+ All Categories
Home > Documents > Life Cycle Assessment of Ethanol produced from ...

Life Cycle Assessment of Ethanol produced from ...

Date post: 26-Jan-2022
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
212
Life Cycle Assessment of Ethanol produced from Lignocellulosic Biomass: Techno-economic and Environmental Evaluation by Poritosh Roy A Thesis presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy in Engineering Guelph, Ontario, Canada © Poritosh Roy, August, 2014
Transcript

i

Life Cycle Assessment of Ethanol produced from Lignocellulosic Biomass:

Techno-economic and Environmental Evaluation

by

Poritosh Roy

A Thesis

presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Doctor of Philosophy

in

Engineering

Guelph, Ontario, Canada

© Poritosh Roy, August, 2014

ii

ABSTRACT

LIFE CYCLE ASSESSMENT OF ETHANOL PRODUCED FROM LIGNOCELLULOSIC

BIOMASS: TECHNO-ECONOMIC AND ENVIRONMENTAL EVALUATION

Poritosh Roy Advisor:

University of Guelph, 2014 Dr. Animesh Dutta

The life cycle (LC) of ethanol derived from lignocellulosic biomass (hereafter referred to

biomass: wheat straw, sawdust & miscanthus) by both enzymatic hydrolysis and thermochemical

[gasification-biosynthesis pathway; torrefied/non-torrefied, with/without chemical looping

gasification (CLG)] conversion processes has been evaluated, considering various scenarios. A

follow-up study has also been conducted to identify the potential locations for ethanol industries

in Ontario. Life cycle assessment (LCA) methodologies have been used to evaluate the LC of

ethanol to determine if environmentally preferable and economically viable ethanol can be

produced in Ontario, Canada. A novel continuous stirred tank bioreactor has also been developed

(consisting of an innovative gas supply and an effluent extraction process) for syngas

fermentation into ethanol.

The net energy consumption, GHG emissions and production cost of ethanol were found to

be dependent on ethanol yield, feedstock cost, processing plant capacity and assumptions. This

study revealed that environmental benefit can be gained from biomasses, the economic viability

and biomass logistics of agri- and forest residues remain doubtful, unless a nominal subsidy (for

example FiT) is provided. The CLG process seems to be useful to reduce net energy

consumption and GHG emissions for both torrefied and non-torrefied miscanthus. Consequently,

iii

miscanthus has emerged as a promising feedstock for ethanol industries (both enzymatic

hydrolysis and biosynthesis) even if it is grown on marginal land in Ontario, avoids any sort of

competition with food crops for higher quality land, avoids the food vs fuel debate, and improves

farm income and the rural economy. Eastern Ontario has emerged as the best option for

miscanthus based ethanol industry.

This study also revealed that syngas can be fermented with Clostridium Ljungdahlii into

ethanol by using the developed bioreactor. It is worthy to note that careful consideration has to

be placed on the land use changes, soil quality and their rebound effects if biomass, especially

agri-residues are to be put to use in the ethanol industry. The information generated in this study

has emerged to be novel and may help the stakeholders in their decision making processes, help

meeting the ethanol demand, and achieving GHG emissions target of Canada.

iv

This thesis is dedicated to my loved Parents (Late Sahadeb Roy & Late Bimola Roy), who are

always remembered and in the center of all kind of inspiration in the life of the author.

v

Acknowledgements

The author wishes to express his deepest sense of gratitude to his advisor Dr. Animesh

Dutta, School of Engineering, University of Guelph, Ontario, Canada for the institutional support

and scholastic supervision, constructive criticism and constant encouragement during the entire

period of this study.

He extend heartfelt gratitude to Dr. Bill Deen, Department of Plant Agricultural, Dr.

Brajesh Dubey and Dr. Shohel Mahmud, School of Engineering, University University of

Guelph, not only for serving in the Advisory Committee and extending their valuable time to

review this manuscript, but also for their thoughtful remarks, useful suggestions and constructive

criticism during this study. He also likes to thank Prof. Amar Mohanty and Dr. Fantahun M.

Defersha for their service in the qualifying examination committee and constructive suggestions.

The author is also thankful to Dr. Douglas M. Joy, Graduate Coordinator for his fruitful

suggestion during the qualifying examination. The author also likes to thank Dr. Sheng Chang

and Mr. Richard Chen for their support in preparing the membrane separator used in this study.

The author wishes to extend his appreciation to Mr. Michael Speagle for his ever-ready-to-

help in the laboratory activities and experimental setups. He also likes to thank Mrs. Carly

Fennell and Mrs. Joanne Ryks, Mr. Ryan Smith, and Mr. John Whiteside for extending their

helping hand, especially for the syngas fermentation study and computer related issues,

respectively. He would like to thank all the members of Dr. Dutta’s research team: Dr. Bimal

Acharya, Mr. Harpreet Kambo, Mr. Stefan Goupal, Mr. Jamie Minarat, Mr. MD Tushar, Dr.

Mathias Leon and Mr. Subhash Paul for their generous help and friendly companion during this

study. He is also indebted to Ontario Graduate Scholarship (OGS) Program for awarding the

prestigious scholarship during this study. The author also acknowledges for the Dean’s

scholarship from the University of Guelph.

The author feels proud to acknowledge his beloved wife Rita Roy who made a fruitful life

for the author during the entire period of this study and for her constant encouragement. He is

also thankful to his daughter Riya Roy for her love, which gives all kind of inspiration. The

author is also grateful to all of his family members and relatives for their moral support, enduring

patience and positive encouragement throughout his study in the University of Guelph, Ontario,

Canada. Finally, he wishes to express his sincere appreciation to those who have contributed

directly or indirectly for the successful completion of this study.

vi

Table of Contents

Cover ................................................................................................................................................ i

Abstract ........................................................................................................................................... ii

Dedication ...................................................................................................................................... iv

Acknowledgements ........................................................................................................................ v

Table of Contents .......................................................................................................................... vi

List of Tables ............................................................................................................................... xii

List of Figures .............................................................................................................................. xiii

Chapter 1: Introduction ...................................................................................................................... 1

1.1. Rationale .......................................................................................................................................... 1

1.2. Objectives ......................................................................................................................................... 5

1.3. Scope and limitation of this research ............................................................................................... 6

1.4 Novelty of the research .................................................................................................................... 7

1.4.1 Bioreactor development............................................................................................................ 7

1.4.2 Chemical looping gasification (CLG) ...................................................................................... 7

1.4.3 Life cycle assessment ............................................................................................................... 8

1.5 Contribution of this research ............................................................................................................ 8

1.6 Publications from this research ...................................................................................................... 10

1.6.1 Publications in peer reviewed journals ................................................................................... 10

1.6.2 Submitted manuscripts ........................................................................................................... 10

1.6.3 Publications: Research presentations ..................................................................................... 10

Chapter 2: Literature Review ........................................................................................................... 11

2.1. Ethanol production via biochemical conversion process (enzymatic hydrolysis).......................... 11

2.1.1. Pretreatment ............................................................................................................................ 11

2.1.2. Fermentation ........................................................................................................................... 12

2.1.3. Distillation and purification .................................................................................................... 13

2.1.4. Waste management ................................................................................................................. 14

2.2. Life cycle assessment (LCA) of ethanol produced by biochemical conversion process ............... 16

2.2.1. LCA of ethanol produced from agri-residues ......................................................................... 16

2.2.2. L CA of ethanol from energy crops, woody biomass and forest residues .............................. 19

2.2.3. Land, water and other approaches in LCA of ethanol ............................................................ 22

vii

2.3. Ethanol production via gasification process................................................................................... 24

2.3.1. Gasification ............................................................................................................................ 24

2.3.2. Gas cleanup ............................................................................................................................ 28

2.3.3. Syngas synthesis into ethanol ................................................................................................. 28

2.4 Life cycle cost analysis (LCCA) ....................................................................................................... 32

2.4.1 Life cycle costing of ethanol produced by biochemical process ............................................ 33

2.4.2 Life cycle costing of ethanol produced by thermochemical process ...................................... 37

Chapter 3: Life Cycle Assessment (LCA) Methodologies .............................................................. 41

3.1 LCA Methodologies ....................................................................................................................... 41

3.1.1 Goal definition and scoping.................................................................................................... 42

3.1.2 Life cycle inventory (LCI) analysis ........................................................................................ 45

3.1.3 Impact assessment .................................................................................................................. 46

3.1.4 Interpretation .......................................................................................................................... 46

3.2 Life cycle cost analysis (LCCA) .................................................................................................... 47

Chapter 4: Life Cycle Assessment of Ethanol produced from Wheat Straw ............................... 49

4.1 Introduction .................................................................................................................................... 49

4.2 Materials and methods ................................................................................................................... 49

4.2.1 System boundary .................................................................................................................... 49

4.2.2 Biochemical conversion process ............................................................................................ 50

4.2.3 Cost analysis ........................................................................................................................... 52

4.2.4 Data collection ........................................................................................................................ 52

4.3 Results and discussion ...................................................................................................................... 53

4.3.1 Energy consumption, CO2 emission and production cost ...................................................... 53

4.3.2 Sensitivity analysis ................................................................................................................. 56

4.4 Conclusion ..................................................................................................................................... 61

Chapter 5: Life cycle assessment of ethanol derived from sawdust .............................................. 62

5.1 Introduction .................................................................................................................................... 62

5.2 Methodology .................................................................................................................................. 62

5.2.1 System boundary and assumptions ......................................................................................... 62

5.2.2 Ethanol production ................................................................................................................. 64

5.2.3 Cost analysis ........................................................................................................................... 65

5.2.4 Data collection ........................................................................................................................ 65

viii

5.3 Results and discussion.................................................................................................................... 66

5.3.1 Net energy consumption and CO2 emission ........................................................................... 66

5.3.2 Production cost ....................................................................................................................... 68

5.3.3 Sensitivity analysis ................................................................................................................. 70

5.4 Conclusion ..................................................................................................................................... 73

Chapter 6: Evaluation of the Life Cycle of Ethanol derived from Miscanthus in Ontario......... 74

6.1 Introduction .................................................................................................................................... 74

6.2 Methodology .................................................................................................................................. 75

6.2.1 Study area, system boundary and assumptions ...................................................................... 75

6.2.2 Miscanthus cultivation ........................................................................................................... 78

6.2.3 Transportation ........................................................................................................................ 80

6.2.4 Ethanol production ................................................................................................................. 80

6.2.5 Cost analysis ........................................................................................................................... 82

6.3 Results and discussion ...................................................................................................................... 82

6.3.1 Net energy consumption ............................................................................................................ 82

6.3.2 Greenhouse gas emission (CO2e) ........................................................................................... 83

6.3.3 Net production cost ................................................................................................................ 84

6.3.4 Sensitivity analysis ................................................................................................................. 86

6.4 Conclusion ..................................................................................................................................... 90

Chapter 7: Identification of suitable plant location for ethanol industry in Ontario, Canada ... 91

7.1 Introduction ............................................................................................................................... 91

7.2 Materials and methods ................................................................................................................... 91

7.2.1 Study area ............................................................................................................................... 91

7.2.2 System boundary .................................................................................................................... 91

7.2.3 Transportation, ethanol production and cost analysis ............................................................ 92

7.3 Results and discussion.................................................................................................................... 93

7.3.1 Net energy consumption ......................................................................................................... 93

7.3.2 Greenhouse gas emission (CO2 e) ........................................................................................... 94

7.3.3 Production cost ....................................................................................................................... 97

7.3.4 Sensitivity analysis ................................................................................................................. 98

7.4 Conclusion ................................................................................................................................... 102

Chapter 8: Development of a Continuous Stirred Tank Bioreactor for Syngas Fermentation 103

ix

8.1 Introduction .................................................................................................................................. 103

8.2 Materials and Methods ................................................................................................................. 104

8.2.1 Reactor development ............................................................................................................ 104

8.2.2 Microorganism and media .................................................................................................... 106

8.2.3 Syngas fermentation ............................................................................................................. 106

8.2.4 Analytical method ................................................................................................................ 109

8.3 Results and discussion.................................................................................................................. 109

8.3.1 pH and temperature profile during syngas fermentation ...................................................... 109

8.3.2 Ethanol and other alcoholic compounds ............................................................................... 110

8.4 Conclusion ................................................................................................................................... 111

Chapter 9: Evaluation of the Life Cycle of Ethanol derived from Biosyngas Fermentation .... 112

9.1 Introduction .................................................................................................................................. 112

9.2 Materials and methods ................................................................................................................. 113

9.2.1 System boundary and assumptions ....................................................................................... 113

9.2.2 Pretreatment (torrefaction) ................................................................................................... 113

9.2.3 Ultimate analysis .................................................................................................................. 114

9.2.4 Gasification and syngas cleaning ......................................................................................... 114

9.2.5 Syngas fermentation ............................................................................................................. 115

9.2.6 Separation (distillation & purification) ................................................................................ 116

9.2.7 Waste management ............................................................................................................... 117

9.2.8 Cost analysis ......................................................................................................................... 117

9.2.9 Data collection ...................................................................................................................... 117

9.3 Results and discussion.................................................................................................................. 117

9.3.1 Net energy consumption ....................................................................................................... 117

9.3.2 GHG emission (CO2e) .......................................................................................................... 119

9.3.3 Production cost ..................................................................................................................... 120

9.3.4 Sensitivity analysis ............................................................................................................... 121

9.4 Conclusion ...................................................................................................................................... 127

Chapter 10: Conclusions and Recommendations ......................................................................... 128

10.1 Conclusions .................................................................................................................................. 128

10.1.1 Evaluation of the LC of ethanol produced by enzymatic hydrolysis process ....................... 128

10.1.2 Evaluation of the LC of ethanol produced by gasification-biosynthesis process ................. 129

x

10.1.3 Continuous stirred tank bioreactor ....................................................................................... 130

10.2 Recommendations ........................................................................................................................ 130

10.2.1 Life cycle assessment ........................................................................................................... 130

10.2.2 Improvement of bioreactor ................................................................................................... 130

Chapter 11: References ................................................................................................................... 132

Appendices ........................................................................................................................................ 170

A-2-1 The schematic diagram of chemical looping gasification (CLG) system ................................... 170

A-2-2 Brief summary of microorganisms identified and used for syngas fermentation ....................... 171

A-2-3 Syngas fermentation parameters and ethanol yield .................................................................... 175

A-6-1 Land classification in Ontario .................................................................................................... 177

A-6-2 On-farm inputs for miscanthus cultivation in different regions ................................................. 178

A-6-3 On-farm energy and other inputs for miscanthus cultivation ..................................................... 180

A-6-4 Estimated emission from farm input and carbon sequestration .................................................. 182

A-6-5 Calculation of energy consumption and material cost of enzyme production ............................ 183

A-8-1 Membrane separator ................................................................................................................... 184

A-8-2 Membrane support ...................................................................................................................... 184

A-8-3 List of materials/accessories for the developed bioreactor ......................................................... 185

A-8-4 List of materials/accessories for anaerobic gas chamber ............................................................ 186

A-8-5 List of chemicals and their amount used for broth media .......................................................... 186

A-8-6 Photograph of the incubator ....................................................................................................... 187

A-8-7 Calibration curve of the pump .................................................................................................... 187

A-8-8 Photographs of overall experimental setup ................................................................................ 188

A-9-1 Experimental setup of torrefaction process adopted in this study .............................................. 189

A-9-2 Composition of flue gas from biomass torrefaction process ...................................................... 189

A-9-3 Energy consumption in the torrefaction of biomass (for 45 min) .............................................. 190

A-9-4 Flash 200 CHNS-O, Organic Elemental Analyzer ..................................................................... 190

A-9-5 Photograph of thermo gravimetric analyzer (TGA) ................................................................... 191

A-9-6 Photograph of Fourier transform infrared spectroscopy (FT-IR) ............................................... 191

A-9-7 TGA/FT-IR experimental parameters ........................................................................................ 192

A- 9-8 Comparison among various raw biomasses............................................................................... 192

xi

A- 9-9 Comparison among various torrefied biomasses ....................................................................... 193

A- 9-10 Comparison among various raw biomasses degraded with CaO ............................................. 193

A- 9-11 Comparison among various torrefied biomasses degraded with CaO ..................................... 194

A- 9-12 Comparison among raw and torrefied with or without CaO (miscanthus) .............................. 194

A-9-13 Cold gas efficiency (CGE) calculation for steam gasification . ............................................... 195

A-9-14 Summary of ASPEN simulation parameters ............................................................................ 195

A-9-15 Summary of ASPEN simulation parameters and CLG block diagram..................................... 196

A-9-16 CLG simulation flowsheet ....................................................................................................... 197

A-9-17 Product gas compositions (simulated) and CGE ...................................................................... 198

xii

List of Tables

Table 1.1 Projected biofuel production in major biofuel producing countries and in the world ....................... 2

Table 1.2 Lignocellulosic ethanol plants in Canada and their capacity ............................................................. 3

Table 2.1 Pretreatment processes of biomass .................................................................................................. 12

Table 2.2 Brief summary of energy consumption in distillation processes ..................................................... 15

Table 2.3 The LC GHG emission/energy consumption of ethanol produced by thermochemical

conversion process ................................................................................................................................... 33

Table 2.4 Tax credits on ethanol in various provinces in Canada .................................................................... 34

Table 2.5 Summary of the reported cost of ethanol produced from different feedstocks (biochemical

conversion) ............................................................................................................................................... 38

Table 2.6 Summary of the reported cost of ethanol from different feedstock and energy efficiency

(thermochemical conversion) ................................................................................................................... 40

Table 3.1 Mill residues production in Canada in 2004 (ODt: Oven dry tonnes) ............................................. 43

Table 3.2 Volatile matter, fixed carbon, and ash content in selected biomass (dry basis) ............................... 43

Table 3.3 Potential feedstocks and their major components ............................................................................ 44

Table 3.4 Chemical composition of different feedstock .................................................................................. 44

Table 4.1 Summary of parameters for which data are collected from the literature ........................................ 53

Table 4.2 Summary of the reported cost of ethanol produced from different feedstock ................................. 55

Table 4.3 Ethanol yield from wheat straw ....................................................................................................... 56

Table 5.1 Scenarios of this study. .................................................................................................................... 63

Table 5.2 Summary of parameters for which data are collected from literature .............................................. 66

Table 6.1 Land areas in Ontario, ha ................................................................................................................. 77

Table 6.2 Land classes, soil types and miscanthus yield ................................................................................. 77

Table 6.3 Scenarios of this study. .................................................................................................................... 78

Table 6.4 Summary of parameters for which data are collected from literature .............................................. 79

Table 7.1 Land area under different tillable land classes and various regions in Ontario, ha.......................... 92

Table 7.2 Scenarios of this study ..................................................................................................................... 92

Table 9.1 Components of different feedstock ................................................................................................ 114

Table 9.2 Ethanol yield from biosyngas fermentation ................................................................................... 116

Table 9.3 Summary of parameters for which data are collected from literature or estimated ....................... 118

xiii

List of Figures

Figure 1.1 Contribution of this study ................................................................................................................. 9

Figure 2.1Schematic diagram of ethanol production process from syngas...................................................... 27

Figure 3.1 Stages of an LCA (ISO, 2006) ........................................................................................................ 42

Figure 3.2 System boundary of this study ........................................................................................................ 45

Figure 3.3 Structure of the LCIA method based on endpoint modeling (LIME2) ........................................... 47

Figure 4.1 Schematic diagrams of the life cycle of wheat straw and the system boundary of this study ........ 50

Figure 4.2 Schematic diagram of ethanol production process from biomass .................................................. 51

Figure 4.3 Energy, emission and cost breakdown of the life cycle of ethanol produced from wheat straw. ... 54

Figure 4.4 Effect of ethanol yield on net energy consumption, emission and production cost of ethanol ...... 57

Figure 4.5 Effect of feedstock cost on the production cost of ethanol ............................................................. 58

Figure 4.6 Effect of plant capacity on the production cost and emission of the life cycle of ethanol ............. 58

Figure 4.7 Effect of feedstock cost on the emission of the life cycle of ethanol ............................................. 60

Figure 4.8 Effect of system boundary and the ethanol yield on life cycle GHG emission of ethanol ............. 60

Figure 4.9 Effect of carbon sequestration and ethanol yield on the life cycle GHG emission ........................ 61

Figure 5.1 Schematic diagrams of the LC of sawdust and the system boundary of this study ........................ 64

Figure 5.2 Energy breakdown of the life cycle of ethanol ............................................................................... 67

Figure 5.3 Emission breakdown of the life cycle of ethanol ............................................................................ 68

Figure 5.4 Effect of carbon sequestration on the net emission of the life cycle of ethanol ............................. 69

Figure 5.5 Cost breakdown of the life cycle of ethanol ................................................................................... 70

Figure 5.6 Effect of the change in energy consumption at different stages on net energy consumption. ........ 71

Figure 5.7 Effect of the change in energy consumption at different stages on net emission (kg-CO2 e/L) ..... 71

Figure 5.8 Effect of the change in energy consumption at different stages on net cost ($/L) .......................... 72

Figure 5.9 Effect of the changes in feedstock- and fixed cost on the production cost of ethanol .................... 72

Figure 6.1 Transportation fuel consumption and contribution of ethanol in Canada ...................................... 75

Figure 6.2 Different regions in Ontario, Canada ............................................................................................. 76

Figure 6.3 Different regions and land classes in Ontario, Canada ................................................................... 76

Figure 6.4 Schematic diagrams of the life cycle of sawdust and the system boundary of this study .............. 80

Figure 6.5 Energy breakdown of the life cycle of ethanol derived from miscanthus ...................................... 83

Figure 6.6 Emission breakdown of the life cycle of ethanol derived from miscanthus ................................... 84

Figure 6.7 Cost breakdown of the life cycle of ethanol ................................................................................... 85

Figure 6.8 Effect of the variation in transportation distance and pretreatment energy consumption on the

net energy consumption (MJ/L) ............................................................................................................... 87

Figure 6.9 Effect of the variation in transportation distance and pretreatment energy consumption on the

net emission and production cost ............................................................................................................. 88

xiv

Figure 6.10 Effect of feedstock and fixed cost (S/L) ....................................................................................... 88

Figure 6.11 Effect of carbon dynamics on the net emission of the life cycle of ethanol ................................. 89

Figure 7.1 Feedstock transportation distance at different location in Ontario ................................................. 93

Figure 7.2 Energy breakdown of the life cycle of ethanol (Southern Ontario) ................................................ 94

Figure 7.3 Net energy consumption at different location in Ontario ............................................................... 95

Figure 7.4 Emission breakdown of the life cycle of ethanol (Southern Ontario) ............................................ 96

Figure 7.5 Net emissions at different location in Ontario ................................................................................ 96

Figure 7.6 Cost breakdown of the life cycle of ethanol (Southern Ontario) .................................................... 97

Figure 7.7 Net production cost at different location in Ontario. ...................................................................... 98

Figure 7.8 Effect of ethanol plant capacity on production cost and emission ................................................. 99

Figure 7.9 Effect of ethanol plant capacity on production cost and emission ............................................... 100

Figure 7.10 Effect of the variation of different parameters on net energy consumption (MJ/L) ................... 101

Figure 7.11 Effect of the variation of different parameters on net emission ................................................. 101

Figure 8.1 Photograph of the developed reactor ............................................................................................ 104

Figure 8.2 Schematic diagram of the gas chamber (not to scale) .................................................................. 105

Figure 8.3 Photograph of the developed gas chamber ................................................................................... 105

Figure 8.4 Photograph of the experimental setup .......................................................................................... 108

Figure 8.5 Schematic diagram of the experimental setup of this study ......................................................... 108

Figure 8.6 pH and temperature profile of the fermentation broth .................................................................. 110

Figure 8.7 Mass spectra of the effluent .......................................................................................................... 111

Figure 9.1 Schematic diagram of the system boundary of this study. ........................................................... 113

Figure 9.2 Energy consumption at various stages of the LC of ethanol ........................................................ 119

Figure 9.3 Emission at different stages of the LC of ethanol ......................................................................... 120

Figure 9.4 Production cost at different stages of the LC of ethanol .............................................................. 121

Figure 9.5 Effect of transportation and pretreatment on net energy consumption ......................................... 122

Figure 9.6 Effect of transportation and pretreatment on emission and cost ................................................... 123

Figure 9.7 Effect of the variation of gasification and heat recovery on net energy consumption (MJ/L) ..... 123

Figure 9.8 Effect of the variation of gasification and heat recovery on emission and cost ........................... 124

Figure 9.9 Effect of the variation of fermentation and distillation on net energy consumption (MJ/L) ........ 125

Figure 9.10 Effect of variation of fermentation and distillation on emission and cost .................................. 125

Figure 9.11 Effect of variation of fixed and feedstock cost on production cost ($/L) ................................... 126

Figure 9.12 Effect of CGE on GHG emissions and production cost ............................................................. 126

1

Chapter 1

Introduction

1.1. Rationale

The global energy demand was 424 EJ/year in 2000 and is increasing at the rate of 2.2%

per year (Lal, 2009) and the world’s total primary energy supply was reported to be 479 EJ in

2005 (GBEP, 2007). With existing technologies and consumption patterns, global energy

demand could double by 2050 due to the combination of population and economic growth

(UNEP, 2007). Greenhouse gas (GHG) emissions, which have increased remarkably due to

tremendous energy use liable for global warming, and perhaps the most serious problem that

humankind faces today. The growing concerns about climate change, rising costs of fossil fuels

and the geo-political uncertainty associated with possible interruption of current fossil fuel-based

energy supplies have motivated individuals, organizations and nations to seek clean and

renewable substitutes. Liquid biofuels (ethanol and biodiesel) are widely recognized alternatives

to fossil fuels. Table 1.1 represents the projected biofuel production in different regions.

Renewable energy not only reduces the reliance on foreign oil and improves energy security, but

also provides significant environmental benefits and enhances rural economies (Kim & Dale,

2003; Spatari et al., 2005; Farrell et al., 2006). In contrast, this rapid expansion affects virtually

every aspect of the field crop sectors and there remain an inevitable conflict between the

increasing diversion of crops or crop land for fuel instead of food (Liew et al., 2013). Biofuels

contain low sulphur, noted to be nontoxic, biodegradable and can reduce harmful GHG, carbon

monoxide, hydrocarbons and particulate matter (Mata et al., 2010).

The transportation sector in Canada accounts for about 25% of the nation's energy use and

the major part of this energy (99%) comes from fossil fuels (NRC, 2013). In 2009, Canada

committed to reduce total GHG emissions by 17% from 2005 levels by 2020. The Renewable

Energy Regulation (SOR/2010-189) was enacted (came-into-force from July 1, 2011) in Canada

which requires fuel producers and importers of gasoline to have renewable fuel content of at

least 5% distillates (by volume) that they produce and import yearly (Environment Canada,

2010), which generates a considerable demand on biofuels. The Government of Canada also

offers $0.10/L and $0.26/L operating incentive for ethanol and biodiesel, respectively, for up to

seven consecutive years (Biofuelnet, 2013). Consequently, the interest in biofuels is expanding.

The cellulosic ethanol processing plants in Canada and their production capacity is reported in

2

Table 1.2. First generation biofuels are produced from food or feed grains, and thus compete

with food or feed and contribute to higher food prices (Yang et al., 2009; Kazi et al., 2010a;

Mueller et al., 2011; Liew et al., 2013). Consequently, production of second/third generation

biofuels from lignocellulosic biomass (hereafter referred to biomass) has been emphasized,

because it does not compete with food or feed (Zaldivar et al., 2001; Gray et al., 2006; Hahn-

Hägerdal et al., 2006; Cardona & Sánchez, 2007; Sánchez & Cardona, 2008).

Table 1.1 Projected biofuel production in major biofuel producing countries and in the world

Country/

Region

Biofuels

Projected production or consumption in different years, L (×106)

2013 2015 2017 2019 2021

World Ethanol

113853.8 129015.6 140902.1 150728.4 162664.6

Biodiesel

28507.8 29879.9 33314.7 36161.5 39636.9

Brazil Ethanol

28684.5 37323.3 43483.7 45644.3 47929.6

Biodiesel

2587.0 2744.0 2902.8 3069.9 3245.5

USA Ethanol

55769.8 60536.2 63784.9 68825.1 75889.1

Biodiesel

6057.5 5138.5 5192.0 5191.7 6390.9

Canada Ethanol

1605.0 1512.5 1430.4 1459.6 1486.2

Biodiesel

487.8 452.2 413.9 380.5 355.2

Europe Ethanol

7048.5 7625.1 8528.8 9978.9 11565.8

Biodiesel

11287.6 12151.7 14211.2 16374.1 17784.2

Argentina Ethanol

497.3 625.9 742.2 853.2 963.6

Biodiesel

2697.1 2956.4 3171.4 3136.4 3300.5 Data source: OECD-FAO, 2013

Several methods have been used in producing biofuels from biomass. Ethanol has been

produced by either biochemical processes (hydrolysis) or thermochemical processes (gasification

/pyrolysis of biomass to syngas followed by biosynthesis or chemical synthesis). Traditionally,

ethanol production from biomass involves different steps of pretreatment, acid/enzymatic

hydrolysis (saccharification), fermentation and distillation. However, the commercialization of

ethanol production from biomass has been hindered mainly by the prohibitive cost of the

expensive and inefficient pretreatment, and saccharification and distillation methods. Each

process has its strengths and weaknesses. Biosynthesis of syngas results in poor mass transfer

3

properties of gaseous substrates and low ethanol yield (Munasinghe & Khanal, 2010).

Conversely, higher ethanol yield was reported in this process (Clausen & Gaddy, 1993).

Biosynthesis is a two-stage process consisting of biomass gasification followed by microbial

fermentation of syngas into ethanol. This process may offer distinct advantages (utilization of

whole biomass including lignin, irrespective of biomass quality, elimination of complex

pretreatment steps and costly enzymes, higher specificity of biocatalysts, independence of the

H2/CO ratio, aseptic operation of syngas fermentation, bioreactor operation at ambient

conditions, no issue with inorganic catalyst poisoning due to trace sulphur containing gas) over

both hydrolysis-fermentation and gasification-catalytic synthesis (high pressure, temperature,

expensive metallic catalyst and complex gas cleaning). Traditional fermentations rely on

carbohydrates as the carbon and energy sources for the microbial growth; however, syngas

fermentation utilizes microorganisms (especially, the Clostridium family) capable of

metabolizing syngas into ethanol and other valuable chemicals in an inexpensive liquid substrate.

However, poor mass transfer between gaseous and liquid substrates is one of the most significant

challenges for this process.

Table 1.2 Lignocellulosic ethanol plants in Canada and their capacity

Company Location Feedstock

Capacity, 106

L/y Remarks

Enerkem Inc.

Westbury,

QC

Wood waste from used

utility poles, RDF 5 Existing

Iogen Corp. Ottawa, ON

Wheat Straw, Oat Straw,

Barley Straw, Bagasse 2 Existing

Enerkem/GreenField -

Varennes cellulosic

Ethanol LP

Varennes,

QC

RDF, C&D debris

38

Proposed

Mascoma Corp.

Drayton Valley

Drayton

Valley, AB

Hardwood

72

Proposed

Nipawin Biomass

Ethanol New

Generation Co-

operative Ltd.

Nipawin,

SK

Wood waste, Straw

100

Proposed

Woodland Biofuel Inc. Sarnia, ON Wood waste 0.5 Proposed

Source: EPM, 2014

4

Although the thermochemical process produces ethanol in large quantities, it requires

expensive catalysts and high operating pressure (Subramani and Gangwal, 2008). Many

researchers have studied ethanol production processes from syngas either by biosynthesis or

catalytic synthesis process (Ruth, 2005; Martchamadol, 2007; Clausen & Gaddy, 2008;

Munasinghe & Khanal, 2010) except for few examples (Foust et al., 2009; Mu et al., 2010).

Bioenergy systems have also been modeled with Aspen Plus and evaluated to estimate

production cost and environmental impacts of bioenergy. GHG emissions and production cost of

biofuels were reported to be dependent on both technical and economic parameters, such as the

cost and choice of feedstock, conversion technologies and value of coproducts/byproducts

(Wyman, 1994; Ballerini et al., 1994; Wooley et al., 1999a; Aden et al., 2002; Mabee et al.,

2006; Aden, 2008; Dutta et al., 2010a; Balat, 2011). Thus, a wide variation was reported in the

case of GHG emissions and production cost. The cost of cellulase was the major expense when

producing lignocellulosic ethanol with conventional technology (Singh & Kumar, 2010) and

contributed about 40–55% of the enzymatic ethanol production cost. Distillation, enzyme

production and pretreatment were reported to be the main contributor to energy consumption and

GHG emissions in the LC of ethanol produced by conventional technology (Roy et al., 2012a,b;

Orikasa et al., 2009). The cost effective and innovative fermentation strategies integrated in the

technology chain of gasification and gas cleaning combined with syngas fermentation, and

catalytic synthesis could significantly improve the overall economics of ethanol from biomass.

Life cycle assessment (LCA) is a tool for evaluating environmental effects of a product,

process, or activity throughout its LC. The key objective of an LCA study is to provide as

complete a portrait as possible of energy consumption, environmental impacts, economic

viability and their rebound effects, hence enable effective planning for a sustainable society. This

study evaluated the LC of ethanol to estimate net energy consumption and GHG emissions to

identify the hotspots, improve the production process, and determine if environmentally friendly

ethanol can be produced from biomass. It is also known that environmental information

generated through an LCA, while useful, does not always provide a sufficient basis for making a

sound decision on an investment. The cost analysis along with the estimation of GHG emissions

broadens the process of making sound decisions. Therefore, the production cost of ethanol has

also been estimated with both fixed costs (straight line depreciation on installation, labor,

maintenance and interest on investment) and variable costs (feedstock, yeast/bacteria, utilities

5

and waste management). The Aspen Plus process modeling has also been used for steady-state

simulation (for syngas composition and syngas production to calculate the cold gas efficiency).

Sensitivity analyses were also conducted to determine the acceptability, profitability and risk on

investment if any. Finally, the results were interpreted to communicate to the stakeholder,

environmental activist and policy makers, which may help investor and policy maker’s decision

and draw more investment in this sector.

1.2. Objectives

The goal of this research was to investigate the technical feasibility of ethanol derived from

biomass (crop/forest residues and energy crops) by hybrid thermochemical and

biochemical/chemical processing, considering innovative technologies [enzymatic hydrolysis:

pretreatment (CaCCO: calcium capturing by carbonation), vacuum extractive fermentation and

distillation; biosynthesis: with/without chemical looping gasification (CLG) of torrefied and non-

torrefied biomass]. At first, the conventional ethanol production processes were evaluated to

have the baseline information and then Aspen Plus simulation software has been used to gather

required data (especially for the gasification processes). Finally, the LC of ethanol produced by

the above mentioned innovative technologies have been evaluated by using the LCA

methodologies (centering on net energy consumption, GHG emissions and production cost) to

determine the environmentally preferable and economically viable energy pathway for Ontario,

Canada. This study developed/identified environmentally preferable and economically viable

innovative technology (either biochemical or thermochemical) for lignocellulosic ethanol, which

is renewable and clean, thus help in reducing GHG emissions from the present energy sector in

Canada, may enable to compete economically and technologically in the world ethanol markets,

and contribute to improve rural economy in Ontario, Canada.

The specific objectives of this study were:

i. Evaluate the LC of ethanol produced by conventional /traditional technology (enzymatic

hydrolysis).

ii. Develop a bioreactor for syngas fermentation into ethanol.

iii. Evaluate the LC of ethanol produced from syngas (biosynthesis). Both torrefied and

non-torrefied biomass were used for a novel CLG and steam gasification (without CLG)

processes.

iv. Determine the LC cost of ethanol.

6

1.3. Scope and limitation of this research

This research consists of four individual objectives, the scope of those are explained

bellow.

i. Three types of feedstocks (wheat straw, sawdust and miscanthus) have been selected for

enzymatic hydrolysis process. The LC of ethanol produced from these feedstocks by

enzymatic hydrolysis process was evaluated based on the estimated and literature data.

Ontario specific data were collected from the literature for miscanthus cultivation. All

those data were neither site nor country specific nor from the same plant size. The plant

capacity was considered to be 20000 kL/yr. The yearly operation period and life span of

the processing plant was assumed to be 350 days and 20 years, respectively (Dutta et al.,

2011; Huang et al., 2009; Wu et al., 2006). Some optimistic literature data has also been

used in this study, especially the enzyme production.

ii. Mass transfer between liquid and gaseous substrates hinders the syngas fermentation

into ethanol. The bioreactor has been developed as a part of this study to improve the

mass transfer by incorporating an innovative gas supply system. In addition to this, a

membrane separation system has also been added to the reactor to extract the effluent

excluding the microorganism. The reactor has been tested with only CO instead of

syngas (60%CO; 35%H2 and 5%CO2) to prove the concept.

iii. The torrefied and non-torrefied biomass have been considered for gasification (with or

without CLG) processes. Thermal degradation experiments have been conducted with

TGA to determine the effectiveness of CLG. Aspen Plus simulation at equilibrium

condition has been used to generate data for syngas composition and volume to

calculate the cold gas efficiency, thus the ethanol yield. The estimated and literature

data have been used to evaluate the LC of ethanol produced by gasification-biosynthesis

process considering the plant capacity of 20000 kL/yr.

iv. The production cost has been determined based on both the fixed and variable costs.

The life span of an ethanol plant and an operation period was assumed to be 20 years

and 350 days per year, respectively (Dutta et al., 2011; Huang et al., 2009; Wu et al.,

2006). Both the estimated and literature data have been used to determine the

production cost.

7

1.4 Novelty of the research

This research consists of both analytical and theoretical analysis (LCA: policy analysis)

which are thought to be a novel approach.

1.4.1 Bioreactor development

A novel continuous stir-tank bioreactor (consists of innovative gas supply and effluent

extraction processes) has been developed to produce ethanol from syngas by using

microorganism. Innovative systems for syngas supply and ethanol extraction have also been

developed to improve gas-liquid mass transfer and reuse of microorganism. The research also

attempted to produce ethanol from syngas by using the developed bioreactor.

1.4.2 Chemical looping gasification (CLG)

Biomass characteristics have been determined at the laboratory, and those properties were

applied for further studies. A macro-reactor (QWM: Quartz wool matrix) was used for the

torrefaction process. The thermal degradation properties were also determined. The weight loss

and heat flow during the degradation process were monitored in a micro-reactor (TGA-FTIR).

The CLG concept consists of two fluidized bed reactors (operate parallely), which were

functioning as calciner and gasifier. In the calciner, limestone was calcined. Lime was then

circulated to the reactor (gasifier), where the lime sorbent was carbonated in parallel with the

gasification reactions, tying up most of the CO2 as CaCO3. The CaCO3 was then recirculated to

the calciner, acting as a regenerator. This looping concept has definite economic promise, and

was particularly attractive in the gasification processes, where it has added benefits-shifting the

equilibrium to yield more hydrogen, releasing heat in the reactor (where the main reactions were

endothermic), reducing loss of activity, and decreasing tar formation, at the same time capturing

CO2 produced during gasification.

Aspen Plus (V7.3) simulation software has been used to determine the syngas productivity

and gas composition at equilibrium condition for treated (torrefied) and untreated feedstock with

(CLG: chemical looping gasification) or without CO2 capture. The novelty of the CLG process

lies in the generation of relatively pure H2 from biomass on a continuous basis, while CO2

produces as a byproduct using steam as the gasifying agent. Another unique feature of the

process is internal regeneration of the sorbent, fouled in the gasifier. The technology served the

twin purpose of regenerating the sorbent, and generation of relatively clean syngas which is a

novel approach. Utilization of torrefied biomass in ethanol production process is also known to

8

be a novel idea and is a novelty of this research. Thus, it will open up a new area of research on

H2-enriched gas production from biomass with in-process CO2 capture. The specific new

information generated by this study are as follows:

Torrefied and non-torrefied feedstock has been used in the CLG process.

Torrefied feedstock produced relatively better quality syngas.

The use of torrefied feedstock in ethanol production process.

1.4.3 Life cycle assessment

The research evaluated the life cycle (LC) of ethanol produced from various feedstocks

(wheat straw, sawdust and miscanthus), identified the feedstock and locations for ethanol

industry in Ontario, Canada; especially, from energy crop (miscanthus) by using hybrid

enzymatic hydrolysis and biosynthesis (syngas fermentation) processes. Both the

technoeconomic and environmental evaluation were carried out by adopting the life cycle

assessment (LCA) methodologies. The LCA methodologies also identify hotspots, and help to

improve the production process. The main hotspots identified are either pretreatment, feedstock

or gasification in the case of GHG emissions, depending on the type of feedstock, and scenario

or the conversion technology. The LCA study on the ethanol production process which

incorporated the torrefaction and the chemical looping gasification (CLG) is also a novel

approach. The novel information generated in this study would be useful to researchers, investors

and policy makers which might help Ontario compete economically and technologically in the

world ethanol markets, and contribute to improve rural economies in Canada. The specific new

and novel information generated by this study are as follows:

This study generated new information on lignocellulosic ethanol in the context of

Ontario.

Identified the potential location for miscanthus based ethanol industry in Ontario.

Life cycle of ethanol from torrefied and non-torrefied feedstocks with or without CLG

has also been evaluated, which is a novel work of this kind.

1.5 Contribution of this research

The following diagram briefly represents the plan, background and contribution of this

study (Fig. 1.1).

9

Figure 1.1 Contribution of this study

Research on

torrefied biomass

thermal degradation

with or without CaO

is limited.

Syngas fermentation

with microorganism

in an innovative

reactor is scarce.

Life cycle of ethanol

from lignocellulosic

biomass, wheat straw,

sawdust and

miscanthus which

received limited

attention.

LCA of syngas

fermentation

(biosynthesis) is scarce

Ste

am g

asif

icat

ion w

ith o

r

wit

hout

CL

G o

f tr

eate

d o

r

untr

eate

d

bio

mas

s is

sca

rce

Stu

die

s on i

nnovat

ive

gas

su

pply

and e

fflu

ent

extr

acti

on

is l

imit

ted

LCA of ethanol derived from lignocellulosic biomass

Pla

nned

B

ackgro

und

Experiments Simulation LCA & LCCA Bioreactor

Contr

ibuti

on

Biomass

characterization

Torrefaction of

biomass in QWM

fluidized bed reactor

Thermal degradation

with or without CaO

in a micro-gasifier

(TGA)

Syngas fermentation

with microorganism

for ethanol

Syngas

com

posi

tion a

nd s

yngas

pro

duct

ivit

y a

re s

imula

ted b

y A

SP

EN

Plu

s si

mula

tion s

oft

war

e

A n

ovel

bio

reac

tor

has

bee

n d

edel

oped

whic

h c

onsi

sts

of

innovat

ive

gas

supply

and e

fflu

ent

extr

acti

on m

ethods

LCA and LCCA of

ethanol produced from

wheat straw, sawdust,

miscanthus by

enzymatic hydrolysis

process

LCA and LCCA of

ethanol derived from

syngas (gasification-

biosynthesis path) from

treated or untreated

biomass with or without

CLG.

