4-methyl-

syringol and 4-(1-propenyl)-guaiacol. The unstable and undesired acetic acid,

furfural and levoglucosan in the bio-oil were totally converted to acetic acid

ethyl esters, and cyclopentanone in the heavy oil via hydrogenation,

decarbonylation, dehydroxylation, and ring opening, or low molecular gas

Page 162: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

141

phase. Additionally, heavy oil obtained from Ni/Al-SBA-15 yielded low levels

of acetic acid ethyl ester, and a high level of cyclopentanone. Therefore, Ni/Al-

SBA-15 might be the most effective support to convert unstable carbohydrate

derivatives into a stable form (Ochoa-Hernández et al. 2013).

Page 163: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

142

Table 3-4. Quantitative analysis of major micromolecular components in the bio-oil and heavy oils by GC/MS analysis

No. Compound

Amount (mg/ml bio-oil)

Bio-oil Pt/C Ni/C Ni/SBA-15 Ni/Al-SBA-15

Acids 1 Acetic acid 28.0 8.2 - - -

Esters 2 Acetic acid, ethyl ester - 14.6 24.5 28.1 7.7

Ketones 3 Cyclopentanone - 2.0 0.9 1.8 3.3

4 2-Methyl-cyclopentanone - 0.7 - - -

Phenols

5 Phenol - 9.5 0.5 0.3 0.5

6 p-Cresol - 3.4 1.8 2.3 1.3

7 Guaiacol 3.3 16.7 11.9 12.2 11.9

8 4-Ethyl-phenol 0.4 - - - -

9 Creosol 1.1 3.0 - 1.9 10.0

10 4-Vinyl-phenol 0.4 - - - -

11 2-Propyl-phenol 1.2 - - - -

12 Catechol - 1.6 3.1 1.6 1.4

13 4-Ethyl-guaiacol 1.1 6.6 4.0 5.8 5.5

14 3,4-Dihydroxyacetophenone 0.7 - - - -

15 Syringol 0.4 7.7 1.6 1.4 2.6

16 4-Propyl-guaiacol 2.1 7.1 3.9 5.6 3.5

17 4-Methyl-syringol 0.5 - - - -

18 4-(1-Propenyl)-guaiacol 0.7 - - - -

19 4-Propyl-syringol - 9.1 4.1 7.2 3.6

Sum of Phenols 11.9 64.7 30.9 38.3 40.3

Aldehydes 20 Furfural 7.6 - - - -

21 Vanillin 1.2 - - - -

Sugars 22 Levoglucosan 27.0 - - - -

Total 75.7 90.2 56.3 68.2 51.3

Page 164: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

143

Table 3-5. Quantitative analysis of major micromolecular components in the bio-oil and light oils by GC/MS analysis

No. Compound

Amount (mg/ml bio-oil)

Bio-oil Pt/C Ni/C Ni/SBA-15 Ni/Al-SBA-15

Acids 1 Acetic acid 28.0 58.3 93.3 107.3 133.0

Esters 2 Acetic acid, ethyl ester - 5.5 75.1 91.7 82.4

Ketones 3 Cyclopentanone - 0.2 1.8 4.4 5.5

4 2-Methyl-cyclopentanone - 3.5 - - -

Phenols

5 Phenol - 0.4 - - -

6 p-Cresol - 3.0 0.8 2.2 0.8

7 Guaiacol 3.3 7.1 5.9 7.8 7.1

8 4-Ethyl-phenol 0.4 - - - -

9 Creosol 1.1 - 7.1 - -

10 4-Vinyl-phenol 0.4 - - - -

11 2-Propyl-phenol 1.2 - - - -

12 Catechol - 9.1 2.9 3.3 2.9

13 4-Ethyl-guaiacol 1.1 1.8 2.6 2.6

14 3,4-Dihydroxyacetophenone 0.7 - - - -

15 Syringol 0.4 3.7 5.1 1.4 1.0

16 4-Propyl-guaiacol 2.1 - - - 2.5

17 4-Methyl-syringol 0.5 - - - -

18 4-(1-Propenyl)-guaiacol 0.7 - - - -

19 4-Propyl-syringol - - - 3.6 1.8

Sum of Phenols 11.9 23.3 23.6 20.9 18.7

Aldehydes 20 Furfural 7.6 - - - -

21 Vanillin 1.2 - - - -

Sugars 22 Levoglucosan 27.0 - - - -

Total 75.7 90.8 193.8 224.3 239.6

Page 165: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

144

3.7. Gas composition

Gas phase analysis was performed, and the distribution is shown in Table 3-

6. The results revealed the formation of CH4, CO, C2H2, and C3H8 when HDO

was performed with Ni-based catalysts, and supports. Compared with Ni-based

catalysts, the gas phase obtained from HDO with supports contained more

hydrocarbon gas. For effective hydrogenation, and deoxygenation, the metal

active sites are essential. The CH4 yield was lower with Ni-based catalysts

compared to that with Pt/C. HDO with Ni/SBA-15 produced 2.3% CH4, while

Ni/Al-SBA-15 and Ni/C produced 5.8 and 7.1%, respectively. However, CO

concentrations increased significantly when solvent Ni-based catalysts were

used, especially Ni/SBA-15 (94.9%), possibly due to release through

decarboxylation (Şenol et al. 2005). The observed 2.3-7.1 % of CH4 generally

originated from demethylation and slight hydrogenation (Furimsky et al. 1986).

The highest CH4 concentration (7.1%) was observed when active charcoal was

used as a support, whereas the use of mesoporous silica supports allowed the

release of a larger amount of CO. The non-acidic SBA-15 support especially

allowed hydrogenation, and inhibited dealkylation according to the low

concentrations of CH4, C2H2, and C3H8 that likely originated from dealkylation

of phenolic side chains.

Page 166: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

145

Table 3-6. Distribution of representative gas components determined by TCD-FID after HDO reaction

Concentration

(%) Active charcoal SBA-15 Al/SBA-15 Pt/C Ni/C Ni/SBA-15 Ni/Al-SBA-15

CH4 16.7

(2.3)

12.4

(1.7)

13.6

(1.3)

24.1

(1.0)

7.1

(0.0)

2.3

(0.0)

5.8

(0.0)

CO 70.6

(1.4)

63.3

(2.1)

77.9

(1.8)

70.9

(1.4)

88.4

(0.7)

94.9

(1.8)

90.1

(1.5)

C2H4 6.5

(0.0)

19.6

(0.3)

3.2

(0.0)

0.3

(0.0) - - -

C2H2 2.5

(0.0)

1.9

(0.0)

1.8

(0.0) -

1.5

(0.3)

0.6

(0.0)

1.5

(0.1)

C3H6 - - - 3.2

(0.0) - - -

C3H8 3.7

(0.4)

2.8

(0.0)

3.6

(0.1)

1.6

(0.0)

3.1

(0.0)

2.2

(0.0)

2.7

(0.0)

Page 167: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

146

4. Summary

Mesoporous SBA-15, and Al-SBA-15 were synthesized via the sol-gel

method, and Ni-based catalysts (Ni/C, Ni/SBA-15, and Ni/Al-SBA-15) were

prepared by a wet impregnation method. The characterization of the catalysts

showed that Al-SBA-15 had a highly-ordered hexagonal structure, and Ni

particles were well dispersed on the supports. The catalytic properties of the

bio-oil HDO were investigated in an autoclave reactor. HDO was affected by

the catalyst supports, and bio-oil was converted to gas, char, light oil, and heavy

oil. The results of the HDO indicated that mesoporous silica supports (SBA-15,

and Al-SBA-15) exhibited greater activity than Ni supported on active charcoal

(Ni/C). The higher activity of the catalyst could be correlated to its high surface

area, and pore size, which uniformly dispersed the Ni particles. The

physicochemical properties of bio-oil, such as water content, viscosity, acidity,

oxygen level, and HHV, improved after HDO, with differences in catalyst

support type. Based on the results, hydrogenation (active charcoal),

deoxygenation (SBA-15), and dehydration (Al-SBA-15) can be effectively

enhanced during the reaction and affect heavy oil properties.

Page 168: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

147

Chapter 4

Enhancement of bio-oil hydrodeoxygenation

reactivity over Ni-based bimetallic catalysts

supported on SBA-15

Page 169: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

148

1. Introduction

In response to the problems resulting from increasing petroleum usage,

worldwide biofuel production has also increased. The bio-oil obtained from fast

pyrolysis is one of the most promising biofuels, although improvement of its

quality and stability is required before widespread usage. Hydrotreating might

become a major process for enhancing fuel quality. Hydrotreatment processes

under development include the use of Ru-based catalysts and conventional

petroleum hydrotreatment catalysts in fixed beds (Snowden-Swan et al. 2016).

However, such processes are difficult to control owing to the many factors

involved, such as temperature, pressure, catalyst, solvent, and reactor type.

Numerous data are available in the literature concerning hydroprocessing of

model compounds, but most studies only provided data related to the differences

in catalyst composition. Homogeneous catalysts are more effective than

heterogeneous ones, but the latter are preferred because they offer easier catalyst

recovery. Generally and widely used sulfided metal catalysts such as NiMo and

CoMo have shown good performance in enhancement of bio-oil quality;

however, the inclusion of sulfur leads to the problems of rapid catalytic

deactivation owing to catalytic poisoning, and contamination of water used to

wash the reactor. Liquid acid catalysts such as H3PO4 are possible alternatives

that can for example catalyze the dehydration of cycloalkanol to cycloalkene.

However, recovery of the liquid acid catalyst from the reaction mixture is

difficult (Ohta et al. 2011). Other acid catalysts have also been considered as

hydrotreating catalyst, though their use introduces the problems of increasing

instability and acidity. Thus, non-sulfided and acid-free catalysts are under

active investigation, and for example, noble metal, or bimetallic catalysts with

activated carbon and aluminum oxide–supported metal catalysts have shown

promise (Wildschut et al. 2010). Noble metal catalysts are advantageous in

Page 170: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

149

hydrodeoxygenation (HDO) because they are highly active and stable during

HDO and also strongly activate hydrogenation. Also, noble metal catalysts are

less easily deactivated than the commonly used metal sulfides. Noble metal

catalysts, which are generally dispersed on zeolites such as ZSM-5 or on active

charcoal supports, are efficient in the catalysis of bio-oil HDO. However, the

noble metals used are very expensive, and their use strongly affects the

economy of the HDO reaction. Supported non-noble metal catalysts, such as

Fe/SiO2, Co/SiO2, Pt/TiO2, and Ni-based materials, have recently been used in

HDO studies. Although Ni-based catalysts show greater HDO activity than

other transition metal catalysts, coke deposition remains a problem for bio-oil

upgrading with these catalysts (Moulijn et al. 2001; Huynh et al. 2014).

According to Wang (Wang and Rinaldi 2012), Raney nickel and other nickel

catalysts can catalyze the transfer hydrogenation of ketones, phenols, aldehydes,

olefins, and aromatic alcohols with 2-PrOH as an H-donor. Mesoporous silica

materials are notable catalyst supports owing to their high specific surface areas,

narrow pore size distributions, and large pore volumes, and are desirable

alternative silica supports owing to their utility in controlling catalyst particle

size and influencing product selectivity through transport effects (Ghampson et

al. 2012). Alumina is occasionally added to catalytic materials on silica

mesoporous supports as a means to introduce acidic sites that can act as active

sites. To create catalytic active sites on mesoporous silica supports, the supports

are doped with metal. Among the mesoporous silicate supports, SBA-15 has the

largest pores, with regioregular hexagonally arranged mesopores, adjustable

pore size, and high thermal stability (Lu et al. 2010; Lestari et al. 2010;

Choudhary and Phillips 2011). Moreover, SBA-15 has been found as more

effective in deoxygenation catalysts than HMS. Supporting the active phases in

mesostructured supports might contribute to more complete utilization of their

catalytic properties by promoting dispersion and increasing the number of

active sites. Based on catalysis research, Pacific Northwest National Laboratory

Page 171: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

150

(PNNL) and other entities have promoted the development of a general

multistage catalytic hydrotreating scheme for the successful upgrading of bio-

oil into transportation fuels (Zacher et al. 2014). Including conventional

petroleum hydrotreating catalysts such as NiMoS and CoMoS, and newer

catalysts such as Ru/C, several catalysts have been studied to estimate their

suitability for such a process (Zacher et al. 2014; Elliott 2007). The research at

PNNL has focused on improving reactor stability and testing and developing

improved catalysts to achieve optimal technical and cost performance

(Snowden-Swan et al. 2016). Generally, Cu is applied as a promoter, which

could reduce the reduction temperature of Ni and prevent excessive carbon

deposition on Ni, and NiCu catalyst has been tested in the hydrodechlorination

of chlorinated hydrocarbons. However, recent screening studies of bimetallic

NiCu catalyst have indicated its potential for HDO upgrading of anisole and fast

pyrolysis oils as well as vegetable oil (Yakovlev et al. 2009; Ardiyanti et al.

2012a; De Rogatis et al. 2009; Zhang et al. 2013). According to previous studies

(Bui et al. 2012; Zhao et al. 2011; Lee et al. 2012), metal phosphides (Ni2P, CoP,

FeP, MoP, and WP) supported on silica are also effective in converting guaiacol

into aromatic compounds.

Therefore, in the present work, transition metals were used that were easily

obtainable and had high catalytic activity such as ring opening via C–O

cleavage. Ni and transition metal particles were distributed uniformly on

mesostructured SBA-15 support material to investigate the effect of the

resulting Ni-based bimetallic catalysts upon the HDO of bio-oil.

Page 172: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

151

2. Materials and methods

2.1. Synthesis of SBA-15 supports by a sol–gel method

We previously reported that Ni/SBA-15 was the most active catalyst for bio-

oil hydrodeoxygenation, yielding a high amount of biofuel among Ni on three

various supports (Ni/C, Ni/SBA-15, and Ni/Al-SBA-15), and presenting not

that significant different HHV compared with heavy oil from Ni/Al-SBA-15.

Therefore, we used SBA-15 in the present work as the support material for

bimetallic catalysts for HDO of bio-oil. Mesoporous structured silica support

(SBA-15) was synthesized by a means of sol–gel method (Mohan et al. 2012);

pluronic P123 (poly(ethylene-glycol)-block-poly(propylene-glycol)-block-

poly(ethylene-gly col), average Mn ~5800, Aldrich, catalog number : 435465)

was used as the structure-directing agent and tetraethyl orthosilicate (TEOS,

Aldrich, catalog number : 131903) was used as the silicate source. Pluronic 123

of 8 g was dissolved in 250 ml of 1.9 mol/L HCl. The solution was heated to

40 °C, and TEOS was mixed into the solution to initiate the reaction. The

solution was stirred for 20 h at 40 °C and then aged at 110 °C for 24 h. The aged

solid product was recovered by filtration, dried overnight and calcined at 550 °C

for 5 h under static air.

Page 173: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

152

2.2. Selection of transition metals and preparation

bimetallic catalysts

The synthesized SBA-15 was used as the support material for various Ni-

based catalysts. Three transition metals (Cu, Mn, and Zn) were used with Ni to

prepare bimetallic catalysts. The reasons for selecting these three other metals

are as follows. Cu is known as a promoter of Ni (Ardiyanti et al. 2016; Ardiyanti

et al. 2012b; Saidi et al. 2014). Mn can rapidly react with carbocyclic rings such

as benzofuran (Zhao et al. 2014; Zhang et al. 1999; Shen 2015). Zn can

selectively cleave C–O bonds in a variety of monomeric, dimeric, and

polymeric lignin molecules under relatively mild conditions; especially, NiZn

can reduce the formation of oligomeric compounds by decreasing the binding

energy of acetylene, thereby boosting ethylene selectivity (Parsell et al. 2013;

Spanjers et al. 2015). Bimetallic NiCu, NiMn, and NiZn on SBA-15 were each

prepared by means of the following wet impregnation method.

Catalytically active transition metal sites (Cu, Mn, or Zn) were added by

loading metals (Ni) onto the synthesized SBA-15 support by means of a wet

impregnation method. Equal amounts of nitrate-form transition metal (Ni and

Cu, Mn, or Zn) solution and SBA-15 support were stirred for 1 hr. After the

mixture was homogenized, it was dried overnight at 80 °C and calcined in air

at 550 °C for 3 h to obtain NiCu/SBA-15, NiMn/SBA-15, or NiZn/SBA-15. The

catalysts were reduced at 800 °C for 3 h under an H2 atmosphere before use.

Page 174: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

153

2.3. Characterization of catalysts

2.3.1. BET specific surface area

The specific surface areas of the support and catalysts were measured by

means of a Brunauer-Emmett-Teller (BET) method, and the pore distributions

and cumulative pore volumes were calculated by means of the Barret–Joyner–

Halenda (BJH) method, based upon the desorption branches of the N2 isotherms

measured at 77 K. Samples were degassed at 250 °C for 4 h before these tests.

2.3.2. NH3-TPD surface acidity measurements

NH3-TPD measurements were conducted in a Cirrus 2 Mass Spectrometer

equipped with an AutoChem 2950 analyzer. Samples were placed in a U-tube

made of quartz glass between two layers of quartz wool. The temperature was

monitored by using a type K thermocouple fixed to the outside of the tube next

to the sample. Measurements were carried out in the following four steps: (1)

drying (temperature ramp to 500 °C; heating rate: 5 °C/min; hold time: 100 min;

helium flow: 30 ml/min), (2) NH3 adsorption (time: 60 min; NH3 concentration:

5% in helium; gas flow: 15 ml/min; temperature: 100 °C), (3) NH3 flushing with

He (time: 100 min; helium flow: 30 ml/min; temperature: 120 °C), and (4)

desorption (temperature ramp to 600 °C; heating rate: 10 °C/min, helium flow:

30 ml/min; hold time: 40 min). Each result was divided by the sample mass to

allow comparison among results. Before and after this measurement, each

sample was weighed to determine its dry mass.

Page 175: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

154

2.3.3. Surface bonding energy of transition metal

The surface properties of the catalysts were analyzed using an AXIS-HIS

surface analyzer (Kractos Inc.). X-ray photoelectron spectroscopy (XPS) was

performed using Mg in an Mg/Al dual anode (15 mA, 12 kV), and the spectra

were analyzed using Casa XPS software.

2.3.4. Crystallinity of transition metal active sites

X-ray powder diffraction patterns (XRD) were acquired using a Bruker D8

Advance with a Cu Kα radiation source (λ = 1.5418 Å), operated at 40 kV and

40 mA with the scan speed of 0.5°/min over the range 10–70° (2θ).

2.3.5. Degree of transition metal dispersion on SBA-15 support by SEM-

EDS

The morphology of each of the three catalysts (NiCu/SBA-15, NiMn/SBA-

15 and NiZn/SBA-15) was investigated by means of FE-SEM (field-emission

scanning electron microscopy, AUGIRA) and EDS (energy-dispersive X-ray

spectroscopy) mapping.