(QWM: quartz wool matrix, TGA: Thermo gravimetric anatysis, LCCA: life cycle cost analysis)

10

1.6 Publications from this research

1.6.1 Publications in peer reviewed journals

Life cycle assessment of ethanol derived from sawdust. Poritosh Roy & Animesh Dutta,

2013. Bioresource Technology, 150(December), 407–411.

A review of life cycle of ethanol produced from bio syngas. Poritosh Roy & Animesh

Dutta, 2013. Bioethanol, 1(1), 9–19.

Life cycle assessment of ethanol produced from wheat straw. Poritosh Roy & Animesh

Dutta, 2012. Journal of Biobased Materials and Bioenergy, 6(3), 276–282.

1.6.2 Submitted manuscripts

Evaluation of the life cycle of ethanol derived from miscanthus in Ontario, Canada.

Poritosh Roy, Animesh Dutta & Bill Deen, 2014. (Biomass and Bioenergy).

Review on syngas fermentation processes for bioethanol. Bimal Acharya, Poritosh Roy &

Animesh Dutta, 2014. (Biofuels).

1.6.3 Publications: Research presentations

The potential location for lignocellulosic ethanol processing plant in Ontario. Poritosh

Roy, Animesh Dutta & Bill Deen, 2014. 13th

International Symposium on Bioplastics,

Biocomposites & Biorefining: Moving towards a Sustainable Bioeconomy, Guelph,

Ontario, Canada, May 19–24.

Miscanthus: A promising feedstock for ethanol in Ontario. Poritosh Roy, Animesh

Dutta & Bill Deen, 2013. Bioeconomy Research Highlights Poster Showcase, Guelph,

Ontario, Canada, November 27.

Evaluation of the life cycle of ethanol produced from agri-residues (wheat straw).

Poritosh Roy & Animesh Dutta, 2012. Growing the Margins/Canadian Farm & Food

Biogas Conferences, London, Canada, March 5–7.

Life cycle assessment of bioethanol produced from biomass. Poritosh Roy & Animesh

Dutta, 2011. Bioeconomy Research Highlights Poster Showcase, Guelph, Ontario,

Canada, December 7.

11

Chapter 2

Literature Review

2.1. Ethanol production via biochemical conversion process (enzymatic hydrolysis)

2.1.1. Pretreatment

Pretreatment (either physical or chemical or both) is a prerequisite for biological

conversion of biomasses (Lynd et al., 2008; Yang & Wyman, 2008) to make them more

amenable to cellulose hydrolysis. Table 2.1 summarizes some pretreatment processes of biomass.

Physical pretreatment refers to the size reduction of feedstock to increase enzyme-accessible

surface areas (Zhu et al., 2009b) and chemical pretreatments remove or modify key chemical

components that interfere with biomass cellulose saccharification, mainly hemicelluloses and

lignin (Zhu et al., 2009a; Zhu & Pan, 2010). The potential pretreatment methods are: acid

hydrolysis (concentrated or diluted), liquid hot water extraction, steam explosion, dilute acid-

steam explosion, ammonia fiber explosion, lime pretreatment, etc. (Holtzapple et al., 1991;

Mosier, et al., 2005a,b; Wyman et al., 2005; Yang & Wyman, 2008; Huang et al., 2009; Banerjee

et al., 2010a; Manzanares et al., 2012). The acid pretreatments are reported to be toxic,

hazardous, and corrosive, and require expensive reactors resistant to corrosion and also causes

difficulties in waste management streams (Yang & Wyman, 2008).

High energy requirement and inhibitors generation are reported to be major drawback of

steam explosion method (Hendriks & Zeeman, 2009; Banerjee et al., 2010a). The cost of

ammonia, its handling, and recovery, and high energy consumption in recompression are the

main bottlenecks in the process (Banerjee et al., 2010a). On the other hand, the alkaline

pretreatment is reported to be more suitable and effective for herbaceous crops and agricultural

residues (Bjerre et al., 1996; Chang et al., 2001; Rabelo et al., 2009). The hot water pretreatment

avoids the formation of inhibitors and catalyze hydrolysis of cellulosic materials (Mosier et al.,

2005a; Yu et al., 2010). It is also reported that cost of lime is relatively low and safer reagent

compared with other alkalis and ammonia (Kaar & Holtzapple, 2000; Saha & Cotta, 2008). The

CaCCO (calcium capturing by carbonation) process has also been developed to facilitate the

pretreatment of biomass (Park et al., 2010; Shiroma et al., 2011). The author also noted that this

pretreatment can also be applied at room temperature (7 days and 10% lime) to facilitate the

conversion process. Pretreatment process contributes about 18–20% of the total cost of

12

biological production of cellulosic ethanol, which is greater than any other single step of the LC

of ethanol (Aden et al., 2002; Yang & Wyman, 2008; Wooley et al., 1999b). Although the

pretreament to biomasses may vary depending on their characteristics (for example sawdust from

softwood vs hardwood), the hotwater pretreatment (with 10% lime) has been adopted for this

study (Park et al., 2010; Shiroma et al., 2011).

Table 2.1 Pretreatment processes of biomass

Methods Processes Remarks

Physical

Milling: ball, hammer, two-roll, colloid, vibro etc. High energy demand

Irradiation: microwave, electron-beam, gamma ray Cannot remove lignin

Others: torrefaction, extrusion, pyrolysis, high pressure

steam, hotwater etc.,

No chemicals are used

Physico-

chemical

Explosion: steam, ammonia fiber (AFEX), CO2, SO2

Alkali: NaOH, NH4, (NH4)2S, Ca(OH)2, RT-CaCCO,

ammonia recycled percolation (ARP), liquid ammonia

Low cost

Gas: NO2, ClO2, SO2

Acid: H2SO4, HCl, H3PO4 Complex downstream

Oxidizing: Hydrogen peroxide, wet oxidation, ozone

Solvent extraction: Ethanol-water, benzene-water,

ethylene glycol, butanol-water, swelling agents

SPORL: Acid with sulfite or bisulfite

Biological Fungi, actinomycetes Low energy requirement,

low treatment rate

Source: Taherzadeh & Karimi, 2008; Zhu et al., 2009a; Shiroma et al., 2011; Wilkins, 2011; Yoon et al., 1995.

2.1.2. Fermentation

Fermentation is the chemical decomposition process of a substance by bacteria, yeasts, or

other microorganisms. This process is usually used in the preparation of alcohol, wine and

liquor. The complex organic compounds, such as glucose, are broken down by the action of

enzymes into simpler compounds in an anaerobic environment, known as an energy generating

process, where organic compound act as both electron donor and acceptors (Stanbury, 2000).

Microbial fermentation is classified into five groups based on the produce: microbial cell or

biomass, microbial metabolites, microbial enzymes, recombinant products and biotransformation

13

(Stanbury, 2000). Biochemical conversion process of biomass consists of four biological steps:

enzyme production, enzymatic hydrolysis, hexose fermentation and pentose fermentation.

Fermentation of the pretreated biomass can be carried out in a variety of ways: separate

hydrolysis and fermentation (SHF), simultaneous saccharification and fermentation (SSF),

simultaneous saccharification and co-fermentation (SSCF), and consolidated bioprocessing

(CBP) (Bisaria & Ghose, 1981; Boyle et al., 1997; McAloon et al., 2000; Lynd et al., 2005;

Olson et al., 2012). Sugars produces in cellulose hydrolysis or saccharification are

simultaneously fermented into ethanol in the SSF which greatly reduces inhibition to hydrolysis

(Boyle et al., 1997; Krishna et al., 1998). Among these process SSF and SSCF are preferred

because both unit operation can be completed in the same tank, reducing the cost (Wright et al.,

1988; Olofsson et al., 2008; Balan et al., 2012). The conventional alcoholic fermentation is a

typical inhibitory process, with cells growth rate affected by cellular, substrate and product

concentration (Rivera et al., 2006) and must be maintained between 7–10°GL to prevent

inhibitory effects (Junqueira et al., 2009a,b), beyond that reduce yield and productivity of the

process (Silva et al., 1999). Vacuum extractive fermentation process allows simultaneous

removal of produced ethanol from the fermentor, yields a highly concentrated wine, as a result

reduces the amount of vinasse and energy consumption in fermentation and the subsequent

distillation steps (Silva et al., 1999; Junqueira et al., 2009a,b).

2.1.3. Distillation and purification

Distillation is a process of separating a mixture of liquids based on the difference in boiling

point temperature of components in a liquid mixture. Azeotropes are formed when the mixture

has a vapor pressure lower than that of either binary component; such as ethanol boils at 78.5°C

and water at 100°C and the azeotrope at 78.5°C. This technology as such cannot be used to

separate azeotropes because they have the same composition in the vapour and liquid phase.

Usually, azeotrops are separated based on the pressure-swing or extractive distillation using an

additive, which are noted to be energy intensive. Pervaporation process significantly reduces the

investment and operating cost (Sommer et al., 2002; Van Hoof et al., 2004; Yuan et al., 2011).

Retrofit-extractive distillation achieved significant energy and cost savings compared to the

conventional extractive distillation process (Duc Long & Lee, 2013). Li uid-li uid e traction

and e tractive distillation saves both energy and cost ( vil s art ne , 20 )

14

The hydrous ethanol produced in the distillation undergoes purification stages to achieve

anhydrous ethanol (99.5%). The industrial separation methods are azeotropic distillation with

cyclohexane/benzene/pentane, extractive distillation with monoethyleneglycol (MEG)/gasoline/

glycerol/salt-solvent mixtures, and adsorption with molecular sieves and processes that include

the use of pervaporation membranes (Lynn & Hanson, 1986; Ulrich & Pavel, 1988; Pinto et al.,

2000; Fu, 2004a,b; Gil et al., 2008; Dias et al., 2009; Gil et al., 2012). Some of these methods are

no longer in use due to the high operating costs, operative problems and high energy

consumption (Gil et al., 2008). MEG is a fossil and toxic solvent. Bioglycerol as a byproduct of

biodiesel production process reported to be cheap and is not harmful to humans or to the

environment, a suitable agent for the separation of ethanol–water mixtures (Lee & Pahl, 1985).

Bioglycerol can be safely used to produce anhydrous ethanol for use in food or pharmaceutical

industries (Dias et al., 2009). Zacchi and Axelsson (1989) noted that energy consumption in the

distillation process significantly depend on the ethanol concentration in the feed material to a

certain concentration and seems have no effect beyond 7.5% (wt).

Internally heated integrated distillation column (HIDiC) is noted to be a promising option

to reduce energy consumption in the distillation process and reduce 60% of distillation energy

compared with the conventional column system without raising capital cost (Olujic et al., 2003;

Nakaiwa & Ohmori, 2009). Self-heat recuperation technology in azeotropic distillation process

also reduced distillation energy consumption compared with the conventional azeotropic

distillation (Kansha et al., 2009). Membrane-assisted vapor stripping process reduced at least

43% distillation energy re uirement Energy re uirement varied from 2 5−8 9 J/kg of fuel

grade depending on the ethanol concentration in the solution (Vane & Alvarez, 2008). Table 2.2

represents a brief summary of distillation parameters, energy consumption and cost.

2.1.4. Waste management

The stillage from the distillation column is sent to multieffect evaporator for partial

dewatering. The main residual solid from lignocellulosic ethanol industry is lignin, and amount

dependent of the feedstock, and pretreatment conditions. Usually, the waste stream is used to be

separated into three feed streams: solids (lignin), biogas and syrup high in solids (Wooley et al.,

1999; McAloon et al., 2000; Aden et al., 2002; Contreras et al., 2009; Greer, 2011). Lignin can

be used to produce coproducts, such as high-octane hydrocarbon fuel additives and replace

phenol in phenol formaldehyde resins (Hamelinck et al., 2005). Anaerobic digestion of the

15

wastewater produces a biogas high in methane. Biogas production is noted to be dependent on

the COD level in the effluent entering the digester. The COD level is estimated to be 4800 mg/l

(Shafiei et al., 2011). De Paoli et al. (2011) noted that biogas production is also dependent on the

pretreatment conditions of biomass (178–554 LN/kg volatile solids). The author also argued that

ethanol residues can contribute about 5% of total energy consumption in Brazil. Methane

production rate is reported to be 0.31–0.66 m3/kg COD (Barta et al., 2010a; Greer, 2011; Gyenge

et al., 2013). Biogas production from fermentation wort is reported to be 678 mL/kg slurry

(Ofoefule et al., 2013). Burning these byproducts streams to generate either the heat or electricity

required in the processes reduces not only the waste management costs, but may also lead to

profit.

Table 2.2 Brief summary of energy consumption in distillation processes

Methods Ethanol

concentration

Energy

consumption

Cost Remarks Reference

Membrane-assisted

vapor stripping

1 wt%

5 wt%

8.9 MJ/kg

2.5 MJ/kg

$0.098/L

$0.042/L

Simulation Vane & Alvarez,

2008

Hybrid system

(combining with

pervaporation)

- - €0.130/kg Experiment

&

simulation

Van Hoof et al.,

2004

Vacuum extractive

fermentation &

distillation

>40 ºGL 7.525 MJ/kg

hydrous

ethanol

- Simulation Junqueira, et al.,

2009a

Extractive

distillation

(purification)

93 wt% 1.085MJ/kg - Simulation Dias et al., 2009

Conventional

distillation

5.0 wt%

7.5 wt%

7.500 MJ/L

5.969 MJ/L

- Simulation Zacchi &

Axelsson, 1989

The concentrated solids in syrup can be sent to the burner minimizing the load to the

wastewater treatment. Utilization of excess solid residue for heat and power production had a

considerable effect on the process economics (McAloon et al., 2000; Sassner & Zacchi, 2008;

16

Sassner et al., 2008; Dutta et al., 2010a; Klein-Marcuschamer et al., 2010). Residual solid

burning in the boiler produces ash as a waste. Ash recovered form biomass power industries can

be dispersed on the field as fertilizer (Daugherty, 2001; Mani et al., 2010; Mandre et al., 2010).

On the other hand ash disposal cost is reported to be $0.157–$2.2/t (Frederick Jr et al., 2008;

Mani et al., 2010) depending on the transportation distance. Ash from biomass combustion in

modern boilers or stoves primarily consists of non-combustible mineral constituents such as

oxides or salts. Both fly ash and bottom ash have an economic value and could be used in

cement, brick manufacturing, construction of roads and embankments (IRRI, 2007). Bottom ash

from biomass can be used either as fertilizer or manufacturing of lightweight blocks (Pitman,

2006; BEC, 2011; Pérez-Villarejo et al., 2012). Nutrients and heavy metals leaching are reported

in case of wood ash fertilization on peatland (Piirainen, 2005). On the other hand, fly ash is noted

to be not only useful in the cement industry but also improve thermal and acoustic properties as

well as the durability of concrete (Youngquist et al., 1996; Bhatty & Miller, 2003; Naik et al.,

2006; Salas et al., 2009; Chatveera & Lertwattanaruk, 2011).

2.2. Life cycle assessment (LCA) of ethanol produced by biochemical conversion process

The enzymatic processes reduce global warming, acidification, eutrophication,

photochemical ozone formation and energy use compared to traditional process, and LCA

reveals in depth environmental properties of the processes have been studied (Jegannathan &

Nielsen, 2013).

2.2.1. LCA of ethanol produced from agri-residues

Agri-residues are identified as the most abundant feedstock for lignocellulosic ethanol,

have gained increasing attention as a renewable energy source. The life cycle (LC) of

lignocellulosic ethanol extensively studied using the LCA methodology. Those studies noted that

lignocellulosic ethanol can improve energy security and contributes significantly to abate GHG

emissions (Fleming et al., 2006; González-García et al., 2009; Mabee & Saddler, 2010; Spatari et

al., 2005; Vliet et al., 2009; Williams et al., 2009a; Wyman, 1994). The reduction in GHG

emission is reported to be dependent on feedstock, conversion technology, utilization of

coproducts and allocation methods (Spatari et al., 2005; Tilman et al., 2009; Luo et al., 2009a,b;

Kaufman et al., 2010; Roy et al., 2012a,b). In contrast, greater GHG emissions are also reported

compared to fossil alternative when the energy used to feed the biomass conversion process

comes from carbon-intensive fossil sources (Fu et al., 2003).

17

Pourhashem et al. (2013) studied the alternate use of ethanol byproduct (lignin) and noted

that the lowest emission from the LC of ethanol in case the lignin is used as a land amendment to

replace the soil organic carbon followed by the replacement for coal and the highest for onsite

electricity production. Enzyme is an important contributor to the net GHG emission of the LC of

ethanol. The GHG emission from enzyme production is reported to be 0.26–0.40 kg CO2 e/L for

onsite and offsite production (Hong et al., 2013).

Although, biofuel production is beneficial to reduce fossil energy consumption and the

global warming potential if biomass from cropping systems are utilized. The utilization of

biomass for biofuels would also tend to increase acidification and eutrophication, primarily

because significant nitrogen and phosphorus related environmental burdens are released from the

soil during cultivation, if additional measures are not in place such as planting the cover crops

(Kim & Dale, 2005). The lignocellulosic residues of banana fruit are also reported to be

energetically feasible for ethanol production (Velásquez-Arredondo et al., 2010). Ethanol

produced from grass clippings, corn stalks and other plants using future techniques is noted to be

beneficial (Stolman, 2005). Bioenergy production from sugarcane bagasse revealed that the

cogeneration option results in lower energy-related emissions (i.e. lower global warming,

acidification and eutrophication potentials), however the fuel ethanol option is preferred in terms

of resource conservation (since it is assumed to replace oil not coal), and scores better in terms of

human and eco-toxicity if lead-bearing oxygenates are replaced (Botha and Blottnitz, 2006).

Ethanol produced from stover avoids 86–113% of GHG emissions, if E85 is used in fuel

flexible vehicles instead of gasoline (Dutta et al., 2010a; Stoeglehner & Narodoslawsky, 2009).

The carbon neutrality of biomass and the use of residues may play an important role to abate

more than 100% GHG emissions. Although fossil energy consumption is 102% lower and

whereas hydrocarbon ozone precursors are reduced, emissions of CO, NOx, and SOx increased

(Stoeglehner & Narodoslawsky, 2009). Emissions from stover ethanol are noted to be 65% lower

for the near-term scenario (2010) due to the sharing of emissions with corn grains (Spatari et al.,

2005). Emissions would be about 25–35% lower than the near-term scenario if the mid-term

scenario (2020) is considered. The use of corn stover as a feedstock results in lower GHG

emissions relative to conventional corn-grain ethanol (Kim & Dale, 2005,Williams et al., 2009a),

although this reduction is dependent on the allocation method used (Kaufman et al., 2010; Kim

& Dale, 2005; Williams et al., 2009a). The carbon intensity of ethanol is reportedly 10–44% that

18

of gasoline (Kaufman et al., 2010). The equipment capable of performing a single-pass harvest

of stover becomes commercially available which help abating GHG emissions from stover

harvest (Stoeglehner & Narodoslawsky, 2009). It is worthy to note that biomass combustions are

assumed to be carbon neutral in all these studies.

Ozone layer and abiotic resources depletion decrease if gasoline is replaced by stover

ethanol fuels (E10 and E85), which is not relevant to the allocation method. However, other

impacts are larger except the global warming potential (GWP). The GWP reduces when

mass/energy allocation is applied, but increases in the case of economic allocation (Luo et al.,

2009a). In LCA studies, system boundaries cause a considerable variation since they not only

vary according to start and end points (e.g. well to tank and well to wheel) but also over space

and time in a way that can significantly affect energy and GHG balances (Botha & Blottnitz,

2006). The GWP of the lignocellulosic ethanol plant is noted as significantly (two fold) worse

than that of the gasoline refinery, but its improved eco-efficiencies make it superior in terms of

abiotic and ozone layer depletion potentials (Luo et al., 2010). In contrast, GHG savings from

ethanol and Ethyl Tertiary Butyl Ether (ETBE) blending are reported to be positive, even the

modification of the refinery sector is included (Croezen & Kampman, 2009).

Lignocellulosic ethanol produced by an enzymatic hydrolysis process shows that El0

improves the environmental performance in GHG emissions if the energy used in the steam

generation process is derived from biomass instead of fossil fuel for pretreatment of feedstock,

but has inferior performances in terms of acidification, eutrophication, winter smog, summer

smog, carcinogenic substances, heavy metals, ozone layer depletion and solid waste (Fu et al.,

2003). Bio-based products and fuels from straw may also be associated with environmental

disadvantages due to land use or water eutrophication (Uihlein & Schebek, 2009). The

environmental impacts predominantly result from the provision of hydrochloric acid and, to a

smaller extent, from the provision of process heat. The optional acid and heat recoveries yield

environmental impacts that are approximately 41% lower than those of the fossil counterparts.

The net energy ratios (output energy divided by input energy from fossil fuels) of ethanol

production systems from high yield rice plants are also noted to be positive, where whole rice

plants are used (Koga & Tajima, 2011; Saga et al., 2010). Whole rice plant-based ethanol

production systems improved energy efficiency and reduce GHG emission, because straw

removals notably mitigate CH4 emissions from the paddy field. The use of straw for energy

19

(CHP: combined heat and power) in ethanol production from wheat grains has significant

benefits, but the eutrophication and atmospheric acidification impact categories were slightly

unfavorable in some cases (Gabrielle & Gagnaire, 2008). The use of agricultural residues in a

biorefinery saves GHG (50%) and reduces demand of fossil fuels (80%), where the best

management practices are employed (Cherubini & Ulgati, 2010). However, biomass harvest rates

must be carefully established to avoid any negative consequences on stability and productivity of

land.

Although agri-residues are identified as abundant biomass resources, there debate is prevail

regarding the actual amounts of residues could be removed from arable soils without any loss of

quality, as well as the potential trade-offs in the overall energy chain compared to the use of

fossil energy. The removal of wheat straw had little influence on environmental emissions from

the field, and incorporating it in soil resulted in sequestration of only 5–10% of its C in the long

term (Gabrielle & Gagnaire, 2008). It is noted that a certain portion of crop residues can be

removed to produce ethanol without degrading the soil quality, which is dependent on the

season, location, tillage and soil types (Nelson, 2002; Reijnders, 2008). Selecting residues that

contain relatively high levels of available cellulose and hemicellulose for removal or returning

suitable crop residues that are rich in refractory compounds may increase the scope for removal

of crop residues for ethanol production (Reijnders, 2008). Sheehan et al. (2002) argued that up to

60% of the stover can be collected and converted into ethanol. However, Blanco-Canqui and Lal

(2009) suggested that stover removal rate should be as low as 25%, beyond which soil fertility

and structural stability would be negatively affected. In current agricultural practice, only 28%

of the stover is harvested, and the rest is left on the field for soil fertility (Graham et al., 2007).

Although biomass is recognised to be the most promising feedstock considering its great

availability and low cost, the large-scale commercial production of lignocellulosic ethanol has

yet to be implemented (Balat, 2011; Jensen et al., 2010), due to challenges and obstacles (cost,

technology and environmental issues) needing to be overcome for the commercial production of

lignocellulosic ethanol (Hatti-Kaull et al., 2007; Tan et al., 2008).

2.2.2. L CA of ethanol from energy crops, woody biomass and forest residues

The contribution of energy crops to total biomass energy is set to grow in the near future.

The majority of LCA studies noted that bioenergy from energy crops reduces GWP and fossil

energy consumption if the most common transportation biofuels are used to replace their

20

counterpart (Blottnitz & Curran, 2007; Cherubini & Jungmeier, 2010; Guo et al., 2010; Schmer

et al., 2008; Sims et al., 2006) in all but a few studies (Pimente & Patzek, 2005; Searchinger et

al., 2008). In contrast, acidification and eutrophication increased (Cherubini & Jungmeier, 2010),

and including land use change effects in GHG balances, biofuels substituting fossil fuels may

lead to increased negative impacts (Searchinger et al., 2008). Pimentel & Patzek (2005) reported

that ethanol production from switchgrass and woody biomass requires 50 and 57% more fossil

energy than the ethanol fuel produced, respectively. Cherubini et al. (2009) argued that these

limitations could be partially overcome by developing second generation biofuels, derived from

various lignocellulosic non-food crops and residues.

The estimated GHG emissions from cellulosic ethanol were 94% lower than those of

gasoline, while genetic and agronomical improvement may further enhance the energy

sustainability and biofuel yield of switchgrass (Schmer et al., 2008). Switchgrass fields are noted

to be near-GHG neutral depending on the agricultural inputs (mainly N fertilization) and biomass

yields. The use of ligneous biomass in cellulosic biorefinery is the main key to abetting GHG

emissions rather than biofuels from annual crops, where processing energy is derived from fossil

fuels (Farrell et al., 2006). Spatari et al. (2005) noted that emissions from energy crop

(switchgrass) ethanol were 57% in the case of the near-term scenario (2010) and lower for an

E85-fueled automobile compared to gasoline, on a CO2 equivalent per kilometer basis. It could

be 25–35% lower than those of the near-term scenario if the mid-term scenario (2020) were

considered. Net energy gains from each hectare of biofuels are affected by the crop yield,

conversion rate, and energy inputs to produce, deliver and process feedstock. The yearly net

energy gain is noted to be greater than low-input switchgrass grown in small plots (Schmer et al.,

2008).

Switchgrass is reported to be effective at storing soil organic carbon (SOC), not just near

the soil surface, but also at depths below 30 cm where carbon is less susceptible to

mineralization and loss (Liebig et al., 2005; Schmer et al., 2008; Wu al., 2008). Haney et al.

(2010) noted that perennial grass systems had higher SOC and water extractable organic C

(WEOC) than the annual corn system. Among perennial grass systems, switchgrass had the

lowest SOC and WEOC. Nitrogen leaching is reported to be less for switchgrass than corn, but

greater than in alfalfa–corn cropping systems (Vadas et al., 2008). Monti et al. (2009) analyzed

the energy crops (switchgrass, giant reed and cynara) production in terms of energy and hectares,

21

and compared them with conventional wheat and maize rotation. The authors concluded that on

average, 50% lower environmental impacts can be achieved by substituting conventional rotation

with perennial crops. The benefits are reportedly dependent on biomass yield and the preference

to a specific energy crop strongly depends on weighting sets that may change considerably in

terms of space and time. In contrast, the switchgrass biochar-pyrolysis system is noted to be a net

GHG emitter (+36 kg-CO2e/t feedstock) (Roberts et al., 2010).

Tilman et al. (2006) reported that biofuels derived from low-input high-diversity (LIHD)

mixtures of native grassland perennials can provide more usable energy, greater GHG emission

abatement, and less agri-chemical pollution than that of corn grain ethanol. LIHD biofuels are

carbon negative because net ecosystem carbon dioxide sequestration (4.4 t/ha/year of carbon

dioxide in soil and roots) is reported to be greater that the release during biofuel production (0.32

t/ha/year). LIHD biofuels can also be produced on agriculturally degraded lands and thus neither

displaces food production nor cause any loss of biodiversity via habitat destruction. The

environmental performance of ethanol produced from poplar biomass considering three ethanol

applications (E10, E85 and E100) revealed that the impact potentials per kilometer driven by a

mid-size passenger car, may help ease the exacerbation of global warming, and depletion of

abiotic resources and the ozone layer by up to 62, 72 and 36%, respectively. However,

acidification and eutrophication would intensify.

The economic and environmental aspects of high yield cropping systems are not

necessarily conflicting, whereas under or over supply of nitrogen fertilizers leads to a decline in

resource use efficiency (Brentrup et al., 2004a,b; Haas et al., 2001). Pedersen et al. (2005) reveal

that in the USA, some long-term breeding of switchgrass has achieved large yields and may

begin to contribute significantly to biofuel production. Genetically modified (GM) herbicide

tolerant energy crops (sugar beet) are reported to be less harmful to the environment and human

health than growing conventional crops, largely due to lower emissions from herbicide

manufacture, transport and field operations (Bennett et al., 2004). These studies indicate that the

social and environmental co-benefits, including carbon sequestration opportunities, will be

drivers of future energy cropping uptake, although they must also be ecologically sustainable,

environmentally acceptable and economically competitive with fossil fuels (Sims et al., 2006).

22

2.2.3. Land, water and other approaches in LCA of ethanol

The global population continues to grow geometrically, exerting great pressure on arable

land, water, energy and biological resources to provide an adequate food supply while

maintaining the ecosystem. The availability of land on which to grow biofuel crops without

affecting food production or GHG emissions from land conversion is limited, hence land use

efficiency should be maximized to achieve climate change goals. Although lignocellulosic

ethanol supply chains are considered feasible for making GHG savings relative to gasoline, an

important caveat is that if lignocellulosic ethanol production uses feedstock that cause indirect

land-use change, or other resulting significant impacts, any benefit may be greatly offset (Slade

et al., 2009). The effects of land use changes were noted as having a significant influence on the

final GHG balance (about 50%) (Cherubini & Ulgiati, 2010). Jegannathan & Nielsen (2013)

reported that land use savings can be achieved in industries where enzymatic processes save

agricultural raw materials. On the other hand it becomes a trade-off where only fossil fuels

and/or inorganic chemicals are conserved.

It is noted that converting croplands or grasslands to produce energy crops may actually

lead to an increase rather than fall in GHG emissions (Fargione et al., 2008; Searchinger et al.,

2008). The carbon debt (CO2 emission) increases 17–420 folds compared to the gasoline if

rainforests, peatlands, savannas, or grasslands are converted to produce food crop–based

biofuels. In contrast, a little or no carbon debt is resulted if biofuels are produced from waste

biomass or biomass grown on degraded and abandoned agricultural lands (Fargione et al., 2008).

Brandão et al. (2011) studied the different land use systems for energy crops and noted that

miscanthus is the optimal choice in terms of GHG emissions and soil quality compared to oilseed

rape, short-rotation coppice willow and forest residues, but performed worse in the categories of

acidification and eutrophication, while oilseed rape showed the worst performance across all

categories. Stephenson et al. (2010) revealed that if willows are grown on idle arable land in the

UK, or in Eastern Europe, and imported as wood chips into the UK to produce ethanol, this saves

about 70–90% of GHG emissions compared to fossil-derived gasoline on an energy basis. In

contrast, Searchinger et al. (2008) estimated GHG emissions from land-use changes by using a

global agricultural model and reported that corn-based ethanol, instead of achieving 20%

savings, nearly doubles GHG emissions over 30 years and increases GHGs for 167 years.

Biofuels from switchgrass also increase emissions by 50%, if grown on U.S. corn lands. The

23

bioelectricity pathway outperforms the cellulosic ethanol across a range of feedstock, conversion

technologies, and vehicle classes; producing 81% more transportation kilometers and 108% more

emission offsets per unit area of cropland (Campbell et al., 2009).

Stoeglehner et al. (2009) noted that biofuels will only be able to contribute to a certain–

may be relatively limited-extent, to an overall sustainable energy supply that will vary widely

between regions, and the sustainability of biofuel production depends on the amount of land

available. Direct land use changes, the choice of calculation methods, utilization of coproducts

and the technical design of production systems affect the GHG balances and eutrophication for

all biofuels (Börjesson & Tufvesson, 2010). The enhanced demand for biofuel crops under the

EU Biofuel Directive has a strong impact on agriculture at a global and European level, while the

incentive to increase production in the EU tends to increase land prices and farm income there

and in other regions (Banse et al., 2011).

Several competing factors have need to be balanced, such as changes in land use (clearing

tropical forests or using peatlands for crop cultivation) to negate any of the intended future

climate benefits, and impacts on biodiversity. Also, developments in the agricultural sector for

food and non-food crops will have important implications for water usage and its availability.

The opportunity costs and rebound effects of land use changes must be addressed while

considering any decision to assign land to biofuel feedstock (Pickett et al., 2008). Although

biomass residues have been identified as a potential feedstock for bioenergy, the global mature

forest area will decrease by 24% between 1990 and 2100, due to both population growth and

wood biomass demand in developing regions, and may even disappear by 2100 in some

developing regions, such as Centrally Planned Asia, Middle East and North Africa, and South

Asia (Yamamoto et al., 2001). Consequently, the sustainability of biofuels depends on the

selection of land on which feedstock are grown.

Reith et al. (2002) reported that the gross water consumption in the lignocellulosic ethanol

production processes is 28–54 liters per liter of ethanol. The high water consumption results

from the process water used in the Ca(OH)2 pretreatment, washing of solids prior to enzymatic

hydrolysis. In contrast, water consumption is noted to be only 0.3 L per liter of ethanol produced

from agri-residues (corn stover or wheat straw), because the water requirement for crop

production was attributed only to grains (Singh & Kumar, 2010). Biochemical or

thermochemical conversion process of biomass into ethanol is expected to reduce GHG and air

24

pollutant emissions, but involve similar or potentially greater water demands and solid waste

streams than conventional ethanol biorefineries. Despite current expectations, significant

uncertainty remains regarding how well next-generation biofuels will fare in terms of different

environmental and sustainability factors when derived on a commercial scale in the U.S.

(Williams et al., 2009a). Although ethanol production consumes huge amounts of water, its

impact on water resources is seldom included. The land to man ratio in developing countries is

not as favorable as in developed countries, with far scarcer land resources creating serious

problems in land resources management and possibly resulting in land degradation in such

developing countries. The use of bioenergy also involves environmental challenges, for instance

increased mono-cropping practices and greater fertilizer and pesticide use, which may jeopardize

water and soil quality. Perhaps the main concern over land use change is the risk of large areas of

natural forests and grasslands being converted to energy crop production, which not only

threaten biodiversity and ecosystems, but also result in a possible increase in GHG emissions.

2.3. Ethanol production via gasification process

Thermochemical conversion technologies, including gasification and pyrolysis, heat

biomass feedstock under low oxygen conditions to produce synthesis gas, or ‘syngas,’ which can

be converted into various biofuels and biochemicals efficiently (Phillips et al., 2007; Brown,

2007; Henstra et al., 2007; Weber et al., 2010). The feedstock flexibility of gasification process

gives advantages over other ethanol production processes from biomass. The bacterial

fermentation noted to have advantages over catalytic conversion. The microbes are reported to be

less sensitive to syngas impurities like sulfur and normally produce specific alcohols instead of

mixtures (Munasinghe & Khanal 2010). The thermochemical processes are more effective

especially in the case of plants with a high content of lignin (Möller et al., 2006).

2.3.1. Gasification

Gasification is reported to be very effective at converting non-carbohydrate biomass

fractions and all other components of biomass into syngas with nearly equal efficiency and

effectiveness (Phillips et al., 2007; Brown, 2007; Henstra et al., 2007; Weber et al., 2010; Aden

et al., 2002; Wang & Yan, 2008; Pereira et al., 2012). Usually, feedstock is heated up to 700–

1000°C in a gasification process where all components of feedstock are decomposed into syngas

primarily containing H2, CO, CO2, CH4, with very few residues (tar and ash). The syngas

produced under anaerobic conditions is composed of 15–30% H2, 10–65% CO, 1–20% CO2, 0–

25

8% CH4, and trace amounts of other gases depending on the type of feedstock (Wei et al., 2009;

He & Zhang, 2011); however, with air gasification, the syngas may contain 40–50% N2 (Hu et

al., 2007; Wei et al., 2006; Rao et al., 2004; Eriksson et al., 2004; Datar et al., 2004; Clausen and

Gaddy, 1993). The composition and quality of syngas from biomass are dependent on the type of

gasifier, feedstock, gasifying agent, steam to biomass ratio, temperature, pressure, catalyst, etc.

(He & Zhang, 2011; Carpenter et al., 2010). Gasification affects the calorific value of syngas.

The increasing gasification pressure is noted to be economically more feasible than increasing

the syngas pressure in downstream equipment (Passandideh-Fard et al., 2008). Low temperature

operation results in higher selectivity (of Rh-Mn/SiO2) to ethanol and lower methane formation

(Hu et al., 2007).

Gasifiers are categorized based on the type of bed and type of flow. Bed type can be fixed-

bed or fluidized bed and the flow types can be downdraft, updraft or cross-flow gasifiers.

Fluidized bed gasifiers can be either bubbling bed or circulating fluidized bed. Downdraft

gasifiers produce lower amounts of tar than do updraft (Kumar et al., 2009). Various bed

materials (silica, alumina) and catalyst have also been used, which improve heat transfer and

conversion rate of biomass in gasification processes (Acharya et al., 2009; Calvo et al., 2012).

Low temperature operation results in higher selectivity (of Rh-Mn/SiO2) to ethanol and lower

methane formation (Hu et al., 2007). Increasing gasification pressure is noted to be economically

more feasible than increasing the syngas pressure in downstream equipment (Passandideh-Fard

et al., 2008).

Tars can deactivate catalysts/sorbents used for reforming and gas cleaning, are considered

to be the major bottleneck of industrial biomass gasification (Bain et al., 2005). Tars formation is

dependent on the type of gasifier, and gasification conditions. Tar concentration is reported to be

1–15, 20–100 and 0.01–1.5g/n-m3 for turbulent-bed, fixed-bed updraft and fixed-bed downdraft

gasifier, respectively (Stevens, 2001). Catalytic tar cleaning process needs no additional energy

input, efficiency and heating value losses are also kept at a minimum, and generates no tarry

waste streams that need to be disposed of or recycled to the gasifier (Bridgwater, 1994, 1995;

Simell et al., 1996). Solid oxide fuel cells (SOFCs) with Ni/GDC anodes have been successfully

tested with biosyngas cleaning (Aravind & de Jong, 2012). A catalytic fluidized-bed steam

reformer can be used to convert tars and hydrocarbons into syngas (Phillip, 2007).

26

Both the fixed-bed gasifier and fluidized-bed gasifier concepts are seen to be suitable for

small to medium scale of thermal input operation (Aravind & de Jong, 2012). The fluidized bed

gasifier requires a catalytic reformer at the downstream to clean the syngas (Bessou et al., 2011).

Although entrained flow gasifier can produces syngas without a reformer, require energy

intensive pretreatement (torrefaction/pyrolysis) of biomass in order to reach a sufficient

conversion rate (Bessou et al., 2011). It is also noted that the entrained flow reactors are not

easily downscaled efficiently and need much larger capacities (Aravind & de Jong, 2012).

Chemical looping gasification (CLG) process is noted to be able to produce relatively pure

product gas (nearly 0% CO2) with in situ CO2 capturing in a circulating fluidized-bed (CFB)

steam gasifier using CaO as the sorbent (Acharya, 2011; Acharya et al., 2009). This system

consists of a bubbling fluidized bed (BFB), CFB and a cyclone (A-2-1). BFB and CFB have

worked as a gasifier and regenerator, respectively. The cyclone was used to capture the solids.

The regenerator is connected to the gasifier through the cyclone. The riser section of the CFB

acted as a regenerator where the catalyst, CaCO3 converted into CaO and CO2 by the application

of external heat typically at about 900°C. The product CaO and CO2 is separated in the cyclone,

and CaO fed to the loopseal (gasifier) through the standpipe, and CO2 is collected and supplied

to the riser (regenerator) after filtration for continuous fluidization. The loop-seal acted as a BFB

gasification system where biomass and steam were supplied and CaO fed from the cyclone. The

loopseal is operated at 650°C where gasification with in-process CO2 captured. The CaO is

converted to CaCO3 by absorbing CO2 produced during gasification process was fed back to the

regenerator to continue the cycle. Moreover, the hot CaO delivered to the gasifier/carbonator

provided additional heat for the gasifier/carbonator vessel. In addition, heat releases by the

exothermic carbonization reactions can supply most of the heat required by the endothermic

gasification reactions (Acharya, 2011).

Thermodynamically, biosyngas can be converted into ethanol at 350°C and 30 bar (Spivey

& Egbebi, 2007). The yield of ethanol depends upon the composition of biosyngas.

Theoretically, one-third of the carbon from CO can be converted into ethanol in water gas shift

reaction (Eq. 2.1), however two-thirds of the carbon from CO can be converted in an equimolar

mixture of H2 and CO (Eq. 2.3) (Munashinge & Khanal, 2010). The conversion rate of CO can

be as high as 40%, yielding an alcohol mixture containing 70% ethanol (Cotter, 2007; Morrison,

2004). It is important to note that CO2 can also be used by acetogens if H2 is present (Eq. 2.2).

27

The overall stoichiometric reaction for alcohol synthesis is summarized in the following equation

(Eq. 2.4). In a two-stage syngas fermentation process, stoichiometric evaluation revealed that the

carbon and hydrogen recovery from the supplied carbon monoxide and hydrogen into ethanol is

28% and 74%, respectively (Richter et al., 2013) where the growth stage is operated at 5.5 pH

and the production stage had lower pH than that of growth stage. Figure 2.1 depicts the stages of

ethanol production process via biomass gasification and catalytic and microbial fermentation.

6CO + 3H2O → C2H5OH+ 4CO2 . . . . . (Eq. 2.1)

6H2 + 2CO2 →C2H5OH + 3H2O . . . . . (Eq. 2.2)

6CO + 6H2 → 2C2H5OH + 2CO2 (Eq. 3)

nCO + 2nH2 →CnH2n+1OH + (n-1) H2O . . . . (Eq. 4)

Biosynthesis

Feedstock B

iosy

ngas

Pre - treatment (Drying/size reduction)

Gasification

Gas cleanup/

Purification Coproducts/ byproducts

Chemical synthesis

conditioning

Catalyst Bacteria

Source: Roy & Dutta, 2013

Figure 2.1Schematic diagram of ethanol production process from syngas

28

2.3.2. Gas cleanup

Gas cleanup steps are crucial for preventing both catalyst fouling and poisoning in the

subsequent alcohol synthesis steps (Gonzalez et al., 2012). Gas cleanup and conditioning remove

the problematic tars, chars, particulate matters and other contaminants which cause slagging and

downstream process inhibitions. Cyclones, adsorption columns, water or oil scrubbers and

various types of filters are some of the common syngas refining units (Munasinghe & Khanal,

2010; Pereira et al., 2012). Catalytic steam reforming or thermal cracking has also been used for

gas cleanup (Damartzis & Zabaniotou, 2011). Catalytic tar cleaning process needs no additional

energy input, efficiency and heating value losses are also kept at a minimum, and generates no

tarry waste streams that need to be disposed of or recycled to the gasifier (Bridgwater, 1994,

1995; Simell et al., 1996). Solid oxide fuel cells (SOFCs) with Ni/GDC anodes have been

successfully tested with biosyngas cleaning (Aravind & de Jong, 2012).