Page 176: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

155

2.4. Hydrodeoxygenative activity test of prepared catalyst

The Bioliq group at the Karlsruhe Institute of Technology provided bio-oil

for HDO activity testing of the synthesized catalysts. A stainless steel reactor

equipped with a thermocouple, a pressure sensor, and a system for controlling

the temperature and pressure was used as the test chamber. Before testing, the

catalysts were activated directly in the reactor by reduction under an H2

atmosphere at 550 °C for 3 hr. After activation, 4 wt% of the catalysts were

added to a mixture of bio-oil and ethanol (4:1, w/w) and then placed into the

sealed reactor and flushed with N2. The reactor was then pressurized to 3 MPa

H2, and the temperature was increased to the reaction temperature (250, 300, or

350 °C), as mentioned in Chapter 2. After the reaction, three phases of four

products were obtained: liquid (immiscible light oil and lower heavy oil), char,

and gas. The upper solvent-based water-rich light oil and lower organic heavy

oil were recovered after the reaction and separated using a separatory funnel.

The yields of char, light oil and heavy oil were calculated according to the

following equations.

Yield of char (%) = [Solid (g) − Catalyst (g)] / [Bio-oil (g) + Ethanol (g)] × 100

Yield of light oil (%) = [Light oil (g)] / [Bio-oil (g) + Ethanol (g)] × 100

Yield of heavy oil (%) = [Heavy oil (g)] / [Bio-oil (g) + Ethanol (g)] × 100

Yield of gas (%) = 100 − [Yield of light oil + Yield of heavy oil + Yield of char]

Page 177: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

156

2.5. Measuring fuel quality of the products

2.5.1. Physicochemical properties

To measure the water content and acidity (total acid number, TAN) of the bio-

oil and heavy oils, Karl Fischer titration and ASTM D664 TAN testing were

performed. Each sample was analyzed three times; values reported in tables

herein represent the averages of the triplicate analyses. Viscosities were

determined at 40 °C using a capillary-type viscometer and a ViscoClock

automatic analyzer (Schott Instruments). Elemental analysis of carbon,

hydrogen, and nitrogen was carried out using a Vario EL Cube elemental

analyzer (Elementar). A bio-oil sample of about 5 mg was mixed with tungsten

trioxide, purged with oxygen, and sealed in a tin capsule with a capsule press

(Elementar). The higher heating value (HHV) was calculated using the

following formula based on the elemental analysis results (Friedl et al. 2005).

HHV = 3.55 × C2 - 232 × C - 2230 × H + 51.2 × H × C + 131 × N + 20,600

Atomic O/C and H/C were calculated to compare the hydrogenation and

deoxygenation degree of each heavy oil; the degree of deoxygenation (DOD)

was estimated as follows.

Degree of deoxygenation (%) = (MObio-oil − MOheavy oil) / MObio-oil × 100

Here, MOheavy oil and MObio-oil are the molar oxygen/carbon ratios of heavy oil

and bio-oil.

Page 178: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

157

2.5.2. GC/MS analysis of low molecular compounds

For measuring low-molecular-weight compounds in bio-oil and heavy oils

qualitatively and quantitatively, GC/MS analysis was performed with

fluoranthene as an internal standard. The analysis was conducted with an

Agilent 7890B coupled with a 5975C mass selective detector (MSD) and a

flame ionization detector (FID) equipped with a DB-5 capillary column (60 m

× 0.25 mm × 0.25 um). The oven temperature was maintained at 50 °C for 5

min, followed by heating at a rate of 3 °C/min to 280 °C and holding for 20 min.

The injector and FID detector temperatures were 250 and 300 °C, respectively.

2.5.3. Gas composition

After the HDO reaction, the gas phase was captured in a gas sampling

cylinder. A GC equipped with a thermal conductivity detector (Agilent 7890A)

was used to analyze the compositions of the gases produced by HDO with the

bimetallic catalysts. The oven temperature was maintained at 50 °C for 10 min,

followed by heating at a rate of 3 °C/min to 90 °C, then heating at 20 °C/min to

150 °C and holding for 16 min. Finally, the oven was heated to 230 °C at

50 °C/min and held for 10 min. The injector and FID detector temperatures were

both 250 °C.

Page 179: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

158

3. Results and Discussion

3.1. Characterization of bimetallic catalysts

3.1.1. Surface area

Figure 4-1 shows N2 adsorption–desorption isotherms and pore size

distribution curves of the catalysts. The pore size distributions of the catalyst

were calculated from the adsorption branches of the isotherms, showing narrow

distributions. The isotherms each exhibited an inflection in the range from 0.43

to 0.98 P/P0. This broad distribution suggested the presence of inter-particle

space (Gao et al. 2015). The catalysts exhibited similar type-IV isotherms with

H4-type hysteresis loops, indicative of hexagonally structured mesoporous

particles with slit- and/or panel-shaped pores (Wang et al. 2012; Ryu et al. 1999).

Table 4-1 summarizes the textural and structural characteristics of the support

and catalysts. Compared with the support material (SBA-15), NiZn/SBA-15 and

NiCu/SBA-15 showed the greatest decreases in surface area, falling from 786.7

to 585.3 and 586.0 m2/g, respectively, whereas NiMn/SBA-15 differed much

less from the support (726.1 m2/g). The pore volumes and pore diameters of the

catalysts showed no significant change after wet impregnation.

Page 180: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

159

Figure 4-1. N2 adsorption-desorption isotherms and pore size distribution curves of bimetallic catalysts

Page 181: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

160

Table 4-1. Characterization of synthesized Ni-based catalysts

Sample SBET

a

(m2/g)

Vtb

(cm3/g)

Vmesc

(cm3/g)

Vmicd

(cm3/g)

dBJHe

(nm)

SBA-15 800.2 1.1 0.7 0.1 6.0

Ni/SBA-1 508.7 0.7 0.7 0.1 6.1

NiCu/SBA-15 586.0 0.7 0.6 0.1 6.2

NiMn/SBA-15 726.1 0.9 0.8 0.1 6.2

NiZn/SBA-15 585.3 0.7 0.6 0.1 6.2

a: Calculated by the BET method b: The total pore volume was obtained at a relative pressure of 0.99 c: Calculated by BJH method d: Calculated using the t-plot method e: Mesopore diameter calculated using the BJH method

Page 182: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

161

3.1.2. Acidity

The NH3-TPD profiles for Ni/SBA-15 exhibited one peak around 320 °C,

whereas bimetallic catalysts (NiCu/SBA-15, NiMn/SBA-15, and NiZn/SBA-15)

exhibited two peaks in the range of 220–400 °C (Figure 4-2), which resembled

literature reports (Lónyi and Valyon 2001; Liu et al. 2003). The peak at lower

temperature is generally attributed to desorption of weakly adsorbed NH3 at

weak acid sites and formation of NH4+(NH3)n groups (Lónyi and Valyon 2001).

The two peaks indicate the presence of weak acid sites originating from silicon

centers (Ambursa et al. 2017). The high temperature peak at around 600 °C was

clearly attributable to water desorption.

Page 183: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

162

Figure 4-2. NH3-TPD profiles for Ni- and bimetallic catalysts of SBA-15

Page 184: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

163

3.1.3. XRD patterns of the bimetallic catalysts

A small-angle XRD pattern of the SBA-15 support and XRD patterns of Ni-

based bimetallic catalysts supported on SBA-15 are shown in Figure 4-3. A

broad peak at 23° attributed to SiO2, the component of SBA-15 (Liu et al. 2012),

was observed in all bimetallic catalysts. The peaks at 37°, 43°, and 63°, which

respectively represent the metal oxide forms from the (100), (101), and (110)

planes, respectively, are commonly observed in all Ni-based bimetallic catalysts

(Günter et al. 2001; Richardson et al. 2003). During the XRD analysis, the

catalysts were exposed to air, causing oxidation that resulted in the metal oxide

peak at around 37°. According to the small angle XRD pattern of the SBA-15

support, the peak at 2θ = 0.47° was indexed to the (100) plane of the highly

ordered hexagonal (p6mm) lattice and had a well-ordered mesoporous structure

(Zhao et al. 1998).

Page 185: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

164

Figure 4-3. Small angle XRD pattern of SBA-15 support and XRD peaks of Ni- and bimetallic catalysts on SBA-15

Page 186: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

165

3.1.4. Surface binding energies of the Ni-based catalysts

To elucidate the structure and chemical state of the Ni-based catalysts, we

performed XPS characterization of Ni/SBA-15, NiCu/SBA-15, NiMn/SBA-15,

and NiZn/SBA-15 (Figure 4-4). The spectra showed that the Ni was mostly

oxidized. For Ni, the XPS signals can be assigned to Ni0 (853.7 eV) and to NiO

and Ni2+ (855.6 eV); the latter peak can also indicate Ni(OOH) or Ni2O3 (Zhou

et al. 2017; Luo et al. 2017).

Page 187: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

166

Figure 4-4. XPS of (a)Ni/SBA-15; (b) NiCu/SBA-15; (c) NiMn/SBA-15; (d) NiZn/SBA-15 (853 eV : Ni0; 856-857 eV : Ni2+

ions or NiO particles; 860 eV : shake-up satellite peak). The shake-up satellite peak only observed in Ni/SBA-

15 catalyst.

Page 188: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

167

3.1.5. The morphologies of support and Ni-based bimetallic catalysts

For SEM-EDS investigation, the SBA-15 support and Ni-based bimetallic

catalysts were deposited on a stub with carbon tape and coated with Pt. SEM

images of the support and the Ni-based catalysts showed hexagonal structure

(Figure 4-5) (Gao et al. 2015). The smooth surface of the SBA-15 support

(Figure 4-5 (a)) became rough and uneven after the impregnation of metal

particles. Along with the observed increase in pore diameter (Table 4-1), this

observation suggested that the metal particles were incorporated between Ni

and the other transition metals, and that this collapsed part of the support during

the impregnation. SEM-EDS micrographs (Figure 4-6) showed that the metal

particles were homogeneously distributed over the surface of the SBA-15

support. However, the metal particles formed agglomerates or dispersed

multilayers after impregnation (compared with Figure 4-6 (a) and others).

Page 189: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

168

Figure 4-5. SEM micrographs of (a) SBA-15 support; (b) Ni/SBA-15; and Ni-

base bimetallic catalysts ((c) NiCu/SBA-15; (d) NiMn/SBA-15; (e)

NiZn/SBA-15)

Page 190: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

169

Figure 4-6. SEM-EDS images of (a) SBA-15 support; (b) Ni/SBA-15; and Ni-

base bimetallic catalysts ((c) NiCu/SBA-15; (d) NiMn/SBA-15; (e)

NiZn/SBA-15)

Page 191: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

170

3.2. Effect of bimetallic catalysts on HDO yields

The performances of bio-oil HDO over catalysts are closely related to the

physicochemical properties of the support, the dispersion of the active metals,

and the acid properties of the catalysts. The morphology of SBA-15, especially

its mesoporous nature, is a crucial factor for bio-oil HDO. The bio-oil was

hydrodeoxygenated and four phases of products (heavy oil, light oil, char, and

gas) were obtained. The product yield and fuel properties greatly differed with

changes in the active metal composition used (Table 4-2). The yields of heavy

oil improved considerably from 40.3–51.2% (Ni/SBA-15) to 40.8–60.6% when

the HDO reaction was conducted over bimetallic catalysts on SBA-15 supports.

The NiCu/SBA-15 and NiMn/SBA-15 bimetallic catalysts gave significantly

enhanced heavy oil yield and properties compared with NiZn/SBA-15.

Especially, heavy oil yield increased with increasing temperature when these

two bimetallic catalysts were used. We concluded that the supported metal had

great effects on both product yield and properties. These results were supported

by previous studies (Nakagawa et al. 2014; Schutyser et al. 2016) in which high

temperature and metal loading proved beneficial for enhancing catalytic sites

active for HDO. The conversion of bio-oil via HDO over the three bimetallic

catalysts studied followed the order NiZn/SBA-15 < NiCu/SBA-15 <

NiMn/SBA-15. The higher performance of the NiMn/SBA-15 catalyst was due

its high BET surface area (Table 4-1). In addition to surface area, the

incorporated mesoporous structure of the SBA-15 support enhanced the

dispersion of the active phases and increased the efficiency of their collision

with compounds in the bio-oil. This indicates that even non-acidic supported

catalysts can show activity for the HDO of bio-oil, a finding that is supported

by previous research on the HDO of model compounds over acidic and non-

acidic supports (Selvaraj et al. 2014; Nava et al. 2009; Zhu et al. 2011).

Page 192: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

171

Table 4-2. Mass balance of the products obtained from HDO reaction with Ni-catalysts

Yield (%) Char Light oil Heavy oil Gas

Ni/SBA-15

250 10.4 (0.2) a 30.4 (0.3) 51.2 (1.3) 8.0 (0.7)

300 16.8 (0.4) 28.3 (0.5) 47.7 (1.3) 7.1 (0.7)

350 13.2 (0.2) 20.3 (0.6) 40.3 (0.7) 26.2 (1.3)

NiCu/SBA-15 250 8.2 (0.1) 9.7 (0.1) 50.2 (1.0) 31.9 (2.0)

300 10.0 (0.3) 9.3 (0.1) 55.3 (0.8) 25.4 (1.1)

350 22.3 (0.0) 3.7 (0.0) 50.4 (1.2) 23.7 (0.8)

NiMn/SBA-15 250 10.2 (0.2) 5.4 (0.2) 40.8 (0.7) 43.7 (0.8)

300 9.3 (0.1) 4.1 (0.1) 55.2 (0.5) 31.4 (1.2)

350 11.4 (0.7) 11.7 (0.5) 60.6 (0.7) 16.3 (0.8)

NiZn/SBA-15 250 10.6 (0.1) 9.2 (0.3) 48.1 (1.2) 32.1 (1.5)

300 12.8 (0.0) 8.3 (0.2) 41.4 (1.5) 37.6 (1.2)

350 18.0 (0.1) 3.6 (0.0) 44.1 (0.8) 34.3 (2.0)

a: standard deviation

Page 193: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

172

3.3. Fuel properties of heavy oil with Ni-based bimetallic

catalysts on mesoporous SBA-15

Typical fuel properties of bio-oil such as acidity, heating value, water content,

and viscosity are improved by the hydrocracking, hydrogenolysis, dehydration,

hydrogenation, and deoxygenation that occur during HDO (Furimsky et al.

1986). To measure the improvements in these properties, these properties were

measured in the heavy oil products and compared with the source bio-oil; Table

4-3 lists the results. The water content, which is related to the heating value and

to combustion properties in an engine, decreased from 11.7 wt% in the bio-oil

to 2.0–8.8 wt% in the heavy oils, due to dehydration of the organic phase and

separation of water into the light oil fraction (which included over 85 wt% of

the water content). The dehydration capacities of the Ni-based catalysts were

similar to those of the commercial noble metal catalyst Pt/C, with higher metal

loading. Moreover, the mesoporous hexagonal structure of the SBA-15 support

effectively catalyzed the dehydration owing to its acidic sites (Figure 4-2). The

TAN of the heavy oil was significantly reduced (49.2-59.8 mg KOH/g oil)

compared to the source bio-oil (152.8 mg KOH/g oil) and the heavy oil obtained

from monometallic Ni/SBA-15 (62.2 mg KOH/g oil). The acidity level was

49.2–59.8 mg KOH/g in the heavy oils, primarily determined by phenolic

compounds, which measured as weak acids by TAN method. The compounds

derived from carbohydrates, which affected the acidity, were converted into

esters and ketones (chemical conversion will be discussed further in the

following section). Moreover, phenolic compounds might also be converted,

and any such conversion plays an important role in acidity reduction. Reducing

the viscosity is an important point of reforming bio-oil into fuel because high

fuel viscosity produces many disadvantages during injection into an engine

(Ochoa-Hernández et al. 2013). In this experiment, the viscosity of the original

Page 194: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

173

bio-oil was not detected, but the heavy oils decreased to 3.2–4.3 cSt. This lower

viscosity of heavy oils is likely due to how the solvent reacted as a co-reactant.

Moreover, the heavy oils viscosity obtained from bimetallic catalysts was lower

than that of the heavy oils obtained from monometallic Ni/SBA-15 (6.8-7.1 cSt).

Page 195: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

174

Table 4-3. Physicochemical properties of heavy oil with bimetallic catalysts under 250, 300, and 350 °C

Bio-oil Ni/SBA-15

NiCu/SBA-15 NiMn/SBA-15 NiZn/SBA-15

Hydrodeoxygenation temperature (°C)

250 300 350 250 300 350 250 300 350

Water content

(wt%) 11.7 11.4 4.5 2.4 2.9 3.7 6.3 8.1 2.0 6.5 8.8

TAN

(mg KOH/g oil) 152.8 62.2 59.8 54.3 54.8 50.3 51.0 49.2 55.2 54.8 53.5

Viscosity (cSt) a - 7.1 3.5 4.3 4.2 3.2 3.7 4.0 4.1 4.0 4.0

C (wt%) 45.8 58.9 58.8 58.6 60.8 56.8 60.6 62.4 64.5 63.2 64.5

H (wt%) 6.8 8.8 7.3 7.4 7.4 7.4 7.8 7.6 8.5 8.2 8.8

N (wt%) 0.9 0.1 0.9 0.9 0.5 0.9 0.3 0.6 0.2 0.1 0.2

O b (wt%) 46.2 32.2 32.7 32.8 31.0 34.6 31.1 29.2 26.6 28.3 26.2

S (wt%) 0.3 0.0 0.3 0.3 0.3 0.3 0.2 0.2 0.2 0.2 0.3

HHV(MJ/kg) c 18.3 22.8 25.0 24.0 29.5 25.0 26.4 28.4 26.2 27.4 29.9

Energy

efficiency (%) - 62.8 68.6 75.5 72.1 53.5 79.6 90.5 77.6 64.1 71.9

Degree of

deoxygenation - 54.9 44.9 44.5 49.5 39.6 49.1 53.6 59.1 55.6 59.7

a: measured at 40°C b: calculated by difference c: high heating value was calculated following formula,

HHV = 3.55 × C2 - 232 × C - 2230 × H + 51.2 × H × C + 131 × N + 20,600

Page 196: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

175

Because high oxygen levels result in high acidity, high viscosity, and low

heating values, deoxygenation is vital in adding value to bio-oils. HDO

decreases oxygen levels in bio-oil by means of deoxygenation processes,

including demethoxylation, dehydroxylation, and/or dehydration. The

elemental compositions of the bio-oil and heavy oils were determined and are

presented in Table 4-3. Compared to the carbon (45.8 and 58.9 wt%), hydrogen

(6.8 and 8.8 wt%), and oxygen (46.2 and 32.2 wt%) contents of the bio-oil and

heavy oil from monometallic Ni/SBA-15, the carbon and hydrogen levels

increased (56.8–64.5 and 7.3–8.8 wt%, respectively) and the oxygen level

significantly decreased (26.2–34.6 wt%) in the heavy oils from the bimetallic

catalysts. Use of the bimetallic catalysts also yielded heavy oils with greater

HHV (24.0–29.9 MJ/kg) than the source bio-oil (18.3 MJ/kg) and the heavy oil

produced over the monometallic Ni/SBA-15 catalyst (22.8 MJ/kg). A Van

Krevelen diagram of the bio-oil and heavy oils (Figure 4-7, dry basis), based on

the elemental analyses, showed a decrease of the O/C ratio from 0.7 for the bio-

oil to 0.3 for NiZn/SBA-15, 0.3–0.4 for NiMn/SBA-15, and 0.4 for NiCu/SBA-

15. These changes were due to the removal of oxygenated compounds in the

bio-oil to the gas (NiZn/SBA-15) or light oil phase (NiCu/SBA-15 and

NiMn/SBA-15) via deoxygenation, as well as to dehydration. Especially, Zn

had a significant tendency to enhance hydrogenation and deoxygenation, due to

its capacity to selectively cleave C–O bonds in a variety of monomeric, dimeric,

and polymeric lignin molecules under relatively mild conditions (Parsell et al.