2.3.3. Syngas synthesis into ethanol

The clean biomass syngas (biosyngas) is then converted into ethanol either by catalytic

conversion or bacterial fermentation (biosynthesis).

2.3.3.1 Catalytic synthesis into ethanol

Both homogeneous and heterogeneous catalytic synthesis processes are used in the

conversion of syngas into ethanol. The homogeneous catalytic synthesis process produces

ethanol in large quantities, requires expensive catalysts, high operating pressure and tedious

separation process (for catalyst). On the other hand, the heterogeneous catalytic processes suffer

from low yield and poor selectivity (Subramani & Gangwal, 2008). The main catalyst groups are

natural catalyst (dolomite and olivine), alkali (KOH, Na2CO3, CaCO3, CsCO3, ZnCl2, NaCl etc.)

and nickel-based catalysts (Mohammed et al., 2011; Acharya et al., 2009). A series of reactions

take place (with syngas and catalyst) in the reactor yielding a mixture of alcohols that varies

depending on the catalyst and reaction temperature and pressure. Reaction rates are noted to be

quick and completed in seconds or minutes (Cotter, 2007; Morrison, 2004). The low specificity,

high operating temperature and pressure, high sensitivity to toxic gases are the major setback for

the chemical catalysts (Phillips et al., 1994; Vega et al., 1990; Worden et al., 1991).

2.3.3.2 Biosynthesis into ethanol

Bacterial fermentation has some advantages because microbes are less sensitive to syngas

impurities like sulfur and normally produce specific alcohols instead of mixtures (Henstra et al.,

29

2007; Munasinghe & Khanal, 2010; Bredwell et al., 1999; Kundiyana et al., 2010), but result in

poor mass transfer properties of gaseous substrates and lower ethanol yield (Munasinghe &

Khanal, 2010; Wei et al., 2009). The yield is also dependent on the partial pressure of CO and

the size of the fermentor (Kundiyana et al., 2010; Younesi et al., 2005; Hurst et al., 2005). In this

route, mesophilic microorganisms produces short-chain fatty acids and alcohols from CO and H2,

and hydrogen can be produced by carboxydotrophic hydrogenogenic bacteria which converts CO

and H2O to H2 and CO2 (Henstra et al., 2007). The use of Clostridium ljungdahlii in the

fermentation process of syngas improves mass transfer properties (BRI, 2008). Ethanol

production with the Clostridium ljungdahlii is usually a mixture of two processes that occur at

temperatures 37–39°C. The acetate production inhibits bacteria and lowers the pH that makes the

bacteria shift to produce solely ethanol. The ideal pH and ethanol concentration for Clostridium

ljungdahlii is noted to be 6 and 3% (greater strength is toxic for the bacteria) (Hensra et.al.

2007). Other microorganisms (Clostridium autoethanogenum, Acetobacterium woodii,

Clostridium carboxidivorans and Peptostreptococcus productus) are also used to ferment syngas

into liquid fuel (Henstra et al., 2007; Heiskanen et al., 2007; Rajagopalan et al., 2002). Nontoxic

surfactants and novel dispersion devices can enhance mass transfer (Worden et al., 1997), and

resulting in higher ethanol yield in this process (Clausen & Gaddy, 1993). Some of the biological

catalysts (Clostridium ljungdahlii, Clostridium autoethanogenum, Acetobacterium woodii,

Clostridium carboxidivorans and Peptostreptococcus productus) are able to ferment syngas into

liquid fuel more effectively than that of chemical catalysts (e.g., iron, copper or cobalt)

(Heiskanen et al., 2007; Henstra et al., 2007). A brief list of different types of bacteria identified

and used for biofuel production from syngas, and syngas fermentation parameters and ethanol

yield are reported in the appendix (A-2-2 & A-2-3).

2.3.3.3 LCA of ethanol produced from biomass syngas (biosyngas)

The growing concern about sustainability of ethanol produced from biomass syngas

prompted researchers to evaluating the LC of ethanol. The majority LCA studies on biosyngas to

ethanol noted that thermochemical conversion process reduces GHG emissions (Tonini &

Astrup, 2012; Mu et al., 2010; Jungmeier et al., 2007). Mu et al. (2010) studied the LC of ethanol

from biosyngas (from wood chips) produced by indirectly-heated dual fluidized bed gasification

technology and a modified Fischer-Tropsch catalyst molybdenum disulfide (MoS2) at

atmospheric pressure. The LC material and energy required were modeled by Aspen Plus

30

simulation software and then the LC inventory data of material and energy have been extracted

from SimaPro 7.1 and Ecoinvent 2.0 database to determine the environmental impacts of the

thermochemical conversion process [gasification-synthesis combination has been modeled based

on NREL report (Phillips et al., 2007) and different technological targets. The thermochemical

process noted to have negative net fossil fuel consumption, consequently negative net GHG

emissions (Mu et al., 2010; Grossmann & Martín, 2010) if plant is designed as energy self-

sustained and no mixed alcohol separation unit installed (i.e., some syngas is diverted for heat

and electricity generation). The co-product credit (mainly from electricity export) offsets the

energy consumption and GHG emissions in the LC of ethanol. Otherwise thermochemical

conversion process resulted in positive fossil fuel consumption and GHG emission

(approximately 0.4–0.5 kg CO2 e/L) (Mu et al., 2010).

The ethanol production processes were modeled with the mixed-integer nonlinear

programming (MINLP) and implemented in the GAMS modeling system to optimize the systems

(Grossmann & Martín, 2010; Martín & Grossmann, 2011). Ethanol yield was reported to be

greater in low pressure gasification process than that of high pressure thus reduce production

cost. Ethanol obtained from switchgrass via gasification (low pressure indirect gasification with

steam: 0.26–0.75 bar; high pressure direct gasification with steam and oxygen: 2.1 bar; operating

conditions are considered to be that of Phillips et al., 2007) and catalytic reaction/fermentation

consumes either no energy or even produces it (Grossmann & Martín, 2010; Martín &

Grossmann, 2011) because of the implementation of multieffect columns followed by heat

integration (hot and cold streams) in the process. MINLP approach is also used to evaluate the

LC (cradle to grave) of ethanol from wood chips considering the environmental and social

criteria. The synthesis gas is cleaned, cooled and fermented by using Clostridium ljungdahlii

microorganisms. This study concluded that ethanol production from forest wood waste is not

sustainable because of low ethanol yield and high emissions mainly from fermentation and waste

wood collection (Čuček & Kravanja, 2010).

The thermochemical conversion of forest residues into ethanol is energy self-sufficient

where forest residues are assumed to be a waste product (i.e., only chipping and loading are

attributed to forest residue-based E85) and no land use change is considered. The SimaPro

(v.7.2) and Ecoinvent database (v.2.1) were used to evaluate the LC of ethanol and addressed the

impact potentials of ethanol (E85) used in a mid-size passenger car and concluded that 43–57%

31

of GHG emission can be reduced compared to that of conventional gasoline if CO2 sequestration

is considered (i.e., approximately about 0.83 kg CO2 e/L) (Hsu et al., 2010). In an updated study,

ethanol produced by gasification of biomass reduced 65% of GHG emission with reference to

gasoline if only the ethanol portion is considered. The GHG emission is estimated to be about

0.21 kg CO2 e/L ethanol produced in that updated study (0.01 kg CO2 e/MJ of ethanol; 21 MJ/L

ethanol is assumed) (Hsu, 2012). Although the biorefinery system reduces CO2 and CH4, releases

more N2O emissions compared with a fossil fuel system thus has higher impacts in acidification

and eutrophication (Cherubini & Jungmeier, 2010; Kim & Dale, 2005).

He and Zhang (2011) has designed, simulated and optimized thermochemical conversion

process mainly by using Aspen Plus simulation software. The estimated energy requirement was

0.28 kWh/L ethanol. The syngas cleanup is an energy intensive process and one of the key issues

in commercialization of the thermochemical conversion technology. It is noted that bioenergy

system produced from forest residues (i.e., gasification and FT synthesis process) can save up to

88% GHG emission compared with a fossil fuel (Jungmeier et al., 2007). Corn stover ethanol

blends (E85) offer substantial energy savings (94–95%) relative to those fueled with regular

fossil gasoline and the cellulosic ethanol pathway mitigates 86–89% of GHG emissions (Wu et

al., 2006). Farrell et al. (2006) also noted that cellulosic ethanol (switchgrass) reduces GHG by

88%. In contrast, energy shortage is reported in the thermochemical conversion processes

(Zhang, 2008).

Tonini and Astrup (2012) evaluated the future energy systems for Denmark (fossil fuels;

rapeseed based biodiesel; Fischer–Tropsch based biodiesel for 2030 and 2050) using the LCA

methodology. The authors concluded that GHG emissions could be significantly reduced (from

68 to 17 Gg CO2 e/PJ) by increased use of wind and residual biomass resources as well as by

electrifying the transport sector. Increased share of wind power and replacement of fossil fuels

with domestically available biomasses, and reduction of energy demand led to GHGs emissions

savings in the future energy scenarios. It is also noted that introduction of energy crops for

biofuels and the use of biofuels for heavy terrestrial transportation were responsible for most

environmental impacts in the 2050 scenarios. Biodiesel production via Fischer–Tropsch is noted

to be comparable with fossil diesel only for the global warming. In contrast, fossil diesel noted to

be preferable over biodiesel for acidification, aquatic eutrophication and land occupation except

global warming. Also, land occupation increased to a range of 600–2100×106 m

2 per PJ

32

depending on the amounts and types of energy crops introduced (Tonini & Astrup, 2012)). A

comparative study of six different assessment scales and metric calculation techniques against

the common biomass demand scenario revealed that assessment scale and metric calculation

technique strongly influence the net GHG balance in woody biomass to energy conversion

process (Galik & Abt, 2012).

Although same gasification technology is used in most of the reviewed studies a significant

difference in GHG emission is observed. Table 2.3 represents a brief summary of the LCA

studies concerning system boundary, conversion processes, energy consumption and GHG

emissions. These studies indicate that the environmental impacts of ethanol are dependent on

research targets, system boundary and assumptions, feedstock and conversion technology. It

seems the environmental and social co-benefits, including land use and carbon sequestration

opportunities will be drivers of future lignocellulosic ethanol, although they must also be

ecologically sustainable, environmentally acceptable and economically competitive with fossil

fuels.

2.4 Life cycle cost analysis (LCCA)

The production cost of ethanol is dependent on both technical and economic parameters,

such as the subsidies/fit in tariff, cost of feedstocks, choice of feedstocks, energy consumption,

conversion technology and efficiency, and the value of coproducts (Ballerini et al., 1994;

Wyman, 1994; Aden et al., 2002; Mabee et al., 2006; Wu et al., 2006, 2008; Aden, 2008;

Nechodom et al., 2008; Mark et al., 2009; Balat, 2011; Haro et al., 2012; Phillips et al., 2007;

Dutta et al., 2010a; He & Zhang, 2011). The production cost of lignicellulosic ethanol is reported

to be considerably higher than the market price of gasoline (Ballerini et al., 1994; Wu et al.,

2008; Huang et al., 2009; Luo et al., 2010; Orikasa et al., 2009; Roy et al., 2012a,b; Roy &

Dutta., 2013). It also noted that the lignocellulosic ethanol technology is not yet mature enough

to be profitable on its own merits, consequently this industry is being subsidised to encourage its

production achieving the GHG emission abatement targets. It is reported that biochemical

conversion of biomass into ethanol requires lower investment compared to the thermochemical

conversion process. Total investment for 150 MWth pant is reported to be MM$281 for

biochemical conversion and MM$580 to MM$760 for thermochemical conversion process (Tunå

& Hulteberg, 2014). Table 2.4 depicts the tax credit applied in Canada. The tax credit to liquid

33

biofuels is also prevailed in many other countries. Federal tax credit in the USA is reported to be

$0.45/gallon (Murse, 2011).

Table 2.3 The LC GHG emission/energy consumption of ethanol produced by thermochemical

conversion process

Authors

System Feedstock, feed rate, cost & yield Gasifier

type

Catalyst/

Bacteria

GHG

emission

boundary Feedstock Rate,

t/d

Yield,

L/DT

kg CO2 e/L

(kWh/L)

He & Zhang,

2011

Field to

gate

Lignocellulo

sic biomass 2455 332

Indirect dual-

bed gasifier Co-Mo (0.46)

Mu et al.,

2010

Field to

wheel

Wood chips,

wheat straw,

waste paper,

corn stover

2000 270–359

Indirect-heating

fluidized bed

gasifier

Molybden

um

disulfide

(MoS2)

0.4–0.5

Hsu et al.,

2010

Field to

wheel

(E85)

Forest

residues 2000 387–417

Indirect

gasification - 0.83^

*Hsu, 2012

Field to

wheel

(E85)

Forest

residues 2000 387–417

Indirect

gasification - 0.21^^

Spath &

Dayton, 2003

Field to

wheel Wood chips 2000 0.27-0.39 Gasification Bacteria -

t/d: ton/day; DT: dry ton; ^estimated based on the emission data from the graph and heating value of E85; ^^0.01 kg

CO2 eq./MJ; *ethanol portion only (Source: Roy & Dutta, 2013).

2.4.1 Life cycle costing of ethanol produced by biochemical process

Vadas et al. (2008) noted that net energy production per hectare is greater for switchgrass

than that of alfalfa-corn cropping systems, but may not return the potential income to farmers

that alfalfa-corn could. The costs of cellulase and capital are the major expenses when producing

lignocellulosic ethanol (Reith et al., 2002), while industrial cellulase contributes about 40–55%

of the enzymatic cellulosic ethanol production cost. The estimated costs of producing ethanol

from lignocellulosic residues (verge grass, wheat milling residues and woody energy

crop/willow) are 0.75–0 99 €/L The authors noted that the cellulase cost (assumed 0 5 €/L)

would have to be reduced at least tenfold and the capital cost by 30% to achieve ethanol

production costs comparable to those of ethanol from starch crops.

34

Table 2.4 Tax credits on ethanol in various provinces in Canada

Province Tax credits, ¢/L Criteria for credit Duration

Alberta 9.0

No restriction on the source of

ethanol

Five years from the start of

an ethanol processing plant

British

Columbia 14.5

Ethanol must be produced in BC;

E85 to E100 and E5 to E25. -

Ontario 14.7

No restriction on the source of

ethanol Until 2010

Saska-

tchewan 15.0

Ethanol must be produced in

Saskatchewan 5 years

Quebec ~20 Ethanol must be produced in Quebec 1999-2012

Manitoba

20 (until August

2007)

15 (2007- 2010)

10 (2010- 2013)

Ethanol must be produced in

Manitoba

No specific duration

Federal 10 - No specific duration

Source: Olar et al., 2004

It is also noted that the production cost of cellulosic ethanol depends on feedstocks and

their composition as well as plant capacity. The estimated production cost varies from about

0.38–0.48 US$/L (plant size: feedstocks consumption is 2000 t/day) (Huang et al., 2009; Searcy

& Flynn, 2010). For the same plant capacity the production cost of ethanol from corn stover is

reported to be 0.71–0.87 US$/L dependent on the assumed scenarios (Dutta et al., 2010a). The

production cost is noted to be 0.56–0.77 US$/L depending on the feedstock and plant sizes

(Gnansounou & Dauriat, 2010). The simulated production cost of ethanol is reported to be 0.94–

1.20 US$/L which depends on the ethanol yield (Klein-Marcuschamer et al., 2010). The

production cost is also noted to be dependent on the market price of fuels (Roy et al., 2012a;

Zafeiriou et al., 2014).

The economic viability, GHG emission and economic performance of lignocellulosic

ethanol under extreme weather conditions are also reported to be dependent on the availability of

feedstock (weather condition) and the use of single or multiple feedstocks (Kou & Zhao, 2010.

Wingren et al. (2003) noted that the production cost is also dependent on enzymatic processes.

35

The cost of ethanol produced from softwood based on simultaneous saccharification and

fermentation (SSF), and separate hydrolysis and fermentation (SHF) are reported to be 0.57 and

0.63 US$/L, respectively. The main reason for SSF being lower was the lower capital cost and

the overall higher ethanol yield. Major economic improvement in both SSF and SHF could be

achieved by boosting income from the solid fuel coproduct, reducing energy consumption and

recycling process streams. A techno-economic evaluation of the spruce-to-ethanol process, based

on SO2-catalysed steam pretreatment followed by simultaneous saccharification and

fermentation with various process configurations, achieved an ethanol cost of about 0.38–0.50

€/L naerobic digestion of the stillage with biogas upgrading was a demonstrably favorable

option in terms of both energy efficiency and ethanol production cost (Barta et al., 2010a,b) and

the contribution of enzyme is reported to be 0.04–0 05 €/L(Barta et al., 2010b).

Ballerini et al. (1994) concluded that technical and economic optimization of the

pretreatment step, the total substitution of lactose by pentose hydrolysate as the main carbon

source for enzyme production, and the recycling of a fraction of the enzyme, the incorporation of

pentose in ethanol fermentation, and the utilization of by-products all reduce the production cost

of lignocellulosic ethanol. The authors also noted that since ethanol from biomass is tax-

exempted, it could compete with gasoline assuming a crude oil price of around US$50. In

contrast, it is noted that with current technology, the production cost of cellulosic ethanol (0.75

$/L) is almost double compared to the market price of oil (0.48 $/L) and much of the optimism

surrounding cellulosic ethanol has faded (Service, 2010). The externality (environmental and

health) cost of ethanol is also reported to be dependent on the feedstock (Kusiima & Powers,

2010).

Hamelinck et al. (2005) stated that the combined effect of higher hydrolysis-fermentation

efficiency (68%), lower specific capital investments, increased scale (5 times) and

lignocellulosic and woody biomass feedstock costs reduced to about 67% could slash ethanol

production costs to 59–40% of the current level in 10–20 years or more. The production cost is

reported to be slightly higher for wood-produced ethanol compared to that of switchgrass

(Pimentel & Patzek, 2005). The production costs of ethanol from energy crops vary widely due

to the complex characteristics of the resource, their site specificity, national policies, labor costs

and efficiency of the conversion technologies, but are expected to decline over time (Sims et al.,

2006) and it is noted to have clear socioeconomic benefits (Guo et al., 2010).

36

The coproduct revenue and utilization of the excess solid residue for heat and power

production had a considerable effect on the process economics, and improved ethanol yield and

reduced energy demand resulted in significant production cost reductions (0.41–0 50 €/L)

(Sassner & Zacchi, 2008; Sassner et al., 2008). Sassner et al. (2008) also concluded that the

utilization of pentose fractions for ethanol production helped achieve good process economy,

especially in the case of Salix or corn stover. It is also noted that ethanol produced from

softwood and sold as a low percentage blend with gasoline could ultimately be cost competitive

with gasoline without requiring subsidy, but that production from straw would generally be less

competitive (Slade et al., 2009). Despite the environmental benefits of ethanol produced from

coppice willow, its economic viability remains doubtful at present (Stephenson et al., 2010). The

author argued that the production cost could be reduced significantly if the willow were altered

by breeding to improve its suitability for hydrolysis and fermentation. A techno-economic

assessment of lignocellulosic ethanol also revealed that commercial success of pilot plants (0.3–

67 MW) remains pending, although cost-competitive ethanol can be produced with efficient

equipment, optimized operation, cost-effective syngas cleaning technology, inexpensive raw

material with low pretreatment cost, high performance catalysts, off-gas and methanol recycling,

an optimal systematic configuration and heat integration, and a high value by-product with a

plant capacity of 200 MW (He & Zhang, 2011). The estimated cost of ethanol from wood varies

between 0.50–0.76 US$/L depending on the plant capacity (AEA Technology, 2003; Galbe &

Zacchi, 2002; S&T2, 2004).

The reported enzyme cost of lignocellulosic ethanol varied widely, with the on-site enzyme

production/purchase cost reported to date perhaps the most contentious or dubious estimation. In

the USA, the costs associated with dedicated cellulase production are reported to be 0.1–0.5

US$/gal ethanol (Aden et al., 2002; Aden, 2008; Lynd et al., 2002; Wu et al., 2008). It is also

predicted that in future, less cellulase will be necessary, due to increased specific enzyme

activity: threefold in 2005 and tenfold in 2010 (Wu et al., 2008). The present enzyme production

cost is estimated as 265 $/m3 (1 $/gal), but with recent investments and continuous research

efforts, this value may drop to130 $/m3 (0.5 $/gal) by 2010 (Bryant, 2009; Seabra et al., 2010).

The most astonishing prediction seems to be of enzyme productivity: 600–2000 FPU/g

glucose+Xylose between 2005 and 2010 (Wu et al., 2008), which is subject to considerable

doubt. Presently, the enzyme productivity achieved is reported to be 333 FPU/g glucose (NFRI,

37

20 0; unpublished data) Conversely, the cost of cellulase is reported to be 0 5 €/L (Reith et al.,

2002). The reported enzyme cost (production/purchase) is Canadian dollar (CAD) 12/million

FPU (enzyme loading: 10 FPU/g cellulose) (Gregg et al., 1989). These studies reported a wide

variation of the cost of cellulase, hence the ethanol.

The enzyme production is noted to be dependent on enzyme processing plant, site of

production, and the method of transportation and the protein yield. Onsite enzyme production

reduces production cost compared to the off-site production. The production cost varied from

$3.8–$6.8 and $4.0 –$8.8/kg protein for onsite and offsite, respectively. The enzyme production

cost is noted to be $0.12/L-ethanol for that ethanol processing plant capacity of 150MML/year

using 11.5 mg enzyme/g substrate (Hong et al., 2013). It is also noted that despite the low cost of

biomass, enzyme is a significant contributor to the production cost of ethanol (Ensinas et al.,

2013). Table 2.5 shows a summary of the reported cost of cellulase and the production cost of

ethanol produced from cellulosic biomass by different authors.

2.4.2 Life cycle costing of ethanol produced by thermochemical process

Haro et al. (2012) designed and developed the kinematic laboratory data which involves

biomass pretreatment (poplar wood chips drying, size reduction to 4 cm) and gasification

(atmospheric-pressure indirect circulating fluidi ed bed gasifier) to produce ethanol from wood

chips. The estimated production cost is noted to be 0.55–0.59 $/L. The estimated production cost

of ethanol produced by optimized thermochemical conversion process is reported to be 0.11–

0.25$/L due to the large contribution of coproduct (hydrogen) (Grossmann & Martín, 2010;

Martín & Grossmann, 2011). The production cost is reported to be greater for gasification-

fermentation route compared with gasification-catalytic synthesis (Phillips, 2007). The

production cost of ethanol has also been estimated for different feedstocks using the material and

energy data modeled with Aspen Plus (at the NREL, USA). The forest-based feedstocks

including loblolly pine, natural hardwood and eucalyptus present more attractive financial

returns when compared to switchgrass and corn stover, mainly due to their composition and

alcohol yield (Phillips, 2007; Gonzalez et al., 2012).

38

Table 2.5 Summary of the reported cost of ethanol produced from different feedstocks (biochemical conversion)

Authors Feedstock, feed rate, cost & yield

2Enzyme Enzyme Cost of ethanol for different cases and years, $/L

Remarks Rate, t/d Cost, $/t L/t loading cost, $/L 1999 2000 2002 2005 2010 2012

1Wooley et

al., 1999a *CS, 2000 25

257.38-

355.79

15-20

FPU 0.079 0.38 - - 0.248 0.217 -

Enzyme cost need to be reduced

ten-fold, dollar value in 1997

McAloon

et al., 2000 *CS, 1050 35 272.52 - 0.05

0.396 - - - -

Little information is available on

enzyme production, dollar value

in 1999

1Aden et

al., 2002 *CS, 2000 30

272.52-

339.51

12-17

FPU 0.026 - - 0.346 - 0.283

Buying of enzymes, dollar value

in 2000

1Aden et

al., 2008 *CS, 2000 60.0-46.0 257.38 -

0.085-

0.026 - - 1.11 0.666 - 0.351

Enzyme cost is assumed, dollar

value in 2002

1Dutta et

al., 2010a *CS, 2000 60.1 -

30-40 mg

protein 0.085 - - - - 0.801

Enzyme cost is assumed, dollar

value in 2007

1Eggeman

et al., 2005 CS, 2000 35 - 15 FPU 0.039 - - -

0.262

-

0.441

- - Enzyme cost is assumed

Reith et al.,

2002 LVG, 2000 20 € 152.49 - 0 5 € - - 0 92 € - - - Enzyme cost is assumed

1Shafiei et

al., 2011

Spruce,

200000a

16-82 € - - 1.226

€/kg - - - - 0 44€ - Enzyme cost is estimated

Orikasa et

al., 2009 *RS, 200 15000

¥ 250 - - - -

Enzyme cost is assumed

3Barta et

al., 2010a

Spruce,

200000a

68.15 254.0-

270.0 10 FPU

0.058-

0.073 - - - -

0.548

-

0.722

- Enzyme cost is assumed

4Roy et al.,

2012a,b

LRS, 150-

200 150 250-330

9.1-12

FPU/g-

straw

0.14-0.24 - - - - - 0.85-

1.45 Enzyme cost is estimated

CS: corn stover; RS: rice straw; VG: verse grass; FPU: filter paper unit; 1Plant life: 20 years;

2per g-cellulose; *dilute acid pretreatment;

LLime pretreatment;

3Plant life 15 years;

4Plant life: 9 years; €: cost in Euros;

aAnnually (This table is partly adopted from Roy et al., 2012c).

39

A thermochemical process has also been designed, simulated and optimized mainly with

Aspen Plus (He & Zhang, 2011) and concluded that feedstock and syngas cleaning are the major

contributors in the ethanol production cost followed by the feedstock. It is also noted that cost-

competitive ethanol production can be realized with efficient equipment, optimized operation,

cost-effective syngas cleaning technology, inexpensive raw material with low pretreatment cost,

high performance catalysts, off-gas and methanol recycling, optimal systematic configuration

and heat integration, and high value byproduct with a plant capacity around 200 MW. A wide

variation of production cost was also reported by several authors where the lignocellulosic

ethanol production by thermochemical gasification processes were modeled (most of the studies

modeled with Aspen Plus software) to estimate the production cost. The production cost varied

from 0.27–1.25 $/L (Wu et al., 2006; Gonzalez et al., 2012; Haro et al., 2012; Phillips, 2007;

Phillips et al., 2007; Nechodom et al., 2008; Foust et al., 2009; Mark et al., 2009; Dutta et al.,

2010b; Perales et al., 2011). The feedstock and capital cost are identified to be main hotspots

(Dutta et al., 2010b; Foust et al., 2009).

The ethanol production cost in an integrated process (corn and corn stover: grain 18 kg/s

and stover 10.8 kg/s) is reported to be $0.32/L and noted that production cost decreased if only

stover is used (Čuček et al., 2011). Although most of the theoretical studies (modeled with Aspen

Plus software) noted that the thermochemical conversion of biomass to ethanol would be

competitive with the fossil ethanol, in depth studies at bench/pilot plants are required for any

future investment and commercial production. A summary of the reported production cost and

other assumptions are reported in Table 2.6.

Moisture and ash content in preprocessed feedstock are observed to be highly sensitive to

the LC GHG emission of ethanol (Daystar et al., 2013; Tan & Dutta, 2013). The GHG emission

reduction is noted to be more than 13% and 7%, if moisture and ash content reduced from 50%

to 30% (wt) by field drying and 7% to 1% (wt), respectively (Tan & Dutta, 2013). The authors

also noted that the thermochemical pathway reduces 83% GHG emission compared with 2005

gasoline baseline. Daystar et al. (2013) reported the highest ethanol yield in the case of loblolly

pine and the lowest for corn stover among hardwood, loblolly pine, eucalyptus, miscanthus, corn

stover, and switchgrass. The baseline GHG emission is noted to be 2.8 kg CO2 e/L from the

thermochemical conversion of these feedstocks into ethanol.

40

Table 2.6 Summary of the reported cost of ethanol from different feedstock and energy efficiency (thermochemical conversion)

Authors Feedstock, feed rate, cost & yield Gasifier

types

Catalyst/

Bacteria Cost,

$/L Energy efficiency, % Feedstock Rate, t/d Cost, $/t Yield, L/DT

1Haro et al., 2012^ Wood chip 2140 66 DT - Circulating fluidized bed CoeMo/ZnO 0.555–0.592 44.35-45.53

Nechodom et al., 2008 Forest waste 5000 45 DT 284–322 Fluidized-bed gasifier Iron-based catalyst 0.17–0.22 -

1Phillips et al., 2007 Wood chip 2000 35 DT 217–430 Entrained flow gasifier MoS2 0.267 –0.533 47.4

1Phillips, 2007 Wood chip 2000 39 DT 280 Circulating fluidized bed MoS2 0.267 43

1Dutta et al., 2010b Wood chip 2000 50.5 DT 241 Shell entrained flow Co-Mo 0.63 37.1-38.3

3He & Zhang, 2011^

Lignocellulosic

biomass 2455 58 66 € 332 Dual -bed gasifier Co-Mo 0 33 € -

1Perales et al., 2011^ Wood chip 2140 66 DT 206 Entrained flow

Rh-Mn/SiO2;

KCoMoS2 0.90 – 1.25

- 2Piccolo & Bezzo,

2009^ Wood 700000* 36-63 € 228-282 Fluidized bed

Clostridium

ljungdahlii 0.41-0 80 €

62

2Gonzalez et al., 2012^

Softwood,

hardwood 453597* 69.4-80.3 373-450 Indirect gassification

molybdenum

disulfide 0.54-0.71 -

1Spath & Dayton, 2003 Softwood 2204 25 DT 256-397 BCL/FERCO gasifer - 0.35 -

artı´n & Grossmann,

2011 Switchgrass 22 kg/s - 0.33-0.39 High/low pressure Bacteria 0.11-0.27 -

1Foust et al., 2009 Wood chip

46 DT 356 L

Circulating fluidized bed

indirect gasification

Nickel-based

catalyst 0.32 47.4

Mu et al., 2010^

Wood chips,

wheat straw,

waste paper, corn

stover

2000 - 270-359 Indirect-heating

fluidized bed gasifier

molybdenum

disulfide (MoS2) 36-51

1Plant life span 20 years; 2Plant life span 15 years; 3Steam plant 20 year and others 7 years; *t/year; t/d: ton/day; DT: dry ton ^Either centering the NREL studies or energy and mass balance modeled with AspenPlus

41

Chapter 3

Life Cycle Assessment (LCA) Methodologies

3.1 LCA Methodologies

Life cycle assessment (LCA) is a tool for evaluating environmental effects of a product,

process, or activity throughout its life cycle (LC) or lifetime, which is known as a ‘from cradle to

grave’ analysis lthough the concept of LC evolved in the 960s and there have been several

efforts to develop LCA methodology since the 1970s, it has received much attention from

individuals in environmental science fields since the 1990s. Since then the LCA concept was

promoted, sponsored and developed by the various national and international organization

(SETAC: Society of Environmental Toxicology and Chemistry, USEPA: United States

Environmental Protection Agency, ISO: International Organization for Standardization, ILCAJ:

Institute of Life Cycle Assessment, Japan etc.), and LCA practitioners. Consequently, consensus

has been achieved on an overall LCA framework and a well-defined inventory methodology

(ISO, 1997).

The LCA method is rapidly developed into an important tool for authorities, industries, and

individuals in environmental sciences. The UNEP (United Nations Environment Programme)-

SETAC initiative includes methods for the evaluation of environmental impacts associated with

water consumption and land use (Jolliet et al., 2004). A common methodological framework

(“Version Zero”) has also been developed by the Global Bioenergy Partnership (GBEP) Task

Force on GHG Methodologies that could be applied to the LCA of bioenergy production and

compared to the full life cycle of its fossil fuel equivalent to improve the transparency and

acceptance of the results (GBEP, 2009). The LCA methodology consists of four components:

Goal definition and scoping, Inventory analysis, Impact assessment and Interpretation. Figure 3.1

shows the stages of an LCA (ISO, 2006). The purpose of an LCA can be: (1) comparison of

alternative products, processes or services; (2) comparison of alternative life cycles for a certain

product or service; (3) identification of parts of the life cycle where the greatest improvements

can be made. Accordingly, the LCA methodologies (ISO, 2006) will be used to evaluate the life

cycle of ethanol from biomass.

42

Inte

rpre

tati

on

Life cycle assessment framework

Goal and scope

definition

Inventory analysis

Impact assessment

- Product development and improvement

- Strategic planning

- Public policy making

- Marketing

- Other

Direct applications:

Figure 3.1 Stages of an LCA (ISO, 2006)

3.1.1 Goal definition and scoping

Goal definition and scoping is perhaps the most important component of an LCA because

the study is carried out according to the statements made in this phase, which defines the purpose

of the study, the expected product of the study, system boundaries, functional unit (FU) and

assumptions. The system boundary of a system is often illustrated by a general input and output

flow diagram. All operations that contribute to the LC of the product, process, or activity fall

within the system boundaries. The purpose of functional unit (FU) is to provide a reference unit

to which the inventory data are normalized. The definition of FU depends on the environmental

impact category and aims of the investigation. The functional unit is often based on the mass

(kg) or volume (L) of the products under study; however, the distance (km), land area (ha),

energy (MJ) and economic values of products are also used.

Biomass especially, agri-residues and forest residues are identified as abundant biomass

resources for bioenergy. Forest residues are broadly categories into two groups: mill residues

(bark, sawdust and shavings), and forest residues (tops, branches and leaves from harvest and

thinning operations). Mill residue productions in various provinces of Canada are reported in

Table 3.1. The price of sawdust, bark and bark piles varies from 0–22.5, 2.9–32.5 and 0–7.8

$/ODt, respectively, in 2007, Eastern Ontario, Canada (Bradley, 2010). These residues are noted

to have low commercial value (Talebnia et al., 2010; sawdust), high in cellulose content, thus a

suitable raw material for ethanol production (Zabihi et al., 2010; sawdust; miscanthus). The

major components of different residues and the chemical compositions are presented in Table

3.2, Table 3.3 and Table 3.4. In Canada, the amount of biomass is reported to be approximately

9.4×106 MT/yr (Mabee et al., 2006). The potential sources of renewable biomass include

43

agricultural residues, municipal solid waste, forestry byproducts and energy crops. The technical,

economic and sustainability constraints in Ontario conditions also limits their supply to ethanol

industry (Kludze et al., 2010). The use of forestry wastes for liquid biofuels are also restricted

due to its enormous demand by the solid biofuels industries. Miscanthus is a promising energy

crop with high yield and energy content, which can be grown on low quality or marginal land,

and add carbon to the soil and safeguard it against erosion (Somerville et al., 2010; Kludze et al.,

2013), has an important role in sustainable energy production (Sørensen et al., 2008; Khanna et

al., 2008; Bocquého & Jacquet, 2010). Therefore, the goal of this study is to evaluate the LC of

ethanol produced from agri-residues/forest residues/energy crops. The FU of this study is defined

as 1 L of anhydrous ethanol produced from biomass. Figure 3.2 depicts the system boundary of

this study (‘cradle to gate’ scenario)

Table 3.1 Mill residues production in Canada in 2004 (ODt: Oven dry tonnes)

Province Production Consumption Export Surplus

Alberta 2406 1924 0 481

British Columbia 6554 4338 350 1815

New Brunswick 1373 1223 150 0

Manitoba 225 212 0 13

Ontario 2602 2480 1 121

Quebec 6669 6400 169 100

Saskatchewan 580 416 0 164

Source: Bradley, 2010

Table 3.2 Volatile matter, fixed carbon, and ash content in selected biomass (dry basis)

Feedstock Volatile matter, % Fixed carbon, % Ash, %

Wheat straw 80.04 15.31 4.65

Miscanthus 87.50 12.16 0.80

Sawdust 85.63 12.98 1.39

Source: Kambo, 2014 (unpublished data).

44

Table 3.3 Potential feedstocks and their major components

Feedstock Components, % (w/w)

Cellulose Hemicellulose Lignin Ash Reference

48.57 27.7 8.17 6.68 Saha & Cotta, 2008

39 32 14 8 Alinia et al., 2010

32.1 26.9 20.3* 11.2 Thompson et al., 2003

Wheat straw 33-40 20-25 15-29 - Prasad et al., 2007

52.04 30.12 7.79 4.63 Mark, 2010

40.5 26.3 18.2 - Xu et al., 2011

37.4 33.8 19.4 - Ruiz et al., 2011

35.44 24.56 19.8 - Leistritz et al., 2006

30 50 15 - Sun & Cheng, 2002

Sawdust 55 14 21 -

Olsson & Hahn-Hägerdal,

1996

Miscanthus 36.96 22.12 23.31 2.84 Han et al., 2011

39.21 23.47 21.36 2.87 Li et al., 2013

38.2 24.3 23.0 2.0 Vrije et al., 2002

40.0 18.0 21.0 5.9 Sørensen et al., 2008

40.2 22.4 24.4 - Yoshida et al., 2008

* with extractives

Table 3.4 Chemical composition of different feedstock

Feedstock

Chemical composition, g/kg Reference

C N P S K Ca Mg

Wheat

straw 420 5.5 0.4 0.9 10.4 2.9 0.6 Kaboneka et al., 2004

455 8 - - - - - Sule, 2012

Sawdust 490 0.1 - 0.1 - - - Acharya, 2011

562 10 - - - - - OMAFRA, 2011

Miscanthus 453 2.1 - - - - - Sule, 2012

480 1.0 - 0 - - - Heo et al., 2010

471 0.44 - 0.06 - - - Cullura et al., 2006

45

Figure 3.2 System boundary of this study

3.1.2 Life cycle inventory (LCI) analysis

The inventory analysis involves collecting data on raw materials and energy consumption,

emissions to air, water and soil, and generation of solid waste. This phase is the most work

intensive and time consuming compared to other phases in an LCA, mainly because of data

collection,. The data collection can be less time consuming if good databases are available and if

customers and suppliers are willing to help. Consequently, LCA allows the use of quality data, if

that are not product specific. Nowadays, many LCA databases exist and can normally be bought

together with LCA software. Data on transport, extraction of raw materials, processing of

materials, production of usually used products like plastic; cardboard etc. and disposal can

normally be found in an LCA-database. Data from databases can be used for processes that are

not product specific, such as general data on the production of electricity, coal or packaging. For

product specific data, site-specific data are required. The data should include all inputs and

outputs from the processes. Inputs are energy (renewable and non-renewable), water, raw

materials etc. Outputs are the products and co-products, and emission (CO2, CH4, SO2, NOx, and

CO) to air, water and soil (total suspended solids: TSS, biological oxygen demand: BOD,

chemical oxygen demand: COD and chlorinated organic compounds: AOXs) and solid waste

generation (municipal solid waste: MSW and landfills). For the simplicity of this study, available

quality data are collected from the literature and different database. Rest of the required data are

generated experimentally at the laboratory wherever possible or simulation technic (Aspen Plus

V7.3) has been used to gather the required data. The collected and generated data are analyzed

by applying LCA methodologies.

With/without CO2 capturing

SSF

Dry

ing/p

re-

trea

tmen

t

Fermentation Gasification

Bacteria

Syngas cleanup &

conditioning

Cu

ltiv

atio

n

Co

llec

tio

n &

tr

ansp

ort

atio

n

Sep

arat

ion

Pretreatment Waste

water

Distillation &

purification

Enzyme & Yeast Residues Boiler/turbin

e

Byproducts/

waste

An

hyd

rou

s e

than

ol

Energy & materials

Emissions & waste

With/without torrefaction

46

3.1.3 Impact assessment

The life cycle impact assessment (LCIA) aims to understand and evaluate environmental

impacts based on the inventory analysis, within the framework of the goal and scope of the

study. In this phase, the inventory results are assigned to different impact categories, based on

the expected types of impacts on the environment. Impact assessment of the LCA generally

consists of the following elements: classification, characterization, normalization and valuation.

Classification is the process of assignment and initial aggregation of LCI data into common

impact groups. Characterization is the assessment of the magnitude of potential impacts of each

inventory flow into its corresponding environmental impact (e.g., modeling the potential impact

of carbon dioxide and methane on global warming). Characterization provides a way to directly

compare the LCI results within each category. Characterization factors are commonly referred to

as equivalency factors. Normalization expresses potential impacts in ways that can be compared

(e.g., comparing the global warming impact of carbon dioxide and methane for the two options).

Valuation is the assessment of the relative importance of environmental burdens identified in the

classification, characterization, and normalization stages by assigning them weighting which

allows them to be compared or aggregated. Impact categories include global effects (global

warming, ozone depletion etc.); regional effects (acidification, eutrophication, photo-oxidant

formation etc.); local effects (nuisance, working conditions, effects of hazardous waste, effects of

solid waste etc.).

A common framework consisting ‘midpoint’ and ‘endpoint’ approach is desirable for life

cycle impact assessment (LCIA) because both approaches have their specific strengths and

weakness (Heijungs et al., 2003). The Institute of Life Cycle Assessment, Japan has developed a

LC impact assessment method based on endpoint modeling (LIME) to quantify the

environmental impacts as accurately as possible with a high degree of transparency and to

develop a single central index (Eco-index) (Fig. 3.3). The authors also concluded that a single

index inevitably involves value judgment (pricing) and has a higher degree of uncertainty (Itsubo

& Inaba, 2007).

3.1.4 Interpretation

The purpose of an LCA is to draw conclusions that can support a decision or can provide a

readily understandable result of an LCA. The inventory and impact assessment results are

discussed together in the case of an LCIA, or the inventory only in the case of LCI analysis, and

47

significant environmental issues are identified for conclusions and recommendations consistent

with the goal and scope of the study. This is a systematic technique to identify and quantify,

check and evaluate information from the results of the LCI and LCIA, and communicate them

effectively. This assessment may include both quantitative and qualitative measures of

improvement, such as changes in product, process, and activity design; raw material use,

industrial processing, consumer use, and waste management. Cost and profit are the key

indicators in decision-making on an investment, while costs are what producers or consumers

understand best and an integral part of the decision-making process when identifying

improvements of a product, process or activity, hence LCA results are also interpreted in the

form of LC costing.

Figure 3.3 Structure of the LCIA method based on endpoint modeling (LIME2)

3.2 Life cycle cost analysis (LCCA)

Although an LCA is useful, always may not be sufficient basis for a sound decision

making process, the LCCA can broaden the decision making process. The combination of LCA

Source: Itsubo and Inaba, 2007

48

and LCCA results can be used to guide in the biomass to biofuels strategies (Roy et al., 2007;

Baldwin et al., 2012). Gerber et al. (2011) noted that the integration of environmental impacts in

the optimization procedure can influence the engineering decisions related to the final process

design (thermo-economic) of polygeneration systems for (renewable) energy services. The

uncertainty in LCA can be managed by combining process and economic input–output

approaches (Williams et al., 2009b). Nechodom et al. (2008) also noted that combination of LCA

and LCCA models can be used to determine economic viability, environmental impacts, and

energy efficiency of bioenergy from forest biomass. A mixed-integer linear programing (MILP)

approach revealed that optimal feedstock selection, technologies, intermediate and final product

flows improve economic efficiency and reduces GHG emission (ˇCuˇcek et al , 20 4) Another

MILP modeling frame work for ethanol supply chains, where multiple decision criteria are

considered in an uncertain market scenario revealed that the strategic decisions are dependent on

the trade-off between environmental and economic performance and are also firmly coupled with

investor’s attitude (Giarola et al., 2013).

Gerber et al. (2011) integrated the LCA indicators with an existing computer aided process

engineering platform to optimize thermo-economic design of polygeneration systems for

renewable energy services. The thermo-economic model has also been used to evaluate the LC

of bioenergy from biomass (Gerber et al., 2011; Tock et al., 2010; Gassner & Maréchal, 2009),

which can be used for analyzing both the economics and the environmental impact of bioenergy.

Møller et al. (2014) has studied the environmental and economic consequences of the LC of

ethanol integrating the material based LCA with welfare economic cost benefit analysis (CBA)

which claimed to have broadened the perspective of advantages and disadvantages of biofuels.

Therefore, the LC cost of ethanol is estimated with both fixed costs (straight line depreciation on

installation, labor, maintenance and interest on investment) and variable costs (feedstock,

enzyme production, utilities and waste management). The cost of a product may vary based on

the base year of calculation. The consumer price index (CPI) can be used to convert costs if cost

data are collected from different years. Most of the collected data are from 2010 except the

processing plant construction data, which is collected from a Japanese literature published in

2007. This later dat was not converted to a 2010 basis, since the Japanese CPI was not available.

49

Chapter 4

Life Cycle Assessment of Ethanol produced from Wheat Straw

[Published in the Journal of Biomass Materials and Bioenergy, 6(3): 276–282]

4.1 Introduction

The life cycle (LC) GHG emissions of different forms of bioenergy and their ability to

reduce GHG emissions vary widely, and are dependent on land use changes, choice of feedstock,

agricultural practices, refining and conversion processes with differing socioeconomic and

environmental impacts (Tilman & Lehman, 2006; Luo et al., 2009; Kaufman et al., 2010). It is

thus essential to evaluate the environmental impact and the economic viability of lignocellulosic

ethanol. Life cycle assessment (LCA) methodology has been extensively used to evaluate the LC

of lignocellulosic ethanol (Wooley et al., 1999; Sheehan et al., 2003; Kim & Dale, 2005; Adler et

al., 2007; Spatari et al., 2010; Hsu et al., 2010; Seabra et al., 2010; Orikasa et al., 2009; Roy et

al., 2012b). However, the LCA of ethanol from wheat straw has received only limited attention

(Kaparaju et al., 2009; Hsu et al., 2010), which are mostly confined to pretreatment processes,

production of alternate bioenergy, and utilization of ethanol. This study evaluated the LC of

ethanol to determine if environmentally preferable and economically viable ethanol can be

produced from wheat straw in Ontario, Canada.

4.2 Materials and methods

4.2.1 System boundary

Wheat is one of the most important agricultural commodities in Canada. Wheat was grown

on 9638200 ha in Canada in 2009 (FAO, 2011), and regionally was grown on 475500 ha in

Ontario (OMAFRA, 2011a), which may produce about 2282400 tons straw [4.8 dry-tons/ha

(Evans et al., 2005)]. Wheat straw is noted to be an abundant agricultural residue with low

commercial value (Talebnia et al., 2010), high in cellulose content, thus a suitable raw material

for ethanol production (Zabihi et al., 2010). Cellulose, hemicellulose and lignin contents are

reported to be 48.6, 27.7 and 8.2%, respectively (Saha & Cotta, 2008). An ethanol processing

plant is assumed to be established in the wheat growing area (100% cropland is assumed to be

under wheat crop; farmland is found to be 5.9% in Ontario) for efficient utilization of wheat

straw. The ethanol processing plant capacity is considered to be 20000 kL/year. Figure 4.1

depicts the system boundary of this study (‘cradle to gate’ scenario) The roads and channels are

50

assumed to be 2% of rural land area in Ontario. Crop residue incorporation is important in

maintaining soil organic matter, soil structure, fertility and productivity. The sustainable residue

collection rate is noted to be dependent on soil type, slope, crop rotation and tillage practices. It

is noted that about 50–60% crop residues can be collected without deteriorating the soil quality

and productivity (Jeschke, 2011 Smith, 1986; Sheehan et al., 2002), thus the straw collection is

considered to be 60% that grows in farm without the fear of soil degradation and productivity

loss (Smith, 1986; Sheehan et al., 2002). The straw transportation distance from farms to the

ethanol processing plant is estimated based on a published methodology (Huang et al., 2009).

The baled straw is assumed to be transported to nearby collection center (5 km from the farm)

and then transported to the ethanol processing plant (30 km) with 4 t- and 10 t-truck (loading

capacity is assumed to be 75%), respectively. It has been reported that agricultural LCAs often

exclude production processes of machines, buildings, and roads because of lack of data

(Cederberg & Mattsson, 2000). The environmental impacts related to the construction of the

ethanol processing plant, storage facilities and the production of transportation and other

machinery are not considered.

Figure 4.1 Schematic diagrams of the life cycle of wheat straw and the system boundary of this

study

4.2.2 Biochemical conversion process

Plant biomass contains high cellulose, thus a suitable material for ethanol production. Devi

et al. (2011) noted that plant biomass contains approximately 75% polysaccharides, a rich source

of sugars. In the biochemical conversion process, plant cell wall breaks through the introduction

of enzymes or acid in order to extract the sugars which are then converted to biofuels using

microorganisms. Figure 4.2 shows the schematic diagram of ethanol production process from

biomass (enzymatic hydrolysis process).

Tra

nsp

ort

ati

on

Co

lle

ctio

n

Pre

-tre

atm

en

t

Sa

cch

ari

fica

tio

n

& F

erm

en

tati

on

Dis

till

ati

on

&

pu

rifi

ca

tio

n

Wa

ste

m

an

ag

em

en

t

Cu

ltiv

ati

on

51

Figure 4.2 Schematic diagram of ethanol production process from biomass

4.2.2.1 Pretreatment

Lime pretreatment is be given (at 120°C for 1 h; lime 10%) (Park et al., 2010; Shiroma et

al., 2011) to the crushed (size 3 mm) feedstock. Although boiler efficiencies of between 60–67%

(switchgrass and corn stover; rice straw) (Wooley et al., 1999a: Orikasa et al., 2009) and 75–

90% (firewood, woodchips, and straw; corn stover) (CBT, 2002; Mani et al., 2010) have been

reported, a boiler efficiency of 80% was assumed for this study (Mani et al., 2010). The solid

concentration during pretreatment was considered to be 30% (w/w).

4.2.2.2 Vacuum extractive fermentation and distillation

Conventional alcoholic fermentation is a typical inhibitory process, with cell growth rate

affected by cellular substrate and product concentration (Rivera et al., 2006). The fermentation

must be maintained between 7–10°GL (Gay-Lussac) to prevent inhibitory effects (Junqueira et

al., 2009a) that reduce the yield and productivity of the process (Silva et al., 1999). Conversely,

the vacuum extractive fermentation process allows the simultaneous removal of produced

ethanol from the fermenter, which yields a highly concentrated wine, and as a result, reduces the

amount of vinasse and energy consumption in fermentation and the subsequent distillation steps

(Junqueira et al., 2009a; Silva et al., 1999). The hydrous ethanol produced in the distillation

undergoes purification stages to achieve anhydrous ethanol (99.5%). Bioglycerol is reported to

be a suitable agent for the separation of ethanol–water mixtures (Lee & Pahl, 1985) and be safely

used to produce anhydrous ethanol for use in food or pharmaceutical industries (Dias et al.,

2009). Therefore, vacuum extractive fermentation and distillation, and purification (with

glycerol) processes were adopted for this study (Junqueira et al., 2009a). The enzyme loading is

assumed to be 11 FPU/g straw (Talebnia et al., 2010). The enzyme cost (considering material

and energy consumption) was calculated based on the literature data (Wooley et al., 1999a) and

SSF (C6) Distillation & purification

Pretreatment

Crushing

Boiler

Anhydrous Ethanol

Residues

Heat

LNG

SSF (C5)

Enzyme & yeast

52

the enzyme loading rate of this study. Ethanol yield is considered to be 0.3 L/kg dry-straw (Li et

al., 2011).

4.2.2.3 Waste management

The waste stream is used to be separated into three feed streams: solids (lignin), biogas and

syrup high in solids (Wooley et al., 1999a). Anaerobic digestion of the wastewater produces a

biogas high in methane. Burning these byproducts streams to generate either the heat or

electricity required in the processes reduces not only the waste management costs, but may also

lead to profit. Therefore, it is assumed that energy consumption and cost incurred in the waste

management processes would be offset by biogas and syrup generated in the process, and the

lignin recovered in the waste management process can be used for process heat generation. The

residual lignin is assumed to be dried by utilizing the wasted heat from the boiler. The heat

generation from the residual lignin is estimated based on the heating value of lignin and

efficiency of the boiler (80%; Mani et al., 2010), is assumed to be used for heat.

4.2.3 Cost analysis

The cost of ethanol is estimated with both fixed costs (straight line depreciation on

installation, labor, maintenance and interest on investment) and variable costs (feedstock,

enzyme production, utilities and waste management). The business capital is considered to be

twice of the price of yearly feedstock demand. The yearly interest rate on investment and

maintenance cost is assumed to be 3 and 2%, respectively. The economic life span of an ethanol

processing plant and the yearly operating period are assumed to be 20 years and 350 days,

respectively (Dutta et al., 2011; Huang et al., 2009; Wu et al., 2006). The cost of wheat straw is

considered to be about 143 $/t-dry (OMAFRA, 2011b). [Note: The cost of wheat straw is even

higher at present (August 2014), Approximagtely, $176/t at the farm gate; personal

communication: Prof. B. Deen, Department of Plant Agriculture, University of Guelph,

Canada)].

4.2.4 Data collection

Data collection is the most work intensive and time consuming activities in an LCA, thus

allows the use of quality data, if that are not product specific. For the simplicity of this study,

data have been collected from the literature and different database. Although the inventory data

vary depending on the regions of the study, production process and allocation methods, the data

53

used are assumed to be valid for this study. A summary of the parameters/processes for which

data have been collected from the literature and their sources are reported in Table 4.1.

Table 4.1 Summary of parameters for which data are collected from the literature

Parameters/Systems Sources

Ethanol plant construction & No. of labor Asano & Minowa, 2007

Straw collection JA-Zenno, 2002

Enzyme Wooley et al., 1999a

Yeast Dutta et al., 2010a

Vacuum extractive fermentation & distillation Junqueira, et al., 2009a

Purification Dias et al., 2009

4.3 Results and discussion

4.3.1 Energy consumption, CO2 emission and production cost

The energy consumption, GHG emissions (CO2 e) and production cost in each stage of the

LC of ethanol are worked out to represent the energy-, emissions- and cost breakdown of the

production steps and identify the hotspots (Fig. 4.3). The pretreatment process is observed to be

the main hotspot followed by enzyme production, and others in the case of energy consumption.

It is worthy to note that byproduct (residues) has a strong contribution in net energy

consumption, because it offsets a part of energy consumed in the process. The estimated net

energy consumption is found to be about 15 MJ/L (Fig. 4.3a). The net energy consumption for

corn and cellulosic ethanol is reported to be about 10.6–17.0 and 6.0–31.0 MJ/L, respectively

(Sheehan et al., 2003; Kim & Dale, 2005; Pimentel & Patzek, 2005; Luo et al., 2009a). A slight

variation in net energy consumption is observed compared with some of the earlier studies,

which might be because of different feedstock and processes have been used.

GHG emissions are directly related to energy and resource consumption. Therefore, main

hotspot and the order of other stages are observed to be similar to those observed in the case of

energy consumption. Figure 4.3(b) shows the emission breakdown of the LC of ethanol. The

residues used for heat generation has a strong contribution to offsetting a part of emission of the

production process. The net emission is estimated to be 0.91 kg CO2 e/L. Emission from switch

54

grass and corn stover ethanol is reported to be 0.49 and 0.33 kg CO2 e/L, respectively where

carbon sequestration has been considered (Spatari et al., 2005). The emission is also noted to be

1.6 kg CO2 e/L (Farrell et al., 2006). These results reveal that an environmental benefit can be

achieved relative to gasoline (1.5 kg CO2 e/L-ethanol equivalent in Canada; Environment

Canada, 2010) when ethanol can be produced by using the technologies adopted in this study.

Figure 4.3 Energy, emission and cost breakdown of the life cycle of ethanol produced from

wheat straw.

The production cost of ethanol is also noted to be dependent on both technical and

economic parameters, such as the cost of feedstock, choice of feedstock, energy consumption,

conversion technology and efficiency, and the value of byproducts (Aden et al., 2002; Dutta et

al., 2010a; Balat, 2011). Figure 4.3(c) depicts the cost breakdown of different stages of the LC of

ethanol. The main hotspot is emerged to be feedstock followed by fixed cost and others. The

fixed cost is calculated to be $0.26/L, which may vary, if different processing plant sizes and

operating periods and life spans are considered. The enzyme cost is estimated to be $0.13/L. The

net production cost is estimated to be $1.14/L. However, the net production cost would be

$1.19/L, if different enzyme cost is considered ($0.18/L, i.e., 0.69/gallon ethanol; Kazi et al.,

2010a,b). The use of different input data, functional units, allocation methods, reference systems

-6

0

6

12

18

MJ/

L

a. energy

-0.4

0.0

0.4

0.8

1.2k

g C

O2

e/L

b. emission

-0.2

0.1

0.4

0.7

1.0

1.3

$/L

c. cost

Residues

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Collection

Feedstock

Fixed cost

55

and other assumptions complicates comparison of LCA bioenergy studies (Liska & Cassman,

2008; Cherubini & Strømman, 2011), thus a direct comparison was not attempted with other

studies. A wide variation was observed in the reported production cost of ethanol. The

production cost of ethanol from corn stover and fescue straw is reported to be 0.71–0.87 US$/L

(Dutta et al., 2010a; Kumar & Murthy, 2011). The simulated production cost of ethanol is

reported to be 0.94–1.20 US$/L which depends on the ethanol yield (Klein-Marcuschamer et al.,

2010). The cost is also noted to be dependent on the feedstock and plant sizes (Gnansounou &

Dauriat, 2010). Table 4.2 represents a brief summary of the reported production cost, which

indicates that the variation in cost among various studies might be because of different

processing plant size, types of feedstock, ethanol yield and assumptions. The production cost of

lignocellulosic ethanol is reported to be considerably higher than the market price of gasoline

(Wooley et al., 1999a; Huang et al., 2009). This study also confirmed that despite the

environmental benefits (when agri-residues is assumed to be carbon neutral) of ethanol produced

from wheat straw, its economic viability remains doubtful at present, even if highly optimistic

assumptions are made for the cost calculation, especially in the case of enzyme. The costs of

enzyme and capital are the major expenses when producing lignocellulosic ethanol (Reith et al.,

2002), thus production cost may differ when different enzyme cost is considered.

Table 4.2 Summary of the reported cost of ethanol produced from different feedstock

Authors

Feedstock, feed rate, cost & yield 2Enzyme Enzyme Cost, $/L

Rate, t/d Cost,

$/t L/t loading cost, $/L

1Roy et al.,

2012a,b LRS, 150-200 150 250-330

9.1-12 FPU/g-

straw 0.14-0.24 0.85-1.45

1Wooley et al.,

1999a *CS, 2000 25

257.38-

355.79 15-20 FPU 0.079 0.217-0.38

1Aden et al., 2002 *CS, 2000 30

272.52-

339.51 12-17 FPU 0.026

0.283-

0.346

1Dutta et al.,

2010a *CS, 2000 60.1 -

30-40 mg

protein 0.085 0.801

Reith et al., 2002 LVG, 2000 20 € 152.49 - 0.5 € 0 92 €

3Barta et al.,

2010a

Spruce,

200000a

68.15 254.0-

270.0 10 FPU

0.058-

0.073

0.548-

0.722 CS: corn stover; VG: verse grass; FPU: filter paper unit; 1Plant life: 20 years; 2per g-cellulose;

*dilute acid pretreatment; LLime pretreatment; 3Plant life 5 year; €: cost in Euro; aAnnually

56

4.3.2 Sensitivity analysis

The LCA results of lignocellulosic ethanol are noted to be more sensitive to the changes in

parameters related to the biomass and ethanol yield. A wide variation in ethanol yield from

wheat straw (Table 4.3) (Maas et al., 2008; Li et al., 2011; Xu et al., 2011), feedstock- and

enzyme cost of lignocellulosic ethanol are also reported (Wooley et al., 1999a; Dutta et al.,

2010a; Aden et al., 2002; Reith et al., 2002; Barta et al., 2010a). Therefore, the effect of ethanol

yield, feedstock- and plant capacity on the emission and production cost has been determined.

Carbon sequestration by the biomass also plays an important role in the LC of agri-products

(Sanscartier et al., 2013). Consequently, the effect of carbon sequestration has also been

evaluated.

Table 4.3 Ethanol yield from wheat straw

Reference Conversion method Pretreatment Yield, g/kg–wheat

straw

Maas et al., 2008 SSF; GC 220 Lime 102

Li et al., 2011 SSF; Celluclast 1.5 L Dilute acid 230–240

Xu et al., 2011 SSF; Cellubrix L NaOH 149

Xu et al., 2011 SSF; Cellubrix L untreated 27

Panagiotou et al., 2011 SSF; Celluclast 1.5 L

FG & Novozym 188

Steam

explosion 40*

*Pretreated wheat straw &corn cob

It is important to note that the construction- and labor cost for different plant sizes are

calculated based on published methodologies (Huang et al., 2009; Asano & Minowa, 2007). The

scaling factor for estimating the construction and labor cost is considered to be 0.70 and 0.27,

respectively. Figures 4.4 and 4.5 depict that net energy consumption, emission and production

cost dependent on the ethanol yield and feedstock cost. The lower the ethanol yield, the greater is

the energy consumption, emission and production cost. Production cost is also increased with an

increase in feedstock cost. Figure 4.6 confirmed that the production cost decreased when

processing plant capacity is increased because of lower fixed cost in the case of greater capacity.

Although the production cost reduced with the increase of processing plant capacity, emission

rises because of longer transportation distance resulted from greater feedstock demand. Figure 6

57

also depicts that 26% of production cost can be reduced if the plant capacity rose from 5 to 200

ML/year; however, only about 4% of GHG emission would be increased. It is worthy to note that

energy efficiency was assumed to be the same for different processing plant sizes i.e., energy

consumption in processing steps are constant for different plant sizes. If the energy efficiency

varied with the plant size the emission will be changed.

12

15

18

21

0.30 0.25 0.20

Ener

gy,

MJ/

L

Net energy consumption

0.0

0.5

1.0

1.5

0.30 0.25 0.20

kg C

O2

e/L

Emission

1.0

1.2

1.4

1.6

0.30 0.25 0.20

$/L

Production cost

Ethanol yield, L/kg straw

Figure 4.4 Effect of ethanol yield on net energy consumption, emission and production cost of ethanol

58

Figure 4.5 Effect of feedstock cost on the production cost of ethanol

Figure 4.6 Effect of plant capacity on the production cost and emission of the life cycle of

ethanol

0.5

1.0

1.5

2.0

0 40 80 120 160

Pro

duct

ion c

ost

, $

/L

Feedstock cost, $/t

0.30 L/kg 0.25 L/kg 0.20 L/kg

0.8

0.9

1.0

0.9

1.0

1.1

1.2

1.3

1.4

0 50 100 150 200 250

Pro

duct

ion c

ost

, $/L

Processing plant capacity, ML/Year

Cost Emission

Em

issio

n, kg

CO

2 e

/L

59

Wheat straw is considered as carbon neutral, because it is a byproduct of wheat cultivation.

However, any marginal inputs and emissions are allocated to wheat straw. Emission from wheat

cultivation in Ontario is reported to be 1.995 t CO2e/ha (Dyer et al., 2010). Allocating this

emission between wheat grains and straw (based on the economic value of grains and straw) the

emission from the LC of ethanol has also been calculated. The price of wheat grains and straw is

reported to be 234.4 and 143 $/t, respectively (OMAFRA, 2011b; Agriculture and Agri-food

Canada; 2011). The emission is dependent on the carbon neutrality of biomass (Fig. 4.7). The

emission remains constant, if biomass is considered to be carbon neutral. Conversely, the

emission is increased with an increase in the feedstock cost because of greater share of emission

resulted from wheat cultivation, where carbon sequestration was not considered. Emission is

found to be 1.5, 1.7 and 2.0 kg CO2 e/L for ethanol yield 0.30, 0.25 and 0.20 L/kg straw,

respectively when feedstock cost considered to be 143$/t (Fig. 4.8). Figure 4.8 also depicts the

effect of system boundary or the carbon neutrality of wheat straw. These results confirm that

both environmental and economic viability of ethanol from wheat straw remains doubtful (for a

‘cradle to gate’ analysis) with present technologies However, wheat straw based ethanol should

represent lower GHG emissions for a ‘cradle to grave’ or ‘well to wheel’ LC because CO2

emissions released when using ethanol as a fuel are of biogenic origin (i.e. carbon neutral).

The impacts of soil carbon dynamics on the LC of crops noted to be minimal since most of

the crop shifts are among annual crops (Dyer et al., 2010). Thus soil carbon change is not

accounted, if any. However, carbon sequestration by the crop residues (harvested residues) is

considered to determine its impact on the LC of ethanol derived from wheat straw. The carbon

sequestration is estimated based on the carbon content (fixed carbon only) in wheat straw.

Although carbon content in wheat straw is reported to be about 42%, the fixed carbon is only

15.31%. The net GHG emissions reduced to -0.24 to -0.61 kg CO2 e/L if carbon sequestration

(with only fixed carbon) is considered (Fig. 4.9) which is also dependent on ethanol yield,. This

figure indicates that environmentally viable ethanol can be produced from wheat straw even it is

considered to be carbon non-neutral i.e., emissions from the agricultural operation is shared by

the harvested wheat straw. The net GHG emissions for stover is also reported to be negative (–

864 kg CO2 e/t) where carbon sequestration contributed about 62–66% (Roberts et al., 2010).

Moreover, it is worthy to note that this is a highly optimistic study, and all the data do not

necessarily correspond to the Canadian context or to the same processing plant size. Therefore,

60

in-depth studies are necessary for each step of the LC of ethanol from wheat straw (especially

pretreatment, enzyme and yeast production, SSF, distillation and purification) for any future

investment and commercial production.

Figure 4.7 Effect of feedstock cost on the emission of the life cycle of ethanol

Figure 4.8 Effect of system boundary and the ethanol yield on life cycle GHG emission of

ethanol

0.7

0.9

1.1

1.3

1.5

1.7

0 40 80 120 160

Em

issi

on

, k

g C

O2

e/L

Price of feedstock, $/t

Neutral Non-neutral

0.0

0.5

1.0

1.5

2.0

2.5

0.30 0.25 0.20

Em

issi

on

, kg C

O2

e/L

Ethanol yield, L/kg-straw

Carbon neutral Non-Carbon neutral

61

Figure 4.9 Effect of carbon sequestration and ethanol yield on the life cycle GHG emission

4.4 Conclusion

The net energy consumption, GHG emission and the production cost are found to be 15.0

MJ/L, 0.91 kg-CO2 e/L and 1.14 $/L, respectively (when plant capacity considered to be 20000

kL/year) which depend on the ethanol yield, plant capacity, and system boundaries. This study

also depicts that environmental benefit can be gained with present technologies if wheat straw is

considered to be carbon neutral, otherwise both environmental and economic viability of ethanol

from wheat straw are doubtful. It is also worthy to note that environmental viability of ethanol

from carbon non-neutral wheat straw can be improved if carbon sequestration is considered.

-0.70

-0.60

-0.50

-0.40

-0.30

-0.20

-0.10

0.00

0.30 0.25 0.20

Em

issi

on

, kgC

O2e/

L

Ethanol yield, L/kg-straw

62

Chapter 5

Life cycle assessment of ethanol derived from sawdust

[Published in the Bioresource Technology, 150(December): 407–411]

5.1 Introduction

The potential sources of renewable biomass in Canada include waste products from

forestry and agricultural residues, municipal solid waste, and energy crops. In Canada, the

amount of biomass production is reported to be approximately 9.4×106

MT/year. Residual

lignocellulosic feedstock could provide up to 50% of Canada’s 2006 transportation fuel demand

(Mabee & Saddler, 2010). Forestry products, particularly sawdust, forest thinning and trimming

are potential feedstock for ethanol production (Kadam et al., 2000; Mu et al., 2010). The life

cycle (LC) GHG emissions from biofuels and their ability to reduce GHG emission are

dependent on choice of feedstock, agricultural practices, and conversion technologies with

differing socioeconomic and environmental impacts (Tilman et al., 2009; Luo et al., 2009;

Kaufman et al., 2010). Although many researchers have evaluated the LC of lignocellulosic

ethanol produced by enzymatic hydrolysis process, ethanol from sawdust received only limited

attention (Slade et al., 2009; Sandilands et al., 2009), their research deals with thermochemical

conversion (gasification-synthesis). This study evaluated the LC of ethanol produced by

enzymatic hydrolysis and considered two scenarios to determine if environmentally preferable

and economically viable ethanol can be produced from sawdust in Ontario, Canada.

5.2 Methodology

5.2.1 System boundary and assumptions

The forest products industry produces woody biomass as a byproduct, including bark,

sawdust and shavings. In Canada, sawmill residue production rate is estimated to be

approximately 2.3×106 dry-tons/year (Ackom et al., 2010). The forest area in Ontario is reported

to be 71,067,769 ha in 2008 (MNR, 2011), which produces a considerable amount of residues.

Sawdust is reported to be high in cellulose content, thus a suitable raw material for ethanol

production. Cellulose, hemicellulose and lignin contents are reported to be 55, 14 and 21%,

respectively (Olsson & Hahn-Hägerdal, 1996). Ethanol yield is assumed to be 0.305 L/kg of dry

sawdust (Olsson & Hahn-Hägerdal, 1996). Two scenarios are considered to evaluate the LC of

ethanol from sawdust (Table 5.1). Sawdust from sawmills (base case: S1) and sawdust produced

63

from forest residues (thinning, pruning, shaving etc.: S2) are considered as carbon neutral,

because these are byproducts of timber industry (sawmill) and forest, respectively. However, any

marginal inputs (energy consumption in collection and transportation of sawdust, and preparation

of sawdust from forest residues) and emissions are allocated to sawdust from sawmills or

sawdust produced from forest residues (thinning, pruning, shaving etc.). GHG emission has been

calculated interms of CO2 e (i.e., GWP for a time span of 100 year; IPCC, 2001).

Table 5.1 Scenarios of this study.

Scenario Feedstock source Transport distance, km ^Cost, $/t

Scenario-1 (base case)

Sawdust (sawmill byproduct)

15* 80

Scenario-2

Sawdust (produced from

forest residues: thinning,

pruning, and the logging

residues)

20

60

*Distance estimated based on the Mani et al., 2006; 6 m (20 feet) tailor truck is used for transportation; ^plant gate

price.

An ethanol processing plant is assumed to be established nearby the forest/sawmill area for

efficient utilization of forest and sawmill residues. The ethanol processing plant capacity is

considered to be 20000 kL/year. Biomass is noted to have a low bulk density and higher

moisture content ranging from 10–70%. The moisture content in raw material (sawdust/forest

residues) is considered to be 40% (wb). The transportation distance is assumed to be 15 km (base

case: sawdust transportation distance is estimated based on feedstock demands and the

transportation distance reported by Mani et al., 2006). The bagged sawdust assumed to be

transported by 6 m (20 feet) trailer truck. The transportation capacity is calculated based on the

density of sawdust (417 kg/m3) and volume of the trailers. The loading capacity is assumed to be

75% of the volume of the trailer. Cradle to gate scenario [system boundary of LC of ethanol from

sawdust: sawdust either from sawmills (transportation) or sawdust prepared from forest residues

(collection and transportation) followed by pretreatment, saccharification and fermentation,

distillation and purification, and waste management] has been adopted for this study (Fig. 5.1).

The environmental impacts related to the construction of the ethanol processing plant, storage

facilities and the production of transportation and other machines, building and roads are not

considered. It is also worthy to note that energy input in the form of labor and energy content in

64

the feedstock are not taken into account. Net energy consumption is defined as the difference

between the sum of the energy consumption in each process and the amount of energy recovered

from the lignin byproduct (hereafter referred to byproduct).

Figure 5.1 Schematic diagrams of the LC of sawdust and the system boundary of this study

5.2.2 Ethanol production

5.2.2.1 Pretreatment

Lime pretreatment (calcium capturing by carbonation, i.e., CaCCO process at 120°C for 1

h; lime 10%) is considered for this study (Park et al., 2010; Shiroma et al., 2011). The solid

concentration during pretreatment is considered to be 30% (w/w).

5.2.2.2 Fermentation and distillation

The pretreated sawdust slurry (solid content 10% wt) is then allowed for simultaneous

saccharification and fermentation (SSF) at 33° for 72 h. The enzyme loading is considered to be

14 FPU (filter paper unit)/g-cellulose (McMillan et al., 1999). Vacuum extractive fermentation

and distillation, and purification (using glycerol) processes are adopted (Dias et al., 2009;

Junqueira et al., 2009a). The ethanol concentration in the vacuum extractive fermentation is

assumed to be more than 7.5% (wt).

5.2.2.3 Enzyme (cellulase) and yeast production

Energy consumption in enzyme production process is calculated based on the enzyme

production cost (enzyme loading: 15 FPU/g-cellulose i.e. 19263 FPU/L) and retail electricity

price in the USA in 1997 (Wooley et al., 1999a; EIA, 2010). Then, enzyme cost of this study is

worked out (based on the 2012 electricity price in Ontario). The cost of yeast ($0.01/gallon) is

collected from the literature (Dutta et al., 2010a).

5.2.2.4 Waste management

The waste stream is assumed to be separated into centrifuged solids (lignin) and liquid

streams (waste water). The lignin is assumed to be dried by utilizing the wasted heat from the

C

olle

ctio

n &

pre

par

atio

n

Tran

spo

rtat

ion

Pre

-tre

atm

ent

Sacc

har

ific

atio

n

& F

erm

enta

tio

n

Dis

tilla

tio

n &

p

uri

fica

tio

n

Was

te

man

agem

en

t

65

boiler. Anaerobic digestion of the wastewater produces biogas (Cardona & Sa´nchez, 2006). It is

assumed that energy consumption and cost incurred in waste management processes would be

offset by biogas, and the byproduct recovered in the waste management processes can be used

for process heat generation. Based on the heating value of lignin and the boiler efficiency (80%;

Mani et al., 2010) heat generation from the byproduct is estimated and used to offset some of the

heat supplied by LNG (liquid natural gas). The emission and cost that credited to lignin is

determined with the emission factor and cost of LNG.

5.2.3 Cost analysis

The economic life of the ethanol processing plant and the yearly operating period are

assumed to be 20 years and 350 operation days, respectively (Dutta et al., 2011; Huang et al.,

2009; Wu et al., 2006). The business capital is assumed to be equivalent to twice of the price of

yearly feedstock demand. Both fixed costs (straight line depreciation on installation, labor,

maintenance and interest on investment: $0.3/L) and variable costs (feedstock, enzyme

production, utilities and waste management: $0.78/L) are taken into account to estimate the

production cost of ethanol. The yearly interest rate on investment is assumed to be 3% and

maintenance cost is 2% of processing plant cost. Sawmill residues, particularly sawdust, are

noted to be abundant with low commercial value (about $20/t at processing plant gate; Mani et

al., 2006). However, higher cost of sawdust ($100) has also been reported (Millman, 2008). The

cost of sawdust is considered to be about 80 $/t-wet. The cost of forest residues is assumed to be

75% of the price of sawdust.

5.2.4 Data collection

Both the estimated and literature data are used to evaluate the LC of ethanol. Although the

inventory data vary depending on the regions of the study and technologies used, the data used

are assumed to be valid for this study. A summary of the parameters/processes for which data

have been collected from literature are reported in the Supporting Information (Table 5.2).

66

Table 5.2 Summary of parameters for which data are collected from literature

Parameters/Systems Actual data Sources

Ethanol plant construction $38 million Asano & Minowa, 2007

No. of labor (persons) 23 Asano & Minowa, 2007

*Forest residue collection (L/t) 1.95 JA-Zenno, 2002

Sawdust production from forest residues (kWh/t) 20.56 TianYuan, 2013

Enzyme:

Energy consumption kWh/L 0.802 Wooley et al., 1999a

Material cost ($/L) 0.014 Wooley et al., 1999a

Yeast ($/gallon) 0.01 Dutta et al., 2010a

Vacuum extractive fermentation & distillation

(MJ/kg hydrous ethanol)

7.525 Junqueira, et al., 2009a

Purification (MJ/kg ethanol) 1.085 Dias et al., 2009

Ethanol yield (L/kg-dry sawdust) 0.305 Olsson et al., 1996

Note: Plant capacity is 20000 kL/year; labor cost $46000/person/year; *assumed to be same that of straw

collection; boiler efficiency 80% (Mani et al., 2010).

5.3 Results and discussion

5.3.1 Net energy consumption and CO2 emission

The energy consumption (fossil fuel and electricity) in each stage of the LC of ethanol and

the energy output from byproduct are calculated to estimate the net energy consumption (Fig.

5.2). Utilization of byproduct offsets some of the energy consumption and has a strong

contribution to the net energy consumption. Energy consumption remains the same for different

stages of the LC of ethanol for both scenarios except collection, transportation and pretreatment.

Energy consumption in collection, transportation and sawdust preparation are estimated to be

0.41, 0.15 and 0.64 MJ/L for scenario-2. On the other hand, for scenario-1 energy consumption

in transportation is estimated to be 0.11 MJ/L and no energy is consumed for collection and

preparation, because sawdust is assumed to be readily available at the sawmill gate. Depending

on the scenarios of the study, the estimated net energy consumption varied from 12.29–13.37

MJ/L. A slight variation in net energy consumption is observed because of the difference in

assumed transportation distance, and energy consumption in collection and sawdust preparation

67

when forest residues are used (scenario-2). The pretreatment process is observed to be the main

hotspot followed by distillation, and others in the case of energy consumption. Net energy

consumption is found to be within the range of reported net energy consumption of cellulosic

ethanol (Kim & Dale, 2005; Roy & Dutta, 2012). The variation in net energy consumption might

be because of different assumptions and feedstock (for example, earlier authors has used corn

stover as a feedstock and byproduct is used for electricity generation; however the latter has used

wheat straw as a feed stock and byproduct is used for heat generation) have been used.

Figure 5.2 Energy breakdown of the life cycle of ethanol

Figure 5.3 depicts emission breakdown of the LC of ethanol. The GHG released from

different stages are directly related to energy and resource consumption. The main hotspot is

observed to be the pretreatment followed by distillation and other stages. The byproduct (lignin)

recovered and used for heat generation have a strong contribution to offsetting some of the

emission from different stages. The net emission is estimated to be 0.75–0.92 kg CO2 e/L. Wang

et al., (2012) noted that energy recovery is dependent on the waste management scenarios, thus

the net emission may vary, if different waste management scenarios are considered. A slight

variation in net GHG emission is also observed mainly because of the difference in energy

consumption for collection (forest residues) and pretreatment (includes sawdust preparation from

forest residues) resulting from the assumptions for different scenarios. It seems environmental

-10

-5

0

5

10

15

20

MJ/

L

Byproduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Collection

S1 S2

68

benefit can be gained from sawdust ethanol relative to gasoline (1.5 kg CO2 e/L-ethanol

equivalent in Canada; Environment Canada, 2011), if the ethanol can be produced by using the

technologies considered for this study, which seems to be supported by the findings of other

researcher (Schmer et al., 2008). However, the environmental benefit drastically improved if

carbon sequestration is considered (Fig. 5.4). The estimated net emission varied from -0.54 to -

0.72 kg-CO2 e/L. Figure 5.4 also depicts that significant environmental benefit (GHG) can be

achieved compared with the gasoline. It is also worthy to note that natural forest is considered to

be the source of sawdust for both scenarios. The environmental benefit from the LC of ethanol

derived from sawdust may change if source is considered to be the purpose grown forest. The net

GHG emissions reported to be -885 kg-CO2 e/t in the case of yardwaste and carbon sequestration

contributed about about 62–66% to net GHG emissions reduction (Roberts et al., 2010), which

also support the findings of this study.

Figure 5.3 Emission breakdown of the life cycle of ethanol

5.3.2 Production cost

Figure 5.5 represents the cost breakdown of different stages of the LC of ethanol. The

main hotspot has emerged to be the feedstock, followed by fixed cost and others. The feedstock

cost is estimated to be $0.33–0.44/L depending on the source of feedstock. The fixed cost is

calculated to be $0.25–$0.26/L for a plant capacity of 20000 kL/y, which may vary, if different

processing plant sizes, operating periods and life spans are considered. Although little variation

-0.5

0.0

0.5

1.0

1.5

Em

issi

on, kg C

O2e/

L

Byproduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Collection

S1 S2

69

is observed in feedstock and fixed cost due to different assumptions. The net production cost is

estimated to be about $0.98–$1.04/L because of the variation in feedstock-, fixed-, collection-,

transportation- and pretreatment cost. The estimated production cost indicates that ethanol from

sawdust will have a steep competition with its counterpart (i.e., fossil gasoline). The production

cost of ethanol is noted to be dependent on assumption, feedstock cost, allocation methods and

reference systems (Cherubini & Strømman, 2011). Consequently, a direct comparison was not

attempted with other studies. However, the production cost of this study is observed to be within

the range of reported production cost (0.55–1.45 US$/L) of lignocellulosic ethanol (Barta et al.,

2010a; Dutta et al., 2010a; Roy et al., 2012b). The variation in production cost among different

studies might be because of differences in ethanol yield, feedstock cost, processing plant size,

system boundary and assumptions (Gnansounou & Dauriat, 2010; Cherubini and Strømman,

2011). It is worthy to note that this study has been conducted based on the estimated and

literature data, and all the data do not necessarily correspond to the Canadian context or to the

same processing plant size. Consequently, in-depth studies are needed for any future investment

and commercial production from sawdust/forest residues in Canada.

Figure 5.4 Effect of carbon sequestration on the net emission of the life cycle of ethanol

-0.8

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0

S1 S2

Em

issi

on, kg C

O2e/

L

70

Figure 5.5 Cost breakdown of the life cycle of ethanol

5.3.3 Sensitivity analysis

Results of an LCA are dependent on the energy and resource input at each stage of the LC

of ethanol. It is also noted that LCA results are more sensitive to the changes in parameters

related to the biomass and ethanol yield. Consequently, sensitivity of the variation in energy

consumption at each stage (transportation, pretreatment, SSF, distillation) and ethanol yield is

investigated. The variation of each parameter is assumed to be ±20% compared to the base case.

It is observed that a change in the parameters also resulted in a change in net energy

consumption, emission and production cost are changed (Figs. 5.6–5.8). The net energy

consumption is observed to be dependent on the variation in energy consumption at each stage;

however, the effect of ethanol yield is confirmed to be more sensitive compared to others. Any

increase in considered parameters, net energy consumption, emission and production cost are

also increased except the ethanol yield. The net energy consumption, emission and production

cost decreased with an increase in ethanol yield. Moreover, this study revealed that feedstock and

fixed cost are the main contributors to the production cost of ethanol followed by pretreatment.

Therefore, it is imperative that feedstock and fixed cost need to be mitigated to diminish

production cost of ethanol. At present this can only be achieved with implementing modified

agro-industrial and environmental policy (it may be in the form of FiT program).

-0.3

0.0

0.3

0.6

0.9

1.2

Cost

, $/L

Byproduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Collection

Feedstock

Fixed cost

S1 S2

71

Scenario-1 Scenario-2

12.29

11.54

13.42

12.31

12.29

13.59

11.03

13.24

11.34

13.03

11.55

0

5

10

15Base

Yield increase

20%

Yield decrease

20%

Transport

distance

increase 20%

Transport

distance

decrease 20%

Pretreatment

increase 20%

Pretreatment

decrease 20%

SSF increase

20%

SSF decrease

20%

Distillation

increase 20%

Distillation

decrease 20%

0.75

0.70

0.81

0.75

0.75

0.82

0.67

0.80

0.69

0.80

0.69

0

0.3

0.6

0.9Base

Yield increase 20%

Yield decrease

20%

Transport distance

increase 20%

Transport distance

decrease 20%

Pretreatment

increase 20%

Pretreatment

decrease 20%

SSF increase 20%

SSF decrease 20%

Distillation

increase 20%

Distillation

decrease 20%

Figure 5.6 Effect of the change in energy consumption at different stages on net energy consumption (MJ/L).