2013; Spanjers et al. 2015). Moreover, the well-dispersed NiZn/SBA-15 (Figure

4-5) accelerated hydrogenation and deoxygenation, yielding the highest DOD

level (55.6–59.7%). The decrease in H/C ratio in the heavy oil obtained from

NiMn/SBA-15 (1.2–1.4) compared to that of the bio-oil (1.4) could have

resulted from dehydration and deoxygenation, whereas NiZn/SBA-15 and

NiCu/SBA-15 enhanced the H/C ratio to 1.4–1.5 and 1.4, respectively. We

suppose that the NiCu/SBA-15 and NiZn/SBA-15 effectively activated

Page 197: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

176

hydrogenation in the temperature range 250–350 °C.

Catalytic activity is always associated with several factors, such as metal

composition, support type, metal dispersity, pore diameter, and acidity. The

acidity and support significantly affected the product yield, properties, and

process efficiency. Improving metal dispersity leads to improved catalytic

activity and larger metal particles due to the particle size effect (Martin and

Dalmon 1982). Herein, both total metal loading and pore size were almost

similar. However, the introduction of additional transition metals decreased the

surface area in comparison with the support (Table 4-1). The catalytic activity

of the different catalysts (Table 4-2) suggested that better dispersity with large

pore volume would improve HDO activity, as the catalyst including Mn as the

promoter exhibited better conversion to the targeted heavy oil fraction than the

other two transition metals (Cu or Zn).

Page 198: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

177

Figure 4-7. Van Krevelen diagram of bio-oil, and heavy oils

Page 199: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

178

3.4. Chemical composition of hydrodeoxygenatively

upgraded heavy oil

Twenty-three typical compounds in the source bio-oil and the heavy oil

products were identified from the GC/MS analysis and classified based on their

chemical structure and functional groups. Table 4-4 lists the product

distributions for the various Ni and promoter metals catalysts on mesoporous

SBA-15 supports. The Ni-based catalysts mainly generated guaiacols

compounds in the heavy oil fraction, such as guaiacol, 4-ethyl-guaiacol, and 4-

propyl-guaiacol. Minor compounds such as esters, ketones, and other phenolic

compounds were also present in the heavy oil produced by bio-oil HDO. The

unstable and undesired acetic acid, furfural and levoglucosan in the bio-oil were

totally converted to acetic acid ethyl esters and cyclopentenone in the heavy oil

via hydrogenation, decarbonylation, dehydroxylation, and ring opening, or were

converted to low molecular gases in the gas phase. The NiMn/SBA-15 catalyst

proved excellent for the production of phenolic compounds, with guaiacol as

the main product. Phenol, 4-ethyl-phenol, and cresol, which have much higher

H/Ceff compared with other compounds such as syringol, were highly detected

in NiZn/SBA-15. Generation of phenol, 4-ethyl-phenol and removal of 4-

propenyl-guaiacol suggest that the Ni-based catalysts can accelerate

demethoxylation and hydrogenation. However, the persistence of 4-propyl-

guaiacol after the HDO and the production of 4-propyl-syringol suggest that Ni-

based catalysts might not effectively reduce the activation energy of

dealkylation or decomposition. Based on a previous study (Ambursa et al. 2017),

we investigated the possible reaction pathways over bimetallic catalysts with

similar transition metals. Catechol should be the intermediate compound of the

conversion of guaiacol into cyclohexanone, produced by the respective

cleavages of methyl and methoxy groups from guaiacol and syringol, attributed

Page 200: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

179

to the acid functions of the SBA-15 support (Scheme 1).

Scheme 1. Proposed reaction pathways for conversions of syringol or

4-methyl-guaiacol over Ni-based bimetallic catalysts above 300 °C

The hydrogenation or dehydroxylation were performed by metal active sites

under high temperature and hydrogen pressure. The high activity of the

bimetallic catalysts supported on SBA-15 can be explained by the external

surfaced distribute with mesoporous cavities. Moreover, the characteristics

could make the internal surfaces have high connectivity and high dispersion of

metal particles. The addition of Cu, Mn, and Zn to the monometallic Ni catalyst

increased the ketone content of the bio-oil HDO product. The maximum yield

among the three bimetallic catalysts studied was that of NiMn/SBA-15.

Interestingly, acetosyringone was a product of the bio-oil HDO over

NiMn/SBA-15.

Page 201: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

180

Table 4-4. Low molecular compounds in bio-oil and heavy oils

Compound

Amount (mg compound/ml oil)

Bio-oil Ni/SBA-15

NiCu/SBA-15 NiMn/SBA-15 NiZn/SBA-15

Hydrodeoxygenation temperature (˚C)

250 300 350 250 300 350 250 300 350

Acids Acetic acid 17.4 - - - - - - - - - -

Esters Acetic acid, ethyl ester - 28.1 8.3 10.0 2.6 9.9 14.6 13.5 9.1 5.0 10.5

Hexadecanoic acid, ethyl ester - - 3.0 4.3 0.8 3.0 1.8 3.4 3.0 0.2 2.5

Ketones Cyclohexanone - - - - - - 0.9 1.3 - 0.6 1.3

2-Methyl-2-cyclopenten-1-one - - 0.6 1.5 3.8 1.1 2.7 1.1 0.9 0.4 0.5 2,3-Dimethyl-cyclopenten-1-one 0.6 - - 0.6 0.6 0.6 0.9 1.9 - 1.0 2.3

Acetosyringone - - - - 0.9 1.7 1.0 0.7 - 0.0 0.5

Aldehydes Syringaldehyde 0.8 - - - - - - - - - -

Alcohols Ergosta-5,22-dien-3β-ol - - - - 1.2 - - 0.9 - - 1.0

Sugars Levoglucosan 14.2 - - - - - - - - - -

Phenols Phenol - 0.3 - - - - - - - 0.4 0.6

Cresol 0.4 2.3 - 0.3 0.4 - 0.6 1.4 - - 2.0

Guaiacol 11.0 12.2 7.6 14.4 4.1 10.7 17.8 18.1 7.8 14.0 17.8

4-Ethyl-phenol 0.7 - - 1.2 1.8 - 1.4 2.6 - - 2.7

4-Methyl-guaiacol 1.0 1.9 0.6 0.9 0.7 0.7 1.0 1.4 0.6 1.0 1.3 4-Ethyl-guaiacol 2.3 5.8 2.8 6.8 4.8 3.7 7.2 8.8 3.7 1.9 7.5

Syringol - 1.6 3.0 3.5 1.1 3.3 4.3 1.3 2.8 0.9 0.5

2,4-Dimethoxyphenol 0.7 - 0.9 0.8 2.4 0.8 0.8 - 0.8 0.3 - 4-Propyl-guaiacol - 5.6 - 3.8 6.1 1.1 4.2 5.2 1.3 1.1 3.7

4-Methyl-syringol 9.3 - 1.8 1.8 0.9 1.8 2.2 3.6 1.6 - -

4-Propenyl-guaiacol 3.0 - - - - - - - - - - 4-Propyl-syringol - 7.2 - 3.9 3.2 1.3 4.2 3.2 1.6 1.4 1.6

Total 61.5 65.0 53.6 35.6 39.8 65.5 68.4 33.2 28.2 56.3

Page 202: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

181

3.5. Gas composition

The gas phases produced by the HDO reactions were captured in a gas

cylinder and analyzed by means of GC-TCD; Table 4-5 lists the distributions of

the gas phases. CH4, CO, C2H4, C3H6, and C3H8 were formed when HDO was

conducted with Ni-based monometallic and bimetallic catalysts. Compared

with monometallic Ni/SBA-15, the Ni-based bimetallic catalysts produced gas

phases with more hydrocarbons. The CH4 yield was highest for NiZn/SBA-15

among the bimetallic catalysts studied; HDO with NiZn/SBA-15 produced 53.4%

CH4, compared to 19.6% for NiCu/SBA-15 and 18.0% for NiMn/SBA-15.

Generally, CH4 originated from demethylation and slight hydrogenation

(Furimsky et al. 1986), but the high concentration of CH4 in the gas produced

with NiZn/SBA-15 might be due to the selective cleavage of the C–O bonds of

methoxy groups by Zn under the hydrogen atmosphere (Scheme 1).

Contrastingly, CO concentrations increased significantly when monometallic

Ni/SBA-15 was used (90.1%), possibly due to the release of methoxy groups

through decarboxylation (Şenol et al. 2005).

Page 203: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

182

Table 4-5. Distribution of representative gas components determined by TCD-FID after HDO reaction

Concentration (%) Ni/SBA-15 NiCu/SBA-15 NiMn/SBA-15 NiZn/SBA-15

CH4 5.8 19.6 18.0 53.4

CO 90.1 63.6 61.9 21.2

C2H4 - 6.9 8.7 16.3

C2H2 1.5 5.6 6.4 2.8

C3H6 - - - 4.6

C3H8 2.7 4.2 5.0 1.6

Page 204: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

183

4. Summary

Mesoporous SBA-15 was synthesized via the sol–gel method, and Ni-based

bimetallic catalysts (NiCu/SBA-15, NiMn/SBA-15, and NiZn/SBA-15) were

prepared via wet impregnation. Characterization of the catalysts showed that

the Ni particles were well dispersed on the highly ordered and hexagonally

structured SBA-15 supports. The catalytic activity of the bio-oil HDO was

investigated in an autoclave reactor. HDO was affected by the base Ni particles

and the promoter metals (Cu, Mn, and Zn), and bio-oil was converted to gas,

char, light oil, and heavy oil. Analysis of the resulting heavy oil fractions

indicated that the bimetallic catalysts exhibited greater conversion activity

(heavy oils yielded 40.8–60.6%) than monometallic Ni/SBA-15 (40.3–51.2%).

The higher activity of the catalyst was associated with its high surface area and

pore size, which indicated uniform dispersion of metal particles in the catalyst.

Use of the bimetallic catalysts for HDO improved the resulting bio-oil features

such as water content, viscosity, acidity, oxygen level, and HHV, with various

differences for the three promoter metals used. Deoxygenation and

hydrogenation were enhanced with increasing temperature when NiZn/SBA-15

was used. The maximum yield of high-quality heavy oil (60.6%) was observed

for HDO over NiMn/SBA-15 at 350 °C.

Page 205: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

184

Page 206: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

185

Chapter 5

Screening of the catalysts HDO activity in

continuous reactor and biofuel blending test for

engine application

Page 207: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

186

1. Introduction

Most activities in modern society heavily depend on fossil resources like oil,

natural gas, and coal. In 2012, the world energy consumption was about 549

quadrillion Btu, as same as 255 million barrel per day, and predicted that will

increase to 813 quadrillion Btu in 2040 (Conti et al. 2016). Since most of it is

produced from non-renewable fossil-based resources, high energy consumption

resulted in international concerns including dwindling sources, CO2 emission,

and economic and political problems. Therefore, development of the process

that has capable of producing the high energy density drop-in hydrocarbons are

required for transportation, with minimal processing steps for large-scale with

low-cost (Venkatakrishnan et al. 2015). Among attractive resources (solar, wind,

hydroelectric, geothermal, and biomass), biomass is the only sustainable

resource for the production of fuels, chemical, and carbon-based materials,

especially for liquid hydrocarbon fuels for transportation (Serrano-Ruiz and

Dumesic 2011; Huber et al. 2006). Therefore, the largest sustainable carbon-

based energy source, biomass, are attractive raw materials for fuels, resulting in

more independent fuel supplies than global petroleum supply and price (Madsen

et al. 2011; Torres et al. 2007). Generally, the bio-oil obtained from fast

pyrolysis contains a variety of oxygen-rich compounds, which are chemically

unstable and undergoes oligomerization and polymerization reactions over

times and temperature. Additionally, bio-oil has lower heating value of about

16 MJ/kg compared with 43 MJ/kg for petro-diesel (Brammer et al. 2006).

Therefore, catalytic hydrotreatment is one of the promising steps for upgrading

the bio-oil with simultaneously occurred reactions. The key objectives of

upgrading are deoxygenation, minimization of carbon losses to coke and non-

condensable gases, and hydrogenation, along with C-C coupling reactions, to

shift the product slate toward the distillate-range. Especially, hydro-

Page 208: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

187

desulfurization (HDS), hydrodenitrogenation (HDN), and hydrodeoxygenation

(HDO) to remove S, N, and O are the key processes in modern oil refining,

often accompanying hydrogenation of olefins and aromatics (Furimsky 2000).

Therefore, one of the most promising route to liquid fuels from biomass is the

combination of fast pyrolysis and HDO, which could be effective to

transformation of highly oxygenated biomass into existing fuels (Mortensen et

al. 2011; Bridgwater 1996; Czernik and Bridgwater 2004; Venderbosch et al.

2010; Huber et al. 2006). During HDO, the O content in bio-oils reduced by

saturating C=O bond, cleaving C-O bond, and forming C-H bond

(deoxygenation), as well as C-C hydrogenolysis or cracking. Deoxygenation is

preferable to hydrogenation to minimize H2 consumption and maintain a high

aromatic content with remaining octane number. Moreover, this process could

increase the energy density and stability but reduce the viscosity of bio-oil

(Huber et al. 2006). The development of low-cost, earth-abundant catalysts that

could promote these transformations over extended catalyst lifetimes, with

controlled active sites to promote desired transformations (ex. Deoxygenation,

hydrogenation, and C-C coupling), and resist deactivation (Roberts et al. 2016;

Iliopoulou et al. 2012; Boscagli et al. 2015).

Batch type reactors have been used as a rapid screening of potential catalysts

and possible reaction conditions for the hydrotreatment of bio-oil or model

compounds (Wildschut et al. 2009a; Wildschut et al. 2009b; Elliott and Hart

2008; Wildschut et al. 2010a; Wildschut et al. 2010b; de Miguel Mercader et al.

2011; de Miguel Mercader et al. 2010; Tang et al. 2009; French et al. 2011).

Determination of the intrinsic selectivity for the reaction over a particular

catalyst is complicated by the factors like feedstock, temperature, diffusion of

H2, pressure, or even the catalyst characterizations such as volume or stability

(Luo et al. 2015a; Luo et al.,2015b). The common heterogeneous catalyst for

the upgrading process, transition metal promoted MoS2, exhibits high rate if

deactivation in the presence of water and requires a continuous feed of sulfide

Page 209: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

188

reagents to maintain catalyst stability, leading to sulfur-contained species in the

fuel that should be removed (Baldauf et al. 1994; Badawi et al. 2011; Ruddy et

al. 2014). CoMo- and NiMo-based HDS catalysts and Ni-based catalysts are

also effective. Previous studies (Şenol et al. 2007; Kubicka et al. 2007)

investigated HDO of heptanoic acid, heptanol, and heptanoate methyl and ethyl

esters, and rapeseed oil with sulfide/unsulfided CoMo- and NiMo/Al2O3. It

resulted that sulfide NiMo/Al2O3 showed higher activity than CoMo/Al2O3 or

unsulfided NiMo/ Al2O3. But those catalysts presented lower activity than

Noble metal catalysts (Ferrari et al. 2001; De la Puente et al. 1999; Mortensen

et al. 2013). Noble metal cataylsts, such as Pt, Ru, and Pd, are robust and do not

require supplemental cofeeds to maintain activity. However, they are expensive

and less abundant elements (Boonyasuwat et al. 2013; Gao et al. 2014; Griffin

et al. 2016). Since zeolites offer sufficient thermal stability, high specific area,

and have acidic sites, noble metal and other active metal particles supported on

zeolites have been studied for the upgrading process (Mortensen et al. 2011;

Huber et al. 2006; Gayubo et al. 2004a; Gayubo et al. 2004b; Sharma and

Bakhshi 1993; Pattiya et al. 2008; Vitolo et al. 2001; Murata et al. 2010a, 2010b;

Castaño et al. 2009; Ausavasukhi et al. 2012). Previous studies reported that

1wt Pt/Al2O3 and 1 wt% Pt/MgO conducted intensive HDO of guaiacol via

hydrogenolysis and hydrogenation over metal sites and transalkylation over

support with increasing temperature and pressure (Runnebaum et al. 2012;

Nimmanwudipong et al. 2011; Nimmanwudipong et al. 2012; Nimmanwudipon

g et al. 2014). Resasco et al. (Do et al. 2009) also investigated HDO of methyl

octanoate and methyl stearate over Pt/Al2O3, and resulted in decarbonylation

occurred as a primary reaction, and high H2 pressure prohibited self-

condensation of the compounds. Continuous reactors is used in order to get

further detailed information for design and operation of commercial process for

bio-oil upgrading (Venderbosch et al. 2010; Elliott and Neuenschwander 1997;

Elliott et al. 2006; Elliott et al. 2009; Oasmaa et al. 2010). The liquid hourly

Page 210: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

189

space velocity (LHSV) of the bio-oil feed is one of the most important

parameters for the continuous hydrotreatment reactor operation. Previous

studies conducted bio-oil hydrotreating with Ru/C in continuous flow reactor

and yielded 64-81 % upgraded oil with 32-46 % deoxygenation degree

(Bridgwater and Boocock 2006; Wang et al. 2013). Moreover, they performed

dual stage hydrotreating with more severe hydrocracking step using Pd/C with

varied contact time, changing by the flow rate (Elliott et al. 2009). Therefore,

screening of the catalysts used in batch type reactor was conducted with

continuous reactor for bio-oil HDO at the temperature ranged 250-350 °C with

H2 pressure. These metals were chosen because they are all good hydrogenation

catalysts (Griffin et al. 2016). Ethanol (4:1 w/w with bio-oil) was used as the

solvent as well as a co-reactant, because it is able to form a single-phase mixture

with bio-oil and worked as an H-donor during HDO. This could investigate the

activity of different metal particles in bio-oil HDO with continuous reactor.