Figure 5.7 Effect of the change in energy consumption at different stages on net emission (kg-CO2 e/L)

72

Figure 5.8 Effect of the change in energy consumption at different stages on net cost ($/L)

Figure 5.9 Effect of the changes in feedstock- and fixed cost on the production cost of ethanol

1.04

0.93

1.19

1.04

1.04

1.08 1.00

1.06

1.01

1.05

1.02

0

0.3

0.6

0.9

1.2Base

Yield increase 20%

Yield decrease 20%

Transport distance

increase 20%

Transport distance

decrease 20%

Pretreatment

increase 20%

Pretreatment

decrease 20%

SSF increase 20%

SSF decrease 20%

Distillation increase

20%

Distillation decrease

20%

1.04

0.93

0.82

0.71 0.97

0.91

0.84

0

0.3

0.6

0.9

1.2Base

Feedstock-25%

Feedstock-50%

Feedstock-75%Fixed cost-25%

Fixed cost-50%

Fixed cost-75%

73

The Government of Ontario has launched a Feed-in-Tariff (FiT) program to encourage

people and organizations to develop renewable energy projects, especially producing electricity

from renewable sources (OPA, 2013; Sanscartier et al., 2009). The FiT is varied depending on

the source of renewable energy ($0.1–$0.55/kWh). The price of produced electricity is

guaranteed for 20 years (40 years for waterpower). In Canada, the renewable energy sector is not

only subsidized but also fossil fuel (Milan, 2010). A similar FiT program in renewable liquid

biofuels sector of Ontario would reduce their production costs. Therefore, we assumed that fixed

and feedstock cost can be reduced to 25–75% of the base case with the introduction of FiT

program. The production cost of ethanol significantly reduced with the introduction of FiT

program (Fig. 5.9), which confirmed that economically viable ethanol can be produced from

sawdust if FiT program for liquid biofuels is implemented in Ontario. The production cost is

estimated to be 0.51 $/L (in reference to the base case at plant gate) if $0.025/MJ FiT is

considered. If fossil gasoline can be replaced by this amount of incentive to ethanol industry, the

estimated emission abatement cost would become about $400/t CO2 e. However, any reduction

in this incentive to replace fossil gasoline emission abatement cost will also be reduced. It seems

that ethanol industry will emerge as an economically viable/profitable rural agro-industry for the

producers/investors, if a similar FiT program that of electricity from renewable sources is

implemented. It may also help in achieving the GHG emission reduction target set by Canada,

improving energy security and combating global warming potential, simultaneously creating

rural employment opportunity and enhance rural economy in Ontario. Finally, FiT program

would help Ontario become a leader in the ethanol market.

5.4 Conclusion

This study reveals that despite estimated environmental benefit of ethanol that produced

from sawdust with adopted technologies, its economic viability remains doubtful unless FiT

program is considered. A modified agro-industrial and renewable energy policy that allows FiT

to the lignocellulosic ethanol industry in Ontario not only reduces production cost but also may

encourage future investment and create more green jobs as well as help in achieving committed

GHG emission reduction targets in Canada.

74

Chapter 6

Evaluation of the Life Cycle of Ethanol derived from Miscanthus in Ontario

[Submitted to the Biomass and Bioenergy]

6.1 Introduction

The rising cost of fossil fuels and the geo-political uncertainty associated with their supply

chains, and growing concerns about climate change led to recognize liquid biofuels as an

alternative to fossil fuels for transportation. In 2010, secondary transportation fuel consumption

is reported to be 2595 PJ in Canada, a growth of 38.2% from 1990 level (Natural Resource

Canada, 2013). However, in 2010, ethanol contributed only 1.7% and 3.2% to total

transportation energy and motor gasoline, respectively (Natural Resource Canada, 2013) (Fig.

6.1), which depicts the substantial demand of ethanol in Canada. Renewable energy not only

provides significant environmental benefits but also enhances rural economies (Kim & Dale,

2003; Spatari et al., 2005; Farrell et al., 2006). Production of biofuels from biomass (agricultural

and forest residues, and energy crops) has been emphasized, because it does not compete with

food or feed (Zaldivar et al., 2001; Gray et al., 2006; Hahn-Hägerdal et al., 2006; Sánchez &

Cardona, 2008).

Although ethanol production from biomass has been emphasized, concern about soil

fertility and structural stability and restricts the collection of agricultural residues (Sheehan et al.,

2002; Blanco-Canqui & Lal, 2009). Farmers are also reluctant to the removal of crop residues

from their farms (Tyndall et al., 2010). In Canada, the potential sources of renewable biomass

include agricultural residues, municipal solid waste, forestry byproducts and energy crops. The

technical, economic and sustainability constraints in Ontario conditions also limits their supply

to ethanol industry (Kludze et al., 2010). The use of forestry wastes for liquid biofuels are also

restricted due to its enormous demand by the solid biofuels industries. Miscanthus is a promising

energy crop with high yield and energy content, which can be grown on low quality or marginal

land, and add carbon to the soil and safeguard it against erosion (Somerville et al., 2010; Kludze

et al., 2013), has an important role in sustainable energy production (Sørensen et al., 2008;

Khanna et al., 2008; Bocquého & Jacquet, 2010).

Kludze et al (2013) noted that Ontario has an adequate land base for producing miscanthus

to meeting/surpassing numerous viable uses of biomass without significantly affecting food

75

crops supply. Although life cycle assessment (LCA) methodology has been extensively used to

evaluate the life cycle (LC) of lignocellulosic ethanol, the LCA of ethanol from miscanthus has

received limited attention (Fazio & Monti, 2011; Scown et al., 2012; ), which are mostly deal

with biorefinery, crop location and agricultural practices. This study evaluates the LC of ethanol

produced by enzymatic hydrolysis considering three scenarios to determine if environmentally

preferable and economically viable ethanol can be derived from miscanthus in Ontario, Canada.

6.2 Methodology

6.2.1 Study area, system boundary and assumptions

Ontario is located in east-central region (48˚N to 83˚W) of Canada (60˚N to 95˚W) and

consists of two major regions (southern and northern). The southern region is further divided into

four sub-regions, namely southern-, western-, central-, and eastern Ontario (Fig. 6.2). The land

area in Ontario is 91.8 million ha with 4.4 million ha of tillable land, and only about 3.6 million

ha is arable for growing conventional crops. The land classes are scattered throughout various

regions (Fig. 6.3) and the area under different land classes varies from region to region (Table

6.1).

Figure 6.1 Transportation fuel consumption and contribution of ethanol in Canada

76

Figure 6.2 Different regions in Ontario, Canada

Figure 6.3 Different regions and land classes in Ontario, Canada

Biomass regions

Southern Ontario

Western Ontario

Eastern Ontario

Central Ontario

Land classes

77

Table 6.1 Land areas in Ontario, ha

Regions Land class and tillable land area, ha

Class 1 Class 2 Class 3 Class 4 Class 5 Total

Southern Ontario 238102 876664 414109 37138 39332 1605345

Western Ontario 724831 254067 239130 73112 130091 1421231

Central Ontario 165830 113526 122839 118976 84424 605595

Eastern Ontario 32005 312567 304841 148471 65166 863050

Source: Kludze et al., 2013

Availability of crop residues are reported to be limited because of technical, economic and

sustainability constraints in Ontario conditions (Kludze et al. 2010). On the other hand, Ontario

has adequate land base for producing energy crops, especially miscanthus to meet/surpass

diverse uses (Kludze et al., 2013). The land area in Ontario is 91.8 million ha with 4.4 million ha

of tillable land, and only about 3.6 million ha is arable for growing conventional crops (Statistics

Canada 2008; OMF, 2013). The land in Ontario has been classified into seven classes (A-6-1). It

is also grouped as the prime- (classes 1, 2 & 3) and marginal lands (classes 4 & 5). The

productivity of miscanthus is dependent on the type of land (land classes) (Table 6.2).

Table 6.2 Land classes, soil types and miscanthus yield

Land class

Description Soil type Yield, dry-t/ha

Class 1

Suitable for field crops Silt loam-clay loam 11.14

Class 2 Suitable for field crops with

moderate limitations Silt loam-clay loam 11.14

Class 3 Suitable for field crops with

moderately severe limitations Sandy-clay 10.03

Class 4 Has severe limitations for field

crops Sandy-loam 8.9

Class 5 Very severe limitations for

field crops Loam 7.8 Prime land: classes 1, 2 & 3; Marginal land: classes 4&5 (Source: Kludze et al., 2013)

78

Three scenarios are established to evaluate the LC of ethanol derived from miscanthus

(Table 6.3). Miscanthus yield, net emission from cultivation and feedstock cost for various

scenarios are assumed to be the average yield, emission and cost of corresponding land classes,

respectively. Cradle to gate scenarios is adopted to outline the system boundary of this study

(Fig. 6.4). Both the estimated and literature data are used to evaluate the LC of ethanol (Table

6.4). Infrastructure construction, production of transportation and other machineries, and energy

input in the forms of labor and feedstock are not considered. Emission has been estimated in

terms of CO2 e (i.e., GWP for a time span of 100 year; IPCC, 2001). Net energy consumption is

determined based on the difference between the sum of the energy consumption in each process

and the amount of energy recovered from the lignin (hereafter referred to byproduct). Net

emission is the difference between the emission from input energy and the sum of carbon

sequestration and the amount of emission offset by the byproducts recovered from the system.

Table 6.3 Scenarios of this study.

Scenarios Descriptions Yield, Feedstock cost On farm emission,

tDM/ha $/tDM kg CO2e/tDM

S1

All classes of land are used

for miscanthus cultivation 10.02 66.19

54.62

S2

Only prime land is used for

miscanthus cultivation 10.86 64.56

16.94

S3

Only marginal land is used

for miscanthus cultivation 8.35 71.50

167.67

Prime land: Classes 1, 2 & 3; Marginal land: Class 4 & 5 (Source: Kludze et al., 2013; Sanscartier et al., 2013).

6.2.2 Miscanthus cultivation

Miscanthus has received increasing attention as a source of renewable energy because of

its high productivity, and its proven economics as an energy crop in Europe and USA (Khanna et

al., 2008; Bocqueho & Jacquet, 2010; Sherrington & Moran, 2010; Kludze et al., 2013). After

planting, miscanthus can stands for fifteen to twenty years on the farm, and has a very low agro-

chemical requirement (DEFRA, 2007). Miscanthus grows in tropical, subtropical and temperate

regions; however does not grow at a temperature below 6°C (AEBIOM, 2010). Thus, miscanthus

can be grown in all the regions of Ontario and its cultivation is noted to be economically viable

79

in Ontario (Vyn et al., 2012; Kludze et al., 2013). Miscanthus stands persist for fifteen to twenty

years. Miscanthus yield increases for the first 3 years and then remains constant for the

remaining years. Yield varies from region to region. Miscanthus has very low agro-chemical

requirement (DEFRA, 2007). The hybrid species yield more compared to others.

Table 6.4 Summary of parameters for which data are collected from literature

Parameters/Systems Actual data Sources

Miscanthus cultivation (kg CO2e/tDM)

16.94 to

167.67 Sanscartier et al., 2013

Feedstock cost ($/tDM) 64.56 to 71.5 Kludze et al., 2013

*Crushing (size 3 mm) (kWh/kg) 0.06095 Roy et al., 2012a; b

Ethanol processing plant construction $38 million Asano & Minowa, 2007

No. of labor (persons) 23 Asano & Minowa, 2007

Enzyme:

Energy consumption kWh/L 0.802 Wooley et al., 1999

Material cost ($/L) 0.014 Wooley et al., 1999

Yeast ($/gallon) 0.01 Dutta et al., 2010a

Vacuum extractive fermentation & distillation

(MJ/kg hydrous ethanol)

7.525 Junqueira, et al., 2009

Purification (MJ/kg ethanol) 1.085 Dias et al., 2009

Ethanol yield (L/kg-dry miscanthus)

0.305 Nilsson et al., 2008; Li et

al., 2013; Zhang et al.,

2012; DOE, 2006

Note: Plant capacity is 20000 kL/year; labor cost $46000/person/year; *assumed to be same that of straw;

boiler efficiency 80% (Mani et al., 2009); $: Canadian dollar.

Inputs in miscanthus cultivation and harvesting stage are reported in supporting

information (A-6-2 & A-6-3). Typically, miscanthus is harvested in late winter or early spring;

thus, contains low moisture at harvest (15–20%) (DEFRA, 2007; Kludze et al., 2013; Sanscartier

et al., 2013). A portion of agricultural land in northern Ontario has been excluded in this study

because of its short growing periods (100–145 frost free days). The net emission from agriculture

80

(hereafter referred to feedstock) is estimated based on energy inputs in agriculture and the carbon

dynamics i.e., carbon storage in all pools (A-6-4).

Figure 6.4 Schematic diagrams of the life cycle of sawdust and the system boundary of this study

6.2.3 Transportation

Harvested and baled miscanthus is transported to the ethanol processing plant by 6 m (20

feet) tailor trucks (loading capacity is assumed to be 75%). The moisture content in harvested

miscanthus is considered to be 15% and the bulk density of baled miscanthus is 218 kg/m3 (Vyn

et al., 2012). Fuel consumption in the transportation process is calculated based on the loading

capacity, feedstock moisture content and specific volume of baled miscanthus. The ethanol

processing plant capacity is assumed to be 20000 kL/year. Miscanthus transportation distance

from farms to the ethanol processing plant is estimated based on following equations (Eqs. 1 &

2) (Huang et al., 2009). Roads and channels are assumed to be 2% of tillable land at the rural

area. The transportation distance depends on the scenarios and it varies from is 17.7–48.0 km.

Radius of the area, R km = {F/(πfaflcY)} /2 . . . . . (Eq.6. 1)

where, F = Annual feedstock demand, t

π = constant

fa = fraction of total farmland from which feedstock can be collected or produced

flc = fraction of surrounding farmland containing crops

Y = biomass yield per unit area (dry), t/km2

Transportation distance (collection center–processing plant), D km = 2Rfw/3 . (Eq. 6. 2)

where, fw = road winding factor (assumed to be 1.3; Sokhansanj & Turhollow, 2002)

6.2.4 Ethanol production

6.2.4.1 Pretreatment

Pretreatment, either physicochemical or chemical or biological is a prerequisite to improve

components digestibility and improve the ethanol yield from miscanthus (Brosse et al., 2009).

Was

te

man

agem

ent

Ag

ricu

ltu

re

&co

llec

tio

n

Tra

nsp

ort

atio

n

Dis

till

atio

n &

p

uri

fica

tio

n

Pre

trea

tmen

t

Sac

char

ific

atio

n

& f

erm

enta

tio

n

81

Although different pretreatments are employed to miscanthus [ammonia fiber expansion, acid

hydrolysis, NaOH pretreatment, wet explosion etc. and liquid hot water (LHW)] pretreatment

(Sørensen et al., 2008; Murnen et al., 2008; Han et al., 2011; Vanderghem et al., 2012; Khullar et

al., 2013; Li et al., 2013), the LHW pretreatment with lime is given to the crushed miscanthus (3

mm) in this study (Shiroma et al., 2011). The solid concentration during pretreatment is

considered to be 30% (w/w).

6.2.4.2 Fermentation and distillation

Vacuum extractive fermentation and distillation, and purification (using glycerol as an

additive) processes are employed (Dias et al., 2009; Junqueira et al., 2009a). The pretreated

miscanthus slurry (solid content 10% wt) is allowed for simultaneous saccharification and

fermentation (SSF) at 33°C for 72 h. Although the reported enzyme loading for ethanol

production varies from 10–20 FPU/g-cellulose (Gregg et al., 1996; Brosse et al., 2009), the

enzyme loading is assumed to be 8.04 FPU/g miscanthus. Similarly, the ethanol yield also varies

from 0.189–0.427 L/kg (DOE, 2006; Nilsson, 2008; Zhang et al., 2012; Li et al., 2013). The

ethanol yield is considered to be 0.305 L/kg-dry feedstock.

6.2.4.3 Energy consumption in enzyme and yeast production

Energy consumption in enzyme production process is estimated based on the reported

enzyme production cost (Wooley et al., 1999a). Then, the enzyme cost of this study is estimated

(based on the 2012 electricity price in Ontario) (A-6-5); however, the production cost of yeast

($0.01/gallon) is derived from the literature (Dutta et al., 2010a).

6.2.4.4 Waste management

The waste stream of lignocellulosic ethanol consists of solids (lignin) and liquid

(wastewater). The waste stream is centrifuged into lignin and liquid streams. Anaerobic digestion

of the wastewater produces biogas (Cardona & Sa´nchez, 2006) and the dried lignin (dried by

utilizing wasted heat from boiler) is combusted in boiler for heat recovery (boiler efficiency is

considered to be 80%; Mani et al., 2010). It is assumed that biogas offsets the energy

consumption in waste management processes and heat recovered from lignin offsets some of the

process heat supplied by liquid natural gas (LNG). The emission and cost that credited to

recovered heat is ascertained with the emission factor and cost of LNG.

82

6.2.5 Cost analysis

The production cost of ethanol is estimated based on both fixed costs (straight line

depreciation on installation, labor, maintenance and interest on investment) and variable costs

(feedstock, enzyme production, utilities and waste management). The economic life span of the

processing plant and the yearly operating period are assumed to be 20 years and 350 operation

days, respectively (Dutta et al., 2011; Huang et al., 2009; Wu et al., 2006). The business capital

is assumed to be equivalent to twice of the price of yearly feedstock demand. The yearly interest

rate on investment is assumed to be 3% and maintenance cost is 2% of processing plant cost. The

cost of miscanthus is considered to be about 63–74 $/tDM depending on the scenarios (Kludze et

al., 2013).

6.3 Results and discussion

6.3.1 Net energy consumption

The energy consumption in each stage of the LC of ethanol and the energy recovery from

byproducts are estimated to represent energy breakdown of ethanol production process (Fig. 6.5).

Feedstock cultivation is emerged to be the main hotspot followed by pretreatment, distillation,

enzyme, SSF, yeast, and transportation for each scenario. A slight variation in transportation

energy consumption is observed because of the difference in transportation distance among the

scenarios. The transportation distance is found to be shortest and longest for scenario-1 (S1) and

scenario-3 (S3), respectively, because of the difference among the corresponding land area. The

net energy consumption varies from 13.02–13.34 MJ/L. The byproduct offsets a part of energy

consumed in the process, which has a strong contribution to net energy consumption for all

scenarios. The net energy consumption is noted to be sensitive to coproduct allocations and

assumptions of the study (Pimentel & Patzek, 2005; Roy & Dutta, 2012). Thus, the net energy

consumption varied from other studies because of different feedstock is used and various

assumptions are made in those studies (Sheehan et al., 2003; Kim & Dale, 2005; Roy et al.,

2012a,b; Roy & Dutta, 2012). However, the net energy consumption may vary, if different

pretreatment methods are adopted.

83

Figure 6.5 Energy breakdown of the life cycle of ethanol derived from miscanthus

6.3.2 Greenhouse gas emission (CO2e)

Figure 6.6 shows the emission breakdown of the LC of ethanol for different scenarios. The

emissions are directly related to energy and resource consumption (fossil energy and electricity)

at different stages except feedstock because carbon dynamics offsets a major part in this step.

The pretreatment process is emerged to be the main hotspot followed by distillation, enzyme,

SSF, yeast, and transportation except S3 where feedstock is emerged as the main hotspot because

of the positive carbon dynamics (i.e., carbon released to the atsmosphere). Consequently,

emission from feedstock is found to be 0.18, 0.06 and 0.55 kg/L for S1, S2 and S3, respectively. A

slight variation in CO2 emission is observed in the case of transportation, caused by the

difference in transportation distance among different scenarios. The carbon dynamics is noted to

be dependent on the land classes, crop replacement/rotation and biomass yield (Kludze et al.,

2013; Sanscartier et al., 2013). The emission from the feedstock is the highest in the case of

marginal land (S3) and the lowest for prime land (S2). These variations might not only because of

land type but also region and crop rotation, because various crop rotations are considered for

different scenarios (A-6-4).

-10

-5

0

5

10

15

20

Ener

gy,

MJ/

L

Bybroduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Feedstock

S1 S2 S3

84

Figure 6.6 Emission breakdown of the life cycle of ethanol derived from miscanthus

The net emissions from the LC of ethanol are estimated to be 0.92, 0.79 and 1.31 kg/L for

scenarios S1, S2, and S3, respectively which are seem to be dependent on the carbon dynamics. It

is worthy to mention that the byproduct used for heat recovery, and the carbon dynamics has a

robust contribution to offsetting a part of emission of the LC of ethanol. The estimated net

emissions from the LC of ethanol reveal that environmental benefit can be achieved even

miscanthus is grown on the marginal in Ontario and ethanol produced by adopting the

technologies considered in this study.

6.3.3 Net production cost

The production cost breakdown depicts that fixed cost is the main hotspot followed by the

feedstock, pretreatment, distillation, enzyme, SSF, yeast and transportation (Fig. 6.7). There is a

slight variation in fixed-, feedstock-, and transportation cost which is yielded by the difference in

feedstock production cost and the transportation distance among the scenarios. The fixed-

feedstock-, and transportation cost varied from 0.246–0.247, 0.212–0.235, and 0.006–0.017 $/L,

in the case of scenarios S1, S2, and S3, respectively. The net production cost for S1, S2, and S3 are

estimated to be 0.80, 0.79, and 0.83$/L, respectively. It is important to note that the breakeven

-0.6

0.0

0.6

1.2

1.8

Em

issi

on, kg C

O2e/

L

Bybroduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Feedstock

S1 S2 S3

85

feedstock cost is considered for the production cost estimation, which may vary depending on

the biomass logistics and farmers profit margins.

Figure 6.7 Cost breakdown of the life cycle of ethanol

Although a wide variation in production cost of lignocellulosic ethanol is reported, the

estimated production cost of this study is observed to be reasonable and comparable with those

studies (Barta et al., 2010a; Dutta et al., 2010a; Klein-Marcuschamer et al., 2010; Kumar &

Murthy, 2011; Roy et al., 2012a,b; Roy & Dutta, 2012, 2013). The enzymes cost for converting

biomass into ethanol in enzymatic hydrolysis process is also noted to be a major hindrance to the

development of an economically viable lignocellulosic ethanol industry (Roy et al., 2012b;

Banerjee et al., 2010b; Merino et al., 2007). The production cost of ethanol is also noted to be

dependent on the conversion technology, enzyme loading, feedstock, allocation methods and

plant sizes (Gnansounou & Dauriat, 2010; Roy et al., 2012a,b; Roy & Dutta, 2012);

consequently, production cost may vary, if different conversion methods, allocation, and plant

sizes are considered. This study indicates that miscanthus is a promising feedstock for ethanol in

Ontario, Canada.

-0.2

0.0

0.2

0.4

0.6

0.8

1.0C

ost

, $/L

Byproduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Feedstock

Fixed cost

S1 S2 S3

86

6.3.4 Sensitivity analysis

LCA methodology identifies the potential hotspots of product, production system and

activity. The potential hotspot of this study is identified to be the feedstock in case of energy

consumption and emission; however, fixed cost is emerged to be the main hotspot in the case of

production cost. A wide variation is also observed in the case of transportation distance (10–49

km) and carbon dynamics (-0.5 to 0.03 dry t C/ha year; Sanscartier et al., 2013) among the

scenarios. Liquid biofuels production is rapidly increasing, and affecting land allocation among

crops (Roy & Shiina, 2010), thus, commercial biofuel production may target higher-quality

lands, due to better profit margins and push the cereals and subsistence crops to the low-quality

land. Although the carbon dynamics of scenario S3 is greater than that of others (S1 and S2),

miscanthus can be grown on marginal lands which are not suitable for food crops (Kludze et al.,

2013; Sanscartier et al., 2013) and environmental benefit can be achieved compared to fossil

gasoline.

Carbon sequestration is the process of capture and long-term storage of atmospheric carbon

i.e., CO2. Carbon can be captured from the atmosphere through biological, chemical or physical

processes and depositing it in a reservoir. Carbon sequestration rate in soil depends on soil types,

farming system, soil management, weather, crop displacement, and biomass yield (aboveground

and belowground). Removal of crop residues or harvesting of biomass affect the soil nutrient

balance and depletes soil fertility, and crop productivity (Henao & Baanante, 2006; Haskins et

al., 2006). Lal (2009) also noted that harvesting of crop residues for biofuels feedstock

jeopardize soil and water resources which are already under great stress.

Aboveground minimum source carbon (MSC) requirement is noted to be 1.8±0.44 to

2.5±1.0 t/ha/year depending on the tillage systems (Johnson, 2006). Blanco-Canqui and Lal

(2009) suggested crop residues removal rate as low as 25%, beyond which soil fertility and

structural stability would be negatively affected. In contrast, it is also noted that about 50–60%

crop residues can be collected without deteriorating the soil quality and productivity (Smith,

1986; Sheehan et al., 2002; Jeschke, 2011). Beale & Long (1995) noted that at the end of the

growing season 39% of the biomass is partitioned to roots and rhizomes in the case of

miscanthus. This information indicates that miscanthus can be harvested for biofuels without

deteriorating the soil quality and productivity. Moreover, miscanthus can be economically

cultivated all over Ontario (Kludze et al., 2013; Kludze et al., 2010). The breakeven feedstock

87

cost has also been used for cost estimation, which may vary depending on the profit margin of

stakeholders and biomass logistics. Consequently, sensitivity of the variation in transportation

distance, pretreatment energy consumption, feedstock and fixed cost are investigated (based on

the third scenario: S3).

Figures 6.8 & 6.9 represent the effect of variation in transportation distance and

pretreatment energy consumption (variation are considered to be ±20 to ±60% for both case) on

the net energy consumption, emission and production cost, respectively. The net energy

consumption, emission and production cost varies from 10.7–17.3 MJ/L, 1.1–1.6 kg-CO2 e/L and

0.7–0.9 $/L, respectively, depending on the severity of the variation of different parameters. The

effect of the change in pretreatment energy consumption seems to be severe than that of

transportation distance.

Figure 6.8 Effect of the variation in transportation distance and pretreatment energy consumption

on the net energy consumption (MJ/L)

Figure 6.10 represents the effect of the variation (±10 to ±30%) of feedstock cost on the net

production cost. The net production cost varies from 0.75–0.90 $/L depending on the severity of

the variation. The production cost can be reduced if agri-industrial and environmental policies

are enacted to support the miscanthus based ethanol industry, especially miscanthus grows on the

marginal land.

0

6

12

18Preatment+20%

Preatment+40%

Preatment+ 60%

Preatment-20%

Preatment-40%

Preatment-60%

S3-BaseTransportation+20%

Transportation+40%

Transportation+60%

Transportation-20%

Transportation-40%

Transportation-60%

88

Figure 6.9 Effect of the variation in transportation distance and pretreatment energy consumption

on the net emission and production cost

Figure 6.10 Effect of feedstock and fixed cost (S/L)

0.0

0.5

1.0

1.5Pretreatment +20%

Pretreatment +40%

Pretreatment +60%

Pretreatment -20%

Pretreatment -40%

Pretreatment -60%

S3Transportation +20%

Transportation +40%

Transportation +60%

Transportation -20%

Transportation -40%

Transportation -60%

kg-CO2e/L $/L

0.0

0.3

0.6

0.9Feedstock cost+10%

Feedstock cost+20%

Feedstock cost+30%

Feedstock cost-10%

Feedstock cost-20%

Feedstock cost-30%

S3Fixed cost+10%

Fixed cost+20%

Fixed cost+30%

Fixed cost-10%

Fixed cost-20%

Fixed cost-30%

89

The carbon dynamics is dependent on the land classes, biomass yield, crop

rotation/replacement (Kludze et al., 2013; Sanscartier et al., 2013). Consequently, carbon

dynamics may vary if different crop rotation/replacement or variety of miscanthus is adopted.

The net emission from the LC of ethanol is found to be directly related with the carbon dynamics

(Fig. 6.11). Negative carbon dynamics (carbon stored from the atmosphere)) reduces the net

emission from the LC of ethanol. In contrast positive carbon dynamics (carbon released to the

atmosphere) increases the net emission from the LC of ethanol. Therefore, careful attention

should be in place in selecting miscanthus variety, crop rotation which not only help in reducing

net emission from the LC of ethanol, but may also improve farm economy and attract more

investment in the rural ethanol industry.

Figure 6.11 Effect of carbon dynamics on the net emission of the life cycle of ethanol

This study reveals that miscanthus grows on the marginal land in Ontario could be a

potential feedstock for lignocellulosic ethanol industries and avoid any competition with food

crops for prime land help Ontario improves her food and energy security, and enhance rural

economy. It is also noted that genetically modified (GM) crops can serve various purposes, such

as improving yield and increasing the share of useful components or decreasing the need for

chemical fertilizers or water, and improves farm income (Bennett et al., 2004; ScienceDaily,

2005). Consequently, GM miscanthus may further improve farm income, reduce net emission

0.0

0.4

0.8

1.2

1.6

2.0

-0.5 -0.3 -0.1 0.1

Em

issi

on, kg C

O2e/

L

Carbon dynamics, tC/ha-year

90

from the LC of ethanol and help Canada achieve renewable energy and emission reduction

target.

6.4 Conclusion

Ethanol derived from miscanthus is found to be environmentally preferable and

economically viable in Ontario. Although a slight variation is observed in the case of net energy

consumption and production cost among the scenarios, the variation is robust in the case of net

emission where carbon sequestration plays an important role. It is worthy to mention that both

the environment and economic benefit can be gained, even miscanthus is grown on the marginal

land (S3) in Ontario for ethanol. Thus, miscanthus grown on marginal land is emerged as a

promising feedstock for ethanol industry in Ontario, which may avoid any sort of competition

with food crops for better quality land, improve farm income and rural economy, and help

meeting the ethanol demand and achieving GHG emission target of Canada.

91

Chapter 7

Identification of suitable plant location for ethanol industry in Ontario,

Canada

7.1 Introduction

Evaluation of the life cycle (LC) of ethanol derived from miscanthus revealed that

miscanthus is a promising feedstock for environmentally and economically viable ethanol in

Ontario, Canada (Chapter 6). The emission and production cost are also noted to be dependent

on the feedstock, location, conversion technologies, plant sizes and biomass logistics (Wooley et

al., 1999; McAloon, et al., 2000; Gnansounou & Dauriat, 2010; Roy et al., 2012a,b; Roy &

Dutta, 2012). Consequently, this study attempts to identify the potential locations for the

miscanthus based ethanol plant in Ontario, Canada not only to abate emission but also to

minimize the production cost.

7.2 Materials and methods

7.2.1 Study area

The land area in Ontario is 91.8 million ha with 4.4 million ha of tillable land. The land has

been grouped into seven land classes (A-6-1) (AAFC, 2008; OMAF, 2013). Land in Ontario is

also classified as prime- (classes 1, 2 and 3) and marginal (classes 4 and 5) land, suitable for

cultivation and other classes are not suitable for cultivation (class 6 and class 7). Ontario has two

major regions: northern and southern. Agricultural land in northern region has short growing

period (100–145 frost free days), thus not suitable for miscanthus cultivation. The southern

region consist of four sub-regions: Central-, Southern-, Western- and Eastern Ontario which are

suitable for miscanthus cultivation (Vyn et al., 2012; Kludze et al., 2013; Sanscartier et al., 2013)

and assumed to be the area for this study. The area under different land classes varies from

region to region and scattered throughout the regions (Fig. 6.1). Table 7.1 shows the area under

different land classes and the total area of various regions.

7.2.2 System boundary

The cradle to gate scenario is adopted to outline the system boundary of this study (Fig.

6.2). Although all classes of cultivable lands are suitable for miscanthus, yield, production cost

and net emission from cultivation are varied depending on the land classes, crop rotation and

regions (Kludze et al., 2013; Sanscartier et al., 2013). Consequently, four scenarios have been

92

established for various regions (central-, southern-, western- and eastern Ontario) to evaluate the

LC of ethanol derived from miscanthus (Table 7.2). A summary of parameters/process for which

data have been collected from the literature and their sources are reported in the previous chapter

(Table 6.4), except emission from miscanthus cultivation and cost of feedstock (which are

reported on table 7.2). The procedure of net energy consumption calculation and other

assumptions are also reported in the previous chapter (section 6.2.1).

Table 7.1 Land area under different tillable land classes and various regions in Ontario, ha

Regions *Land class and tillable land area, ha ^Regional

Class 1 Class 2 Class 3 Class 4 Class 5 area, ha

Southern Ontario 238102 876664 414109 37138 39332 13993100

Western Ontario 724831 254067 239130 73112 130091 56311787

Central Ontario 165830 113526 122839 118976 84424 3981400

Eastern Ontario 32005 312567 304841 148471 65166 3529600

*Kludze et al., 2013; ^Wikipedia

Table 7.2 Scenarios of this study

Scenario Land

class

Soil

type

Crop

rotation

Yield,

tDM/ha

Cost,

$/tDM

*Emission,

kg-CO2e/tDM

S1 1 & 2 Silt loam-

clay loom Corn-soy rotation

11.14 62.63

-28.58

S2 3 Sandy-clay

Continuous soybean

rotation 10.03 65.67

-88.77

S3 3 Sandy-clay

Corn-corn-forage-

forage rotation 10.03 65.67

167.47

S4 4 & 5 Sandy loam Long term pasture 8.35

71.50

167.67

Sources: Kludge et al., 2013; Sanscartier et al., 2013; *net emission from miscanthus cultivation is estimated

considering the carbon dynamics.

7.2.3 Transportation, ethanol production and cost analysis

Transportation, ethanol production, and cost analysis processes are described in the

previous chapter (section 6.2.3 to 6.2.5).

93

7.3 Results and discussion

7.3.1 Net energy consumption

The net energy consumption in the LC of ethanol was dependent on the energy

consumption in each stage that considered in the system boundary of the study. A slight variation

in transportation energy consumption was observed because of the difference in transportation

distance among the scenarios. The transportation distance varied from 12.0–71.5 km depending

on the location and scenarios (S1 to S4). It was found to be the shortest for Eastern Ontario and

the longest for Western Ontario for each scenario compared with other locations. The

transportation distance was found to be the shortest and the longest for scenario S1 and S4,

respectively, for each location except Central Ontario where transportation distance for S4 was

found to be lower than that of S2 and S3 (Fig. 7.1). Feedstock cultivation has emerged as the main

hotspot followed by pretreatment, distillation, enzyme, SSF, yeast, feedstock, and transportation

for each scenario (Fig. 7.2). This figure also indicates that energy recovery from byproducts

plays an important role in net energy consumption in the LC of ethanol.

Figure 7.1 Feedstock transportation distance at different location in Ontario

0

10

20

30

40

50

60

70

Southern

Ontario

Western

Ontario

Central

Ontario

Eastern

Ontario

Tra

nsp

ort

atio

n d

ista

nce

, km

S1 S2 S3 S4

94

Figure 7.2 Energy breakdown of the life cycle of ethanol (Southern Ontario)

The byproduct offsets a part of energy consumed in the process, which has a robust

contribution to the net energy consumption for all scenarios. The net energy consumption varied

from 13.0–13.4 MJ/L. The net energy consumption was noted to be sensitive to coproduct

allocations and assumptions of the study (Pimentel & Patzek, 2005; Roy & Dutta, 2012). Thus,

the net energy consumption varied from other studies because of different feedstock has been

used and various assumptions were made in those studies (Kim & Dale, 2005; Roy et al.,

2012a,b; Roy & Dutta, 2012). The net energy consumption may vary, if different miscanthus

variety, cultivation and pretreatment methods are adopted. Similar to the transportation distance,

the energy consumption was found to be the lowest for S1 and highest for S4 among the

scenarios; however a slight variation was observed among the processing plant locations (Fig.

7.3).

7.3.2 Greenhouse gas emission (CO2 e)

The emissions were directly related to energy and resource consumption at different stages

except feedstock because carbon dynamics offsets a major part in the case of feedstock. The

pretreatment process has emerged as the main hotspot followed by distillation, enzyme, SSF,

yeast, and transportation, except S3 and S4 where feedstock has emerged as the main hotspot

because of the positive carbon dynamics (i.e., carbon released to the atmosphere) of feedstock

(Fig. 7.4). On the other hand, negative carbon dynamics were observed in the case of S1 and S3.

-10

-5

0

5

10

15

20

Ener

gy,

MJ/

L

Bybroduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Feedstock

S2 S2 S3 S4

95

Consequently, emissions from feedstock were found to be -0.09, -0.29, 0.55 and 0.55 kg/L for

S1, S2, S3 and S4 (marginal land), respectively. A slight variation in CO2 emission was observed

in the case of transportation, which caused by the difference in transportation distance among the

scenarios and locations. The carbon dynamics was noted to be dependent on the land classes,

crop replacement/rotation and biomass yield (Kludze et al., 2013; Sanscartier et al., 2013).

Consequently, the net emission would be varied if different crop displacement or crop rotation

were considered.

Figure 7.3 Net energy consumption at different location in Ontario

The emission from feedstock was the highest in the case of S4 and the lowest for S1. These

variations indicate that carbon dynamics dependent not only on the land type and region but also

on the crop rotation. Although, same land class was considered for S2 and S3, the carbon

dynamics varied, might be because of different crop displacements were assumed. The net

emission from the LC of ethanol was estimated to be 0.45–1.32 kg/L depending on the scenarios

and processing plant locations, which seems to be mainly dependent on the carbon dynamics and

the emission offsets by the byproduct. The results also confirmed that these variations might not

only because of land type but also region and crop rotation, because various crop rotations were

considered for different scenarios. The net emission was found to be the lowest in the case of S2

for a plant location at Eastern Ontario and the highest in the case of S4 at Western Ontario (Fig.

7.5). These values indicate that environmental benefit can be achieved from miscanthus based

0

3

6

9

12

15

S1 S2 S3 S4

Ener

gy c

onsu

mpti

on

, M

J/L

Southern Western Central Eastern

96

ethanol at all locations that considered in this study, even if miscanthus was cultivated on

marginal land. It is important to mention that the net emission may vary if different crop

displacement is considered.

Figure 7.4 Emission breakdown of the life cycle of ethanol (Southern Ontario)

Figure 7.5 Net emissions at different location in Ontario

-0.8

-0.3

0.2

0.7

1.2

1.7

Em

issi

on, kg C

O2e/

L

Bybroduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Feedstock

S1

S2

S3 S4

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

S1 S2 S3 S4

Em

issi

on, kg C

O2e/

L

Southern Western

Central Eastern

97

7.3.3 Production cost

The production cost breakdown depicts that the fixed-, feedstock-, and transportation cost

were varied from 0.246–0.247, 0.205–0.235, and 0.005–0.020 $/L, respectively depending on the

scenarios and the location of processing plant. A slight variation in fixed- and feedstock cost was

resulted by the difference in feedstock production cost among the scenarios. Similarly, the

difference in transportation distance caused the variation in transportation cost. However, cost of

other stages of the LC of ethanol remained the same because of the same technologies were

applied for various scenarios. Although, there was a slight variation in fixed-, feedstock- and

transportation cost among the scenarios, the fixed cost has emerged as the main hotspot followed

by feedstock, pretreatment, distillation, SSF, yeast and transportation for all locations (Fig. 7.6).

Figure 7.6 Cost breakdown of the life cycle of ethanol (Southern Ontario)

The estimated net production cost varied from 0.79–0.84 $/L depending on the scenarios

and location of the processing plant. Although a slight variation was observed among the

scenarios and locations, the net production cost was found to be the lowest for S2 and the plant

location at Eastern Ontario, and the highest for S4 and the plant location at Western Ontario

because of the variation in fixed-, feedstock-, and transportation cost among the scenarios (Fig.

7.7). The breakeven feedstock cost was considered for the production cost estimation, which may

vary depending on the biomass logistics and farmers profit margins, consequently the net

production cost may also vary, if the feedstock cost is varied. The production cost of ethanol was

-0.20

0.00

0.20

0.40

0.60

0.80

1.00

Cost

, $/L

Byproduct

Distillation

SSF

Enzyme

Yeast

Pretreatment

Transportation

Feedstock

Fixed cost

S1 S2 S3 S4

98

also noted to be dependent on the conversion technology, enzyme loading, feedstock, allocation

methods and plant sizes (Gnansounou & Dauriat, 2010; Roy & Dutta, 2012); consequently,

production cost may vary, if different conversion methods, allocation, and plant sizes are

considered. Although a wide variation in production cost of lignocellulosic ethanol was reported,

the estimated production cost of this study was observed to be reasonable and comparable with

other studies (Dutta et al., 2010a; Klein-Marcuschamer et al., 2010; Kumar & Murthy, 2011;

Roy et al., 2012a; Roy & Dutta, 2012, 2013).

Figure 7.7 Net production cost at different locations in Ontario.

7.3.4 Sensitivity analysis

The LCA results of lignocellulosic ethanol are reported to be more sensitive to the changes

in parameters related to the biomass and ethanol yield (Roy & Dutta, 2013; Roy et al., 2012b). A

wide variation in ethanol yield from miscanthus (DOE, 2006; Nilsson, 2008; Zhang et al., 2012;

Li et al., 2013), feedstock- and enzyme cost of lignocellulosic ethanol are also reported (Dutta et

al., 2010a; Barta et al., 2010a; Aden et al., 2002). The LCA methodology identifies the potential

hotspots of product, production system or activity depending on the goal of the study. The

potential hotspots were identified to be the feedstock, pretreatment and fixed cost in case of

energy consumption, emission and production cost, respectively. This study also indicates that

byproducts have a robust contribution in the LC of lignocellulosic ethanol. The breakeven

0.0

0.2

0.4

0.6

0.8

1.0

S1 S2 S3 S4

Poro

duct

ion c

ost

, $

/L

Southern Western Central Eastern

99

feedstock cost was used to estimate the production cost, which may depend on the biomass

logistics, farmers profit margin and renewable energy policy. Therefore, the effect of plant

capacity (which is mainly sensitive to fixed cost), ethanol yield and feedstock cost on the

production cost have been determined. Then the effect of the variation (±20%) of different

parameters on energy consumption, emission and production has also been worked out for S4.

It seems that the ethanol plant capacity affects the production cost and emission (Fig. 7.8).

The production cost decreased with an increase in plant capacity; however, emission increases

because of the higher feedstock demand resulted in longer transportation distance. The

production cost was found to be reduced from 1.0–0.67 $/L and emission increased from 1.29–

1.37 kg CO2e/L for the increase of plant capacity from 5 to 300 ML/y. Increasing biomass and

ethanol yield may help in reducing biomass demand consequently the transportation distance.

Figure 7.9 depicts that the production cost of ethanol not only dependent on the feedstock cost

but also on the ethanol yield. Although there was a slight variation in the case of fixed cost

because of different feedstock demand, it seems that the higher the ethanol yield the lower the

production cost. Higher ethanol yield not only reduces the feedstock demand, consequently

transportation distance, but also has a positive impact on each stages of the LC of lignocellulosic

ethanol.

Figure 7.8 Effect of ethanol plant capacity on production cost and emission

1.30

1.33

1.36

1.39

0.0

0.3

0.6

0.9

1.2

0 100 200 300 400

Cost

, $/L

Plant capacity, ML/year

Cost EmissionE

mis

sion, kg C

O2 e

/L

100

Figure 7.9 Effect of ethanol plant capacity on production cost and emission

Figure 7.10 shows the severity of the effect of the variation in different parameters on the

net energy consumption. The net energy consumption varied from 12.1–14.8 MJ/L and it seems

that pretreatment has the most robust effect on net energy consumption. Energy recovered from

byproduct contributes to offset part of total energy. The net emission and production cost varied

from about 1.2–1.5 kg-CO2 e/L and 0.8–1.0 $/L, respectively, depending on the severity of the

variation (Figs. 7.11). These figures also depict that ethanol yield has more robust impact on both

net emission and production cost than others. It is worthy to mention that although higher ethanol

yield reduced energy recovery from the byproduct, consequently lesser opportunities for

emission and cost offsetting because of lower amount of byproduct (lignin) recovery from the

system, the net emission and production cost observed to be reduced. It is also important to note

that the production cost can further be reduced if agri-industrial and environmental policies are

enacted to support the miscanthus based ethanol industry, especially for miscanthus grows on the

marginal land and used for ethanol.

This study also support our earlier findings (chapter 6) that miscanthus grows on the

marginal land in Ontario could be a potential feedstock for lignocellulosic ethanol industries and

avoid any competition with food crops for prime land help Ontario improves her food and energy

security, and enhance rural economy. It is important to note that net emission from the LC not

0.0

0.3

0.6

0.9

1.2

1.5

0 50 100 150 200

Co

st, $

/L

Feedstock cost, $/tDM

0.244 kg/L 0.305 L/kg 0.366 L/kg

101

only dependent on the land classes but also on the crop displacement (S2 & S3 where same land

classes are used).

Figure 7.10 Effect of the variation of different parameters on net energy consumption (MJ/L)

Figure 7.11 Effect of the variation of different parameters on net emission

0

5

10

15S4

Yield +20%

Yield-20%

Transportation+

20%

Transportation-

20%

Pretreatment+20

%

Pretreatment-

20%SSF+20%SSF-20%

Distillation+20

%

Distillation-20%

Byproduct+20%

Byproduct-20%

Enzyme+20%

Enzyme-20%

0

0.6

1.2

1.8S4

Yield +20%

Yield-20%

Transportation+20%

Transportation-20%

Pretreatment+20%

Pretreatment-20%

SSF+20%SSF-20%

Distillation+20%

Distillation-20%

Byproduct+20%

Byproduct-20%

Enzyme+20%

Enzyme-20%

kg-CO2e/L $/L

102

The Eastern Ontario has emerged to be the best location for miscanthus based ethanol

industry among the locations studied in Ontario which seems to be environmentally and

economically viable. Bennett et al. (2004) reported that genetically modified (GM) crops can

serve various purposes, such as improving yield, increasing the share of useful components or

decreasing the need for chemical fertilizers or water, and improves farm income. Therefore, net

emission and production cost of ethanol from miscanthus may further be reduced, if GM

miscanthus is considered and promoted for ethanol industry, improve farm income and rural

economy, and future energy security.

7.4 Conclusion

This study identifies the potential locations for miscanthus based ethanol industry in

Ontario, Canada and determines the environmental and economic viability of the miscanthus

based ethanol. A slight variation was observed in the case of net energy consumption, and

production cost; however, the variation was emerged to be robust in the case of net emission

where carbon dynamics plays a key role. The scenario S2 was found to be the best option to abate

GHG emissions, which indicates that GHG emissions were dependent not only on the land

classes but also on the crop displacement. The results indicate that the miscanthus based ethanol

industries are economically and environmentally viable for all scenarios and locations in Ontario,

even if miscanthus was grown on marginal land; consequently, miscanthus arose to be a

promising feedstock for ethanol industry in Ontario. However, Eastern Ontario was appeared as

the best option for miscanthus based ethanol industry in Ontario. The miscanthus based ethanol

industry might need to be regulated to avoid any sort of competition with food crops for higher

quality land; improve farm income and rural economy. The information generated in this study is

emerged to be novel, may help the stakeholders in their decision making processes, and help

meeting the ethanol demand, and help achieving GHG emissions target of Canada.

103

Chapter 8

Development of a Continuous Stirred Tank Bioreactor for Syngas Fermentation

8.1 Introduction

Although both the biochemical and thermochemical conversion technologies are used for

ethanol production from biomass, biochemical dominates over thermochemical process

(Subramani & Gangwal, 2008). Most of the biomass contains a large amount of non-

carbohydrate materials (lignin) that cannot be converted into ethanol by microorganisms in the

biochemical conversion process (Henstra et al., 2007). Several challenges have also been

reported in biochemical routes such as high pretreatment and enzyme cost, low fermentability of

mixed sugar stream (C5) and the generation of inhibitory soluble compounds (Munasinghe and

Khanal, 2010). Conversely, gasification is very effective at converting non-carbohydrate biomass

fractions and all other components of biomass into syngas with nearly equal efficiency and

effectiveness (Pereira et al., 2012; Weber et al., 2010; Wang & Yan, 2008; Phillips et al., 2007;

Brown, 2007; Henstra et al., 2007; Aden et al., 2002), eliminates the complex pretreatment steps

and requirement of costly enzymes (Munasinghe & Khanal, 2010). Some of the biological

catalysts (Clostridium ljungdahlii, Clostridium autoethanogenum, Acetobacterium woodii,

Clostridium carboxidivorans and Peptostreptococcus products) are able to ferment syngas into

liquid fuel more effectively than that of chemical catalysts (iron, copper or cobalt) (Heiskanen et

al., 2007; Henstra et al., 2007).

Till date various types (batch and continuous stirred tank; two stages continuous stirred

tank; trickle-bed, microbubble dispersion, monolithic biofilm, bubble column, membrane-based

system) of reactors have been developed and used for syngas fermentation (Richter et al., 2013;

Mohammadi et al., 2012; Datar et al., 2004). The continuous culture was noted to be

advantageous compared to a batch culture in a fermentation system (Richter et al., 2013). The

mass transfer between substrate and microbes was also dependent on the level syngas mixing

with the fermentation media. Higher agitation speed tend to produce finer bubbles thus slow

rising velocity in fermentation media, improves microbes accessibility to syngas and improves

mass transfer rates (Munasinghe & Khanal, 2010). This study attempts to develop a reactor

employing an innovative gas supply and effluent extraction systems with continuous stirred tank

for ethanol production and evaluate the developed reactor.

104

8.2 Materials and Methods

8.2.1 Reactor development

Bioreactor is a device that supports biologically active aerobic or anaerobic environment in

which biochemical conversion can be taken place involving microbes. The environmental

conditions in a bioreactor such as syngas and media flow rates, temperature, pH and agitation

speed were not only monitor but also controlled. Consequently, a laboratory scale bioreactor

(3L) has been developed with transparent polyvinylchloride (PVC) pipe (Fig. 8.1), which was

coupled with pH meter (PHE-1411, Omega Environmental, Inc., Laval, QC, Canada) and a

pressure gauge (PHH-222, Omega Environmental, Inc., Laval, QC, Canada) to monitor working

temperature, pH and pressure, respectively. The pH meter was equipped with a temperature

probe (TP-07). Membrane separator (PVDF, GE: 0.02 µm) was also put in place to extract the

effluent free of microbes (A-8-1). A membrane support has also been developed to facilitate the

effluent extraction from the top layers (A-8-2). A list of materials and accessories for the reactor

are appended in the appendix (A-8-3). The gas chamber has also been developed and fabricated

at the University of Guelph, Ontario, Canada. The materials used for the gas chamber are listed

in the appendix (A-8-4).

Figure 8.1 Photograph of the developed reactor

105

Figure 8.2 Schematic diagram of the gas chamber (not to scale)

Figure 8.3 Photograph of the developed gas chamber

60 cm

50 cm

50 cm

106

Syngas fermentation into ethanol was noted to be dependent on the gas-liquid mass transfer

(Devarapalli et al., 2013; Lee, 2010). The gas-liquid mass transfer can be improved by increasing

the residence time of gas in the aqueous media (Lee, 2010). The bubbling technology helps in

improving the gas retention time in the aqueous media. Although the size of bubble affects the

retention time, a bubbling tube was incorporated in the bioreactor to improve the gas retention

time in the aqueous media of the reactor, which can easily be replaced with other bubbling

systems. The reactor can be used for both anaerobic and aerobic condition. The media and

effluent (ethanol) were supplied and extracted from the reactor with a micro-pump (GF-

F155001, Gilson Inc., USA), respectively. The tubing (F117938) was selected based on the

desired flow rates. It is noteworthy to mention that the following parameters were considered for

the development of the bioreactor: Innovative, purpose, usefulness, flexibility, easy monitoring,

durability, safety, easy assemble and disassemble, easy cleaning, and inexpensive.

8.2.2 Microorganism and media

American Type Culture Collection (ATCC#55380; Clostridium ljungdhalii) was purchased

from Cedarlane, Burlington, Ontario, Canada and used in this study. The recommended broth

media for ATCC#55380 has been prepared at the laboratory based on the preparation manual

supplied by Cedarlane. The components of broth media and the production procedure are

reported in the appendix (A-8-5). Clostridium ljungdahlii in the fermentation process of syngas

not only improves mass transfer properties, but also capable of producing only ethanol (BRI,

2008). Aseptically the microorganism was transfered into the produced broth media in the test

tubes and cultured. Cells were anaerobically propagated in the prepared broth media at 37oC in

an incubator (Heratherm IGS60, Thermo Electron LED GmbH, Germany) (A-8-6). The

propagated cells were then used in the syngas fermentation process.

8.2.3 Syngas fermentation

In a gasification process all feedstock components are decomposed into syngas (H2, CO,

CO2, CH4) with few residues (tar and ash) and trace amount of other gases (He & Zhang, 2011;

Wei et al., 2009;). Usually, the composition and quality of syngas from biomass are dependent

on the type of feedstock and gasification parameters (He & Zhang, 2011; Carpenter et al., 2010;

Wei et al., 2009). A wide range of syngas composition has been used by the researchers (Richter

et al., 2013; Kundiyana et al., 2011; Sim et al., 2007; Younesi et al., 2006). The composition of

syngas most recently reported was 60% CO; 35% H2 and 5% CO2 (Ricter et al., 2013). However,

107

only the CO has been used in the fermentation process to prove the concept of the developed

reactor (Chang et al., 1998). The working volume of the reactor and planned operating

temperature were 2 L and 37°C, respectively. The reactor was operated under continuous

condition, after the initial two days batch condition. A bypass tube was used to connect the

reactor and the media-jar. Another tube from the media-jar was connected to the exhaust tube to

release the excess gas if any.

The syngas, fresh media and effluent extraction flow rates were 5.0–15.0, 0.25–0.75 and

0.25–0.75 mL/min, respectively. For this purpose the pump was calibrated against different

operating speed (A-8-7). The fresh media and effluent extraction rate was controlled to maintain

the working volume in the reactor. The working pH, temperature, gas and ethanol concentration

were also monitored. Lee (2010) reported the highest growth rate of Clostridium ljungdahlii on

PETC 1754 with 5 g/L fructose and 1 g/L yeast extract and pH 6.5–7.5. The working pH was

reported to be 5.5–6.5 for optimal growth of the cells (Richter et al., 2013; Mohammadi et al.,

2012, Abubackar et al., 2012; Lee, 2010). The optimum microorganism growth was reported at

pH 4.4–6.2 (Richter et al., 2013; Liou et al., 2005). On the other hand the highest ethanol yield

was observed at pH 4.75 (Abubackar et al., 2012); however in the second stage i.e., ethanol

production stage pH was maintained to 4.4–4.8 (Richter et al., 2013). Consequently, the working

pH was controlled to 4.5–5.0 by adding 1N NaOH in the reactor with a syringe, if required. The

agitation speed was also controlled to 300–500 rpm. Figure 8.4 shows the photograph of the

experimental setup (details in A-8-8). Figure 8.5 shows the schematic diagram of the

experimental setup of this study.

108

Figure 8.4 Photograph of the experimental setup

Figure 8.5 Schematic diagram of the experimental setup of this study [1. Reactor (3L); 2 & 18. Glass bottle for media/effluent (2L); 3. Media inlet; 4. Gas inlet; 5. Temperature probe; 6.

Pressure meter; 7. Gas controller; 8. Control valve; 9. Vent; 10. Membrane sampler/extraction port; 11. Gas

sampling port (gas impermeable butyl rubber stopper); 12. pH meter; 13. Aerator; 14. Stirrer; 15. Stirring machine;

16. Pump; 17. Incubator]

1

17

15

4

8 7

10

2

16

13

18

14

5 6

11 3 12 9

109

8.2.4 Analytical method

The effluent has been collected periodically for each experimental condition to determine

their effect on fermentation activities and ethanol productivity. The effluent has been analyzed

by using a Bruker Scion 436 Gas Chromatograph coupled with a quadruple mass analyzer. The

volatiles are separated using a 5% phenyl-methyl silicone bonded-phase fused silica capillary

column (DB-5MS, 30 m × 0.25 mm i.d., film thickness 0.25 mm), operating at 41.4 kPa of

column head pressure, resulting in a flow of 1.0 mL/min at 35°C. The solid phase

microextraction (SPME) fiber is desorbed and maintained in the injection port at 280°C for the

first 5 min of the chromatographic analysis. The injection port was in a splitless mode. The

temperature program was isothermal for 5 min at 35°C, raised to 200°C at a rate of 10°C/min,

and then raised to 250°C at a rate of 15°C/min, and held for 5 min. The transfer line to the mass

spectrometer was maintained at 280°C. The ions were obtained by electronic impact ionization at

positive ion mode at 70 eV, the data collection rate was 1 scan/s over the m/z (mass to charge

ratio) range of 10 to 150. Compounds were tentatively identified by comparing their mass

spectra with those contained in the NIST/EPA/NIH and Wiley libraries.

8.3 Results and discussion

8.3.1 pH and temperature profile during syngas fermentation

The bioreactor was placed in an incubator at 37°C to maintain the fermentation

temperature for optimum growth. However, the temperature found to be risen upto 46°C might

be because of exothermic reaction during syngas fermentation. Once the temperature was found

to be risen above 37.5°C the door of the incubator kept open, which caused the temperature

fluctuation (Fig. 8.6).

110

Figure 8.6 pH and temperature profile of the fermentation broth

The pH value in the fermentation broth was also found to be varied with time, syngas flow

rate and media flow rate. Initially the pH decreased with the fermentation period might be

because of the microbial activities, ethanol, propanol and formaldehyde production (Fig. 8.6).

Many researchers observed higher ethanol yield at lower pH environment, because

microorganism produces less acid which prevents a further pH drop, consequently yielded higher

ethanol (Datar et al., 2004). In contrast, 110% higher ethanol yield was reported at pH 6.8

compared to that of pH 5.5 (Cotter et al., 2009), however, higher cell densities has been reported

at pH 6.8 compared with pH 5.5. A drastic improvement in the ethanol yield with C. ljungdahlii

has also been reported when the pH dropped to 4.0–4.5 (Klasson et al., 1993), which seems to be

supporting the findings of this study.

8.3.2 Ethanol and other alcoholic compounds

The qualitative analysis of the effluent revealed that ethanol and some other alcoholic

compounds (propanol, formaldehyde, isopropyl alcohol etc.) have been produced from syngas

(CO) fermentation process. Figure 8.7 represents the GC-MS results, while syngas and media

flow rate were maintained to 5 mL/min and 0.25 mL/min, respectively after two days of batch

phase. The qualitive analysis confirmed that ethanol was produced during the fermentation

process. Gas-liquid mass transfer was noted to be the main constrain in ethanol production from

syngas fermentation because of low CO and H2 solubility in the fermentation broth (Devarapalli

30

35

40

45

50

4.5

4.7

4.9

5.1

5.3

5.5

5.7

0 50 100 150 200 250

pH

Time, h

pH Temperature

Tem

per

atu

re,

°C

111

et al., 2013). However, in this study only CO has been used in the fermentation process. The

mass transfer can be improved by the reactor configuration, agitation speed, syngas and media

flow rates (Devarapalli et al., 2013; van Kasteren et al., 2005). The agitation speed helps break

the bubble and enhance gas retention time and increasing the interfacial area available for the

mass transfer (van Kasteren et al., 2005). The samples were also collected to determine the effect

of agitation, syngas and media flow rate. The results of this study revealed that ethanol can be

produced by using the developed bioreactor, and it seems gas-liquid mass transfer has been

improved. Alternate syngas supply (bubbling) and bioreactor design may further improve gas-

liquid mass transfer, thus the ethanol yield. In this purpose, different gas supply/bubbling system

can be used to improve the bubbling proceess which may produce smaller bubble, consequently

improve gas retention time in the media and increase mass transfer.

Figure 8.7 Mass spectra of the effluent

8.4 Conclusion

The developed bioreactor was found to be able to produce ethanol form syngas

fermentation process. The ethanol yield appears to be dependent on the experimental parameters.

Ethanol

Propanol

Formaldehyde

112

Chapter 9

Evaluation of the Life Cycle of Ethanol derived from Biosyngas Fermentation

9.1 Introduction

The evaluation of the life cycle (LC) of ethanol derived from miscanthus by enzymatic

hydrolysis process revealed that miscanthus is a promising feedstock for environmentally and

economically viable ethanol in Ontario, Canada (Chapter 6 & 7). Ethanol has been produced by

either biochemical processes (hydrolysis) or thermochemical processes (gasification /pyrolysis of

biomass to syngas followed by biosynthesis or chemical synthesis). Each process has its

strengths and weaknesses. The cost of cellulase is the major expense when producing

lignocellulosic ethanol with conventional technology (Singh & Kumar, 2010) and contributes

about 40–55% of the enzymatic ethanol production cost. Distillation, enzyme production and

pretreatment were also reported to be the main contributors to the LC of ethanol produced by

conventional technology (Roy et al., 2012; Orikasa et al., 2009). Thus, a wide variation was

reported on GHG emissions and production cost. The use of the cost effective and innovative

fermentation strategies integrated in the technology chain of gasification and gas cleaning,

combined with syngas fermentation or catalytic synthesis could significantly improve the overall

economics of ethanol from biomass.

Biosynthesis of syngas results in poor mass transfer properties of gaseous substrates and

low ethanol yield (Munasinghe & Khanal, 2010). Conversely, higher ethanol yield was also

reported in this process (Clausen & Gaddy, 1993). Although the thermochemical process

produces ethanol in large quantities, it requires expensive catalysts and high operating pressure

(Subramani & Gangwal, 2008). Ethanol production from syngas either by biosynthesis or

catalytic synthesis process have been studied by many researchers (Ruth, 2005; Martchamadol,

2007; Clausen & Gaddy, 2008; Munasinghe & Khanal, 2010) except for few examples (Foust et

al., 2009; Mu et al., 2010) where comparative studies between biosynthesis and catalytic

synthesis have been conducted. This study attempts to evaluate the LC of ethanol produced by

biosynthesis of syngas from pretreated (torrefied) and untreated (non-torrefied) miscanthus with

or without chemical looping gasification (CLG).

113

9.2 Materials and methods

9.2.1 System boundary and assumptions

Lignocellulosic biofuels were noted to be environmentally sustainable products (González-

García et al., 2012; He & Zhang, 2011; Hsu et al., 2010). Production cost and emissions from

thermochemical conversion of biomass into ethanol were also reported to be dependent on the

feedstock, processing conditions, plant capacity, byproduct utilization etc. (Chapter 2; Tables 2.3

& 2.6). Moreover, production cost of ethanol from syngas fermentation has also widely varied

among the reported studies (Piccolo & Bezzo, 2009; artı´n & Grossmann, 20 ).

Consequently, this study attempts to evaluate the LC of ethanol produced by syngas (derived

from treated and untreated biomass: miscanthus) fermentation (Fig. 9.1) to determine if

environmentally friendly and economically viable ethanol can be produced by biosynthesis

process in Ontario, Canada. The raw biomass was assumed to be transported to the integrated

processing plant. Although the processing plant cost may vary depending on the type of

processing, it was assumed to be the same for all scenarios of this study.

Figure 9.1 Schematic diagram of the system boundary of this study.

9.2.2 Pretreatment (torrefaction)

Torrefaction is a mild heat treatment process given to biomass (typically from 200–300°C

in an inert atmosphere) that improves thermochemical properties of biomass, producing a more

stable, denser, hydrophobic material with higher energy values and thus reduces the biomass

handling cost. It was also reported that torrefied biomass can be easily transported and fed to the

gasifier (Bessou et al., 2011). The optimum torrefaction temperature was reported to be 275°C

(Acharya, 2013). Consequently, torrefied biomass was produced with a Quartz Wool Matrix

(QWM) reactor at 275°C for 45 min in an inert condition (Acharya, 2013). The experimental

setup is shown in the appendix (A-9-1). Although experiments have been conducted to measure

the energy consumption during the torrefaction process of biomass and flue gas properties were

monitored (A-9-2), the estimated energy consumption (A-9-3) was used for the LC analysis

because the QWM reactor is yet to be optimized. Biomass yield in the torrefaction process was

Treated/untreated

Syngas

cle

anup

& c

ondit

ionin

g

Fer

men

tati

on

Was

te/

by

pro

du

cts

Eth

an

ol

Torrefied

With CaO

Cultivation

Gas

ific

atio

n

Mis

canth

us

114

reported to be about 70% (Tumuluru et al., 2011; Acharya, 2013); however biomass yield was

assumed to be about 85% for this study (Kambo, 2014).

9.2.3 Ultimate analysis

The biomass samples were analyzed with CHNS-O analyzer (Flash 200 CHNS-O, Organic

Elemental Analyzer, Thermo Fisher Scientific, The Netherlands) based on ASTM D5373-08 to

determine the components in biomass (A-9-4). Samples were dried at 105°C for 24 hours prior to

ultimate analysis. Then samples were combusted at 925°C in Helium atmosphere, while

reduction was carried at 650°C. The components of various biomasses are reported in Table 9.1.

Based on the elemental analysis the air requirement in the gasification process was calculated.

Table 9.1 Components of different feedstock

Feedstock Elements, % HHV,

MJ/kg C H N S O Ash

Miscanthus 46.66 6.00 0.21 0.00 45.34 1.80

18.47

T-miscanthus 49.55 5.71 1.16 0.00 42.24 1.34

20.18

Wheat straw 46.33 5.59 0.00 0.00 43.72 4.36

18.87

Willow 47.81 6.07 0.52 0.00 44.55 1.05

20.01

T: torrefied, Mass yield of torrefied biomass: 85% (Kambo, 2014)

9.2.4 Gasification and syngas cleaning

Although different sorbents have been used in the gasification process, calcium oxide was

noted to be cheaper and effective for capturing CO2 at very high temperatures results in a small

fraction of CO2 in the flue gas and minimize auxiliary power consumption (Acharya, 2011). The

author also reported that in a CLG system, heat releases by the exothermic carbonization

reactions can supply most of the heat required by the endothermic gasification reactions.

Therefore, CaO was used as absorbent for this study. Treated and untreated biomass was

thermally degraded with or without CaO in a micro gasifier (TGA-FTIR; TGA: SDT-Q600, TA

Instruments-Waters LLC, New Castle, USA; FT-IR: Thermo Scientific Nicolet 6700, TA

Instruments-Waters LLC, New Castle, USA) (A-9-5 & A-9-6) at 900°C. The air supply was

controlled from 13.0–14.0 mL/min based on the components of biomass to ensure the sub-

stoichiometric oxygen supply. The TGA experimental parameters are also reported in the

appendix (A-9-7). The solid particles in the product gas can be separated in the cyclone

115

(Acharya, 2011; artı´n & Grossmann, 2011, van Kasteren et al., 2005). The product gas can be

cleaned with water scrubber or water wash process (Spath & Dayton, 2003; van Kasteren et al.,

2005). The water scrubber is assumed to be used to remove any tars from the syngas (Han &

Kim, 2008; van Kasteren et al., 2005). The thermal degradation experiment indicates that syngas

composition may also be dependent on the feedstock (A-9-8 to A-9-12).

The syngas compositions were noted to be dependent on the feedstock and gasification

conditions (Wei et al., 2009; He & Zhang, 2011; Kuo et al., 2014; Dutta & Stefan, 2014). The

syngas quality was also reported to be improved in the CLG (i.e., chemical looping gasification)

process (Acharya et al., 2009). Consequently, the lesser efforts are required in the syngas

cleaning process in the case of torrefied biomass and CLG processes. The benefit achieved in

syngas cleaning process was assumed to be offset by the efforts required in the torrefaction and

CLG gasification process. Although thermal degradation experiments have been conducted, the

simulated (with Aspen Plus V7.3) data have been used to estimate cold gas efficiency and the

ethanol yield. The CGE calculation procedure is reported in the appendix (A-9-13). The

simulation block diagram, simulation flowsheet and other parameters are reported in the

appendix (A-9-14 to A-9-16). The simulated syngas composition also confirmed that gas quality

were not only dependent on the feedstock but also on the gasification conditions. The simulated

syngas composition, heating value and cold gas efficiency are reported in the appendix (A-9-17).

9.2.5 Syngas fermentation

Syngas fermentation process adopted in this study was assumed to be same that reported in

the previous chapter (section 8.2.3). Low pressure and temperature fermentation of syngas

reduces the operating costs, however requires large amount of energy because ethanol need to be

separated from water (Martı´n & Grossmann, 20 ). Several authors reported that the ethanol

yield from syngas fermentation was dependent on the syngas compositions and, CO and H2

conversion efficiency. The ethanol yield of this study was estimated based on the following

procedure (Eq. 9.1) (Spath & Dayton, 2003). The estimated ethanol yield was found to be varied

from 0.36–0.39 L/kg-miscanthus depending on the cold gas efficiency and gas to ethanol

conversion efficiency (80%) (Table 9.2). However, ethanol yields were normalized to untreated

feedstock while evaluating the LC of ethanol, and energy, emission and cost breakdown were

worked out. Although mixed alcohol production has been reported by several authors,

116

Clostridium Ljungdhalii can produce only ethanol from syngas. Consequently, ethanol is

assumed to be the only final product in the syngas fermentation process.

. . . . (Eq. 9.1)

where, Y = Ethanol yield, million gal/year

F = Feed rate, tons/day (dry basis)

= Higher heating value of the feedstock, Btu/lb (dry)

= Cold gas efficiency of gasifier + conditioning

= Average conversion of CO and H2 into to ethanol

Table 9.2 Ethanol yield from biosyngas fermentation Authors

Biomass

Gasification

condition CGE, %

Yield, L/kg-

FS

artı´n & Grossmann, 20 Switchgrass Low pressure - 0.32–0.33

artı´n & Grossmann, 20 Switchgrass High pressure - 0.26

van Kasteren et al., 2005 Wood chips - - 0.36

Spath & Dayton, 2003 Wood chips - 70.00 0.26–0.33

Miscanthus (S1) 78.83 0.36

T-miscanthus (S2) 78.65 0.39

This study Miscanthus (S3) CLG 79.06 0.36

T-miscanthus (S4) CLG 78.67 0.39

FS: Feedstock; T: Torrefied; CLG: chemical looping gasification

9.2.6 Separation (distillation & purification)

Ethanol concentration in the fermentation broth affects the microbial activities and in

reducing concentration, microorganism can produce more ethanol (Nielsen & Prather, 2009).

The ethanol concentration in the broth was considered to be 5% ( artı´n & Grossmann, 20 ;

van Kasteren et al., 2005). Ethanol was assumed to be separated by using membrane-assisted

vapor stripping (MAVS) distillation system (Vane & Alvarez, 2008) and then purified by using

glycerol as an additives (Dias et al., 2009).

117

9.2.7 Waste management

Ethanol production from biosyngas (fermentation route) results in two waste streams: solid

(char, tar etc.) and liquid (waste water) ( artı´n & Grossmann, 20 ; van Kasteren et al., 2005).

Char, tar and other solid particles were assumed to be feedback to the gasifier. The good quality

ash can be used, either as fertilizer or raw materials of cement or other construction industries

(Pérez-Villarejo et al., 2012; Chatveera & Lertwattanaruk, 2011; Salas et al., 2009). The revenue

generated from ash was assumed to be used to offset wastewater treatment cost. Although energy

was consumed in the gasification process, a considerable amount of heat was also recovered

from the cooler (van Kasteren et al., 2005). The emission and cost that credited to the recovered

heat in the gasification process was determined with the emission factor and cost of LNG.

9.2.8 Cost analysis

Cost analysis methods were discussed in the previous chapter (section 6.2.5). Ethanol

processing plant cost was workout based on the ratio of construction cost of biochemical and

thermochemical ethanol plant reported by Foust et al. (2009), and the construction cost reported

by Asano & Minowa (2007).

9.2.9 Data collection

In this study, we tried to use site-specific or country specific data wherever possible and

remaining were collected from the literature. Although data for different processes (torrefaction,

gasification, ethanol productivity etc.) were collected from the lab scale experiments, both the

estimated (pretreatment) and the literature data were used in this study, because the equipment

are yet to be optimized or too small to get valued data. A summary of parameters/processes for

which data have been collected from the literature and their sources were reported in Table 9.3.

9.3 Results and discussion

9.3.1 Net energy consumption

Energy consumption at all stages and energy recovery were examined to denote the net

energy consumption in the LC of ethanol (Fig. 9.2). The assessment revealed that there was a

wide variation among the stages, and the gasification process was found to be the main hotspot

followed by the distillation, fermentation, feedstock production and others. Energy consumption

was greater in the case of torrefied feedstock (S2) compared with untreated feedstock (S1).

However, CLG process reduced the energy consumption in the gasification process. Energy

118

content in torrefied feedstock was greater than that of untreated feedstock, thus noted to be

favorable to reduce transportation energy. In contrast, the transportation energy was observed to

be greater when it was normalized to untreated biomass. It was also important to note that an

integrated ethanol processing plant was considered where torrefied feedstock was produced and

fed into the gasifier. Thus, energy consumption in all stages was also found to be slightly greater

except the gasification compared with untreated feedstock because of the difference in ethanol

yield (adjusted).

Table 9.3 Summary of parameters for which data are collected from literature or estimated

Parameters/Systems Actual data Sources

^Miscanthus cultivation (kg CO2e/tDM) 167.67 Sanscartier et al., 2013

Feedstock cost ($/tDM) 71.50 Kludge et al., 2013

*Crushing (size 3 mm) (kWh/kg) 0.06 Roy et al., 2012a; b

#Ethanol processing plant construction $45 million Estimated

No. of labor (persons) 23 Asano & Minowa, 2007

Pre-treatment (torrefaction, MJ/kg) 0.41 Estimated

Gasification

Energy consumption (MJ/kg-ethanol) 1.53 van Kasteren et al., 2005

Heat recovery (MJ/kg-feedstock) 1.92 van Kasteren et al., 2005

Fermentation (MJ/kg-ethanol) 3.28 van Kasteren et al., 2005

Distillation (MJ/kg-hydrous ethanol) 2.80 Vane & Alvarez, 2008

Purification (MJ/kg-ethanol) 1.09 Dias et al., 2009

Ethanol yield (L/kg-dry miscanthus) 0.36-0.39 Estimated

Note: Plant capacity is 20000 kL/year; labor cost $46000/person/year; ^Miscanthus grown on marginal land;

*assumed to be same that of straw; #Calculated based on the ratio of construction cost of biochemical and

thermochemical ethanol plant reported by Foust et al. (2009), and construction cost reported by Asano & Minowa

(2007); $: Canadian dollar.

Although no external energy was required in the case of scenarios S2 & S4 to gasify the

feedstock, energy was required only for steam production for the systems. Consequently, energy

consumption in the gasification process (S2 & S4) was observed to be lower compared with

119

steam gasification process (S1 & S3). The energy consumed in the steam generation process was

found to be the main contributor in the gasification process. Biomass to steam (400°C) ratio was

considered to be one. The net energy consumption was found to be varied from 13.19–14.81

MJ/L. The heat recovered from the gasification process offset a part of energy consumes in the

LC of ethanol and has a significant contribution to the net energy consumption.

Figure 9.2 Energy consumption at various stages of the LC of ethanol

9.3.2 GHG emission (CO2e)

GHG emissions at different stages were also varied among the scenarios and the stages of

the LC of ethanol. Similar to the net energy consumption, the main hotspot was also found to the

gasification; however, in the case of GHG emissions it was followed by feedstock, distillation,

fermentation and others (Fig. 9.3). The net GHG emissions were found to be 1.27, 1.32, 1.19 and

1.24 kg-CO2 e/L for scenarios, S1, S2, S3 and S4, respectively. It is important to note that although

S1 performed better compared with the S2; however, CLG option improved this performance

(i.e., S1>S3 and S2>S4). It is also noteworthy to mention that heat recovered in the gasification

process offset a portion of GHG emissions of the LC of ethanol. These results indicate that an

environmental benefit can be achieved relative to gasoline if ethanol can be produced by using

the technologies and scenarios adopted in this study. GHG emissions from the LC of the ethanol

reported to be varied from 0.27–0.83 kg-CO2 e/L (Hsu et al., 2010; Tan & Dutta, 2013) which

-6

-3

0

3

6

9

12

15

18

21

S1 S2 S3 S4

Ener

gy,

MJ/

L

Heat recovery

Distillation

Fermentation

Gasification

Pretreatment

Transportation

Feedstock

(S1: Raw, S2: Torrefied, S3: Raw-CLG, S4: Torrefied-CLG)

120

might be because of the assumptions, plant capacities and feedstock. It seems that the results of

this study were a bit greater than those reported data because of different feedstock and

assumptions were made. In contrast, GHG emissions for the hybrid poplar was reported to 2.8

kg-CO2 e/L (Daystar et al., 2013), which indicates that the results of this study is reasonable.

Figure 9.3 Emission at different stages of the LC of ethanol

9.3.3 Production cost

The production cost of ethanol was slightly varied among the production pathways (S1, S2,

S3 and S4,) of this study. Figure 9.4 depicts the GHG emissions from different stages of the LC

of ethanol. Figure 9.4 also shows that the hotspot was the fixed cost followed by either

gasification or feedstock cost depending on the production path. Production costs were found to

be 0.78, 0.81, 0.90 and 0.88$/L for S1, S2, S3 and S4, respectively. It is worthy to note that heat

recovery process has robust contribution to the net production cost of ethanol. This study also

confirmed that production cost of ethanol is dependent on the feedstock, conversion

technologies, and assumptions ( artı´n & Grossmann, 20 ; Gonzalez et al., 2012). The

feedstock cost and the production rate of hydrogen reported to be played an important role in the

production cost of ethanol from syngas ( artı´n & Grossmann, 20 ) A wide variation was

observed in the literature, which was dependent on the feedstock, conversion technologies,

allocation methods and plant sizes. The production cost of this study seems to be comparable

with other studies (Perales et al., 2011; Piccolo & Bezzo, 2009). This study revealed that both the

environmentally preferable and economically viable ethanol can be produced from miscanthus

-0.4

-

0.4

0.8

1.2

1.6

S1 S2 S3 S4

Em

issi

on, kg C

O2e/

L

Heat recovery

Distillation

Fermentation

Gasification

Pretreatment

Transportation

Feedstock

121

grown on the marginal land and produced by gasification-biosynthesis process. It is also

noteworthy to mention that this was an optimistic study, and all the data are neither to Canadian

context nor from the same plant size, consequently, a full LC evaluation is required before any

future investment and commercial production.

Figure 9.4 Production cost at different stages of the LC of ethanol

9.3.4 Sensitivity analysis

It was reported that ethanol produced from syngas can be cost-competitive with efficient

equipment, optimized operation, cost-effective syngas cleaning technology, low feedstock and

pretreatment cost, optimal configuration, heat integration, and high value byproduct (He &

Zhang, 2011). CaO based sorbent can capture CO2 at a relatively high temperature, eases the gas

cleaning process compared with the convectional techniques. The syngas compositions were also

noted to be dependent on the feedstock and gasification conditions (Wei et al., 2009; He &

Zhang, 2011; Kuo et al., 2014; Dutta & Stefan, 2014). The gas quality was reported to be

improved if torrefied biomass is used (Kuo et al., 2014). The syngas quality was also improved

in the case of CLG (Acharya et al., 2009). Ethanol yield from syngas was dependent on the gas

quality, cold gas efficiency (CGE) and gas to ethanol conversion efficiency. The investment in

the ethanol industry may also depend on the processes integrated at the processing plant. The H2

enriched syngas production was dependent on the type of biomass, biomass to steam and Ca/C

ratio, and the type of gasifier and gasification temperature (Guoxin & Hao, 2009; Grasa &

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

S1 S2 S3 S4

Cost

, $/L

Heat recovery

Distillation

Fermentation

Gasification

Pretreatment

Transportation

Feedstock

Fixed cost

122

Abanades, 2006; Corella et al., 2006). Consequently, the effect of the variation (±20 to ±60%) at

each stage of the LC of ethanol (transportation, pre-treatment, gasification, fermentation,

distillation, heat recovery processes), and feedstock and fixed cost on the LC of ethanol were

evaluated.

The net energy consumption, GHG emissions and production cost slightly varied, even

variation was considered to be ±20 to ±60% for transportation and pretreatment because of their

little contribution in the LC of ethanol (Figs. 9.5–9.6). Figures 9.7–9.8 show the effect of the

variation in gasification and heat recovery on the LC of ethanol. The net energy consumption,

GHG emissions and production cost were varied from 7.44–21.67 MJ/L, 0.85–1.70 kg-CO2 e/L

and 0.65–0.92 $/L, respectively. It seems that the effect of the variation in gasification was more

robust compared with heat recovery, because contribution of gasification process was greater

than that of heat recovery.

Figure 9.5 Effect of transportation and pretreatment on net energy consumption

0

3

6

9

12

15

Pretreatment

+20%Pretreatment

+40%

Pretreatment

+60%

Pretreatment -

20%

Pretreatment -

40%

Pretreatment -

60%

S1Transportation

+20%

Transportation

+40%

Transportation

+60%

Transportation -

20%

Transportation -

40%

Transportation -

60%

123

Figure 9.6 Effect of transportation and pretreatment on emission and cost

Figure 9.7 Effect of the variation of gasification and heat recovery on net energy consumption

(MJ/L)

0.0

0.7

1.4Pretreatment +20%

Pretreatment +40%

Pretreatment +60%

Pretreatment -20%

Pretreatment -40%

Pretreatment -60%

S1Transportation +20%

Transportation +40%

Transportation +60%

Transportation -20%

Transportation -40%

Transportation -60%

kg-CO2e/L $/L

0

6

12

18

24

Gasification

+20%Gasification

+40%

Gasification

+60%

Gasification -

20%

Gasification -

40%

Gasification -

60%

S1Heat recovery

+20%

Heat recovery

+40%

Heat recovery

+60%

Heat recovery -

20%

Heat recovery -

40%

Heat recovery -

60%

124

Figure 9.8 Effect of the variation of gasification and heat recovery on emission and cost

Figures 9.9–9.10 represents the effect of the variation of fermentation and distillation on

the net energy consumption, GHG emissions and production cost. The estimated net energy

consumption, GHG emissions and production cost were found to be varied from 12.88–16.22

MJ/L, 1.17–1.38 kg-CO2 e/L and 0.72–0.85 $/L, respectively. The effect of distillation on the

net energy consumption seems to be robust compared to that of fermentation, because of the

variation between distillation and fermentation (Fig. 9.9). The variation in GHG emissions and

production cost dependent on the severity of variation (Fig. 9.10). The net production cost was

also observed to be varied from 0.70–0.84 $/L depending on the severity of the variation of fixed

and feedstock cost (Fig. 9.11). It seems that the variation of fixed cost has more impact on the

production cost than that of feedstock cost, because fixed has greater contribution. The LC GHG

emissions and production cost were also found to be dependent on the CGE (Fig. 9.12). GHG

emissions and production cost were found to be about 0.95–1.46 kg-CO2 e/L and 0.64–0.86 $/L,

respectively depending on the CGE. This variation was resulted because of the variation in

ethanol yield which was dependent on the CGE.

0.0

0.6

1.2

1.8Gasification +20%

Gasification +40%

Gasification +60%

Gasification -20%

Gasification -40%

Gasification -60%

S1Heat recovery +20%

Heat recovery +40%

Heat recovery +60%

Heat recovery -20%

Heat recovery -40%

Heat recovery -60%

kg-CO2e/L $/L

125

Figure 9.9 Effect of the variation of fermentation and distillation on net energy consumption

(MJ/L)

Figure 9.10 Effect of variation of fermentation and distillation on emission and cost

-

4

8

12

16

Fermentation

+20%Fermentation

+40%

Fermentation

+60%

Fermentation -

20%

Fermentation -

40%

Fermentation -

60%

S1Distillation

+20%

Distillation

+40%

Distillation

+60%

Distillation -20%

Distillation -40%

Distillation -60%

0.0

0.5

1.0

1.5Fermentation +20%

Fermentation +40%

Fermentation +60%

Fermentation -20%

Fermentation -40%

Fermentation -60%

S1Distillation +20%

Distillation +40%

Distillation +60%

Distillation -20%

Distillation -40%

Distillation -60%

kg-CO2e/L $/L

126

Figure 9.11 Effect of variation of fixed and feedstock cost on production cost ($/L)

Figure 9.12 Effect of CGE on GHG emissions and production cost

This study also supports the earlier findings (chapter 6 & 7) that environmentally and

economically viable ethanol can be produced from miscanthus by gasification-biosynthesis

process with or without CLG, even if miscanthus is grown on the marginal land. The

0.0

0.3

0.6

0.9Feedstock cost+10%

Feedstock cost+20%

Feedstock cost+30%

Feedstock cost-10%

Feedstock cost-20%

Feedstock cost-30%

S1Fixed cost+10%

Fixed cost+20%

Fixed cost+30%

Fixed cost-10%

Fixed cost-20%

Fixed cost-30%

0.95

1.11

1.3

1.32

1.48 0.64

0.68

0.76 0.78

0.86

-

0.5

1.0

1.5CGE +20%

CGE +10%

S1CGE -10%

CGE -20%

kg-CO2e/L $/L

127

environmental and economic benefit from miscanthus based ethanol may lead the stakeholders

target the quality land (prime land) for better profit margin. Consequently, miscanthus

production for ethanol industry needs to be regulated to avoid any sort of competition with food

crops.

9.4 Conclusion

This study reveals that both the environmental and cost benefit can be achieved from

ethanol that produced from miscanthus with adopted technologies. Although a slight variation

was observed among the scenarios of this study, the untreated (non-torrefied) miscanthus is

emerged to be a better option in terms of energy consumption, GHG emissions and production

cost compared with torrefied miscanthus. The CLG process reduced net energy consumption and

GHG emissions for both the untreated and treated miscanthus but production cost was increased

compared with the non-CLG process. Untreated miscanthus used in CLG and non-CLG process

was found to be the best option interms of GHG esmissions and production cost, respectively.