Page 211: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

190

2. Materials and methods

2.1. Preparation and characterization of catalysts

The catalysts were synthesized of purchased and used for the screening of

bio-oil HDO with continuous reactor. Synthesized catalysts were reduced at

800 °C for 3 h under an H2 atmosphere before using.

Metal loading on the catalysts were measured by acid pretreatment using

Microwave (Anton Parr). Then ICP-ES analysis of pretreated liquid was

conducted with ICAP 7400 Duo. Pt, Ni, and Mn was analyzed under the

wavelength of 340.5, 231.6, 257.6 nm, respectively. Nitrogen adsorption-

desorption experiments were performed at 77K in ASAP 2010 to determine the

Brunauer-Emmett-Teller (BET) surface area of support and catalysts. Samples

were degassed at 523K for 6h before the tests. The morphology of each of the

catalysts was investigated by FE-SEM (field-emission scanning electron

microscopy, SIGMA). The acidity of the catalysts was measured by the NH3-

TPD method. X-ray powder diffraction patterns (XRD) was analyzed by a

Bruker D8 Advance with DAVINCI using Cu Kα radiation (λ= 1.5418Å ),

operated at 40kV and 40mA with a scan speed of 0.5 sec/min in a range of 10-

70 degrees (2θ). And small-angle X-ray scattering (SAXS) for mesoporous

hexagonal support was analyzed by a Bruker D8 Advance with VANTEC500

using Cu Kα radiation (λ= 1.5418Å ), operated at 50kV and 1000μA in a range

of 0-9 degrees (2θ). Surface properties of the catalysts were analyzed by AXIS-

HIS (Kractos Inc.). X-ray photoelectron spectroscopy (XPS) was performed

using Mg in an Mg/Al dual anode (15 mA, 12kV), and the spectra were analyzed

using Casa XPS software.

Page 212: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

191

2.2. Hydrodeoxygenative activity test of prepared catalyst

The bio-oil was hydrodeoxygenated for the activity test of prepared catalysts.

The stainless steel reactor equipped with a thermocouple, a pressure sensor, and

a system for controlling the temperature and pressure was used. Before the

reaction, the catalysts were activated directly in the reactor by reduction under

a H2 atmosphere at 550 °C for 3 h.

After activation, 3 g of the catalysts were placed inside the flow reactor to

make the LHSV as 2 h-1 (which showed no plugging and high yield of upgraded

oil in literature reviews). Bio-oil and ethanol mixture (4:1, w/w) was flowed 2

ml/min with 400 ml/min H2. The pressure is maintained 3 MPa. After the

reaction, four phases of products: liquid (immiscible light oil and lower heavy

oil), char, and gas were obtained. The solvent based water-rich light oil and

organic heavy oil were recovered after the reaction and separated using a

separating funnel. The yields of char, light oil and heavy oil were calculated

according to the following equations.

Yield of char (%) = [Solid (g) Catalyst (g)] / [Bio-oil (g) + Ethanol (g)] 100

Yield of light oil (%) = [Light oil (g)] / [Bio-oil (g) + Ethanol (g)] 100

Yield of heavy oil (%) = [Heavy oil (g)] / [Bio-oil (g) + Ethanol (g)] 100

Yield of gas (%) = 100 [Yield of light oil + Yield of heavy oil + Yield of char]

Page 213: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

192

2.3. Measuring fuel quality of the products

Karl-Fischer titration was used to measure the water content, and ASTM

D664 method were performed to measure the acidity (the total acid number,

TAN) of the bio-oil and heavy oils. Each samples were analyzed every three

times, and the average value is given in the Tables. The viscosity was

determined using a capillary-type viscometer and ViscoClock (Schott

Instruments) at 40 °C. Elemental analysis of carbon, hydrogen, and nitrogen

was performed out with a LECO CHNS-932 analyzer. The higher heating value

(HHV) was calculated using the following equation, based on the elemental

analysis (Friedl et al. 2005).

HHV = 3.55 × C2 - 232 × C - 2230 × H + 51.2 × H × C + 131 × N + 20,600

where C, H, and N are wt% of carbon, hydrogen, and nitrogen, respectively.

The effective hydrogen-to-carbon ratio (H/Ceff) is defined in equation (Carlson

et al. 2011);

H/Ceff = (H - 2 × O) / C

where H, C, and O are the moles of hydrogen, carbon, and oxygen, respectively.

Atomic O/C, and H/C was calculated to compare the hydrogenation, and

deoxygenation degree of each heavy oil, and the degree of deoxygenation (DOD)

was estimated as follows:

Degree of deoxygenation (%) = (MObio-oil MOheavy oil) / MObio-oil 100

where MOheavy oil and MObio-oil are the molar oxygen / carbon ratio of heavy oil

Page 214: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

193

and bio-oil, respectively.

For measuring the low-molecular weight compounds in bio-oil and heavy oils

qualitatively and quantitatively, GC/MS analysis was performed with

fluoranthene as internal standard (I.S.). The analysis was conducted with an

Agilent 7890B coupled with a 5975C mass selective detector (MSD) and a

flame ionization detector (FID) equipped with a DB-5 capillary column (60m

×0.25 mm × 0.25 um). The oven temperature was maintained at 50 °C for 5 min,

followed by heating at a rate of 3 °C/min to 280 °C and holding for 20 min. The

injector and FID detector temperatures were 250 °C and 300 °C, respectively.

Page 215: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

194

2.4. Gas composition

GC-TCD (Agilent 7890A) was used to analyze the gas composition of HDO

with bimetallic catalyst. The oven temperature maintained at 50 °C for 10 min,

followed by heating at a rate of 3 °C/min to 90 °C then 20 °C/min to150 °C and

holding for 1 6min. Finally, heated to 230 °C at a rate of 50 °C and holding for

10 min. The injector and FID detector temperatures were both 250 °C.

Page 216: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

195

2.5. Emulsification with commercial petroleum

Upgraded bio-oil (heavy oil obtained from HDO of bio-oil with Pt/C on

continuous flow reactor) and bio-oil were emulsified with two types of

commercial petroleum (diesel and gasoline). Span 60 (Sigma, S7010), Brij 58

(Sigma, P5884), and IGEPAL CO-520 (Aldrich, 238643); which showed the

less emulsion stability mentioned in Table 1-4, were used as emulsifier. Three

types of emulsification (W/O, bicontinuous, and O/W) were conducted with

emulsifier at 60 °C till homogenized. The heavy oil rate and emulsified (Span

60) rate in W/O with diesel was varied to find emulsification condition. After

emulsification, stability was investigated.

Page 217: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

196

3. Results and Discussion

3.1. Yield of the HDO products obtained from continuous

flow reactor

According to Chaiwat et al., the condition of low pressure (less than 10 MPa)

with LHSV 1 h-1 could be plugged due to the coke formation. Reducing LHSV

and temperature could solve the plugging problem, but the pressure also

reduced. However, increased temperature up to 250 °C, hydrogenation and

hydrogenation were accelerated in oil phase (Chaiwat et al. 2013). In this study,

HDO with lower pressure (3 MPa), kept on stream continually, and higher

temperature (300 °C) was conducted and investigated no plugging within 1 h

experiment. Two phases of hydrodeoxygenated liquid products (light oil and

heavy oil) were obtained with continuous reactor. The oil phase with dark brown

color had higher density than the aqueous phase. The products yield, presented

in Table 5-1, differed by the catalysts. NiMn/SBA-15 resulted the highest heavy

oil yield (42.3 wt%), while Pt/C produced the lowest yield of char (2.3 wt%)

among five catalysts used in this experiment. Notably, all of the char yielded in

continuous reactor was lower than that from batch type reactor. Generally,

continuous reactor reduced char (or coke) production than batch type reactor.

However, increase of char from 2.3-4.0 wt% in activated charcoal support to

3.1-11.0 wt% in mesoporous silica support was due to the Lewis or Brønsted

acid sites, which favors the demethylation pathway (Saidi et al. 2014). Based

on the heavy oil yield, Ni/C (27.9 wt%) showed less activity than Pt/C (35.2

wt%). However, when the transition metal on high surface area and mesoporous

supports, their activity enhanced and resulted higher heavy oil yield (29.6-42.3

wt%).

Page 218: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

197

Table 5-1. Mass balance of the products obtained from HDO reaction with continuous flow reactor

Yield (%) Char Oil phase

Gas Light oil Heavy oil

Pt/C 2.3 (0.8) a 28.9 (1.7) 35.2 (2.2) 33.6 (1.8)

Ni/C 4.0 (0.3) 35.0 (1.4) 27.9 (0.9) 33.1 (2.4)

Ni/SBA-15 10.0 (0.1) 27.0 (0.2) 29.6 (1.8) 33.4 (2.0)

Ni/Al-SBA-15 11.0 (1.2) 30.0 (1.3) 32.7 (1.5) 26.3 (1.4)

NiMn/SBA-15 9.1 (0.4) 10.7 (0.2) 42.3 (2.1) 37.9 (2.9)

a: standard deviation

Page 219: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

198

3.2. Fuel properties of obtained biofuel

Water content, acidity, viscosity, and oxygen content, which were typical fuel

properties and could estimate HDO upgrading acidity, were measured and listed

in Table 5-2. The features were significantly improved after HDO via

hydrocracking, hydrogenolysis, dehydration, hydrogenation, and de-

oxygenation (Furimsky et al. 1986), and Pt/C presented the highest activity

among those five catalysts used in these experiments. Viscosity, which occurs

ignition or plugging problems with engine utilization, notably improved on 4.0-

6.7 cSt compared with bio-oil (210.8 cSt) and even with 4:1 (w/w) bio-oil

ethanol mixture (27.8 cSt). As mentioned in Chapter 2, it suggested that the

ethanol worked as co-reactant during the reaction, not only the H-donor.

The water content, related to the combustion properties in the engine,

decreased from 17.7% in the bio-oil to 1.3 wt% (Pt/C), or 9.4-11.5 wt% (Ni-

based transition metal on mesoporous silica support catalysts) in the heavy oils,

due to dehydration of the organic phase. Despite the high Ni loading, the

dehydration capacity was still lower than those of Pt.

Compared to the carbon (40.0 wt%), hydrogen (7.5 wt%), and oxygen (52.2

wt%) contents of the bio-oil, the carbon and hydrogen levels strikingly

increased (68.6-75.4 wt%, and 8.4-9.1 wt%, respectively), while the oxygen

level reduced (19.6-22.4 wt%) in heavy oil. Those values were even better than

those catalysts applied to batch type reactor (54.9-58.9 wt% in carbon, and 32.2-

34.5 wt% in oxygen content). High carbon levels with low oxygen content

resulted in high HHV and DOD. HHV of 15.8 MJ/kg in bio-oil increased to 37.9

MJ/kg in the heavy oil obtained from Pt/C, with the highest DOD (70.3%). It

supported the metal sites activated the deoxygenation, but still required more

improvement to reach the petroleum (44 MJ/kg).

The Van Krevelen diagram of the bio-oil, and heavy oils (Figure 5-1, dry

Page 220: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

199

basis) shows the heavy oils were placed in promising region (low atomic O/C

ratio with high H/C ratio). High O/C ratio of bio-oil (1.4) effectively decreased

to 0.1-0.2 in heavy oils with various deoxygenation reaction, which activated

by the metal sites. However, H/C ratio of heavy oil also decreased (1.4-1.5)

compared with that of bio-oil (2.0) despite of the hydrogenation during the

process. The reduction of hydrogen was attributed to demethylation or further

decomposition to gas phase.

Page 221: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

200

Table 5-2. Physicochemical properties of heavy oil

Bio-oil Pt/C Ni/C Ni/SBA-15 Ni/Al-SBA-15 NiMn/SBA-15

Water content (wt%) 17.7

(0.2)d 1.3 (0.4) 9.7 (0.1) 11.5 (0.8) 10.7 (0.2) 9.4 (0.7)

TAN

(mg KOH/g oil)

164.8

(0.8) 101.8 (1.5) 61.8 (0.7) 64.9 (0.5) 76.2 (1.8) 54.9 (1.6)

Viscosity(cSt)a

210.8

(0.8) 6.2 (0.3) 6.4 (0.2) 6.7 (0.1) 6.0 (0.7) 4.0 (0.1)

27.8 e

(0.2)

C (wt%) 40.0 75.4 70.4 68.6 71.2 69.9

H (wt%) 7.5 9.0 8.4 8.9 9.1 9.0

N (wt%) 0.2 0.1 0.2 0.1 0.1 0.2

Ob (wt%) 52.2 20.1 21.0 22.4 19.6 20.9

HHV(MJ/kg) c 15.8 37.9 33.0 32.4 34.6 33.6

H/Ceff - 1.1 1.0 1.1 1.1 1.1

Degree of deoxygenation

(DOD) - 70.3 59.8 57.1 62.5 60.0

a: measured at 40°C b:calculated by difference c:high heating value (HHV) was calculated following formula (dry basis)

HHV = 3.55 x C2 - 232 x C - 2230 x H + 51.2 x H x C + 131 x N + 20,600 d:Standard deviation e: Viscosity of bio-oil and ethanol mixture (4:1 w/w)

Page 222: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

201

Figure 5-1. Van Krevelen diagram of bio-oil and heavy oils, dry basis

Page 223: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

202

3.3. Converted chemical compounds during HDO

Twenty-two typical compounds in bio-oil and heavy oil were identified from

the GC/MS analysis and classified based on chemical structure and functional

groups. Their amounts are presented in Table 5-3. The distribution drastically

changed after HDO process. Bio-oil mainly consisted of carbohydrate-derived

compounds like acetic acid, levoglucosan, and furfural, while almost of heavy

oils were consisted of phenolic compounds, derived from lignin molecular.

All of the acetic acid in bio-oil converted into acetic acid, ethyl ester with

ethanol. Levoglucosan and furfural, that make bio-oil unstable, was also

converted into stable ketones, like cyclopentanone or 4-methyl-cyclopentanone

via hydrogenation, decarbonylation, dehydroxylation, and ring opening, or low

molecular gas phase. Generally, syringols and guaiacols underwent

demethoxylation, dealkylation, and hydrogenation, to produce phenols or

methyl-guaiacol or syringol. The amount of phenol was the highest in the heavy

oil obtained from Pt/C (18.4 mg/ml oil), since demethoxylation activated with

metal surface acted with oxygen of methoxy group (Scheme 1). The produced

4-propyl-syringol, not detected in bio-oil, supposed to degradation of large

molecular lignin compounds during the process.

Scheme 1. Proposed reaction mechanism for HDO of 4-methyl-guaiacol on

metal surface

Page 224: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

203

Generation of catechol in heavy oil and decrease of 3,4-

dihydroxyacetophenone suggest that intramolecular hydroxyl group transfer

and bimolecular transalkylation of guaiacol (Scheme 2). Since the

intramolecular hydroxyl group transfer performed with acid sites, Ni/Al-SBA-

15 having acidic support presented large amount of catechol or 3,4-

dihydroxyacetophenone (Table 5-3). Total amount of GC/MS detectable

compounds as increased compared with the heavy oils obtained from batch type

reactor. This also related to fewer char yielded in continuous reactor, resulting

inhibition catalyst deactivation.

Scheme 2. Plausible reaction network of guaiacol during HDO with metal

catalyst

Page 225: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

204

Table 5-3. Quantitative analysis of major micromolecular components in the bio-oil and heavy oils by GC/MS analysis

No. Compound

Amount (mg/ml bio-oil)

Bio-oil Pt/C Ni/C Ni/SBA-15 Ni/Al-SBA-15 NiMn/SBA-15

Acids 1 Acetic acid 28.0 - - - -

Esters 2 Acetic acid, ethyl ester - 10.3 14.5 14.8 7.7 12.7

Ketones 3 Cyclopentanone - 2.7 1.4 1.6 1.2 1.9

4 2-Methyl-cyclopentanone - 1.3 0.8 0.2 0.6 0.7

Phenols

5 Phenol - 18.4 16.2 14.3 15.9 16.8

6 p-Cresol - 4.2 3.5 3.9 2.1 1.0

7 Guaiacol 3.3 10.8 15.4 13.9 12.1 14.7

8 4-Ethyl-phenol 0.4 1.0 0.4 0.6 0.5 0.5

9 Creosol 1.1 1.3 2.8 2.1 2.6 2.3

10 4-Vinyl-phenol 0.4 - - - - -

11 2-Propyl-phenol 1.2 - 0.2 0.9 0.5 -

12 Catechol - 1.0 1.2 1.5 1.6 1.5

13 4-Ethyl-guaiacol 1.1 5.4 4.6 4.3 4.9 4.2

14 3,4-Dihydroxyacetophenone 0.7 0.2 0.4 0.6 1.0 0.5

15 Syringol 0.4 8.1 7.5 9.2 8.3 8.8

16 4-Propyl-guaiacol 2.1 10.2 8.4 9.2 8.0 8.2

17 4-Methyl-syringol 0.5 - 0.1 0.1 - 0.1

18 4-(1-Propenyl)-guaiacol 0.7 - - - - -

19 4-Propyl-syringol - 5.2 5.7 4.2 3.4 4.9

Sum of Phenols 11.9 66.6 11.9 66.6 66.8 64.6

Aldehydes 20 Furfural 7.6 - - - - -

21 Vanillin 1.2 - - - - -

Sugars 22 Levoglucosan 27.0 - - - - -

Total 75.7 80.9 75.7 80.1 83.1 81.4

Page 226: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

205

3.4. Gas composition

Gas phase analysis was performed, and the distribution is shown in Figure 5-

2. The results revealed that CH4, CO, C2H2, C3H6, and C3H8 were formed after

HDO experiments. Compared with Ni-based catalysts, the gas phase obtained

from HDO with Pt/C contained more CO in gas phase, which supported

deoxygenation effectively activated in metal sites.

The CH4 yield was lower with Ni-based catalysts (8.2-12.9%) compared to that

with Pt/C (20.7%), especially 8.2% in Ni/SBA-15. Since the CH4 generally

originated from demethylation and slight hydrogenation, Ni/SBA-15 with

mesoporous structure accelerated demethylation and the gas phase consisted

high yield of CH4. Moreover, C2-C3 hydrocarbon concentration was also higher

in Ni/SBA-15 than other catalysts were used.