128

Chapter 10

Conclusions and Recommendations

10.1 Conclusions

This study is a new approach for the development and determination of a novel and

adoptable renewable energy technology. The techno-economic and environmental evaluation of

the life cycle (LC) of ethanol produced from various biomasses (agri-residue: wheat straw; forest

residue: sawdust; energy crop: miscanthus) is performed by adopting different technological

approaches. Life cycle assessment (LCA) methodologies have been used in the evaluation

processes. The potential plant locations for the miscanthus based ethanol plant in Ontario,

Canada have also been identified to abate greenhouse gas (GHG) emissions and minimize the

production cost. Net energy consumption, GHG emissions and production cost are found to be

dependent on the feedstock, plant location, conversion technology, processing plant size, system

boundaries, biomass logistics and assumptions. This study makes it possible to draw a

comparison among the selected technologies and feedstock, determine the most suitable

pathways, and generate useful and novel information to facilitate the stakeholders involved in

bioenergy sectors. It is worthy to note that although the LCA is a powerful tool for the evaluation

of the environmental effects of a product/process/activities, results are dependent on the data

quality system boundary, process modeling, time horizon and geographical location.

10.1.1 Evaluation of the LC of ethanol produced by enzymatic hydrolysis process

The LCA study depicts that environmental benefit can be gained with present technologies

if wheat straw is considered to be carbon neutral, otherwise both environmental and economic

viabilities of ethanol from wheat straw are doubtful while carbon sequestration is not considered.

The environmental viability of ethanol from wheat straw can be improved if carbon sequestration

is considered (resulted in negative GHG emissions).

The evaluation of the LC of ethanol from sawdust reveals that despite estimated

environmental benefit, its economic viability remains doubtful unless the FiT program is

considered. A modified agro-industrial and renewable energy policy that allows FiT to the

lignocellulosic ethanol industry in Ontario not only reduces production cost but may also

encourage future investment and create more green jobs as well as help in achieving committed

GHG emissions reduction targets in Canada.

129

Ethanol derived from miscanthus is found to be environmentally preferable and

economically viable at all locations in Ontario; however, Eastern Ontario has appeared as the

best option for the miscanthus based ethanol industry, if miscanthus is grown on marginal land.

Although a slight variation is observed in the case of net energy consumption and production

cost among the scenarios, the variation is robust in the case of GHG emissions where carbon

dynamics plays a key role. This study also revealed that GHG emissions are dependent not only

on the land classes but also on the crop displacement. It is worthy to mention that both the

environment and economic benefits can be gained, even if miscanthus is grown on the marginal

land in Ontario for ethanol. Consequently, miscanthus grown on marginal land has emerged as a

promising feedstock for the ethanol industry in Ontario, which may avoid any sort of

competition over food crops for better quality land.

10.1.2 Evaluation of the LC of ethanol produced by gasification-biosynthesis process

This study also confirmed that both the environmental and cost benefit can be gained from

ethanol that is produced from miscanthus with adopted technologies. The untreated (non-

torrefied) miscanthus is emerged to be a better option in terms of energy consumption, GHG

emissions and production cost compared with torrefied miscanthus. The CLG process reduced

net energy consumption and GHG emissions for both the untreated and treated (torrefied)

miscanthus; however, production cost has increased. The miscanthus based ethanol industry

might need to be regulated to avoid any sort of competition for higher quality land. Also a

careful consideration needs to be given to crop rotation/replacement to improve soil carbon

dynamics, to abate any productivity loss, and may improve farm income and rural economy.

This study revealed that GHG emissions and the production cost of ethanol are dependent

on feedstock, conversion technologies, system boundaries, allocation methods, and the utilization

of byproducts. The LCA study also confirmed that both technological pathways are

environmentally and economically viable. Although, the results of this study indicate that similar

benefits can be gained, they seem to be inclined towards the gasification-biosynthesis pathway.

Biotechnological advances, especially in enzyme production would improve the viability of

enzymatic hydrolysis process. The novel information generated in this study may help the

stakeholders in their decision making processes, attract more investment in this sector, help meet

the ethanol demand, and help achieve GHG emission target of Canada.

130

10.1.3 Continuous stirred tank bioreactor

This is a new approach for the development of a continuous stir tank bioreactor. It seems

that the developed bioreactor can be used to generate ethanol from syngas. However, this study

needs to be continued for further evaluation and development.

10.2 Recommendations

10.2.1 Life cycle assessment

The LC of ethanol is evaluated based on both the literature, and estimated and simulated

data. All the literature data are not from the same plant size and specific site. Some optimistic

assumptions are made to evaluate the LC of ethanol, especially for the enzyme production. The

results of this study advocate for a full LCA model based on the bench or pilot scale plant (for

onsite data) that can facilitate an effective decision making process for the lignocellulosic

ethanol industry. Consequently, the following recommendations can be made for further study on

the LC of lignocellulosic ethanol:

i. Site specific or country specific data from the same plant size needs to be collected, if

available for further evaluation.

ii. A bench scale or pilot scale lignocellulosic ethanol plant needs to be built for in-depth

evaluation purposes before any future investment and commercial production.

iii. Integrated and innovative bio-refinery approaches need to be taken into account for

future study.

iv. Alternate use of byproducts or coproducts, especially in the case of CLG where H2

enriched syngas is produced, needs to be considered.

v. The variations of agricultural land price and miscanthus yield and processing plant

capacity also needs to be considered for further study.

vi. The LCA study on LC of ethanol can also be extended to other environmental

paramentes, such as eutrofication, acidification, water consumption, and human toxicity.

10.2.2 Improvement of bioreactor

To improve the gas-liquid mass transfer, thus the ethanol yield the following

recommendations are made:

i. An innovative gas supply system needs to be developed to improve gas retention time in

the liquid.

131

ii. Micro-bubble technology can be used to improve gas retention time in the liquid, thus

the mass transfer.

iii. Alternative/innovative reactor design can also be considered.

iv. Various gas and media flow rates, and stirrer speed can be used to evaluate the mass

transfer and the ethanol yield.

v. The effect of cell density also needs to be evaluated.

Finally, it seems that long term national or international support on research and

development process would help to develop innovative technologies, consequently improve the

viability of the lignocellulosic ethanol.

132

Chapter 11

References

Abubackar, H. N., Veiga, M. C., & Kennes, C. (2011). Biological conversion of carbon

monoxide: rich syngas or waste gases to bioethanol. Biofuels, Bioproducts and

Biorefining, 5(1), 93–114.

Acharya, B. (2011). Chemical looping gasification of biomass for hydrogen-enriched gas

production (Doctoral dissertation). Dalhousie University, Halifax, Nova Scotia, Canada.

Acharya, B. (2013). Torrefaction and pelletization of different forms of biomass of Ontario

(Master Thesis), University of Guelph, Canada.

Acharya, B., Dutta, A., & Basu, P. (2009). Chemical-looping gasification of biomass for

hydrogen-enriched gas production with in-process carbon dioxide capture. Energy & Fuels,

23(10), 5077–5083.

Ackom, E. K., Mabee, W. E., & Saddler, J. N. (2010). Industrial sustainability of competing

wood energy options in Canada. Applied biochemistry and biotechnology, 162(8), 2259–

2272.

Aden, A. (2008). Biochemical production of ethanol from corn stover: State of technology

model. Golden, CO: National Renewable Energy Laboratory, Report No. NREL/TP-510–

43205, < http://www.nrel.gov/docs/fy08osti/43205.pdf>.

Aden, A., Ruth, M., Ibsen, K., Jechura, J., Neeves, K., Sheehan, J., Wallace, B., Montague, L.,

Slayton, A., Lukas, J. (2002). Lignocellulosic biomass to ethanol process design and

economics utilizing co-current dilute acid prehydrolysis and enzymatic hydrolysis for corn

stover. NREL/TP-510-32438. National Renewable Energy Laboratory, Golden, Colorado,

USA.

AEA Technology (2003) International resource costs of biodiesel and bioethanol. Department for

Transport, Government of United Kingdom, London, UK.

AEBIOM (European Biomass Association), (2010). New dedicated energy corps for solid

biofuels.

<http://www.aebiom.org/IMG/pdf/Dedicated_energy_crops_for_solid_biofuels_2008_Janu

ary.pdf>.

133

Alinia, R., Zabihi, S., Esmaeilzadeh, F., & Kalajahi, J.F. (2010). Pretreatment of wheat straw by

supercritical CO2 and its enzymatic hydrolysis for sugar production. Biosystems

Engineering, 107(1), 61–66.

Aravind, P. V., & de Jong, W. (2012). Evaluation of high temperature gas cleaning options for

biomass gasification product gas for solid oxide fuel cells. Progress in Energy and

Combustion Science, 38(6), 737–764.

Asano, K., Minowa, T., (2007). An analysis of bioethanol production costs and CO2 reduction

costs in Japan. Journal of the Japan Institute of Energy, 86(11), 957–963 (In Japanese with

English abstract).

vil s art ne , , Saucedo-Luna, J., Segovia-Hernandez, J. G., Hernandez, S., Gomez-Castro,

F. I., & Castro-Montoya, A. J. (2011). Dehydration of bioethanol by hybrid process liquid–

liquid extraction/extractive distillation. Industrial & Engineering Chemistry Research,

51(17), 5847–5855.

Bain, R. L., Dayton, D. C., Carpenter, D. L., Czernik, S. R., Feik, C. J., French, R. J., Kimberly

A. Magrini-Bair... & Phillips, S. D. (2005). Evaluation of catalyst deactivation during

catalytic steam reforming of biomass-derived syngas. Industrial & Engineering Chemistry

Research, 44(21), 7945–7956.

Balan, V., Kumar, S., Bals, B., Chundawat, S., Jin, M., & Dale, B., (2012). Biochemical and

thermochemical conversion of switchgrass to biofuels In “Switchgrass: Valuable

Biomass Crop for Energy” Edited by ndrea onti, Springer

Balat, M. (2011). Production of bioethanol from lignocellulosic materials via the biochemical

pathway: A review. Energy Conversion and Management, 52(2), 858–875.

Baldwin, R.M., Magrini-Bair, K.A., Nimlos, M.R., Pepiot, P., Donohoe, B.S., Hensley, J.E., &

Phillips, S.D. (2011). Current research on thermochemical conversion of biomass at the

National Renewable Energy Laboratory. Applied Catalysis B: Environmental. 115, 320–

329.

Ballerini, D., Desmarquest, J.P., Pourquie, J., Nativel, F., & Rebeller, M. (1994). Ethanol

production from lignocellulosics: Large scale experimentation and economics. Bioresource

Technology, 50(1), 17–23.

Banerjee, G., Scott-Craig, J. S., & Walton, J. D. (2010b). Improving enzymes for biomass

conversion: a basic research perspective. BioEnergy Research, 3(1), 82–92.

134

Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., & Pandey, R. A.

(2010a). Commercializing lignocellulosic bioethanol: technology bottlenecks and possible

remedies. Biofuels, Bioproducts and Biorefining, 4(1), 77–93.

Banse, M., van Meijl, H., Tabeau, A., Woltjer, G., Hellmann, F., & Verburg, P. H. (2011).

Impact of EU biofuel policies on world agricultural production and land use. Biomass and

Bioenergy, 35(6), 2385–2390.

Barta, Z., Kovacs, K., Reczey, K., & Zacchi, G. (2010b). Process design and economics of on-

site cellulase production on various carbon sources in a softwood-based ethanol plant.

Enzyme Research, Article ID734182.

Barta, Z., Reczey, K. & Zacchi, G. (2010a) Techno-economic evaluation of stillage treatment

with anaerobic digestion in a softwood-to-ethanol process. Biotechnology for Biofuels,

3(1), 21.

Beale, C. V., & Long, S. P. (1995). Can perennial C4 grasses attain high efficiencies of radiant

energy conversion in cool climates?. Plant, Cell & Environment, 18(6), 641–650.

Bennett, R., Phipps, R., Strange, A., & Grey, P. (2004). Environmental and human health

impacts of growing genetically modified herbicide‐tolerant sugar beet: a life‐cycle

assessment. Plant Biotechnology Journal, 2(4), 272–278.

Bessou, C., Ferchaud, F., Gabrielle, B., & Mary, B. (2011). 2 Biofuels, greenhouse gases and

climate change: A review. Agronomy for Sustainable Development, 31(1), 1–79.

Bhatty, J. I., Gajda, J., & Miller, F. M. (2003). Commercial use of high-carbon fly ash in cement

manufacturing. netl. doe. gov/publications/proceedings/02/ubc/bhattysummary. pdf>(14

April 2003).

Biofuelnet, (2013). An overview of Canadian federal biofuels programs. Biofuelnet, <

http://www.biofuelnet.ca/2013/10/03/an-overview-of-canadian-federal-biofuels-programs/

>.

Biomass Energy Centre (BEC), (2011). Ash,

<http://www.biomassenergycentre.org.uk/portal/page?_pageid=77,109209&_dad=portal&

_schema=PORTAL>.

Bisaria, V. S., & Ghose, T. K. (1981). Biodegradation of cellulosic materials: substrates,

microorganisms, enzymes and products. Enzyme and Microbial Technology, 3(2), 90–104.

135

Bjerre, A. B., Olesen, A. B., Fernqvist, T., Plöger, A., & Schmidt, A. S. (1996). Pretreatment of

wheat straw using combined wet oxidation and alkaline hydrolysis resulting in convertible

cellulose and hemicellulose. Biotechnology and Bioengineering, 49(5), 568–577.

Blanco-Canqui, H., & Lal, R. (2009). Corn stover removal for expanded uses reduces soil

fertility and structural stability. Soil Science Society of America Journal, 73(2), 418–426.

Bocquého, G., & Jacquet, F. (2010). The adoption of switchgrass and miscanthus by farmers:

Impact of liquidity constraints and risk preferences. Energy Policy, 38(5): 2598–2607.

Börjesson, P., & Tufvesson, L. M. (2011). Agricultural crop-based biofuels–resource efficiency

and environmental performance including direct land use changes. Journal of Cleaner

Production, 19(2): 108–120.

Botha, T., & Von Blottnitz, H. (2006). A comparison of the environmental benefits of bagasse-

derived electricity and fuel ethanol on a life-cycle basis. Energy Policy, 34(17): 2654–

2661.

Boyle, M., Barron, N., & McHale, A. P. (1997). Simultaneous saccharification and fermentation

of straw to ethanol using the thermotolerant yeast strain Kluyveromyces marxianus imb3.

Biotechnology Letters, 19(1), 49–51.

Bradley, D. (2010). Canada report on bioenergy 2010. IEA, editor. Bioenergy Task, 40.

Brandão, M., Milà i Canals, L., & Clift, R. (2011). Soil organic carbon changes in the cultivation

of energy crops: Implications for GHG balances and soil quality for use in LCA. Biomass

and Bioenergy, 35(6), 2323–2336.

Bredwell, M. D., Srivastava, P., & Worden, R. M. (2008). Reactor design issues for synthesis‐

gas fermentations. Biotechnology Progress, 15(5), 834–844.

Brentrup, F., Küsters, J., Kuhlmann, H., & Lammel, J. (2004a). Environmental impact

assessment of agricultural production systems using the life cycle assessment

methodology: I. Theoretical concept of a LCA method tailored to crop production.

European Journal of Agronomy, 20(3), 247–264.

Brentrup, F., Küsters, J., Lammel, J., Barraclough, P., & Kuhlmann, H. (2004b). Environmental

impact assessment of agricultural production systems using the life cycle assessment

(LCA) methodology II. The application to N fertilizer use in winter wheat production

systems. European Journal of Agronomy, 20(3), 265–279.

136

BRI (2008). BRI energy bioenergy process – technology summary. BRI Energy Inc.,

http://www.brienergy.com/.

Bridgwater, A. V. (1994). Catalysis in thermal biomass conversion. Applied Catalysis A:

General, 116(1), 5–47.

Bridgwater, A. V. (1995). The technical and economic feasibility of biomass gasification for

power generation. Fuel, 74(5), 631–53.

Brosse N, Sannigrahi P, Ragauskas A: Pretreatment of Miscanthus x giganteus Using the Ethanol

Organosolv Process for Ethanol Production. Ind Eng Chem Res 2009, 48, 8328–8334.

Brown, R. C. (2007). Hybrid thermochemical/biological Processing: Putting the cart before the

horse? Applied Biochemistry and Biotechnology, 137(1), 947–956.

Bryant, C. (2009). Step change in cellulosic ethanol–the future is moving closer. In National

Ethanol Conference, May 2009, San Antonio, TX.

Calvo, L. F., Gil, M. V., Otero, M., Morán, A., & García, A. I. (2012). Gasification of rice straw

in a fluidized-bed gasifier for syngas application in close-coupled boiler-gasifier systems.

Bioresource Technology, 109, 206–214.

Campbell, J.E., Lobell, D.B. & Field C.B. (2009) Greater transportation energy and GHG offsets

from bioelectricity than ethanol. Science, 324, 1055–1057.

Cardona, C. A., & Sánchez, O. J. (2006). Energy consumption analysis of integrated flowsheets

for production of fuel ethanol from lignocellulosic biomass. Energy, 31(13), 2447–2459.

Cardona, C. A., & Sánchez, Ó. J. (2007). Fuel ethanol production: process design trends and

integration opportunities. Bioresource Technology, 98(12), 2415–2457.

Carpenter, D. L., Bain, R. L., Davis, R. E., Dutta, A., Feik, C. J., Gaston, K. R., Jablonski, W.,

Phillips, S. D., & Nimlos, M. R. (2010). Pilot-scale gasification of corn stover, switchgrass,

wheat straw, and wood: 1. Parametric study and comparison with literature. Industrial &

Engineering Chemistry Research, 49(4), 1859–1871.

Carpenter, D.L,, Bain, R.L, Davis, R.E., Dutta, A., Feik, C.J., Gaston, K.R., Jablonski, W.,

Phillips, S.D., Nimlos, M.R., (2010). Pilot-scale gasification of corn stover, switchgrass,

wheat straw, and wood: 1. Parametric study and comparison with literature. Industrial &

Engineering Chemistry Research. 49(4), 1859–1871.

Cederberg, C., & Mattsson, B. (2000). Life cycle assessment of milk production—a comparison

of conventional and organic farming. Journal of Cleaner Production, 8(1), 49–60).

137

Chang, I. S., Kim, D. H., Kim, B. H., Shin, P. K., Sung, H. C., & Lovitt, R. W. (1998). CO

fermentation of Eubacterium limosum KIST612. Journal of Microbiology and

Biotechnology, 8(2), 134–140.

Chang, V. S., Nagwani, M., Kim, C. H., & Holtzapple, M. T. (2001). Oxidative lime

pretreatment of high-lignin biomass. Applied Biochemistry and Biotechnology, 94(1), 1–

28.

Chatveera, B., & Lertwattanaruk, P. (2011). Durability of conventional concretes containing

black rice husk ash. Journal of Environmental Management, 92(1), 59–66.

Cherubini, F., & Jungmeier, G. (2010). LCA of a biorefinery concept producing bioethanol,

bioenergy, and chemicals from switchgrass. The International Journal of Life Cycle

Assessment, 15(1), 53–66.

Cherubini, F., & Strømman, A.H. (2011). Life cycle assessment of bioenergy systems: State of

the art and future challenges. Bioresource Technology, 102(2), 437–451.

Cherubini, F., & Ulgiati, S. (2010). Crop residues as raw materials for biorefinery systems–A

LCA case study. Applied Energy, 87(1), 47–57.

Cherubini, F., Bird, N. D., Cowie, A., Jungmeier, G., Schlamadinger, B., & Woess-Gallasch, S.

(2009). Energy-and greenhouse gas-based LCA of biofuel and bioenergy systems: Key

issues, ranges and recommendations. Resources, Conservation and Recycling, 53(8), 434–

447.

Clausen, E.C. & Gaddy, J.L. (1993). Ethanol from biomass by gasification/fermentation.

<www.anl.gov/PCS/acsfuel/preprint%20archive/Files/38_3_CHICAGO_08> (Accessed on

18/09/2011).

Collura, S., Azambre, B., Finqueneisel, G., Zimny, T., & Weber, J. V. (2006). Miscanthus×

Giganteus straw and pellets as sustainable fuels. Environmental chemistry letters, 4(2), 75–

78.

Consonni, S., Katofsky, R. E., & Larson, E. D. (2009). A gasification-based biorefinery for the

pulp and paper industry. Chemical Engineering Research and Design, 87(9), 1293–1317.

Contreras, A. M., Rosa, E., Pérez, M., Van Langenhove, H., & Dewulf, J. (2009). Comparative

life cycle assessment of four alternatives for using by-products of cane sugar production.

Journal of Cleaner Production, 17(8), 772–779.

138

Cotter, J. L., Chinn, M. S., & Grunden, A. M. (2009). Influence of process parameters on growth

of Clostridium ljungdahlii and Clostridium autoethanogenum on synthesis gas. Enzyme

and Microbial Technology, 44(5), 281–288.

Cotter, J. L. (2007). Ethanol and acetate production from synthesis gas using microbial catalysts

(Master Thesis, North Carolina State University).

Cotter, J. L., Chinn, M.S., & Grunden, A.M. (2008) Ethanol and acetate production by

Clostridium ljungdahlii and Clostridium autoethanogenum using resting cells. Bioprocess

and Biosystems Engineering, 32, 369–380.

Cowie, A. L., Smith, P., & Johnson, D. (2006). Does soil carbon loss in biomass production

systems negate the greenhouse benefits of bioenergy?. Mitigation and Adaptation

Strategies for Global Change, 11(5), 979–1002.

Croezen, H., & Kampman, B. (2009). The impact of ethanol and ETBE blending on refinery

operations and GHG-emissions. Energy Policy, 37(12), 5226–5238.

Čuček, L , & Kravanja, Z (20 0) Sustainable LC -based MINLP synthesis of bioethanol

processes. Computer Aided Chemical Engineering, 27, 1889–1894.

Čuček, L , art n, , Grossmann, I E , & Kravanja, Z (20 ) Energy, water and process

technologies integration for the simultaneous production of ethanol and food from the

entire corn plant. Computers & Chemical Engineering, 35(8), 1547–1557.

Čuček, L , art n, , Grossmann, I E , & Kravanja, Z (20 4) ulti-period synthesis of

optimally-integrated biomass and bioenergy supply network. Computers & Chemical

Engineering.

da Silva, F. L. H., Rodrigues, M. I., & Maugeri, F. (1999). Dynamic modelling, simulation and

optimization of an extractive continuous alcoholic fermentation process. Journal of

Chemical Technology and Biotechnology, 74(2), 176–182.

Datar, R.P., Shenkman, R.M., Cateni, B.G., Huhnke, R.L., Lewis, R.S. (2004). Fermentation of

biomass-generated producer gas to ethanol. Biotechnology and Bioengineering, 86, 587–

594.

Daugherty, E.C. (2001). Biomass energy systems efficiency: Analyzed through a LCA study.

Master of Science Thesis, Lund University.

139

Daystar, J. S., Venditti, R. A., Gonzalez, R., Jameel, H., Jett, M., & Reeb, C. W. (2013). Impacts

of feedstock composition on alcohol yields and greenhouse gas emissions from the NREL

thermochemical ethanol conversion process. BioResources, 8(4).

De Paoli, F., Bauer, A., Leonhartsberger, C., Amon, B., & Amon, T. (2011). Utilization of by-

products from ethanol production as substrate for biogas production. Bioresource

Technology, 102(11), 6621–6624.

De Vrije, T. D., De Haas, G. G., Tan, G. B., Keijsers, E. R. P., & Claassen, P. A. M. (2002).

Pretreatment of miscanthus for hydrogen production by Thermotoga elfii. International

journal of hydrogen energy, 27(11), 1381–1390.

DEFRA (Department for Environment Food & Rural Affairs), 2007. Planting and growing

miscanthus: Best practice guidelines for applicants to Defra’s energy crops scheme

<http://www.defra.gov.uk/erdp/regions/default.htm>.

Devarapalli, M., Atiyeh, H. K., Phillips, J. R., Orgillb, J. J., Lewis, R. S., & Huhnke, R. L.

(2013). Comparison of syngas fermentation reactors for biological alcohol production. <

http://sungrant.tennessee.edu/NR/rdonlyres/DDF120E1-C312-4065-B095-

6EC87BD11DA8/3731/Atiyeh_Hasan.pdf>.

Devi, S., Suhag, M., Dhaka, A., Singh, J. (2011). Biochemical Conversion Process of Producing

Bioethanol from Lignocellulosic Biomass. International Journal of Microbial Resource

Technology, 1(1), 28–32.

Dias, M. O., Junqueira, T. L., Maciel Filho, R., Maciel, M. R., & Eduardo Vaz Rossell, C.

(2009). Anhydrous bioethanol production using bioglycerol–simulation of extractive

distillation processes. Computer Aided Chemical Engineering, 26, 519–524.

DOE (2006) Breaking the Biological Barriers to Cellulosic Ethanol: A Joint Research Agenda

(DOE/SC-0095). U.S. Department of Energy, Rockville, MD.

Duc Long, N., & Lee, M. (2013). Optimal retrofit design of extractive distillation to energy

efficient thermally coupled distillation scheme. AIChE Journal, 59(4), 1175–1182.

Dutta, A., Bain, R.L., Biddy, M.J. (2010b). Techno-economics of the production of mixed

alcohols from lignocellulosic biomass via high-temperature gasification. Environmental

Progress & Sustainable Energy, 29(2), 163–174.

140

Dutta, A., Dowe, N., Ibsen, K. N., Schell, D. J., & Aden, A. (2010a). An economic comparison

of different fermentation configurations to convert corn stover to ethanol using Z. mobilis

and Saccharomyces. Biotechnology Progress, 26(1), 64–72.

Dyer, J. A., Vergé, X. P. C., Desjardins, R. L., Worth, D. E., & McConkey, B. G. (2010). The

impact of increased biodiesel production on the greenhouse gas emissions from field crops

in Canada. Energy for Sustainable Development, 14(2), 73–82.

EIA (Energy Information Administration). Official Energy Statistics from the US Govt., 2010; <

http://www.eia.doe.gov/cneaf/electricity/epa/epat7p4.html>(Accessed: 06.8.2010).

Ensinas, A. V., Codina, V., Marechalb, F., Albarelli, J., & Silva, M. A. (2013). Thermo-

Economic Optimization of Integrated First and Second Generation Sugarcane Ethanol

Plant. Chemical Engineering, 35.

Environment Canada (2010). Renewable Fuels Regulations. <http://www.ec.gc.ca/energie-

energy/default.asp?lang=En&n=0AA71ED2-1> (accessed on 2/11/2011)

Eriksson, G., Kjellström, B., Lundqvist, B., & Paulrud, S. (2004). Combustion of wood

hydrolysis residue in a 150kW powder burner. Fuel, 83(11), 1635–41.

Ethanol Producer Magazine (EPM), 2014. Canadian ethanol plants. Etahnol Producer Magazine,

<http://www.ethanolproducer.com/plants/listplants/Canada/All/Cellulosic>.

Evans, S., Patriquin, D., & Scott, J. (2005). Small plot comparisons of phenology, yield, disease,

and weed tolerance of three heritage and two newer cultivars of bread wheat under a high

fertility organic regime in eastern Canada. URL: http://dp.biology.dal.ca/wheat/ (accessed

on 22/10/2011).

FAO (Food and Agriculture Organization of the United Nations) (2011). Food and agricultural

commodities production. [online database], URL:

http://faostat.fao.org/site/339/default.aspx (accessed on 4/10/2011).

Fargione, J., Hill, J., Tilman, D., Polasky, S., & Hawthorne, P. (2008). Land clearing and the

biofuel carbon debt. Science, 319(5867), 1235–1238.

Farrell, A. E., Plevin, R. J., Turner, B. T., Jones, A. D., O'hare, M., & Kammen, D. M. (2006).

Ethanol can contribute to energy and environmental goals. Science, 311(5760), 506–508.

Fazio, S., & Monti, A. (2011). Life cycle assessment of different bioenergy production systems

including perennial and annual crops. Biomass and Bioenergy, 35(12), 4868–4878.

141

Fleming, J.S., Habibi, S., & MacLean, H.L. (2006). Investigating the sustainability of

lignocellulose-derived fuels for light-duty vehicles. Transportation Research Part D:

Transport and Environment, 11(2), 146–159.

Foust, T.D., Aden, A., Dutta, A., & Phillips, S. (2009). An economic and environmental

comparison of a biochemical and a thermochemical lignocellulosic ethanol conversion

processes. Cellulose, 16(4), 547–565.

Frederick Jr, W. J., Lien, S. J., Courchene, C. E., DeMartini, N. A., Ragauskas, A. J., & Iisa, K.

(2008). Production of ethanol from carbohydrates from loblolly pine: A technical and

economic assessment. Bioresource Technology, 99(11), 5051–5057.

Fu, G.Z., Chan, A.W., & Minns, D.E. (2003). Life cycle assessment of bio-ethanol derived from

cellulose. The International Journal of Life Cycle Assessment, 8(3), 137–141.

Fu, J. (2004a). Simulation of salt-containing extractive distillation for the system of

ethanol/water/ethanediol/KAc. 1. Calculation of the vapor-liquid equilibrium for the salt-

containing system. Industrial & Engineering Chemistry Research, 43(5), 1274–1278.

Fu, J. (2004b). Simulation of salt-containing extractive distillation for the system of

ethanol/water/ethanediol/KAc. 2. Simulation of salt-containing extractive distillation.

Industrial & Engineering Chemistry Research, 43(5), 1279–1283.

Gabrielle, B., & Gagnaire, N. (2008). Life-cycle assessment of straw use in bio-ethanol

production: a case study based on biophysical modelling. Biomass and Bioenergy, 32(5),

431–441.

Galbe, M., & Zacchi, G. (2002). A review of the production of ethanol from softwood. Applied

microbiology and biotechnology, 59(6), 618–628.

Galik, C. S., & Abt, R. C. (2012). The effect of assessment scale and metric selection on the

greenhouse gas benefits of woody biomass. Biomass and Bioenergy, 44(), 1–7.

Gassner, M., & Maréchal, F. (2009). Thermo-economic process model for thermochemical

production of Synthetic Natural Gas (SNG) from lignocellulosic biomass. Biomass and

Bioenergy, 33(11), 1587–1604.

GBEP (Global Bioenergy Partnership), 2007. A review of the current state of bioenergy

development in G8+5 countries. GBEP, <http://www.globalbioenergy.

org/uploads/media/0711_GBEP_Report.pdf>.

142

Gerber, L., Gassner, M., & Maréchal, F. (2011). Systematic integration of LCA in process

systems design: Application to combined fuel and electricity production from

lignocellulosic biomass. Computers & Chemical Engineering, 35(7), 1265–1280.

Giarola, S., Bezzo, F., & Shah, N. (2013). A risk management approach to the economic and

environmental strategic design of ethanol supply chains. Biomass and Bioenergy, 58, 31–

51.

Gil, I. D., Gómez, J. M., & Rodríguez, G. (2012). Control of an extractive distillation process to

dehydrate ethanol using glycerol as entrainer. Computers & Chemical Engineering, 39,

129–142.

Gil, I. D., Uyazan, A. M., Aguilar, J. L., Rodríguez, G., & Caicedo, L. A. (2008). Separation of

ethanol and water by extractive distillation with salt and solvent as entrainer: process

simulation. Brazilian Journal of Chemical Engineering, 25(1), 207–215.

Global Bioenergy Partnership (GBEP) (2009). The GBEP common methodological framework

for GHG lifecycle analysis of bioenergy—Version Zero.

<http://www.globalbioenergy.org/fileadmin/user_upload/gbep/docs/2009_events/7th_SC_

NY/GBEP_GHG_report_2306.pdf>.

Gnansounou, E. (2010). Production and use of lignocellulosic bioethanol in Europe: Current

situation and perspectives. Bioresource Technology, 101(13), 4842–4850.

Gnansounou, E., & Dauriat, A. (2010). Techno-economic analysis of lignocellulosic ethanol: A

review. Bioresource Technology, 101(13), 4980–4991.

Gonzalez, R., Daystar, J., Jett, M., Treasure, T., Jameel, H., Venditti, R., & Phillips, R. (2012).

Economics of cellulosic ethanol production in a thermochemical pathway for softwood,

hardwood, corn stover and switchgrass. Fuel Processing Technology, 94(1), 113–122.

González-García, S., Iribarren, D., Susmozas, A., Dufour, J., & Murphy, R. J. (2012). Life cycle

assessment of two alternative bioenergy systems involving Salix spp. biomass: Bioethanol

production and power generation. Applied Energy, 95, 111–122.

González-García, S., Luo, L., Moreira, M.T., Feijoo, G., & Huppes, G. (2009). Life cycle

assessment of flax shives derived second generation ethanol fueled automobiles in Spain.

Renewable and Sustainable Energy Reviews, 13(8), 1922–1933.

143

Gopaul, S. G., Dutta, A., & Clemmer, R. (2014). Chemical looping gasification for hydrogen

production: A comparison of two unique processes simulated using ASPEN Plus.

International Journal of Hydrogen Energy, 39(11), 5804–5817.

Graham, R.L., Nelson, R., Sheehan, J., Perlack, R.D., & Wright, L.L. (2007). Current and

potential US corn stover supplies. Agronomy Journal, 99(1), 1–11.

Gray, K. A., Zhao, L., & Emptage, M. (2006). Bioethanol. Current opinion in chemical biology,

10(2), 141–146.

Greer, D. (Ed.). (2011). Digesters help fuel ethanol facilities. Biocycle, 52(2), 41–42.

Gregg, D. J., & Saddler, J. N. (1996). Factors affecting cellulose hydrolysis and the potential of

enzyme recycle to enhance the efficiency of an integrated wood to ethanol process.

Biotechnology and Bioengineering, 51(4), 375–383.

Gregg, D.J., Boussaid, A. & Saddler, J.N., (1989). Techno-economic evaluations of a generic

wood-to-ethanol process: effect of increased cellulose yields and enzyme recycle.

Bioresource Technology, 63, 7–12.

Grigal, D. F., & Berguson, W. E. (1998). Soil carbon changes associated with short-rotation

systems. Biomass and Bioenergy, 14(4): 371–378.

Guo, Y., Hu, S.Y., Li, Y.R., Chen, D.J., Zhu, B., & Smith, K.M. (2010). Optimization and

analysis of a bioethanol agro-industrial system from sweet sorghum. Renewable Energy,

35(12), 2902–2909.

Gyenge, L., Raduly, B., Barrera, R., Font, X., Lanyi, S., & Abraham, B. (2013, June). Efficiency

of biogas production from corn bioethanol by-products using different inocula. In Energy

(IYCE), 2013 4th International Youth Conference on (pp. 1–6). IEEE.

Haas, G., Wetterich, F., & Köpke, U. (2001). Comparing intensive, extensified and organic

grassland farming in southern Germany by process life cycle assessment. Agriculture,

Ecosystems & Environment, 83(1), 43–53.

Hahn-Hägerdal, B., Galbe, M., Gorwa-Grauslund, M.F., Liden, G., & Zacchi, G. (2006). Bio-

ethanol–the fuel of tomorrow from the residues of today. Trends in Biotechnology, 24(12),

549–556.

Hamelinck, C.N., Hooijdonk, G.V., & Faaij, A.P. (2005). Ethanol from lignocellulosic biomass:

techno-economic performance in short-, middle-and long-term. Biomass and Bioenergy,

28(4), 384–410.

144

Hammerschlag, R. (2006). Ethanol's energy return on investment: A survey of the literature

1990-Present. Environmental Science & Technology, 40(6), 1744–1750.

Han, M., Choi, G. W., Kim, Y., & Koo, B. C. (2011). Bioethanol production by Miscanthus as a

lignocellulosic biomass: Focus on high efficiency conversion to glucose and ethanol.

BioResources, 6(2), 1939–1953.

Han, J., & Kim, H. (2008). The reduction and control technology of tar during biomass

gasification/pyrolysis: an overview. Renewable and Sustainable Energy Reviews, 12(2),

397–416.

Haney, R. L., Kiniry, J. R., & Johnson, M. V. V. (2010). Soil microbial activity under different

grass species: Underground impacts of biofuel cropping. Agriculture, ecosystems &

environment, 139(4), 754–758.

Hansen, E. M., Christensen, B. T., Jensen, L. S., & Kristensen, K. (2004). Carbon sequestration

in soil beneath long-term Miscanthus plantations as determined by 13C abundance.

Biomass and Bioenergy, 26(2), 97–105.

Haro, P., Ollero, P., Villanueva Perales, A.L., & Reyes Valle, C. (2012). Technoeconomic

assessment of lignocellulosic ethanol production via DME (dimethyl ether)

hydrocarbonylation. Energy, 44(), 891–901.

Haskins, J , Wilson, E & Lenard, K (2006) Global leaders launch effort to turn around frica’s

failing agriculture. <http://www.eurekalert.org/africasoil/>.

Hatti-Kaul, R., Törnvall, U., Gustafsson, L., & Börjesson, P. (2007). Industrial biotechnology for

the production of bio-based chemicals–a cradle-to-grave perspective. Trends in

biotechnology, 25(3), 119–124.

He, J., & Zhang, W. (2011). Techno-economic evaluation of thermo-chemical biomass-to-

ethanol. Applied Energy, 88(4), 1224–1232.

Heijungs, R., Goedkoop, M., Struijs, J., Effting, S., Sevenster, M., & Huppes, G. (2003).

Towards a life cycle impact assessment method which comprises category indicators at the

midpoint and the endpoint level. Report of the first project phase: Design of the new

method VROM report, <http://www. leidenuniv. nl/cml/ssp/publications/recipe_phase1.

pdf>.

145

Heiskanen, H., Virkajarvi, I., & Viikari, L. (2007). The effects of syngas composition on the

growth and product formation of Butyribacterium methylotrophicum. Enzyme and

Microbial Technology 41(), 362–367.

Heller, M.C., Keoleian, G.A., & Volk, T.A. (2003). Life cycle assessment of a willow bioenergy

cropping system. Biomass and Bioenergy, 25(2), 147–165.

Henao, J., & Baanante, C. (2006). Agricultural production and soil nutrient mining in Africa;

Implications for resource conservation and policy development. IFDC, Muscle Shoals,

USA.

Hendriks, A. T. W. M., & Zeeman, G. (2009). Pretreatments to enhance the digestibility of

lignocellulosic biomass. Bioresource technology, 100(1), 10–18.

Henstra, A. M., Sipma, J., Rinzema, A., & Stams, A. J. (2007). Microbiology of synthesis gas

fermentation for biofuel production. Current Opinion in Biotechnology, 18(3), 200–206.

Heo, H. S., Park, H. J., Yim, J. H., Sohn, J. M., Park, J., Kim, S. S., Ryu, C., Jeon J. K., & Park,

Y. K. (2010). Influence of operation variables on fast pyrolysis of Miscanthus sinensis var.

purpurascens. Bioresource technology, 101(10), 3672–3677.

Herrera, S. (2006). Bonkers about biofuels. Nature Biotechnology, 24(7), 755–760.

Holtzapple, M. T., Jun, J. H., Ashok, G., Patibandla, S. L., & Dale, B. E. (1991). The ammonia

freeze explosion (AFEX) process. Applied Biochemistry and Biotechnology, 28(1), 59–74.

Hong, Y., Nizami, A. S., Pour Bafrani, M., Saville, B. A., & MacLean, H. L. (2013). Impact of

cellulase production on environmental and financial metrics for lignocellulosic ethanol.

Biofuels, Bioproducts and Biorefining, 7(3), 303–313.

Hsu, D. D. (2012). Life cycle assessment of gasoline and diesel produced via fast pyrolysis and

hydroprocessing. Biomass and Bioenergy, 45, 41–47.

Hsu, D. D., Inman, D., Heath, G. A., Wolfrum, E. J., Mann, M. K., & Aden, A. (2010). Life

cycle environmental impacts of selected US ethanol production and use pathways in 2022.

Environmental science & technology, 44(13), 5289–5297.

Hu, J., Wang, Y., Cao, C., Elliott, D.C., Stevens, D.J., White, J.F. (2007). Conversion of

biomass-derived syngas to alcohols and C2 oxygenates using supported Rh catalysts in a

microchannel reactor. Catalysis Today, 120, 90–95.

146

Huang, H.J., Ramaswamy, S., Al-Dajani, W., Tschirner, U., & Cairncross, R.A. (2009). Effect of

biomass species and plant size on cellulosic ethanol: A comparative process and economic

analysis. Biomass and Bioenergy, 33(2), 234–246.

Hurst, K. M., & Lewis, R. S. (2010). Carbon monoxide partial pressure effects on the metabolic

process of syngas fermentation. Biochemical Engineering Journal, 48(2), 159–165.

International Rice Research (IRRI), (2007). Technologies for energy use of rice straw: a review.

International Rice Research Notes, International Rice Research Institute, Philippines.

ISO (International Organization for Standardization) (1997). ISO 14040 Environmental

management—Life cycle assessment—Principles and framework.

ISO (International Organization for Standardization) (2006). ISO 14040: 2006(E) Environmental

management — Life cycle assessment —Principles and framework.

Itsubo, N., & Inaba, A. (2007). Development of LIME2–towards the establishment of

methodology for decision making. In SETAC-Europe Annual Meeting 2007.

JA-Zenno, 2002. The method for planning mechanization. 49 (in Japanese).

Jegannathan, K. R., & Nielsen, P. H. (2013). Environmental assessment of enzyme use in

industrial production–a literature review. Journal of Cleaner Production, 42, 228-240.

Jensen, J. R., Halvorsen, K. E., & Shonnard, D. R. (2011). Ethanol from lignocellulosics, US

federal energy and agricultural policy, and the diffusion of innovation. Biomass and

Bioenergy, 35(4), 1440–1453.

Jeschke. M., (2011). Sustainable corn stover harvest for biofuel production. Crop Insight, 21.

Johnson, E. (2009). Goodbye to carbon neutral: Getting biomass footprints right. Environmental

impact assessment review, 29(3), 165–168.

Johnson, J., (2006). Minimum C inputs and sustainability of biomass harvest.

<http://crops.confex.com/crops/2006am/techprogram/P20045.HTM>.

Jungmeier, G., Lingitz, A., Spitzer, J., Hofbauer, H., & Fürnsinn, S., (2007). Wood tobiofuels:

feasibility study for a biofuel plant in the Austrian province of Styria. In: Proceedings of

the 15th

European biomass conference and exhibition—from research to market

deployment, 7–11 May 2007, Berlin, Germany.

Junqueira, T. L., Dias, M. O., Filho, R. M., Wolf-Maciel, M. R., Rossell, C. E., & Atala, D. I.