Page 227: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

206

Figure 5-2. Gas composition determined by TCD-FID after HDO reaction

Page 228: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

207

3.5. Blends with commercial petroleum

Generally, bio-oil is not miscible with petroleum and start separating after a

few minutes due to the different density or polarity. Therefore, to estimate the

availability of heavy oil to utilization in diesel engine, the emulsification was

performed with two commercial petroleum (diesel and gasoline). To enhance

the miscibility, the emulsifier (Span60, Brij58, or IGEPAL CO-520) was added

in the blends. Heavy oil (or bio-oil) concentration (0, 25, 50, and 75%) and type

of emulsifier was varied. The emulsion after emulsification at 60 °C were

referred as D-H25-S5, which meant Diesel emulsified with 25% of Heavy oil

addition of 5% Span60 as an emulsifier.

Bio-oil hardly emulsified till 36 h emulsification with both of the diesel and

gasoline, while heavy oils emulsified within 36 h. Especially, heavy oil blended

with gasoline and 5% emulsifier totally emulsified within 12 h (Figure 5-3).

However, heavy oil miscibility with diesel was lower than with gasoline. OW

or bicontinuous emulsification (greater than 50% heavy oil) was completed

within 12 hours, whereas WO emulsification (lower than 25% heavy oil) hardly

form the emulsion with 5% Span60 emulsifier. Therefore, concentration of the

heavy oil Span60 (presented the highest activity in heavy oil emulsification)

was subdivided in the emulsification with diesel. When lower the heavy oil

concentration of 5 or 10% in WO type emulsification, it easily formed emulsion

in 6 h with addition of 5% Span60. Moreover, increasing Span60 concentration

of 10, 20, and 50% also enhance the miscibility of heavy oil and diesel, and

emulsified within 6 h. The stable emulsion formation conditions were presented

in Figure 5-4. The emulsions were investigated for 21 days to confirm their

stability. As shown in Figure 5-5, all the emulsions from heavy oil and

petroleum remained their stability for 21 days. It was supposed by the storage

stable features of heavy oil, investigated in Chapter 2. Based on the results of

Page 229: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

208

stability for 12 weeks storage, the emulsion is expected to remain stable for at

least 12 weeks. Moreover, it was confirmed that Span60 emulsifier, which was

not effective against bio-oil emulsification (Chiaramonti et al. 2003a), showed

good effects in heavy oil emulsification.

Page 230: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

209

Figure 5-3. Ternary diagram of emulsification test of Heavy oil (Upgraded oil) with petroleum; (a) gasoline and (b) diesel

(Yellow dotted region meant the conditions of forming stable emulsion)

Page 231: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

210

Figure 5-4. Images of before and after emulsification, and stable state emulsion (W/O emulsification with Span60 as

emulsifier at 60 °C)

Page 232: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

211

4. Summary

In this study, the activity test of five catalysts for bio-oil HDO in continuous

reactor and potential of heavy oil for fuel application were conducted. The bio-

oil was converted into gas, char, and immiscible liquids (light oil and heavy oil).

Compared with HDO in batch type reactor, continuous reactor produced lower

yield of char, inhibiting catalyst deactivation. The fuel properties of heavy oil,

especially HHV, were even more improved than that produced in batch type

reactor. Despite high activity of noble metal Pt (dehydration and deoxygenation),

bimetallic NiMn/SBA-15 or bifunctional Ni/Al-SBA-15 also presented high

hydrogenation and converted the carbohydrate derived or unstable phenolic

compounds into stable compounds.

In order to form a stable blend, emulsification with an emulsifier should be

considered. Heavy oil could emulsify with 5 wt% of emulsifier in gasoline,

while more emulsifier or less heavy oil concentration was needed for diesel

blending. And the heavy oil and petroleum blends maintained its stability for 21

days.

Page 233: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

212

Page 234: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

213

Chapter 6

Concluding remarks

Page 235: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

214

Biomass is the only sustainable resource for the production of fuels, chemical,

and carbon-based materials, especially for liquid hydrocarbon fuels to

transportation. Among various biomass to liquid conversion, one of the

thermochemical conversion process, fast pyrolysis, could produce high yield of

bio-oil with short and simple way. However, due to the disadvantageous features

of bio-oil, there need upgrading process of bio-oil utilization as a fuel.

Hydrodeoxygenation (HDO), kind of catalytic hydrotreating, is known as an

effective upgrading process to obtain high quality biofuel. Recently, many

researchers have been focused on the HDO mechanisms with model compounds,

but only few of them performed with real bio-oil. Moreover, the further steps

for biofuel application to petroleum engines will be required.

In this study, various Ni-based catalysts were used to HDO for improving

quality and stability of bio-oil to convert into biofuel. The effects of various

parameters, especially the temperature and catalysts on characteristics of

upgraded biofuel (heavy oil) were basically demonstrated. For better

comprehensive understanding of correlation between catalyst characterizations

and chemical conversion mechanisms, changes in fuel properties and chemical

distributions during the HDO were investigated. Furthermore, its potential as a

transportation fuel was evaluated by blending with petroleum fuel.

The solvents as a supercritical fluid state during the process could enhance

heat and mass transfer of the process. According to its polarity, the solvent

affected to the process efficiency, as well as heavy oil properties. Polar protic

ethanol, reacted as an H-donor, highly improved water content (0.5-4.0 wt%)

and acidity (63.2-75.2 mg KOH/g oil), while non-polar ether 59.1-67.5 %

deoxygenation degree. As for storage stability, the heavy oils obtained with

ethanol under Pt/C showed almost no changes within 12 weeks period. It was

due to the conversion to low molecular distribution and degradation with

functional group removal in macromolecules in bio-oil.

In order to assess the correlation between heavy oil properties and catalyst

Page 236: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

215

characterization, various Ni catalysts were synthesized and their activities test

were performed in bath type HDO of bio-oil. At first, monometallic Ni catalysts

on various supports were prepared to compare the effect of structural or

functionalities of supports. According to activity test, hexagonal mesoporous

structured supported catalysts produced high yield of heavy oil ranged in 47.1-

48.1 wt%. It was suggested that their high surface area (508.7-537.7 m2/g) and

large pore volume (0.46-0.65 cm3/g) could maximize the active metal sites.

Moreover, the acid sites, presented in the support affected dehydration during

the reaction.

Based on the results of monometallic Ni catalysts, mesoporous silica SBA-

15 support was chosen for the synthesis of bimetallic catalysts. Three transition

metals (Cu, Mn, and Zn) were selected to support Ni or additional

deoxygenation activity. All of the bimetallic catalysts yielded more heavy oil

(40.8-60.6 wt%) than monometallic catalysts (40.3-51.2 wt%). Especially,

NiMn/SBA-15 enhanced activities with temperature and showed 90.5 % of

energy efficiency at 350 °C.

Screening of those catalysts were performed in continuous HDO reactor. The

results presented similar tendency to those of batch type reactor. However, the

formation of char (coke) was notably reduced in continuous type reactor (2.3-

11.0 wt%, while 7.7-18.6 wt% of char produced in batch type reactor), resulting

in inhibition of catalyst deactivation. Therefore, the deoxygenation degree

(DOD) was increased from 49.1-73.7 % (batch type reactor) to 57.1-70.3 % in

continuous type reactor, with presented higher HHV (32.4-37.9 MJ/kg) heavy

oil. Heavy oil also consisted of the stable compounds like phenol and

cyclopentanone more than that from batch type reactor. From this result, it could

be suggested that the heavy oil produced from continuous reactor showed more

stable features during the storage period than that from batch type reactor.

Furthermore, emulsification of upgraded oil (heavy oil) and petroleum with

emulsifier was performed to estimate the possibility to conventional engine

Page 237: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

216

utilization. Heavy oil and gasoline successfully emulsified with 5 % emulsifier,

while emulsification with diesel required more emulsifier. Obtained emulsion

maintained its stable state for 7 days, while bio-oil emulsions preserved only

few minutes.

This study investigated application of liquid biofuel from hydrotreating

process of bio-oil under various reaction factors. Based on the results,

mesoporous SBA-15 support and bimetallic NiMn catalyst was proven to be an

excellent alternative the most effective noble metal (Pt) by activating of both

hydrogenation and various deoxygenation. Besides, it was demonstrated that

effects of reaction conditions on quality of heavy oil (upgraded oil) also related

to its storage or blending performance on engine utilization. Thus, it can be

possible to control the performance by adjusting the reaction factors, such as

temperature or types of catalysts for an appropriate purpose. The results of this

study suggested that biofuel obtained from hydrodeoxygeation has a great

potential for alternate petroleum based fuel.

Page 238: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

217

References

Page 239: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

218

Aburto, J., Mar-Juarez, E., and Juarez-Soto, C. 2009. Transportation of heavy

and extra-heavy crude oil by pipeline: A patent review for technological

options, Recent Pat. Chem. Eng. 2, 86-97.

Al-Sabawi, M., and Chen, J. 2012. Hydroprocessing of biomass-derived oils

and their blends with petroleum feedstocks: A Review, Energ. Fuel. 26,

5373-5399.

Ambursa, M.M., Sudarsanam, P., Voon, L.H., Hamid, S.B.A., and Bhargava,

S.K. 2017. Bimetallic Cu-Ni catalysts supported on MCM-41 and Ti-

MCM-41 porous materials for hydrodeoxygenation of lignin model

compound into transportation fuels, Fuel Process. Technol. 162, 87-97.

Amer, M., and Daim, T.U. 2010. Application of technology roadmaps for

renewable energy sector, Technol. Forecast. Soc. Chang. 77, 1355-1370.

Anastopoulos, G., Lois, E., Zannikos, F., Kalligeros, S., and Teas, C. 2002.

HFRR lubricity response of an additized aviation kerosene for use in CI

engines, Tribol. Int. 35, 599-604.

Andrews, R., Fuleki, D., Zukowski, S., and Patnaik, P. 1997. Results of

industrial gas turbine tests using a biomass-derived fuel. In 3rd Proc.

Biomass Conf. Am. 425-436. Elsevier Science Ltd.: New York.

Antal, M.J., Allen, S.G., Schulman, D., Xu, X., and Divilio, R.J. 2000. Biomass

gasification in supercritical water, Ind. Eng. Chem. Res. 39, 4040-4053.

Aponte, Y., and de Lasa, H. 2017. The Effect of Zn on Offretite Zeolite

Properties. Acidic Characterizations and NH3-TPD Desorption Models,

Ind. Eng. Chem. Res. 56, 1948-1960.

Ardiyanti, A., Bykova, M., Khromova, S., Yin, W., Venderbosch, R.H.,

Yakovlev, V.A., and Heeres, H.J. 2016. Ni-based catalysts for the

hydrotreatment of fast pyrolysis oil, Energ. Fuel. 30, 1544-1554.

Ardiyanti, A., Gutierrez, A., Honkela, M., Krause, A., and Heeres, H. 2011.

Hydrotreatment of wood-based pyrolysis oil using zirconia-supported

mono-and bimetallic (Pt, Pd, Rh) catalysts, Appl. Catal. A: Gen. 407,

56-66.

Ardiyanti, A., Khromova, S., Venderbosch, R., Yakovlev, V., and Heeres, H.

2012a. Catalytic hydrotreatment of fast-pyrolysis oil using non-sulfided

bimetallic Ni-Cu catalysts on a δ-Al2O3 support, Appl. Catal. B:

Environ. 117, 105-117.

Ardiyanti, A., Khromova, S., Venderbosch, R., Yakovlev, V., and Heeres, H.

2012b. Catalytic hydrotreatment of fast-pyrolysis oil using non-

sulfided bimetallic Ni-Cu catalysts on a δ-Al2O3 support, Appl. Catal.

B: Environ. 117, 105-117.

Ausavasukhi, A., Huang, Y., To, A.T., Sooknoi, T., and Resasco, D.E. 2012.

Hydrodeoxygenation of m-cresol over gallium-modified beta zeolite

catalysts, J. Catal. 290, 90-100.

Badawi, M., Paul, J., Cristol, S., Payen, E., Romero, Y., Richard, F., Brunet, S.,

Lambert, D., Portier, X., and Popov, A. 2011. Effect of water on the

Page 240: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

219

stability of Mo and CoMo hydrodeoxygenation catalysts: a combined

experimental and DFT study, J. Catal. 282, 155-164.

Baldauf, W., Balfanz, U., and Rupp, M. 1994. Upgrading of flash pyrolysis oil

and utilization in refineries, Biomass Bioenerg. 7, 237-244.

Baliban, R.C., Elia, J.A., and Floudas, C.A. 2013. Biomass to liquid

transportation fuels (BTL) systems: process synthesis and global

optimization framework, Energ. Environ. Sci. 6, 267-287.

Bauen, A., Chase, A., Chudziak, C., Ouwens, J.D., Denvir, B., Owen, N.,

Ripken, R., van den Berg, M., and Vuille, F. 2013. A harmonised auto-

fuel biofuel roadmap for the EU to 2030.

Bayerbach, R., and Meier, D. 2009. Characterization of the water-insoluble

fraction from fast pyrolysis liquids (pyrolytic lignin). Part IV: Structure

elucidation of oligomeric molecules, J. Anal. Appl. Pyrol. 85, 98-107.

Boateng, A., Weimer, P., Jung, H., and Lamb, J. 2008. response of

thermochemical and biochemical conversion processes to lignin

concentration in alfalfa stems, Energ. Fuel. 22, 2810-2815.

Bok, J.P., Choi, H.S., Choi, J.W., and Choi, Y.S. 2013. Fast pyrolysis of

Miscanthus sinensis in fluidized bed reactors: characteristics of product

yields and biocrude oil quality, Energy 60, 44-52.

Boonyasuwat, S., Omotoso, T., Resasco, D.E., and Crossley, S.P. 2013.

Conversion of guaiacol over supported Ru catalysts, Catal. Lett. 143,

783-791.

Boscagli, C., Raffelt, K., Zevaco, T.A., Olbrich, W., Otto, T.N., Sauer, J., and

Grunwaldt, J.-D. 2015. Mild hydrotreatment of the light fraction of fast-

pyrolysis oil produced from straw over nickel-based catalysts, Biomass

Bioenerg. 83, 525-538.

Boucher, M., Chaala, A., Pakdel, H., and Roy, C. 2000. Bio-oils obtained by

vacuum pyrolysis of softwood bark as a liquid fuel for gas turbines. Part

II: Stability and ageing of bio-oil and its blends with methanol and a

pyrolytic aqueous phase, Biomass Bioenerg. 19, 351-361.

Boucher, M., Chaala, A., and Roy, C. 2000. Bio-oils obtained by vacuum

pyrolysis of softwood bark as a liquid fuel for gas turbines. Part I:

Properties of bio-oil and its blends with methanol and a pyrolytic

aqueous phase, Biomass Bioenerg. 19, 337-350.

Bradley, T.H., and Frank, A.A. 2009. Design, demonstrations and sustainability

impact assessments for plug-in hybrid electric vehicles, Renew. Sust.

Energ. Rev. 13, 115-128.

Brammer, J., Lauer, M., and Bridgwater, A. 2006. Opportunities for biomass-

derived “bio-oil” in European heat and power markets, Energ. Policy

34, 2871-2880.

Bridgewater, A.V. 2004. Biomass fast pyrolysis, Therm. Sci. 8, 21-50.

Bridgwater, A. 1996. Production of high grade fuels and chemicals from

catalytic pyrolysis of biomass, Catal. Today 29, 285-295.

Page 241: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

220

Bridgwater, A. 1999. Development of advanced fast pyrolysis processes for

power and heat. In JOULE Contractors meeting.

Bridgwater, A., and Boocock, D. 2006. Sci. Therm. Chem. Biomass Conver.

(CPL press).

Bridgwater, A.V. 2012. Review of fast pyrolysis of biomass and product

upgrading, Biomass Bioenerg. 38, 68-94.

Brown, R.C., and Brown, T.R. 2013. Biorenew. Res. Eng. Prod. Agr. (John

Wiley & Sons).

Brownsort, P.A. 2009. Biomass pyrolysis processes: review of scope, control

and variability, Edinburgh: UK Biochar Research Center.

Bui, P., Cecilia, J.A., Oyama, S.T., Takagaki, A., Infantes-Molina, A., Zhao, H.,

Li, D., Rodríguez-Castellón, E., and López, A.J. 2012. Studies of the

synthesis of transition metal phosphides and their activity in the

hydrodeoxygenation of a biofuel model compound, J. Catal. 294, 184-

198.

Bui, V.N., Laurenti, D., Delichère, P., and Geantet, C. 2011.

Hydrodeoxygenation of guaiacol: Part II: Support effect for CoMoS

catalysts on HDO activity and selectivity, Appl. Catal. B: Environ. 101,

246-255.

Carlson, T.R., Cheng, Y.-T., Jae, J., and Huber, G.W. 2011. Production of green

aromatics and olefins by catalytic fast pyrolysis of wood sawdust,

Energ. Environ. Sci. 4, 145-161.

Castaño, P., Gutiérrez, A., Villanueva, I., Pawelec, B., Bilbao, J., and Arandes,

J.M. 2009. Effect of the support acidity on the aromatic ring-opening

of pyrolysis gasoline over Pt/HZSM-5 catalysts, Catal. Today 143, 115-

119.

Centeno, A., Laurent, E., and Delmon, B. 1995. Influence of the support of

CoMo sulfide catalysts and of the addition of potassium and platinum

on the catalytic performances for the hydrodeoxygenation of carbonyl,

carboxyl, and guaiacol-type molecules, J. Catal. 154, 288-298.

Chaiwat, W., Gunawan, R., Gholizadeh, M., Li, X., Lievens, C., Hu, X., Wang,

Y., Mourant, D., Rossiter, A., and Bromly, J. 2013. Upgrading of bio-

oil into advanced biofuels and chemicals. Part II. Importance of holdup

of heavy species during the hydrotreatment of bio-oil in a continuous

packed-bed catalytic reactor, Fuel 112, 302-310.

Chen, H., Zhai, Y., Xu, B., Xiang, B., Zhu, L., Li, P., Liu, X., Li, C., and Zeng,

G. 2015. Feasibility and comparative studies of thermochemical

liquefaction of Camellia oleifera cake in different supercritical organic

solvents for producing bio-oil, Energ. Conver. Manag. 89, 955-962.

Chiaramonti, D., Bonini, M., Fratini, E., Tondi, G., Gartner, K., Bridgwater, A.,

Grimm, H., Soldaini, I., Webster, A., and Baglioni, P. 2003a.

Development of emulsions from biomass pyrolysis liquid and diesel

and their use in engines—Part 1: emulsion production, Biomass

Page 242: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

221

Bioenerg. 25, 85-99.

Chiaramonti, D., Bonini, M., Fratini, E., Tondi, G., Gartner, K., Bridgwater, A.,

Grimm, H., Soldaini, I., Webster, A., and Baglioni, P. 2003b.

Development of emulsions from biomass pyrolysis liquid and diesel

and their use in engines—Part 2: tests in diesel engines, Biomass

Bioenerg. 25, 101-111.

Chiaramonti, D., Oasmaa, A., and Solantausta, Y. 2007. Power generation using

fast pyrolysis liquids from biomass, Renew. Sus. Energ. Rev. 11, 1056-

1086.

Chiranjeevi, T., Kumaran, G.M., and Dhar, G.M. 2008. Synthesis,

characterization, and evaluation of mesoporous MCM-41-supported

molybdenum hydrotreating catalysts, Petrol. Sci. Technol. 26, 690-703.

Choudhary, T., and Phillips, C. 2011. Renewable fuels via catalytic

hydrodeoxygenation, Appl. Catal. A: Gen. 397, 1-12.

Cioffi, M., Hoffmann, A., and Janssen, L. 2001. Rheokinetics and the influence

of shear rate on the Trommsdorff (gel) effect during free radical

polymerization, Polym. Eng. Sci. 41, 595-602.

Conti, J., Holtberg, P., Diefenderfer, J., LaRose, A., Turnure, J.T., and Westfall,

L. 2016. International Energy Outlook 2016 With Projections to 2040.

In.: USDOE Energy Information Administration (EIA), Washington,

DC (United States). Office of Energy Analysis.

Crossley, S., Faria, J., Shen, M., and Resasco, D.E. 2010. Solid nanoparticles

that catalyze biofuel upgrade reactions at the water/oil interface,

Science 327, 68-72.

Czernik, S., and Bridgwater, A. 2004. Overview of applications of biomass fast

pyrolysis oil, Energ. Fuel. 18, 590-598.

Damartzis, T., and Zabaniotou, A. 2011. Thermochemical conversion of

biomass to second generation biofuels through integrated process

design—A review, Renew. Sus. Energ. Rev. 15, 366-378.

De la Puente, G., Gil, A., Pis, J., and Grange, P. 1999. Effects of support surface

chemistry in hydrodeoxygenation reactions over CoMo/activated

carbon sulfided catalysts, Langmuir 15, 5800-5806.

de Miguel Mercader, F., Groeneveld, M., Kersten, S., Way, N., Schaverien, C.,

and Hogendoorn, J. 2010. Production of advanced biofuels: Co-

processing of upgraded pyrolysis oil in standard refinery units, Appl.

Catal. B: Environ. 96, 57-66.

de Miguel Mercader, F., Groeneveld, M.J., Kersten, S.R., Geantet, C., Toussaint,

G., Way, N.W., Schaverien, C.J., and Hogendoorn, K.J. 2011.

Hydrodeoxygenation of pyrolysis oil fractions: process understanding

and quality assessment through co-processing in refinery units, Energ.

Environ. Sci. 4, 985-997.

De Miguel Mercader, F., Koehorst, P., Heeres, H., Kersten, S., and Hogendoorn,

J. 2011. Competition between hydrotreating and polymerization

Page 243: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

222

reactions during pyrolysis oil hydrodeoxygenation, AIChE 57, 3160-

3170.

De Rogatis, L., Montini, T., Cognigni, A., Olivi, L., and Fornasiero, P. 2009.

Methane partial oxidation on NiCu-based catalysts, Catal. Today 145,

176-185.

De, S., Zhang, J., Luque, R., and Yan, N. 2016. Ni-based bimetallic

heterogeneous catalysts for energy and environmental applications,

Energ. Environ. Sci. 9, 3314-3347.

de Wild, P., Van der Laan, R., Kloekhorst, A., and Heeres, E. 2009. Lignin

valorisation for chemicals and (transportation) fuels via (catalytic)

pyrolysis and hydrodeoxygenation, Environ. Prog. Sus. Energ. 28, 461-

469.

Demirbaş, A. 2001. Biomass resource facilities and biomass conversion

processing for fuels and chemicals, Energ. Conver. Manag. 42, 1357-

1378.

Diebold, J.P. 2000. A review of the chemical and physical mechanisms of the

storage stability of fast pyrolysis bio-oils (National Renewable Energy

Laboratory Golden, CO).

Diebold, J.P., and Czernik, S. 1997. Additives to lower and stabilize the

viscosity of pyrolysis oils during storage, Energ. Fuel. 11, 1081-1091.

Difiglio, C. 2014. Oil, economic growth and strategic petroleum stocks, Energ.

Strat. Rev. 5, 48-58.

Do, P.T., Chiappero, M., Lobban, L.L., and Resasco, D.E. 2009. Catalytic

deoxygenation of methyl-octanoate and methyl-stearate on Pt/Al2O3,

Catal. Lett. 130, 9-18.

Doll, K.M., Sharma, B.K., Suarez, P.A., and Erhan, S.Z. 2008. Comparing

biofuels obtained from pyrolysis, of soybean oil or soapstock, with

traditional soybean biodiesel: density, kinematic viscosity, and surface

tensions, Energ. Fuel. 22, 2061-2066.

Elliott, D., and Neuenschwander, G. 1997. Liquid fuels by low-severity

hydrotreating of biocrude, In Dev. Therm. Biomass Conver. (Springer).

Elliott, D.C. 2007. Historical developments in hydroprocessing bio-oils, Energ.

Fuel. 21, 1792-1815.

Elliott, D.C. 2013. Transportation fuels from biomass via fast pyrolysis and

hydroprocessing, WIRES Energ. Environ. 2, 525-533.

Elliott, D.C., and Baker, E.G. 1989. Process for upgrading biomass pyrolyzates.

In Google Patents.

Elliott, D.C., and Hart, T.R. 2008. Catalytic hydroprocessing of chemical

models for bio-oil, Energ. Fuel. 23, 631-637.

Elliott, D.C., Hart, T.R., Neuenschwander, G.G., Rotness, L.J., and Zacher, A.H.

2009. Catalytic hydroprocessing of biomass fast pyrolysis bio‐oil to

produce hydrocarbon products, Environ. Prog. Sus. Energ. 28, 441-449.

Elliott, D.C., Neuenschwander, G.G., Hart, T.R., Hu, J., Solana, A.E., and Cao,

Page 244: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

223

C. 2006. Hydrogenation of bio-oil for chemicals and fuels production.

In.: Pacific Northwest National Laboratory (PNNL), Richland, WA

(US), Environmental Molecular Sciences Laboratory (EMSL).

Ferrari, M., Bosmans, S., Maggi, R., Delmon, B., and Grange, P. 2001.

CoMo/carbon hydrodeoxygenation catalysts: influence of the hydrogen

sulfide partial pressure and of the sulfidation temperature, Catal. Today

65, 257-264.

French, R.J., Stunkel, J., and Baldwin, R.M. 2011. Mild hydrotreating of bio-

oil: effect of reaction severity and fate of oxygenated species, Energ.

Fuel. 25, 3266-3274.

Friedl, A., Padouvas, E., Rotter, H., and Varmuza, K. 2005. Prediction of

heating values of biomass fuel from elemental composition, Anal. Chim.

Acta 544, 191-198.

Furimsky, E. 2000. Catalytic hydrodeoxygenation, Appl. Catal. A: Gen. 199,

147-190.

Furimsky, E., and Massoth, F.E. 1999. Deactivation of hydroprocessing

catalysts, Catal. Today 52, 381-495.

Furimsky, E., Mikhlin, J., Jones, D., Adley, T., and Baikowitz, H. 1986. On the

mechanism of hydrodeoxygenation of ortho substituted phenols, Can.

J. Chem. Eng. 64, 982-985.

Galarneau, A., Cambon, H., Di Renzo, F., Ryoo, R., Choi, M., and Fajula, F.

2003. Microporosity and connections between pores in SBA-15

mesostructured silicas as a function of the temperature of synthesis, J.

Chem. 27, 73-79.

Gao, D., Duan, A., Zhang, X., Zhao, Z., Hong, E., Li, J., and Wang, H. 2015.

Synthesis of NiMo catalysts supported on mesoporous Al-SBA-15 with

different morphologies and their catalytic performance of DBT HDS,

Appl. Catal. B: Environ. 165, 269-284.

Gao, D., Schweitzer, C., Hwang, H.T., and Varma, A. 2014. Conversion of

guaiacol on noble metal catalysts: reaction performance and

deactivation studies, Ind. Eng. Chem. Res. 53, 18658-18667.

Garcia-Perez, M., Adams, T.T., Goodrum, J.W., Das, K., and Geller, D.P. 2010.

DSC studies to evaluate the impact of bio-oil on cold flow properties

and oxidation stability of bio-diesel, Bioresour. Technol. 101, 6219-

6224.

Garcia-Perez, M., Adams, T.T., Goodrum, J.W., Geller, D.P., and Das, K. 2007.

Production and fuel properties of pine chip bio-oil/biodiesel blends,

Energ. Fuel. 21, 2363-2372.

Garcia-Perez, M., Chaala, A., Pakdel, H., Kretschmer, D., Rodrigue, D., and

Roy, C. 2006. Evaluation of the influence of stainless steel and copper

on the aging process of bio-oil, Energ. Fuel. 20, 786-795.

Gayubo, A.G., Aguayo, A.T., Atutxa, A., Aguado, R., and Bilbao, J. 2004.

Transformation of oxygenate components of biomass pyrolysis oil on a

Page 245: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

224

HZSM-5 zeolite. I. Alcohols and phenols, Ind. Eng. Chem. Res. 43,

2610-2618.

Gayubo, A.G., Aguayo, A.T., Atutxa, A., Aguado, R., Olazar, M., and Bilbao, J.

2004. Transformation of oxygenate components of biomass pyrolysis

oil on a HZSM-5 zeolite. II. Aldehydes, ketones, and acids, Ind. Eng.

Chem. Res. 43, 2619-2626.

Ghampson, I.T., Sepúlveda, C., Garcia, R., Fierro, J.G., Escalona, N., and

DeSisto, W.J. 2012. Comparison of alumina-and SBA-15-supported

molybdenum nitride catalysts for hydrodeoxygenation of guaiacol,

Appl. Catal. A: Gen. 435, 51-60.

Gielen, D., Fujino, J., Hashimoto, S., and Moriguchi, Y. 2002. Biomass

strategies for climate policies?, Clim. Policy 2, 319-333.

González-Borja, M.Á ., and Resasco, D.E. 2011. Anisole and guaiacol

hydrodeoxygenation over monolithic Pt–Sn catalysts, Energ. Fuel. 25,

4155-4162.

Gonzalez-Salazar, M.A., Venturini, M., Poganietz, W.-R., Finkenrath, M.,

Kirsten, T., Acevedo, H., and Spina, P.R. 2016. Development of a

technology roadmap for bioenergy exploitation including biofuels,

waste-to-energy and power generation & CHP, Appl. Energ. 180, 338-

352.

Goodrum, J.W., and Geller, D.P. 2005. Influence of fatty acid methyl esters from

hydroxylated vegetable oils on diesel fuel lubricity, Bioresour. Technol.

96, 851-855.

Goyal, H., Seal, D., and Saxena, R. 2008. Bio-fuels from thermochemical

conversion of renewable resources: a review, Renew. Sus. Energ. Rev.

12, 504-517.

Graham, R. 2001. Comparing the benefits and impacts of hybrid electric vehicle

options, Electr. Pow. Res. Ins. Palo Alto, CA, Report, 1000349.

Grahn, M., Azar, C., Lindgren, K., Berndes, G., and Gielen, D. 2007. Biomass

for heat or as transportation fuel? A comparison between two model-

based studies, Biomass Bioenerg. 31, 747-758.

Griffin, M.B., Ferguson, G.A., Ruddy, D.A., Biddy, M.J., Beckham, G.T., and

Schaidle, J.A. 2016. Role of the support and reaction conditions on the

vapor-phase deoxygenation of m-Cresol over Pt/C and Pt/TiO2 catalysts,

ACS Catal. 6, 2715-2727.

Günter, M.M., Ressler, T., Jentoft, R.E., and Bems, B. 2001. Redox behavior of

copper oxide/zinc oxide catalysts in the steam reforming of methanol

studied by in situ X-ray diffraction and absorption spectroscopy, J.

Catal. 203, 133-149.

Gutierrez, A., Kaila, R., Honkela, M., Slioor, R., and Krause, A. 2009.

Hydrodeoxygenation of guaiacol on noble metal catalysts, Catal. Today

147, 239-246.

Hong, D.-Y., Miller, S.J., Agrawal, P.K., and Jones, C.W. 2010.

Page 246: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

225

Hydrodeoxygenation and coupling of aqueous phenolics over

bifunctional zeolite-supported metal catalysts, Chem. Commun. 46,

1038-1040.

Huang, X., Korányi, T.I., Boot, M.D., and Hensen, E.J. 2014. Catalytic

depolymerization of lignin in supercritical ethanol, ChemSusChem 7,

2276-2288.

Huber, G.W., Iborra, S., and Corma, A. 2006. Synthesis of transportation fuels

from biomass: chemistry, catalysts, and engineering, Chem. Rev. 106,

4044-4098.

Hussain, Z., Khan, K.M., Khan, A., Ullah, S., Karim, A., and Perveen, S. 2013.

The conversion of biomass into liquid hydrocarbon fuel by two step

pyrolysis using cement as catalyst, J. Anal. Appl. Pyrol. 101, 90-95.

Huynh, T.M., Armbruster, U., Pohl, M.M., Schneider, M., Radnik, J., Hoang,

D.L., Phan, B.M.Q., Nguyen, D.A., and Martin, A. 2014.

Hydrodeoxygenation of phenol as a model compound for bio‐oil on

non‐noble bimetallic nickel‐based catalysts, ChemCatChem 6, 1940-

1951.

Ikura, M., Stanciulescu, M., and Hogan, E. 2003. Emulsification of pyrolysis

derived bio-oil in diesel fuel, Biomass Bioenerg. 24, 221-232.

Iliopoulou, E.F., Stefanidis, S., Kalogiannis, K., Delimitis, A., Lappas, A., and

Triantafyllidis, K. 2012. Catalytic upgrading of biomass pyrolysis

vapors using transition metal-modified ZSM-5 zeolite, Appl. Catal. B:

Environ. 127, 281-290.

Iwamoto, Y., Noma, K., Nakayama, O., Yamauchi, T., and Ando, H. 1997.

Development of gasoline direct injection engine, In SAE technical

paper.

Jongerius, A.L., Jastrzebski, R., Bruijnincx, P.C., and Weckhuysen, B.M. 2012.

CoMo sulfide-catalyzed hydrodeoxygenation of lignin model

compounds: An extended reaction network for the conversion of

monomeric and dimeric substrates, J. Catal. 285, 315-323.

Joshi, M.D.H., and Patel, T.M. 2012. Parametric optimization of single cylinder

diesel engine for pyrolysis oil & diesel blend for mechanical efficiency

using taguchi method, Int. J. Eng. Res. Technol. 6, 83-88.

Joshi, N., and Lawal, A. 2012. Hydrodeoxygenation of pyrolysis oil in a

microreactor, Chem. Eng. Sci. 74, 1-8.

Kalbande, S., More, G., and Nadre, R. 2008. Biodiesel production from non-

edible oils of jatropha and karanj for utilization in electrical generator,

Bioenerg. Res. 1, 170-178.

Kan, T., Strezov, V., and Evans, T.J. 2016. Lignocellulosic biomass pyrolysis:

A review of product properties and effects of pyrolysis parameters,

Renew. Sus. Energ. Rev. 57, 1126-1140.

Kim, T.-S., Kim, J.-Y., Kim, K.-H., Lee, S., Choi, D., Choi, I.-G., and Choi, J.W.

Page 247: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

226

2012. The effect of storage duration on bio-oil properties, J. Anal. Appl.

Pyrol. 95, 118-125.

Kim, T.-S., Oh, S., Kim, J.-Y., Choi, I.-G., and Choi, J.W. 2014. Study on the

hydrodeoxygenative upgrading of crude bio-oil produced from woody

biomass by fast pyrolysis, Energ. 68, 437-443.

Kratzer, M.B. 1953. Regeneration of hydrocarbon synthesis catalyst by

treatment with ethanol, In Google Patents.

Kruk, M., Jaroniec, M., Ko, C.H., and Ryoo, R. 2000. Characterization of the

porous structure of SBA-15, Chem. Mater. 12, 1961-1968.

Krutof, A., and Hawboldt, K. 2016. Blends of pyrolysis oil, petroleum, and

other bio-based fuels: a review, Renew. Sus. Energ. Rev. 59, 406-419.

Kubicka, D., Simácek, P., Kolena, J., Lederer, J., and Sebor, G. 2007. Catalytic

transformations of vegetable oils into hydrocarbon, In 43rd Int. Petrol.

Con. Bratislava, Slovakia.

Kunitake, M., Kuraya, E., Kato, D., Niwa, O., and Nishimi, T. 2016.

Electrochemistry in bicontinuous microemulsions based on control of

dynamic solution structures on electrode surfaces, Curr. Opin. Colloid

In. 25, 13-26.

Lee, C.R., Yoon, J.S., Suh, Y.-W., Choi, J.-W., Ha, J.-M., Suh, D.J., and Park,

Y.-K. 2012. Catalytic roles of metals and supports on

hydrodeoxygenation of lignin monomer guaiacol, Catal. Commun. 17,

54-58.

Lee, J. 1997. Biological conversion of lignocellulosic biomass to ethanol, J.

Biotechnol. 56, 1-24.

Leech, J. 1997. Running a dual fuel engine on pyrolysis oil, In Proc. Int. Conf.

Gas. Pyrol. Biomass Stuttgart, 9-11.

Lestari, S., Mäki-Arvela, P., Eränen, K., Beltramini, J., Lu, G.M., and Murzin,

D.Y. 2010. Diesel-like hydrocarbons from catalytic deoxygenation of

stearic acid over supported Pd nanoparticles on SBA-15 catalysts, Catal.

Lett. 134, 250-257.

Lewandowski, I., Scurlock, J.M., Lindvall, E., and Christou, M. 2003. The

development and current status of perennial rhizomatous grasses as

energy crops in the US and Europe, Biomass Bioenerg. 25, 335-361.

Li, K., Wang, R., and Chen, J. 2011. Hydrodeoxygenation of anisole over silica-

supported Ni2P, MoP, and NiMoP catalysts, Energ. Fuel. 25, 854-863.

Liu, H., Li, Y., Wu, H., Takayama, H., Miyake, T., and He, D. 2012. Effects of

β-cyclodextrin modification on properties of Ni/SBA-15 and its

catalytic performance in carbon dioxide reforming of methane, Catal.

Commun. 28, 168-173.

Liu, Y.-M., Cao, Y., Yan, S.-R., Dai, W.-L., and Fan, K.-N. 2003. Highly

effective oxidative dehydrogenation of propane over vanadia supported

on mesoporous SBA-15 silica, Catal. Lett. 88, 61-67.

Long, J., Xu, Y., Wang, T., Yuan, Z., Shu, R., Zhang, Q., and Ma, L. 2015.

Page 248: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

227

Efficient base-catalyzed decomposition and in situ hydrogenolysis

process for lignin depolymerization and char elimination, Appl. Energ.

141, 70-79.

Lónyi, F., and Valyon, J. 2001. On the interpretation of the NH3-TPD patterns

of H-ZSM-5 and H-mordenite, Micropor. Mesopor. Mater. 47, 293-301.

Lu, Q., Tang, Z., Zhang, Y., and Zhu, X.-f. 2010. Catalytic upgrading of biomass

fast pyrolysis vapors with Pd/SBA-15 catalysts, Ind. Eng. Chem. Res.

49, 2573-2580.

Luo, J., Arroyo-Ramírez, L., Wei, J., Yun, H., Murray, C.B., and Gorte, R.J.

2015. Comparison of HMF hydrodeoxygenation over different metal

catalysts in a continuous flow reactor, Appl. Catal. A: Gen. 508, 86-93.

Luo, J., Arroyo‐Ramírez, L., Gorte, R.J., Tzoulaki, D., and Vlachos, D.G. 2015.

Hydrodeoxygenation of HMF over Pt/C in a continuous flow reactor,

AIChE 61, 590-597.

Luo, J., Monai, M., Wang, C., Lee, J.D., Duchoň, T., Dvořák, F., Matolín, V.,

Murray, C.B., Fornasiero, P., and Gorte, R.J. 2017. Unraveling the

surface state and composition of highly selective nanocrystalline Ni–

Cu alloy catalysts for hydrodeoxygenation of HMF, Catal. Sci. Technol.

7, 1735-1743.

Lysen, E. 2000. Global Restrictions of biomass availability for import to the

Netherlands, Novem. Utrecht.

Madsen, A.T., Christensen, C.H., Fehrmann, R., and Riisager, A. 2011.

Hydrodeoxygenation of waste fat for diesel production: Study on model

feed with Pt/alumina catalyst, Fuel 90, 3433-3438.

Mane, R.B., and Rode, C.V. 2012. Simultaneous glycerol dehydration and in

situ hydrogenolysis over Cu–Al oxide under an inert atmosphere, Green

Chem. 14, 2780-2789.

Mante, O.D., and Agblevor, F.A. 2012. Storage stability of biocrude oils from

fast pyrolysis of poultry litter, Waste Manage. 32, 67-76.

Marketing, G. 2007. Diesel fuels technical review.

Martin, G.A., and Dalmon, J.A. 1982. Benzene hydrogenation over nickel

catalysts at low and high temperatures: Structure-sensitivity and copper

alloying effects, J. Catal. 75, 233-242.

Mazaheri, H., Lee, K.T., Bhatia, S., and Mohamed, A.R. 2010. Sub/supercritical

liquefaction of oil palm fruit press fiber for the production of bio-oil:

effect of solvents, Bioresour. Technol. 101, 7641-7647.

McKendry, P. 2002a. Energy production from biomass (part 1): overview of

biomass, Bioresour. Technol. 83, 37-46.

McKendry, P. 2002b. Energy production from biomass (part 2): conversion

technologies, Bioresour. Technol. 83, 47-54.

McKendry, P. 2002c. Energy production from biomass (part 3): gasification

technologies, Bioresour. Technol. 83, 55-63.

Meng, J., Moore, A., Tilotta, D., Kelley, S., and Park, S. 2014. Toward

Page 249: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

228

understanding of bio-oil aging: Accelerated aging of bio-oil fractions,

ACS Sus. Chem. Eng. 2, 2011-2018.

Miao, S., and Shanks, B.H. 2009. Esterification of biomass pyrolysis model

acids over sulfonic acid-functionalized mesoporous silicas, Appl. Catal.

A: Gen. 359, 113-120.

Mohan, D., Pittman, C.U., and Steele, P.H. 2006. Pyrolysis of wood/biomass

for bio-oil: a critical review, Energ. Fuel. 20, 848-889.

Mohan, V., Pramod, C., Suresh, M., Reddy, K.H.P., Raju, B.D., and Rao, K.R.

2012. Advantage of Ni/SBA-15 catalyst over Ni/MgO catalyst in terms

of catalyst stability due to release of water during nitrobenzene

hydrogenation to aniline, Catal. Commun. 18, 89-92.

Mortensen, P.M., Grunwaldt, J.-D., Jensen, P.A., and Jensen, A.D. 2013.

Screening of catalysts for hydrodeoxygenation of phenol as a model

compound for bio-oil, ACS Catal. 3, 1774-1785.

Mortensen, P.M., Grunwaldt, J.-D., Jensen, P.A., Knudsen, K., and Jensen, A.D.

2011. A review of catalytic upgrading of bio-oil to engine fuels, Appl.

Catal. A: Gen. 407, 1-19.

Moulijn, J.A., Van Diepen, A., and Kapteijn, F. 2001. Catalyst deactivation: is

it predictable?: What to do?, Appl. Catal. A: Gen. 212, 3-16.

Murata, K., Liu, Y., Inaba, M., and Takahara, I. 2010a. Hydrocracking of

biomass-derived materials into alkanes in the presence of platinum-

based catalyst and hydrogen, Catal. Lett. 140, 8-13.

Murata, K., Liu, Y., Inaba, M., and Takahara, I. 2010b. Production of synthetic

diesel by hydrotreatment of jatropha oils using Pt−Re/H-ZSM-5

catalyst, Energ. Fuel. 24, 2404-2409.

Nakagawa, Y., Ishikawa, M., Tamura, M., and Tomishige, K. 2014. Selective

production of cyclohexanol and methanol from guaiacol over Ru

catalyst combined with MgO, Green Chem. 16, 2197-2203.

Narani, A., Chowdari, R.K., Cannilla, C., Bonura, G., Frusteri, F., Heeres, H.J.,

and Barta, K. 2015. Efficient catalytic hydrotreatment of Kraft lignin to

alkylphenolics using supported NiW and NiMo catalysts in

supercritical methanol, Green Chem. 17, 5046-5057.

Nava, R., Pawelec, B., Castaño, P., Á lvarez-Galván, M., Loricera, C., and Fierro,

J. 2009. Upgrading of bio-liquids on different mesoporous silica-

supported CoMo catalysts, Appl. Catal. B: Environ. 92, 154-167.

Ng, K.S., and Sadhukhan, J. 2011. Techno-economic performance analysis of

bio-oil based Fischer-Tropsch and CHP synthesis platform, Biomass

Bioenerg. 35, 3218-3234.

Nimmanwudipong, T., Aydin, C., Lu, J., Runnebaum, R.C., Brodwater, K.C.,

Browning, N.D., Block, D.E., and Gates, B.C. 2012. Selective

hydrodeoxygenation of guaiacol catalyzed by platinum supported on

magnesium oxide, Catal. Lett. 142, 1190-1196.

Nimmanwudipong, T., Runnebaum, R.C., Block, D.E., and Gates, B.C. 2011.

Page 250: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

229

Catalytic conversion of guaiacol catalyzed by platinum supported on

alumina: Reaction network including hydrodeoxygenation reactions,

Energ. Fuel. 25, 3417-3427.

Nimmanwudipong, T., Runnebaum, R.C., Brodwater, K., Heelan, J., Block,

D.E., and Gates, B.C. 2014. Design of a high-pressure flow-reactor

system for catalytic hydrodeoxygenation: guaiacol conversion

catalyzed by platinum supported on MgO, Energ. Fuel. 28, 1090-1096.

Oasmaa, A., and Czernik, S. 1999. Fuel oil quality of biomass pyrolysis oils

state of the art for the end users, Energ. Fuel. 13, 914-921.

Oasmaa, A., Elliott, D.C., and Korhonen, J. 2010. Acidity of biomass fast

pyrolysis bio-oils, Energ. Fuel. 24, 6548-6554.

Oasmaa, A., and Kuoppala, E. 2003. Fast pyrolysis of forestry residue. 3.

Storage stability of liquid fuel, Energ. Fuel. 17, 1075-1084.

Oasmaa, A., Kuoppala, E., Ardiyanti, A., Venderbosch, R., and Heeres, H. 2010.

Characterization of hydrotreated fast pyrolysis liquids, Energ. Fuel. 24,

5264-5272.

Oasmaa, A., Kuoppala, E., Selin, J.-F., Gust, S., and Solantausta, Y. 2004. Fast

pyrolysis of forestry residue and pine. 4. Improvement of the product

quality by solvent addition, Energ. Fuel. 18, 1578-1583.

Oasmaa, A., Kytö, M., and Sipilä, K. 2001. Pyrolysis oil combustion tests in an

industrial boiler, Prog. Therm. Biomass Conver. 1468-1481.

Oasmaa, A., Peacocke, C., Gust, S., Meier, D., and McLellan, R. 2005. Norms

and standards for pyrolysis liquids. End-user requirements and

specifications, Energ. Fuel. 19, 2155-2163.

Ochoa-Hernández, C., Yang, Y., Pizarro, P., Víctor, A., Coronado, J.M., and

Serrano, D.P. 2013. Hydrocarbons production through hydrotreating of

methyl esters over Ni and Co supported on SBA-15 and Al-SBA-15,

Catal. Today 210, 81-88.

Ohta, H., Kobayashi, H., Hara, K., and Fukuoka, A. 2011. Hydrodeoxygenation

of phenols as lignin models under acid-free conditions with carbon-

supported platinum catalysts, Chem. Commun. 47, 12209-12211.

Ou, X., Zhang, X., and Chang, S. 2010. Scenario analysis on alternative

fuel/vehicle for China’s future road transport: Life-cycle energy

demand and GHG emissions, Energ. Policy 38, 3943-3956.

Overend, R.P., and Chornet, E. 1997. Making a business from biomass in energy,

environment, chemicals, fibers, and materials. In Biomass Conf. Am.

1997: Montreal, Quebec). Pergamon.

Parsell, T.H., Owen, B.C., Klein, I., Jarrell, T.M., Marcum, C.L., Haupert, L.J.,

Amundson, L.M., Kenttämaa, H.I., Ribeiro, F., and Miller, J.T. 2013.

Cleavage and hydrodeoxygenation (HDO) of C–O bonds relevant to

lignin conversion using Pd/Zn synergistic catalysis, Chem. Sci. 4, 806-

813.

Patel, M.K.B., Patel, T.M., and Patel, M.S.C. 2012. Parametric Optimization of

Page 251: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

230

Single Cylinder Diesel Engine for Pyrolysis Oil and Diesel Blend for

Specific Fuel Consumption Using Taguchi Method, IOSR J. Mech.

Civil Eng. 6, 83-88.

Pattiya, A., Titiloye, J.O., and Bridgwater, A.V. 2008. Fast pyrolysis of cassava

rhizome in the presence of catalysts, J. Anal. Appl. Pyrol. 81, 72-79.

Prakash, R., Singh, R., and Murugan, S. 2013. Experimental investigation on a

diesel engine fueled with bio-oil derived from waste wood–biodiesel

emulsions, Energ. 55, 610-618.

Pütün, A.E., Apaydın, E., and Pütün, E. 2004. Rice straw as a bio-oil source via

pyrolysis and steam pyrolysis, Energ. 29, 2171-2180.

Qi, G., Dong, P., Wang, H., and Tan, H. 2008. Study on biomass pyrolysis and

emulsions from biomass pyrolysis oils and diesel. In 2008 2nd Int. Conf.

Bio. Bio. Eng. 4735-4737. IEEE.

Qin, Z., Shen, B., Yu, Z., Deng, F., Zhao, L., Zhou, S., Yuan, D., Gao, X., Wang,

B., and Zhao, H. 2013. A defect-based strategy for the preparation of

mesoporous zeolite Y for high-performance catalytic cracking, J. Catal.

298, 102-111.

Rezaei, P.S., Shafaghat, H., and Daud, W.M.A.W. 2014. Production of green

aromatics and olefins by catalytic cracking of oxygenate compounds

derived from biomass pyrolysis: A review, Appl. Catal. A: Gen. 469,

490-511.

Richardson, J.T., Scates, R., and Twigg, M.V. 2003. X-ray diffraction study of

nickel oxide reduction by hydrogen, Appl. Catal. A: Gen. 246, 137-150.

Roberts, E.J., Habas, S.E., Wang, L., Ruddy, D.A., White, E.A., Baddour, F.G.,

Griffin, M.B., Schaidle, J.A., Malmstadt, N., and Brutchey, R.L. 2016.

High-throughput continuous flow synthesis of nickel nanoparticles for

the catalytic hydrodeoxygenation of guaiacol, ACS Sus. Chem. Eng. 5,

632-639.

Ronning, J.J. 1997. The viable environmental car: the right combination of

electrical and combustion energy for transportation, In SAE Technical

Paper

Ruddy, D.A., Schaidle, J.A., Ferrell III, J.R., Wang, J., Moens, L., and Hensley,

J.E. 2014. Recent advances in heterogeneous catalysts for bio-oil

upgrading via “ex situ catalytic fast pyrolysis”: catalyst development

through the study of model compounds, Green Chem. 16, 454-490.

Runnebaum, R.C., Nimmanwudipong, T., Block, D.E., and Gates, B.C. 2012.

Catalytic conversion of compounds representative of lignin-derived

bio-oils: a reaction network for guaiacol, anisole, 4-methylanisole, and

cyclohexanone conversion catalysed by Pt/γ-Al2O3, Catal. Sci. Technol.

2, 113-118.

Ryu, Z., Zheng, J., Wang, M., and Zhang, B. 1999. Characterization of pore size

distributions on carbonaceous adsorbents by DFT, Carbon 37, 1257-

1264.

Page 252: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

231

Saidi, M., Samimi, F., Karimipourfard, D., Nimmanwudipong, T., Gates, B.C.,

and Rahimpour, M.R. 2014. Upgrading of lignin-derived bio-oils by

catalytic hydrodeoxygenation, Energ. Environ. Sci. 7, 103-129.

Samolada, M., Baldauf, W., and Vasalos, I. 1998. Production of a bio-gasoline

by upgrading biomass flash pyrolysis liquids via hydrogen processing

and catalytic cracking, Fuel 77, 1667-1675.

Sazhina, E., Sazhin, S., Heikal, M., Babushok, V., and Johns, R. 2000. A

detailed modelling of the spray ignition process in Diesel engines,

Combust. Sci. Technol. 160, 317-344.

Schutyser, W., Van den Bossche, G., Raaffels, A., Van den Bosch, S., Koelewijn,

S.-F., Renders, T., and Sels, B.F. 2016. Selective conversion of lignin-

derivable 4-alkylguaiacols to 4-alkylcyclohexanols over noble and non-

noble-metal catalysts, ACS Sus. Chem. Eng. 4, 5336-5346.

Selvaraj, M., Shanthi, K., Maheswari, R., and Ramanathan, A. 2014.

Hydrodeoxygenation of guaiacol over MoO3-NiO/mesoporous silicates:

Effect of incorporated heteroatom, Energ. Fuel. 28, 2598-2607.

Şenol, O., Ryymin, E.-M., Viljava, T.-R., and Krause, A. 2007. Reactions of

methyl heptanoate hydrodeoxygenation on sulphided catalysts, J. Mol.

Catal. A: Chem. 268, 1-8.

Şenol, O., Viljava, T.-R., and Krause, A. 2005. Hydrodeoxygenation of methyl

esters on sulphided NiMo/γ-Al2O3 and CoMo/γ-Al2O3 catalysts, Catal.

Today 100, 331-335.

Serrano-Ruiz, J.C., and Dumesic, J.A. 2011. Catalytic routes for the conversion

of biomass into liquid hydrocarbon transportation fuels, Energ. Environ.

Sci. 4, 83-99.

Shaddix, C.R., and Hardesty, D.R. 1999. Combustion properties of biomass

flash pyrolysis oils, USA: Sandia National Laboratory, 71.

Sharma, R.K., and Bakhshi, N.N. 1993. Conversion of non-phenolic fraction of

biomass-derived pyrolysis oil to hydrocarbon fuels over HZSM-5 using

a dual reactor system, Bioresour. Technol. 45, 195-203.

Shemfe, M., Gu, S., and Fidalgo, B. 2017. Techno-economic analysis of biofuel

production via bio-oil zeolite upgrading: An evaluation of two catalyst

regeneration systems, Biomass Bioenerg. 98, 182-193.

Shen, Y. 2015. Carbothermal synthesis of metal-functionalized nanostructures

for energy and environmental applications, J. Mater. Chem. A 3, 13114-

13188.

Sipilä, K., Kuoppala, E., Fagernäs, L., and Oasmaa, A. 1998. Characterization

of biomass-based flash pyrolysis oils, Biomass Bioenerg. 14, 103-113.

Sluiter, A., Hames, B., Ruiz, R., Scarlata, C., Sluiter, J., Templeton, D., and

Crocker, D. 2008. Determination of structural carbohydrates and lignin

in biomass, Laboratory analytical procedure, 1617, 1-16.

Smith, W., and Consultancy, T. 2007. Mapping the development of UK

biorefinery complexes, UK National Non Food Crops Centre (NNFCC).

Page 253: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

232

Snowden-Swan, L.J., Spies, K.A., Lee, G., and Zhu, Y. 2016. Life cycle

greenhouse gas emissions analysis of catalysts for hydrotreating of fast

pyrolysis bio-oil, Biomass Bioenerg. 86, 136-145.

Soriano, F.H., and Mulatero, F. 2011. EU research and innovation (R&I) in

renewable energies: the role of the strategic energy technology plan

(SET-Plan), Energ. Policy 39, 3582-3590.

Spanjers, C.S., Sim, R.S., Sturgis, N.P., Kabius, B., and Rioux, R.M. 2015. In

Situ Spectroscopic Characterization of Ni1–x Zn x/ZnO Catalysts and

Their Selectivity for Acetylene Semihydrogenation in Excess Ethylene,

ACS Catal. 5, 3304-3315.

Suarez, P.A., Moser, B.R., Sharma, B.K., and Erhan, S.Z. 2009. Comparing the

lubricity of biofuels obtained from pyrolysis and alcoholysis of soybean

oil and their blends with petroleum diesel, Fuel 88, 1143-1147.

Sulek, M., Kulczycki, A., and Malysa, A. 2010. Assessment of lubricity of

compositions of fuel oil with biocomponents derived from rape-seed,

Wear 268, 104-108.

Tang, Z., Lu, Q., Zhang, Y., Zhu, X., and Guo, Q. 2009. One step bio-oil

upgrading through hydrotreatment, esterification, and cracking, Ind.

Eng. Chem. Res. 48, 6923-6929.

Tang, Z., Zhang, Y., and Guo, Q. 2010. Catalytic hydrocracking of pyrolytic

lignin to liquid fuel in supercritical ethanol, Ind. Eng. Chem. Res. 49,

2040-2046.

Torres, W., Pansare, S.S., and Goodwin Jr, J.G. 2007. Hot gas removal of tars,

ammonia, and hydrogen sulfide from biomass gasification gas, Catal.

Rev. 49, 407-456.

Torri, C., Reinikainen, M., Lindfors, C., Fabbri, D., Oasmaa, A., and Kuoppala,

E. 2010. Investigation on catalytic pyrolysis of pine sawdust: Catalyst

screening by Py-GC-MIP-AED, J. Anal. Appl. Pyrol. 88, 7-13.

Tucker, S.C. 1999. Solvent density inhomogeneities in supercritical fluids,

Chem. Rev. 99, 391-418.

Van de Beld, B., Holle, E., and Florijn, J. 2013. The use of pyrolysis oil and

pyrolysis oil derived fuels in diesel engines for CHP applications, Appl.

Energ. 102, 190-197.

Varisli, D., Dogu, T., and Dogu, G. 2007. Ethylene and diethyl-ether production

by dehydration reaction of ethanol over different heteropolyacid

catalysts, Chem. Eng. Sci. 62, 5349-5352.

Venderbosch, R., Ardiyanti, A., Wildschut, J., Oasmaa, A., and Heeres, H. 2010.

Stabilization of biomass‐derived pyrolysis oils, J. Chem. Technol.

Biotechnolog. 85, 674-686.

Venkatakrishnan, V.K., Delgass, W.N., Ribeiro, F.H., and Agrawal, R. 2015.

Oxygen removal from intact biomass to produce liquid fuel range

hydrocarbons via fast-hydropyrolysis and vapor-phase catalytic

hydrodeoxygenation, Green Chem. 17, 178-183.

Page 254: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

233

Vitolo, S., Bresci, B., Seggiani, M., and Gallo, M. 2001. Catalytic upgrading of

pyrolytic oils over HZSM-5 zeolite: behaviour of the catalyst when

used in repeated upgrading–regenerating cycles, Fuel 80, 17-26.

Wang, H., Male, J., and Wang, Y. 2013. Recent advances in hydrotreating of

pyrolysis bio-oil and its oxygen-containing model compounds, ACS

Catal. 3, 1047-1070.

Wang, N., Yu, X., Shen, K., Chu, W., and Qian, W. 2013. Synthesis,

characterization and catalytic performance of MgO-coated Ni/SBA-15

catalysts for methane dry reforming to syngas and hydrogen, Int. J.

Hydrogen Energ. 38, 9718-9731.

Wang, W., Liu, P., Zhang, M., Hu, J., and Xing, F. 2012. The pore structure of

phosphoaluminate cement, J. Compos. Mater. 2, 104.

Wang, X., and Rinaldi, R. 2012. Solvent effects on the hydrogenolysis of

diphenyl ether with Raney nickel and their implications for the

conversion of lignin, ChemSusChem 5, 1455-1466.

Wang, Y., De, S., and Yan, N. 2016. Rational control of nano-scale metal-

catalysts for biomass conversion, Chem. Commun. 52, 6210-6224.

Weerachanchai, P., Tangsathitkulchai, C., and Tangsathitkulchai, M. 2009.

Phase behaviors and fuel properties of bio-oil-diesel-alcohol blends,

World Acad. Sci., Eng. Technol. 56, 387-393.

White, L.P., and Plaskett, L.G. 1981. Biomass as fuel (Academic Press Ltd.).

Wildschut, J., Arentz, J., Rasrendra, C., Venderbosch, R., and Heeres, H. 2009.

Catalytic hydrotreatment of fast pyrolysis oil: model studies on reaction

pathways for the carbohydrate fraction, Environ. Prog. Sus. Energ. 28,

450-460.

Wildschut, J., Iqbal, M., Mahfud, F.H., Cabrera, I.M., Venderbosch, R.H., and

Heeres, H.J. 2010. Insights in the hydrotreatment of fast pyrolysis oil

using a ruthenium on carbon catalyst, Energ. Environ. Sci. 3, 962-970.

Wildschut, J., Mahfud, F.H., Venderbosch, R.H., and Heeres, H.J. 2009.

Hydrotreatment of fast pyrolysis oil using heterogeneous noble-metal

catalysts, Ind. Eng. Chem. Res. 48, 10324-10334.

Wildschut, J., Melian-Cabrera, I., and Heeres, H. 2010. Catalyst studies on the

hydrotreatment of fast pyrolysis oil, Appl. Catal. B: Environ. 99, 298-

306.

Williams, P.T., and Horne, P.A. 1994. Characterisation of oils from the fluidised

bed pyrolysis of biomass with zeolite catalyst upgrading, Biomass

Bioenerg. 7, 223-236.

Wu, S.-K., Lai, P.-C., and Lin, Y.-C. 2014. Atmospheric hydrodeoxygenation of

guaiacol over nickel phosphide catalysts: effect of phosphorus

composition, Catal. Lett. 144, 878-889.

Xu, X., Zhang, C., Liu, Y., Zhai, Y., and Zhang, R. 2013. Two-step catalytic

hydrodeoxygenation of fast pyrolysis oil to hydrocarbon liquid fuels,

Chemosphere 93, 652-660.

Page 255: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

234

Xu, X., Zhang, C., Zhai, Y., Liu, Y., Zhang, R., and Tang, X. 2014. Upgrading

of bio-oil using supercritical 1-butanol over a Ru/C heterogeneous

catalyst: Role of the solvent, Energ. Fuel. 28, 4611-4621.

Xu, Y., Wang, Q., Hu, X., Li, C., and Zhu, X. 2010. Characterization of the

lubricity of bio-oil/diesel fuel blends by high frequency reciprocating

test rig, Energy 35, 283-287.

Yakovlev, V., Khromova, S., Sherstyuk, O., Dundich, V., Ermakov, D.Y.,

Novopashina, V., Lebedev, M.Y., Bulavchenko, O., and Parmon, V.

2009. Development of new catalytic systems for upgraded bio-fuels

production from bio-crude-oil and biodiesel, Catal. Today 144, 362-366.

Yan, N., Yuan, Y., Dykeman, R., Kou, Y., and Dyson, P.J. 2010.

Hydrodeoxygenation of Lignin‐Derived Phenols into Alkanes by Using

Nanoparticle Catalysts Combined with Brønsted Acidic Ionic Liquids,

Angew. Chem. Int. Edit. 49, 5549-5553.

Yang, Y., Brammer, J., Samanya, J., Hossain, A., and Hornung, A. 2013.

Investigation into the performance and emissions of a stationary diesel

engine fuelled by sewage sludge intermediate pyrolysis oil and

biodiesel blends, Energy 62, 269-276.

Yu, W., Tang, Y., Mo, L., Chen, P., Lou, H., and Zheng, X. 2011. Bifunctional

Pd/Al-SBA-15 catalyzed one-step hydrogenation–esterification of

furfural and acetic acid: a model reaction for catalytic upgrading of bio-

oil, Catal. Commun. 13, 35-39.

Yu, Y., Li, X., Su, L., Zhang, Y., Wang, Y., and Zhang, H. 2012. The role of

shape selectivity in catalytic fast pyrolysis of lignin with zeolite

catalysts, Appl. Catal. A: Gen. 447, 115-123.

Zacher, A.H., Olarte, M.V., Santosa, D.M., Elliott, D.C., and Jones, S.B. 2014.

A review and perspective of recent bio-oil hydrotreating research,

Green Chem. 16, 491-515.

Zhang, J., Luo, Z., Dang, Q., Wang, J., and Chen, W. 2011. Upgrading of bio-

oil in supercritical ethanol, In Elect. Commun. Control (ICECC), 2011

International Conference on, 4613-4616. IEEE.

Zhang, J., Teo, J., Chen, X., Asakura, H., Tanaka, T., Teramura, K., and Yan, N.

2014. A series of NiM (M= Ru, Rh, and Pd) bimetallic catalysts for

effective lignin hydrogenolysis in water, ACS Catal. 4, 1574-1583.

Zhang, L., Liu, R., Yin, R., and Mei, Y. 2013. Upgrading of bio-oil from

biomass fast pyrolysis in China: A review, Renew. Sus. Energ. Rev. 24,

66-72.

Zhang, M., and Wu, H. 2014. Phase behavior and fuel properties of bio-

oil/glycerol/methanol blends, Energ. Fuel. 28, 4650-4656.

Zhang, W.-B. 2012. Review on analysis of biodiesel with infrared spectroscopy,

Renew. Sus. Energ. Rev. 16, 6048-6058.

Zhang, W., Yu, D., Ji, X., and Huang, H. 2012. Efficient dehydration of bio-

based 2, 3-butanediol to butanone over boric acid modified HZSM-5

Page 256: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

235

zeolites, Green Chem. 14, 3441-3450.

Zhang, X., Wang, T., Ma, L., Zhang, Q., Yu, Y., and Liu, Q. 2013.

Characterization and catalytic properties of Ni and NiCu catalysts

supported on ZrO2–SiO2 for guaiacol hydrodeoxygenation, Catal.

Commun. 33, 15-19.

Zhang, X., Watson, E.J., Dullaghan, C.A., Gorun, S.M., and Sweigart, D.A.

1999. Activation of a Carbon–Oxygen Bond of Benzofuran by

Precoordination of Manganese to the Carbocyclic Ring: A Model for

Hydrodeoxygenation, Angew. Chem. Int. Ed. 38, 2206-2208.

Zhao, B., Guo, Q., and Fu, Y. 2014. Electrocatalytic Hydrogenation of Lignin-

Derived Phenol into Alkanes by Using Platinum Supported on Graphite,

Electrochemistry 82, 954-959.

Zhao, C., He, J., Lemonidou, A.A., Li, X., and Lercher, J.A. 2011. Aqueous-

phase hydrodeoxygenation of bio-derived phenols to cycloalkanes, J.

Catal. 280, 8-16.

Zhao, C., Kou, Y., Lemonidou, A.A., Li, X., and Lercher, J.A. 2009. Highly

Selective Catalytic Conversion of Phenolic Bio‐Oil to Alkanes, Angew.

Chem. 121, 4047-4050.

Zhao, C., and Lercher, J.A. 2012. Selective Hydrodeoxygenation of Lignin‐Derived Phenolic Monomers and Dimers to Cycloalkanes on Pd/C and

HZSM‐5 Catalysts, ChemCatChem 4, 64-68.

Zhao, D., Feng, J., Huo, Q., Melosh, N., Fredrickson, G.H., Chmelka, B.F., and

Stucky, G.D. 1998. Triblock copolymer syntheses of mesoporous silica

with periodic 50 to 300 angstrom pores, Science 279, 548-552.

Zhao, D., Huo, Q., Feng, J., Chmelka, B.F., and Stucky, G.D. 1998. Nonionic

triblock and star diblock copolymer and oligomeric surfactant

syntheses of highly ordered, hydrothermally stable, mesoporous silica

structures, J. Am. Chem. Soc. 120, 6024-6036.

Zhao, H., Li, D., Bui, P., and Oyama, S. 2011. Hydrodeoxygenation of guaiacol

as model compound for pyrolysis oil on transition metal phosphide

hydroprocessing catalysts, Appl. Catal. A: Gen. 391, 305-310.

Zhou, M., Ye, J., Liu, P., Xu, J., and Jiang, J. 2017. Water-assisted selective

hydrodeoxygenation of guaiacol to cyclohexanol over supported Ni and

Co bimetallic catalysts, ACS Sus. Chem. Eng. 5, 8824-8835.

Zhu, X., Lobban, L.L., Mallinson, R.G., and Resasco, D.E. 2011. Bifunctional

transalkylation and hydrodeoxygenation of anisole over a Pt/HBeta

catalyst, J. Catal. 281, 21-29.

Page 257: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

236

Page 258: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

237

초 록

니켈 촉매를 이용한 바이오오일의 수첨탈산소 개질 및

활용 가능성 평가

오신영

환경재료과학전공

산림과학부

서울대학교 대학원

본 연구에서는 다양한 니켈 촉매를 이용하여 급속열분해

공정에서 생산된 바이오오일의 수첨탈산소 개질 공정을 수행

하였으며, 촉매 종류에 따른 특성 및 공정 조건에 따라 개질된

바이오연료의 물성, 안정성, 공정 효율을 비교분석 하였고 더

나아가 산업분야로의 발전을 위해 석유와의 혼합 가능성을

평가하고자 하였다.

첫 번째 장에서는 용매가 수첨탈산소 개질 공정에 미치는

영향을 확인하기 위해 급속열분해 공정에서 생성된 바이오오일을

아세톤, 에탄올, 에테르 등 다양한 용매와 혼합하여 수첨탈산소

공정을 수행하였다. 이후 생성된 바이오연료의 안정성을 평가하기

위하여 12주 간의 저장기간 동안의 노화특성을 관찰, 바이오오일과

비교 하였다. 수첨탈산소 개질 공정을 통해 바이오오일은 가스, 탄,

light oil과 개질된 바이오연료인 heavy oil로 변환되었다.

수첨탈산소 개질 공정에 사용한 용매의 종류에 따라 생성물의 수율

Page 259: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

238

및 물성에 차이가 나타났다. 특히 점도의 경우, 29.6 cSt를 나타낸

바이오 오일과 비교하여 크게 감소하였으며, 에탄올을 사용하여

개질 공정을 수행한 결과 2.2-4.4 cSt로 감소하였다. 극성 양자성

용매인 에탄올의 경우 수첨탈산소 개질 공정 중 조용매의 역할을

하여 바이오오일 내에 존재하는 유기산을 에스테르 화합물로

변환시키는 것으로 관찰되었다. 12주 동안의 저장기간 동안 heavy

oil의 물성은 초기 상태를 유지하였으며, 노화현상이 관찰된

바이오오일과 비교하여 수첨탈산소 공정을 통해 안정성이 확연히

향상된 것으로 나타났다.

두 번째 장에서는 귀금속 촉매의 경제성 문제를 해결하기 위해

전이금속인 니켈 촉매를 선정하여, 메조세공을 가진 다공성

담지체(SBA-15, Al-SBA-15)가 미치는 영향을 조사하였다.

300 °C, 수소분압 3MPa 하에서 바이오 오일의 수첨탈산소 개질

반응을 수행한 결과, heavy oil의 수율은 45.8-48.1 wt%로 촉매

종류에 따라 큰 차이를 나타내지 않았지만, 높은 비표면적 (508.7-

537.7 m2/g)과 넓은 부피의 기공(0.46-0.65 cm3/g)을 가진

메소세공 다공성 담지체에 담지한 니켈 촉매(Ni/SBA-15, Ni/Al-

SBA-15)의 경우 활성탄에 담지된 니켈 촉매(Ni/C)를 이용한

경우(8.5 wt%)보다 높은 탄 수율(16.3-18.6 wt%)을 나타내었다.

특히, Ni/SBA-15 촉매를 사용한 경우, 발열량(25.4 MJ/kg)과

에너지효율(62.8%), 탈산소정도(54.9%)가 가장 뛰어난 것으로

조사되었다.

따라서 Ni/SBA-15을 기본 촉매로 하여 이중 전이금속

촉매(NiCu/SBA-15, NiMn/SBA-15, NiZn/ SBA-15)를 제조하여

수첨탈산소 반응성을 평가하였다. 단일 전이금속 촉매(Ni/SBA-

Page 260: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

239

15)와 비교하여 이중 전이금속 촉매는 heavy oil의 수율을 40.4-

60.6 wt%로 증가 시켰으며, light oil은 3.6-11.7 wt%로 감소시키고

가스 수율은 16.3-43.7 wt%로 증가시켰다. 이중 전이금속 촉매를

사용한 경우, Ni/SBA-15 촉매를 사용한 경우와 비교하여

전반적으로 heavy oil의 물성이 향상되었다. 특히 NiMn/SBA-15를

사용한 경우 높은 온도에서 28.4 MJ/kg으로 heavy oil의 발열량이

향상되었으며, 공정 효율(90.5%) 또한 높은 것으로 조사되었다.

마지막으로, 회분식 반응기에 사용하였던 촉매 중 일부를

선정하여 연속식 반응기에 적용하여 촉매의 반응성을 시험하였다.

300 °C, 수소분압 하에서 바이오오일의 수첨탈산소 공정을 수행한

결과, 가스, light oil, heavy oil과 탄이 생성되었으며, heavy oil

(27.9-42.3 wt%)과 char (2.3-11.0 wt%)는 회분식 반응기를

이용하였을 때 보다 적은 수율을 나타내었다. 바이오오일의 물성

역시 눈에 띄게 개선되었으며, 특히 탈산소반응이 효과적으로

일어나 발열량이 32.8-38.0 MJ/kg까지 향상되었다. 이는 바이오

오일 (15.7 MJ/kg) 및 동일 촉매를 사용한 회분식 공정에서 획득한

heavy oil (24.4-34.5 MJ/kg)보다도 높은 수치이다. 추가적으로,

heavy oil의 활용 가능성을 평가하기 위해, 에멀전화를 통해 시중에

판매되는 두 종류의 석유 제품(경유 및 휘발유)과의 혼합특성을

관찰하였다. 바이오오일은 24시간 이상의 반응 시간 동안 매우 적은

양의 에멀전을 생성하였으나 heavy oil은 반응 조건에 따라 6-

24시간 이내에 안정한 에멀전을 생성하였다. 생성된 에멀전은 7일

간 안정한 상태를 유지하여 개질된 오일의 경우 석유 제품과

혼합하여 사용 가능함을 보였다.

Page 261: Disclaimer - Seoul National Universitys-space.snu.ac.kr/bitstream/10371/143156/1/Improvement of... · 2019-11-14 · catalysts, via demethoxylation, hydrogenation, dehydration, as

240

본 연구에서는 바이오매스의 열화학적 변환 공정인 급속열분해

공정에서 생산되는 바이오오일의 바이오연료로의 활용을 위해

다양한 특성의 촉매 하에서 수첨탈산소 개질 공정을 수행하였고

물성 및 안정성이 향상된 바이오연료를 획득하였다. 수첨탈산소

개질 공정에 사용하기 위해 제조한 촉매의 표면 및 구조적 특성과

획득한 바이오연료의 물리화학적 물성 분석을 통해 촉매의 특성이

바이오오일의 물성 및 안정성 향상에 미치는 영향을 이해할 수

있었다. 더 나아가 현재 시중에 유통되는 석유계 연료와의 혼합

테스트를 통해 수첨탈산소 개질 공정을 통해 생산된 바이오연료의

잠재적 활용 가능성을 확인하였다.

키워드 : 급속열분해, 수첨탈산소, 바이오오일 니켈촉매, 개질공정,

연속식 반응기

학 번 : 2014-30384