(2009b). Proposition of alternative configurations of the distillation columns for bioethanol

147

production using vacuum extractive fermentation process. Chemical Engineering

Transaction, 17, 1627–1632.

Junqueira, T. L., Dias, M. O., Maciel, M. R., Maciel Filho, R., Rossell, C. E., & Atala, D. I.

(2009a). Simulation and optimization of the continuous vacuum extractive fermentation for

bioethanol production and evaluation of the influence on distillation process. Computer

Aided Chemical Engineering, 26, 827–832.

Kaar, W. E., & Holtzapple, M. T. (2000). Using lime pretreatment to facilitate the enzymic

hydrolysis of corn stover. Biomass and Bioenergy, 18(3), 189–199.

Kaboneka, S., Nivyiza, J. C., & Sibomana, L. (2004). Effects of nitrogen and phosphorus

fertilizer addition on wheat straw carbon decomposition in a Burundi acidic soil. Managing

Nutrient Cycles to Sustain Soil Fertility in Sub-Saharan Africa, 151.

Kadam, K. L., Wooley, R. J., Aden, A., Nguyen, Q. A., Yancey, M. A., Ferraro, F. M., 2000.

Softwood forest thinnings as a biomass source for ethanol production: A feasibility study

for California. Biotechnology Progress 16, 947–957.

Kambo, H. (2014). Upgrading lignocellulosic biomass via hydrothermal carbonization and

torrefaction (Masters Thesis; unpublished).

Kansha, Y., Tsuru, N., Fushimi, C., & Tsutsumi, A. (2009). A new design methodology of

azeotropic distillation processes based on self-heat recuperation. Chemical Engineering

Transactions, 18, 51–56.

Kaufman, A.S., Meier, P.J., Sinistore, J.C., & Reinemann, D.J., (2010). Applying life-cycle

assessment to low carbon fuel standards—How allocation choices influence carbon

intensity for renewable transportation fuels. Energy Policy, 38(9), 5229–5241.

Kazi, F.K., Fortman, J., Anex, R., Kothandaraman, G., Hsu, D., Aden, A. and Dutta, A. (2010b).

Techno-economic analysis of biochemical scenarios for production of cellulosic ethanol.

Report No. TP 6A2-46588, National Renewable Energy Laboratory. Golden Colorado

USA (2010).

Kazi, F.K., Fortman, J.A., Anex, R.P., Hsu, D.D., Aden, A., Dutta, A., & Kothandaraman, G.

(2010a). Techno-economic comparison of process technologies for biochemical ethanol

production from corn stover. Fuel, 89(), S20–S28.

Khanna, M., Dhungana, B., & Clifton-Brown, J. (2008). Costs of producing miscanthus and

switchgrass for bioenergy in Illinois. Biomass and Bioenergy, 32(6), 482–493.

148

Khullar, E., Dien, B.S., Rausch, K.D., Tumbleson, M.E., Singh, V. (2013). Effect of particle size

on enzymatic hydrolysis of pretreated Miscanthus. Industrial Crops and Products, 44(), 11–

17.

Kim, S., & Dale, B. E. (2003). Cumulative energy and global warming impact from the

production of biomass for biobased products. Journal of Industrial Ecology, 7(3‐4), 147–

162.

Kim, S., & Dale, B. E. (2005). Life cycle assessment of various cropping systems utilized for

producing biofuels: Bioethanol and biodiesel. Biomass and Bioenergy, 29(6), 426–439.

Klasson, K. T., Ackerson, M. D., Clausen, E. C., & Gaddy, J. L. (1993). Biological conversion of

coal and coal-derived synthesis gas. Fuel, 72(12), 1673–1678.

Klein-Marcuschamer, D., Oleskowicz-Popiel, P., Simmons, B.A., & Blanch, H.W., (2010).

Technoeconomic analysis of biofuels: A wiki-based platform for lignocellulosic

biorefineries. Biomass and Bioenergy, 34(12), 1914–1921.

Kludze, H., Deen, B., Weersink, A., van Acker, R., Janovicek, K., & De Laporte, A. (2013).

Impact of land classification on potential warm season grass biomass production in

Ontario, Canada. Canadian Journal of Plant Science, 93(2), 249–260.

Kludze, H., Deen, B., Weersink, A., van Acker, R., Janvicek, K. and De Laporte, A. (2010).

Assessment of the availability of agricultural biomass for heat and energy production in

Ontario. Final report submitted to the Ontario Ministry of Agriculture, Food and Rural

Affairs, Toronto, ON. 136 pp.

Koga, N., & Tajima, R., (2011). Assessing energy efficiencies and greenhouse gas emissions

under bioethanol-oriented paddy rice production in northern Japan. Journal of

Environmental Management, 92(3), 967–973.

Kou, N., & Zhao, F. (2011). Effect of multiple-feedstock strategy on the economic and

environmental performance of thermochemical ethanol production under extreme weather

conditions. Biomass and Bioenergy, 35(1), 608–616.

Krishna S. H., Prasanthi, K., Chowdary, G.V., & Ayyanna, C., (1998). Simultaneous

saccharification and fermentation of pre-treated sugarcane leaves to ethanol. Process

Biochemistry, 33(8), 825–830.

149

Kumar, D., & Murthy, G. S. (2011). Impact of pretreatment and downstream processing

technologies on economics and energy in cellulosic ethanol production. Biotechnology for

Biofuels, 4(1), 1–19.

Kundiyana, D. K., Wilkins, M. R., Maddipati, P., & Huhnke, R. L. (2011). Effect of temperature,

pH and buffer presence on ethanol production from synthesis gas by Clostridium ragsdalei.

Bioresource Technology, 102(10), 5794–5799.

Kundiyana, D.K., Huhnke, R.L., & Wilkins, M.R. (2010). Syngas fermentation in a 100-L pilot

scale fermentor: Design and process considerations. Journal of Bioscience and

Bioengineering, 109(5), 492–498.

Kunnakorn, D., Rirksomboon, T., Siemanond, K., Aungkavattana, P., Kuanchertchoo, N.,

Chuntanalerg, P., Hemra, K., Kulprathipanja, S., James, R. B., & Wongkasemjit, S. (2013).

Techno-economic comparison of energy usage between azeotropic distillation and hybrid

system for water–ethanol separation. Renewable Energy, 51, 310–316.

Kuo, P. C., Wu, W., & Chen, W. H. (2014). Gasification performances of raw and torrefied

biomass in a downdraft fixed bed gasifier using thermodynamic analysis. Fuel, 117, 1231–

1241.

Kusiima, J. M., & Powers, S. E. (2010). Monetary value of the environmental and health

externalities associated with production of ethanol from biomass feedstocks. Energy

Policy, 38(6), 2785–2796.

Lal, R. (2004). Soil Carbon Sequestration Impacts on Global Climate Change and Food Security.

Science, 304(5677), 1623–1627.

Lal, R. (2009). Soil quality impacts of residue removal for bioethanol production. Soil and tillage

research, 102(2), 233–241.

Lee, F. M., & Pahl, R. H. (1985). Solvent screening study and conceptual extractive distillation

process to produce anhydrous ethanol from fermentation broth. Industrial & Engineering

Chemistry Process Design and Development, 24(1), 168–172.

Lee, P. H. (2010). Syngas fermentation to ethanol using innovative hollow fiber membrane.

Graduate Theses and Dissertations, Paper 12377, Iowa State University, USA.

Leistritz, F.L., Senechal, D.M., Stowers, M.D., McDonald, W.F., Saffron, C.M., & Hodur, N.M.

(2006). Preliminary feasibility analysis for an integrated biomaterials and ethanol

150

biorefinery using wheat straw feedstock (No. 23500). North Dakota State University,

Department of Agribusiness and Applied Economics.

Li, H. Q., Li, C. L., Sang, T., & Xu, J. (2013). Pretreatment on Miscanthus lutarioriparious by

liquid hot water for efficient ethanol production. Biotechnology for Biofuels, 6(1), 1–10.

Liebig, M.A., Johnson, H.A., Hanson, J.D., & Frank, A.B. (2005). Soil carbon under switchgrass

stands and cultivated cropland. Biomass and Bioenergy, 28(4), 347–354.

Liew, F , K pke, , & Simpson, S D (20 3) Gas fermentation for commercial biofuels

production In “Li uid, Gaseous and Solid Biofuels - Conversion Techni ues”, Edited by

Zhen Fang, InTech, Rijeka, Croatia.

Liou, J. S. C., Balkwill, D. L., Drake, G. R., & Tanner, R. S. (2005). Clostridium

carboxidivorans sp. nov., a solvent-producing clostridium isolated from an agricultural

settling lagoon, and reclassification of the acetogen Clostridium scatologenes strain SL1 as

Clostridium drakei sp. nov. International journal of systematic and evolutionary

microbiology, 55(5), 2085–2091.

Liska, A.J., & Cassman, K.G. (2008). Towards standardization of life-cycle metrics for biofuels:

Greenhouse gas emissions mitigation and net energy yield. Journal of Biobased Materials

and Bioenergy, 2(3), 187–203.

Luo, L., van der Voet, E. & Huppes, G. (2009a). An energy analysis of ethanol from cellulosic

feedstock-corn stover. Renewable and Sustainable Energy Reviews, 13(8), 2003–2011.

Luo, L., van der Voet, E., & Huppes, G., (2010). Biorefining of lignocellulosic feedstock–

Technical, economic and environmental considerations. Bioresource Technology, 101(13),

5023–5032.

Luo, L., van der Voet, E., Huppes, G., & Udo de Haes, H.A., (2009b). Allocation issues in LCA

methodology: a case study of corn stover-based fuel ethanol. The International Journal of

Life Cycle Assessment, 14(6), 529–539.

Lynd, L. R., Laser, M. S., Bransby, D., Dale, B. E., Davison, B., Hamilton, R., Himmel, M.,

Keller, M., McMillan, J.D., Sheehan, J., & Wyman, C. E. (2008). How biotech can

transform biofuels. Nature Biotechnology, 26(2).

Lynd, L. R., Van Zyl, W. H., McBRIDE, J. E., & Laser, M. (2005). Consolidated bioprocessing

of cellulosic biomass: an update. Current Opinion in Biotechnology, 16(5), 577–583.

151

Lynd, L. R., Weimer, P. J., Van Zyl, W. H., & Pretorius, I. S. (2002). Microbial cellulose

utilization: fundamentals and biotechnology. Microbiology and molecular biology reviews,

66(3), 506–577.

Lynn, S., & Hanson, D. N. (1986). Multieffect extractive distillation for separating aqueous

azeotropes. Industrial & Engineering Chemistry Process Design and Development, 25(4),

936–941.

Maas, R. H. W., Bakker, R. R., Boersma, A. R., Bisschops, I., Pels, J. R., de Jong, E., Weusthuis,

R. A., & Reith, H., (2008). Pilot-scale conversion of lime-treated wheat straw into

bioethanol: quality assessment of bioethanol and valorization of side streams by anaerobic

digestion and combustion. Biotechnology for Biofuels, 1(1), 1–13.

Mabee, W.E., & Saddler, J.N., (2010). Bioethanol from lignocellulosics: Status and perspectives

in Canada. Bioresource Technology, 101(13), 4806–4813.

Mabee, W.E., Gregg, D.J., Arato, C., Berlin, A., Bura, R., Gilkes, N., Mirochnik, O., Pan, X.,

Pye, E.K., & Saddler, J.N., (2006). Updates on softwood-to-ethanol process development.

In Twenty-Seventh Symposium on Biotechnology for Fuels and Chemicals. Humana Press,

pp. 55–70.

andre, , Pärn, H , Klõšeiko, J , Ingerslev, M., Stupak, I., Kört, M., & Paasrand, K. (2010).

Use of biofuel ashes for fertilisation of Betula pendula seedlings on nutrient-poor peat soil.

Biomass and Bioenergy, 34(9), 1384–1392.

Mani, S., Sokhansanj, S., Bi, X., & Turhollow, A. (2006). Economics of producing fuel pellets

from biomass. Applied Engineering in Agriculture, 22(3), 421–426.

Mani, S., Sokhansanj, S., Tagore, S., & Turhollow, A.F., (2010). Techno-economic analysis of

using corn stover to supply heat and power to a corn ethanol plant–Part 2: Cost of heat and

power generation systems. Biomass and Bioenergy, 34(3), 356–364.

Manzanares, P., Ballesteros, I., Negro, M. J., Oliva, J. M., Gonzalez, A., & Ballesteros, M.

(2012). Biological conversion of forage sorghum biomass to ethanol by steam explosion

pretreatment and simultaneous hydrolysis and fermentation at high solid content. Biomass

Conversion and Biorefinery, 2(2), 123–132.

Mark, L., Jin, H., Jayawardhana, K., Dale, B. E., & Lynd, L. R., (2009). Projected mature

technology scenarios for conversion of cellulosic biomass to ethanol with coproduction

152

thermochemical fuels, power, and/or animal feed protein. Biofuels, Bioproducts and

Biorefining, 3(2), 231–246.

Mark, W.J. (2008). Development of a model compound-based reaction mechanism for lignin

thermochemical conversion. A PhD thesis, University of Colorado, MI, USA.

Martchamadol, J. (2007). Biofuels from energy crops and lignocellulosic materials in thailand:

technology and potential. A Thesis (unpublished), Asian Institute of Technology, Thailand.

Martín, M., & Grossmann, I. E. (2011). Energy optimization of bioethanol production via

gasification of switchgrass. AIChE Journal, 57(12), 3408–3428.

Mata, T. M., Martins, A. A., Caetano, N. S. 2010. Microalgae for biodiesel production and other

applications: A review. Renewable and Sustainable Energy Reviews, 14, 217–232.

McAloon, A., Taylor, F., Yee, W., Ibsen, K., & Wooley, R. (2000). Determining the cost of

producing ethanol from corn starch and lignocellulosic feedstocks. National Renewable

Energy Laboratory Report.

McMillan, J.D., Newman, M.M., Templeton, D.W., Mohagheghi, A. (1999). Simultaneous

saccharification and cofermentation of dilute-acid pretreated yellow poplar hardwood to

ethanol using xylose-fermenting Zymomonas mobilis. Applied Biochemistry and

Biotechnology, 79, 649–665.

Merino, S. T., & Cherry, J. (2007). Progress and challenges in enzyme development for biomass

utilization. In Biofuels (pp. 95-120). Springer Berlin Heidelberg.

Milan, 2010. Fossil fuel subsidies in Canada. <http://burycoal.com/blog/2010/11/05/fossil-fuel-

subsidies-in-canada/>.

Millman, J., 2008. Housing bust dries up sawdust market.

<http://www.championshavings.com/pdf/housing_bust_dries_up_saw_dust_market.pdf>.

MNR (Ministry of Natural Resources, Ontatio), 2011. The forest resources of Ontario 2011.

<http://www.mnr.gov.on.ca/stdprodconsume/groups/lr/@mnr/@forests/documents/docume

nt/stdprod_092922.pdf>.

Mohammadi, M., Younesi, H., Najafpour, G., & Mohamed, A. R. (2012). Sustainable ethanol

fermentation from synthesis gas by Clostridium ljungdahlii in a continuous stirred tank

bioreactor. Journal of Chemical Technology and Biotechnology, 87(6), 837–843.

153

Möller R., Pauly M., & Hake S. (2006). Cell wall saccharification, Outputs from the EPOBIO

project, CPLpress Science Publishers, CNAP, University of York, ISBN 10: 1-872691-04-

8, ISBN 13: 978-1- 872691-04-6, 67 p.

Møller, F., Slentø, E., & Frederiksen, P. (2014). Integrated well-to-wheel assessment of biofuels

combining energy and emission LCA and welfare economic cost benefit analysis. Biomass

and Bioenergy, 60, 41–49.

Monti, A., Fazio, S., & Venturi, G. (2009). Cradle-to-farm gate life cycle assessment in perennial

energy crops. European Journal of Agronomy, 31(2), 77–84.

Morrison, C.E. (2004). Production of ethanol from the fermentation of synthesis gas (Master

Thesis, Mississippi State University).

Mosier, N., Hendrickson, R., Ho, N., Sedlak, M., & Ladisch, M. R. (2005a). Optimization of pH

controlled liquid hot water pretreatment of corn stover. Bioresource technology, 96(18),

1986–1993.

Mosier, N., Wyman, C., Dale, B., Elander, R., Lee, Y. Y., Holtzapple, M., & Ladisch, M.

(2005b). Features of promising technologies for pretreatment of lignocellulosic biomass.

Bioresource technology, 96(6), 673–686.

Mu, D., Seager, T., Rao, P. S., & Zhao, F. (2010). Comparative life cycle assessment of

lignocellulosic ethanol production: biochemical versus thermochemical conversion.

Environmental Management, 46(4), 565–578.

Mueller, S.A., Anderson, J.E., & Wallington, T.J. (2011). Impact of biofuel production and other

supply and demand factors on food price increases in 2008. Biomass and Bioenergy, 35(5),

1623–1632.

Munasinghe, P. C., & Khanal, S. K. (2010). Biomass-derived syngas fermentation into biofuels:

opportunities and challenges. Bioresource Technology, 101(13), 5013–5022.

Murnen, H. K., Balan, V., Chundawat, S. P., Bals, B., Sousa, L. D. C., & Dale, B. E. (2007).

Optimization of ammonia fiber expansion (AFEX) pretreatment and enzymatic hydrolysis

of Miscanthus x giganteus to fermentable sugars. Biotechnology progress, 23(4), 846–850.

Murse, T. (2011). Understanding the ethanol subsidy. <http://usgovinfo.about.com/od/money-

matters/a/The-Federal-Ethanol-Subsidy.htm>.

154

Naik, T. R., Kraus, R. N., Ramme, B. W., Chun, Y. M., & Kumar, R. (2006). High-carbon fly

ash in manufacturing conductive CLSM and concrete. Journal of materials in civil

engineering, 18(6), 743–746.

Nakaiwa, M., & Ohmori, T. (2009). Innovation in distillation processes: Process intensification

for energy savings through concept of “detuning”from the ideal state Synthesiology–

English edition, 2(1), 55–63.

Natural Resource Canada (NRC), 2013. Energy use data handbook tables (Canada) -

transportation sector. <http://data.gc.ca/data/dataset/cec843a2-b8cb-4168-81f1-

e1ced9daf5da> (accessed 28/09/2013).

Nechodom, M., Schuetzle, D., Ganz, D., & Cooper, A.J. (2008). Sustainable forests, renewable

energy, and the environment. Environmental Science & Technology, 42(1), 13–18.

Nelson, R. G. (2002). Resource assessment and removal analysis for corn stover and wheat straw

in the Eastern and Midwestern United States—rainfall and wind-induced soil erosion

methodology. Biomass and Bioenergy, 22(5), 349–363.

Nielsen, D. R., & Prather, K. J. (2009). Adsorption of second generation biofuels using polymer

resins with in situ product recovery (ISPR) applications. In AIChE Annual Meeting,

Nashville, TN (No. 564f).

Nigam, P. S., & Singh, A. (2011). Production of liquid biofuels from renewable resources.

Progress in Energy and Combustion Science, 37(1), 52–68.

Nilsson, H., 2008. Flexibility in wheat bioethanol plants - Conditions for conversion to

lignocellulosic feedstock,< http://www.chemeng.lth.se/exjobb/E471.pdf>.

OECD-FAO, (2013). Database-OECD-FAO Agricultural Outlook 2013-2022.

<http://www.oecd.org/site/oecd-faoagriculturaloutlook/database-oecd-

faoagriculturaloutlook.htm>.

Ofoefule, A. U., Okoro, U. C., Onukwuli, O. D., & Anyanwu, C. N. (2013). Biogas production

from waste by-products of ethanol production: 3 modeling of gas yield with time. Physical

Review and Research International, 3(1), 18–27.

Olar, M., Romain R., Bergeron N., & Klein, K., 2004. Ethanol Industry in Canada. Research

Series, SR. 04.08, Centre for Research in the Economics of Agrifood, Laval University,

Quebec, Canada.

155

Olofsson, K., Bertilsson, M., & Lidén, G. (2008). A short review on SSF-an interesting process

option for ethanol production from lignocellulosic feedstocks. Biotechnol Biofuels, 1(7),

1–14.

Olson, D. G., McBride, J. E., Joe Shaw, A., & Lynd, L. R. (2012). Recent progress in

consolidated bioprocessing. Current Opinion in Biotechnology, 23(3), 396–405.

Olsson, L., & Hahn-Hägerdal, B. (1996). Fermentation of lignocellulosic hydrolysates for

ethanol production. Enzyme and Microbial Technology, 18(5), 312–331.

Olujic, Z., Fakhri, F., De Rijke, A., De Graauw, J., & Jansens, P. J. (2003). Internal heat

integration–the key to an energy‐conserving distillation column. Journal of Chemical

Technology and Biotechnology, 78(2‐3), 241–248.

OMAFRA (2011b). Hard red spring wheat enterprise budget. URL:

http://www.omafra.gov.on.ca/english/busdev/bear2000/Budgets/Crops/Grains/hrswheat_st

atic.htm (accessed on 11/11/2011).

OMAFRA (Ontario Ministry of Agriculture, Food and Rural Affairs), (2009). Windrow

composting of poultry carcasses. <http://www.omafra.gov.on.ca/english/engineer/facts/09-

017.htm>(accessed on 14/2/2013).

OMAFRA (The Ontario Ministry of Agriculture, Food and Rural Affairs) (2011a). Field crops.

[online database], URL: http://www.omafra.gov.on.ca/english/stats/crops/ (accessed on

18/11/2011).

OMF (Ontario Ministry of Finance), (2013). Ontario fact sheet October 2013, Economic &

Revenue Forecasting & Analysis Branch, Office of Economic Policy, Ontario Ministry of

Finance, Ontario, Canada.

OPA (Ontario Power Authority), (2013). Feed-in Tariff. <http://www.powerauthority.on.ca/>.

Orikasa, T., Tokuyasu, K., Inoue, T., Kojima, K., Roy, P., Nakamura, N., & Shiina, T. (2009).

Effect of ethanol conversion efficiency on cost, CO2 emission and energy balance in the

bioethanol production system from rice straw. Journal of the Japanese Society of

Agricultural Machinery, 71(5), 45–53.

Park, J., Shiroma, R., Al-Haq, M.I., Zhang, Y., Ike, M., Arai-Sanoh, Y., Ida, A., Kondo, M.,

Tokuyasu, K., 2010. A novel lime pretreatment for subsequent bioethanol production from

rice straw – Calcium capturing by carbonation (CaCCO) process. Bioresource Technology,

101(17), 6805–6811.

156

Passandideh-Fard, M., Vaezi, M., & Moghiman, M. (2008). On a numerical modle for

gasification of biomass materials. The 7th High Temperature Air Combustion and

Gasification International Symposium, 13-16 January 2008, Phuket, Thailand.

Pedersen, J. F., Vogel, K. P., & Funnell, D. L. (2005). Impact of reduced lignin on plant fitness.

Crop Science, 45(3), 812–819.

Perales, A.L.V., Valle, C.R., Ollero, P., & Gómez-Barea, A. (2011). Technoeconomic

assessment of ethanol production via thermochemical conversion of biomass by entrained

flow gasification. Energy, 36(7), 4097–4108.

Pereira, E.G., da Silva, J.N., Jofran L. de Oliveira, J.L., Cássio S., & Machado, C.S. (2012).

Sustainable energy: A review of gasification technologies. Renewable and Sustainable

Energy Reviews, 16(7), 4753–4762.

Pérez-Villarejo, L., Eliche-Quesada, D., Iglesias-Godino, F. J., Martínez-García, C., & Corpas-

Iglesias, F. A. (2012). Recycling of ash from biomass incinerator in clay matrix to produce

ceramic bricks. Journal of Environmental Management, 95, S349–S354.

Phillips, J.R., Clausen, E.C., & Gaddy, J.L., (1994). Synthesis gas as substrate for the biological

production of fuels and chemicals. Applied Biochemistry and Biotechnology, 45(1), 145–

157.

Phillips, S. D. (2007). Technoeconomic analysis of a lignocellulosic biomass indirect

gasification process to make ethanol via mixed alcohols synthesis. Industrial &

Engineering Chemistry Research, 46(26), 8887–8897.

Phillips, S., Aden, A., Jechura, J., & Dayton, D. (2007). Thermochemical ethanol via indirect

gasification and mixed alcohol synthesis of lignocellulosic biomass. National Renewable

Energy Laboratory, Technical Report, NREL/TP-510-41168, April 2007.

Piccolo, C., & Bezzo, F. (2009). A techno-economic comparison between two technologies for

bioethanol production from lignocellulose. Biomass and bioenergy, 33(3), 478–491.

Pickett, J., Anderson, D., Bowles, D., Bridgwater, T., Jarvis, P., Mortimer, N., Poliakoff, M., &

Woods, J. (2008). Sustainable biofuels: prospects and challenges. The Royal Society,

London, UK.

Piirainen, S., Domisch, T., Nieminen, M., & Moilanen, M. (2005). Nutrient and heavy metal

leaching after wood ash fertilization from drained peatland. In Regular Recycling of Wood

157

Ash to Prevent Waste Production. RecAsh—A Life-environment Demonstration Project.

RecAsh International Seminar, 8th–9th (10th) of November (pp. 47–55).

Pimentel, D., & Patzek, T. W. (2005). Ethanol production using corn, switchgrass, and wood;

biodiesel production using soybean and sunflower. Natural resources research, 14(1), 65–

76.

Pinto, R. T. P., Wolf-Maciel, M. R., & Lintomen, L. (2000). Saline extractive distillation process

for ethanol purification. Computers & Chemical Engineering, 24(2), 1689–1694.

Pourhashem, G., Adler, P. R., McAloon, A. J., & Spatari, S. (2013). Cost and greenhouse gas

emission tradeoffs of alternative uses of lignin for second generation ethanol.

Environmental Research Letters, 8(2), 025021.

Prasad, S., Singh, A., & Joshi, H. C. (2007). Ethanol as an alternative fuel from agricultural,

industrial and urban residues. Resources, Conservation and Recycling, 50(1), 1–39.

Rabelo, S. C., Maciel Filho, R., & Costa, A. C. (2009). Lime pretreatment of sugarcane bagasse

for bioethanol production. Applied Biochemistry and Biotechnology, 153(1-3), 139–150.

Rajagopalan, S., P Datar, R., & Lewis, R. S. (2002). Formation of ethanol from carbon monoxide

via a new microbial catalyst. Biomass and Bioenergy, 23(6), 487–493.

Rao, M. S., Singh, S. P., Sodha, M. S., Dubey, A. K., & Shyam, M. (2004). Stoichiometric,

mass, energy and exergy balance analysis of countercurrent fixed-bed gasification of post-

consumer residues. Biomass and Bioenergy, 27(2), 155–171.

Reijnders, L. (2008). Ethanol production from crop residues and soil organic carbon. Resources,

Conservation and Recycling, 52(4), 653–658.

Reith, J.H., Den Uil, H., Van Veen, H., De Laat, W.T.A.M., Niessen, J.J., De Jong, E., Elbersen,

H.W., Weusthuis, R., van Dijken, J.P., & Raamsdonk, L. (2002). Co-production of bio-

ethanol, electricity and heat from biomass residues. In Proceedings of the 12th European

Conference on Biomass for Energy, Industry and Climate Protection (pp. 17-21),

Amsterdam, The Netherlands, June 17–21.

Richter, H., Martin, M. E., & Angenent, L. T. (2013). A two-stage continuous fermentation

system for conversion of syngas into ethanol. Energies, 6(8), 3987–4000.

Rivera, E. C., Costa, A. C., Atala, D. I., Maugeri, F., & Maciel, M. R. W. (2006). Evaluation of

optimization techniques for parameter estimation: Application to ethanol fermentation

considering the effect of temperature. Process Biochemistry, 41(7), 1682–1687.

158

Roberts, K. G., Gloy, B. A., Joseph, S., Scott, N. R., & Lehmann, J. (2009). Life cycle

assessment of biochar systems: Estimating the energetic, economic and climate change

potential. Environmental Science & Technology, 44(2), 827–833.

Roy, P., & Dutta, A. (2013). A review of life cycle of ethanol produced from biosyngas.

Bioethanol, 1, 9–19.

Roy, P., & Dutta, A., (2012). Life cycle assessment of ethanol produced from wheat straw.

Journal of Biobased Materials and Bioenergy, 6(3), 276–282.

Roy, P., & Dutta, A., (2013). Life cycle assessment of ethanol derived from sawdust.

Bioresource Technology, 150(), 407–411.

Roy, P., & Shiina, T., (2010). Global Environment, Biofuel: Sustainable Food Production and

Distribution In “Global Environmental Policies: Impact, anagement and Effects”, Edited

by Riccardo Cancilla and Monte Gargano, Nova Science Publishers, pp 29–58.

Roy, P., Orikasa, T., Nakamura, N., & Shiina, T., (2013). Environmental Sustainability in Food

Processing In “Sustainable Food Processing”, Edited by Brijesh K Tiwari; Tomas Norton

and Nicholas M. Holden, Wiley-Blackwell, pp 39–62.

Roy, P., Orikasa, T., Tokuyasu, K., Nakamura, N., & Shiina, T., (2012a). Evaluation of the life

cycle of bioethanol produced from rice straws. Bioresource Technology, 110(April), 239–

244.

Roy, P., Shimizu, N., Okadome, H., Shiina, T., & Kimura, T. (2007). Life cycle of rice:

Challenges and choices for Bangladesh. Journal of Food Engineering, 79(4), 1250–1255.

Roy, P., Tokuyasu, K., Orikasa, T., Nakamura, N., & Shiina, T. (2012c). A review of life cycle

assessment (LCA) of bioethanol from lignocellulosic biomass. Japan Agricultural Research

Quaterly, 46(1), 41–57.

Roy, P., Tokuyasu, K., Orikasa, T., Nakamura, N., & Shiina, T. (2012b). A techno-economic and

environmental evaluation of the life cycle of bioethanol produced from rice straw by RT-

CaCCO process. Biomass and Bioenergy, 37(2), 188–195.

Ruiz, H.A., Ruzene, D.S., Silva, D.P., da Silva, F.F.M., Vicente, A.A., & Teixeira, J.A. (2011).

Development and characterization of an environmentally friendly process sequence

(autohydrolysis and organosolv) for wheat straw delignification. Applied Biochemistry and

Biotechnology, 164(5), 629–641.

159

Ruth, M. (2005). Gridley ethanol demonstration project utilizing biomass gasification

technology: Pilot plant gasifier and syngas conversion testing. NREL, USA.

S&T2, (2004). Economic, financial, social analysis and public policies for fuel ethanol. S&T2

Consultants Inc., Canada.

Saga, K., Imou, K., Yokoyama, S., & Minowa, T. (2010). Net energy analysis of bioethanol

production system from high-yield rice plant in Japan. Applied Energy, 87(7), 2164–2168.

Saha, B. C., & Cotta, M. A. (2008). Lime pretreatment, enzymatic saccharification and

fermentation of rice hulls to ethanol. Biomass and Bioenergy, 32(10), 971–977.

Salas, A., Delvasto, S., de Gutierrez, R. M., & Lange, D. (2009). Comparison of two processes

for treating rice husk ash for use in high performance concrete. Cement and Concrete

Research, 39(9), 773–778.

Sánchez, O´. J. & Cardona, C. A. (2008). Trends in biotechnological production of fuel ethanol

from different feedstocks. Bioresource Technology, 99(13), 5270–5295.

Sandilands, J., Kellenberger, D., Nicholas, I., & Nielsen, P. (2009). Life cycle assessment of

wood pellets and bioethanol from wood residues and willow. NZJ Forest Science, 53(4),

25–33.

Sanscartier, D., Deen, B., Dias, G., MacLean, H. L., Dadfar, H., McDonald, I., & Kludze, H.

(2013). Implications of land class and environmental factors on life cycle GHG emissions

of Miscanthus as a bioenergy feedstock. GCB Bioenergy.

Sanscartier, D., MacLean, H. L., & Saville, B. (2012). Electricity production from anaerobic

digestion of household organic waste in Ontario: Techno-economic and GHG emission

analyses. Environmental Science & Technology, 46(2), 1233–1242.

Sassner, P., & Zacchi, G. (2008). Integration options for high energy efficiency and improved

economics in a wood-to-ethanol process. Biotechnology for Biofuels, 1(1), 4.

Sassner, P., Galbe, M., & Zacchi, G. (2008). Techno-economic evaluation of bioethanol

production from three different lignocellulosic materials. Biomass and Bioenergy, 32(5),

422–430.

Schmer, M. R., Vogel, K. P., Mitchell, R. B., & Perrin, R. K. (2008). Net energy of cellulosic

ethanol from switchgrass. Proceedings of the National Academy of Sciences, 105(2), 464–

469.

160

Schneckenberger, K., & Kuzyakov, Y. (2007). Carbon sequestration under Miscanthus in sandy

and loamy soils estimated by natural C-13 abundance. Journal of Plant Nutrition and Soil

Science – Zeitschrift Fur Pflanzenernahrung Und Bodenkunde, 170, 538–542.

Schubert, C. (2006). Can biofuels finally take center stage?. Nature Biotechnology, 24(7), 777–

784.

ScienceDaily, 2005. Genetically modified rice in china benefits farmers' health, study finds.

<http://www.sciencedaily.com/releases/2005/04/050428181133.htm>.

Scown, C.D., Nazaroff, W. W., Mishra, U., Strogen, B., Lobscheid, A.B., Masanet, E., Santero,

N.J., Horvath, A. & McKone, T. E. (2012). Lifecycle greenhouse gas implications of US

national scenarios for cellulosic ethanol production. Environmental Research Letters, 7(1),

014011.

Seabra, J. E., Tao, L., Chum, H. L., & Macedo, I. C. (2010). A techno-economic evaluation of

the effects of centralized cellulosic ethanol and co-products refinery options with

sugarcane mill clustering. Biomass and Bioenergy, 34(8), 1065–1078.

Searchinger, T., Heimlich, R., Houghton, R. A., Dong, F., Elobeid, A., Fabiosa, J., Tokgoz, S.,

Hayes, D., & Yu, T.H. (2008). Use of US croplands for biofuels increases greenhouse

gases through emissions from land-use change. Science, 319(5867), 1238–1240.

Searcy, E. & Flynn, P. C., (2010). A criterion for selecting renewable energy processes. Biomass

and Bioenergy, 34(5), 798–804.

Service, R. F., (2010). Scaling up alternative energy: Is there a road ahead for cellulosic ethanol?

Science, 329, 784–785.

Shafiei, M., Karimi, K., & Taherzadeh, M. J. (2011). Techno-economical study of ethanol and

biogas from spruce wood by NMMO-pretreatment and rapid fermentation and digestion.

Bioresource Technology, 102(17), 7879–7886.

Shankar Tumuluru, J., Sokhansanj, S., Hess, J. R., Wright, C. T., & Boardman, R. D. (2011).

REVIEW: A review on biomass torrefaction process and product properties for energy

applications. Industrial Biotechnology, 7(5), 384-401.

Sheehan, J., Aden, A., Paustian, K., Killian, K., Brenner, J., Walsh, M., & Nelson, R. (2003).

Energy and environmental aspects of using corn stover for fuel ethanol. Journal of

Industrial Ecology, 7(3‐4), 117–146.

161

Sheehan, J., Aden, A., Riley, C., Paustian, K., Brenner, J., Lightle, D., Nelson, R., Walsh, M., &

Cushman, J. (2002). Is ethanol from corn stover sustainable? Adventures in cyber-farming:

a life cycle assessment of the production of ethanol from corn stover for use in a flex fuel

vehicle. Draft report for peer review, Colorado, NREL.

Sherrington, C., & Moran, D. (2010). Modelling farmer uptake of perennial energy crops in the

UK. Energy Policy, 38(7), 3567–3578.

Shiroma, R., Park, J. Y., Al-Haq, M. I., Arakane, M., Ike, M., & Tokuyasu, K. (2011). RT-

CaCCO process: an improved CaCCO process for rice straw by its incorporation with a

step of lime pretreatment at room temperature. Bioresource technology, 102(3), 2943–

2949.

Sim, J. H., Kamaruddin, A. H., Long, W. S., & Najafpour, G. (2007). Clostridium aceticum—A

potential organism in catalyzing carbon monoxide to acetic acid: Application of response

surface methodology. Enzyme and Microbial Technology, 40(5), 1234-1243.

Simell, P., Kurkela, E., Ståhlberg, P., & Hepola, J. (1996). Catalytic hot gas cleaning of

gasification gas. Catalysis Today, 27(1), 55–62.

Sims, R. E., Hastings, A., Schlamadinger, B., Taylor, G., & Smith, P. (2006). Energy crops:

current status and future prospects. Global Change Biology, 12(11), 2054–2076.

Singh, A., Pant, D., Korres, N. E., Nizami, A. S., Prasad, S., & Murphy, J. D. (2010). Key issues

in life cycle assessment of ethanol production from lignocellulosic biomass: challenges and

perspectives. Bioresource Technology, 101(13), 5003–5012.

Singh, S., & Kumar, A. (2011). Development of water requirement factors for biomass

conversion pathway. Bioresource Technology, 102(2), 1316–1328.

Skinner, R.H., Zegada-Lizarazu, W. & Schmidt, J.P. (2006). Environmental impacts of

switchgrass management for bioenergy production, Switch Grass: A valuable biomass crop

for energy. Edited by Andrea Monti. Springer, London, UK.

Slade, R., Bauen, A., & Shah, N. (2009). The commercial performance of cellulosic ethanol

supply-chains in Europe. Biotechnol Biofuels, 2(3), 1–20.

Smith, J.A., (1986). Distribution of crop residue, a requirement for conservation tillage. G86-

782-A, IANR, U of NE, Lincoln, NE.

Sokhansanj, S., & Turhollow, A. F. (2002). Baseline cost for corn stover collection. Applied

Engineering in Agriculture, 18(5), 525–532.

162

Somerville, C., Youngs, H., Taylor, C., Davis, S. C., & Long, S. P. (2010). Feedstocks for

lignocellulosic biofuels. Science, 329(5993), 790–792.

Sommer, S., Klinkhammer, B., & Melin, T. (2002). Integrated system design for dewatering of

solvents with microporous silica membranes. Desalination, 149(1), 15–21.

Sørensen, A., Teller, P.J., Hilstrøm, T., & Ahring, B. K. (2008). Hydrolysis of miscanthus for

bioethanol production using dilute acid presoaking combined with wet explosion pre-

treatment and enzymatic treatment. Bioresource Technology, 99(14), 6602–6607.

Spatari, S., Zhang, Y., & MacLean, H. L. (2005). Life cycle assessment of switchgrass-and corn

stover-derived ethanol-fueled automobiles. Environmental Science & Technology, 39(24),

9750–9758.

Spath, P. L., & Dayton, D. C. (2003). Preliminary Screening — Technical and economic

assessment of synthesis gas to fuels and chemicals with emphasis on the potential for

biomass-derived syngas. National Renewable Energy Laboratory, Golden, Colorado, USA,

NREL/TP-510-34929.

Spivey, J. J., & Egbebi, A. (2007). Heterogeneous catalytic synthesis of ethanol from biomass-

derived syngas. Chemical Society Reviews, 36(9), 1514–1528.

Stanbury, P.F. (2000). Fermentation technology. Extraction, 2, 1–1.

Statistics Canada. (2008). Total farm area, land tenure and land crops by province (Census of

Agriculture, 1986–2006) (Ontario). <http://www40.statcan.gc.ca/l01/cst01/agrc25g-

eng.htm>.

Stephenson, A. L., Dupree, P., Scott, S. A., & Dennis, J. S. (2010). The environmental and

economic sustainability of potential bioethanol from willow in the UK. Bioresource

Technology, 101(24), 9612–9623.

Stevens, J. D., (2001). Hot gas conditioning: recent progress with larger-scale biomass

gasification systems. NREL/SR-510-29952, National Renewable Energy Laboratory,

Golden, Colorado, USA.

Stoeglehner, G., & Narodoslawsky, M. (2009). How sustainable are biofuels? Answers and

further questions arising from an ecological footprint perspective. Bioresource Technology,

100(16), 3825–3830.

Stolman, M. (2005). Ethanol and the net energy debate.

<www.e2.org/ext/jsp/controller?docId=8461>.

163

Subramani, V., & Gangwal, S. K. (2008). A review of recent literature to search for an efficient

catalytic process for the conversion of syngas to ethanol. Energy & Fuels, 22(2), 814–839.

Sukumaran, R. K., Surender, V. J., Sindhu, R., Binod, P., Janu, K. U., Sajna, K. V., Rajasree,

K.P., & Pandey, A. (2010). Lignocellulosic ethanol in India: prospects, challenges and

feedstock availability. Bioresource technology, 101(13), 4826–4833.

Sule, I. (2012). Torrefaction behaviour of agricultural biomass (Master thesis). University of

Guelph, Guelph, Ontario, Canada.

Sun, Y., & Cheng, J. (2002). Hydrolysis of lignocellulosic materials for ethanol production: a

review. Bioresource technology, 83(1), 1–11.

Swana, J., Yang, Y., Behnam, M., & Thompson, R. (2011). An analysis of net energy production

and feedstock availability for biobutanol and bioethanol. Bioresource Technology, 102(2),

2112–2117.

Taherzadeh, M. J., & Karimi, K. (2008). Pretreatment of lignocellulosic wastes to improve

ethanol and biogas production: a review. International Journal of Molecular Sciences, 9(9),

1621–1651.

Talebnia, F., Karakashev, D., & Angelidaki, I. (2010). Production of bioethanol from wheat

straw: An overview on pretreatment, hydrolysis and fermentation. Bioresource

Technology, 101(13), 4744–4753.

Tan, E. C., & Dutta, A. (2013, June). Sustainability metrics and life cycle assessment for

thermochemical conversion of woody biomass to mixed alcohols. In L. Juan (Ed.), 2nd

International Conference on Environment, Energy and Biotechnology, Kuala Lampur,

Malaysia, June 8 (pp. 69-74). IACSIT Press: Kuala Lampur, Malaysia.

Tan, K. T., Lee, K. T., & Mohamed, A. R. (2008). Role of energy policy in renewable energy

accomplishment: The case of second-generation bioethanol. Energy policy, 36(9), 3360–

3365.

Thompson, D. N., Shaw, P. G., & Lacey, J. A. (2003). Post-harvest processing methods for

reduction of silica and alkali metals in wheat straw. Applied biochemistry and

biotechnology, 105(1-3), 205–218.

TianYuan, (2013). Tianyuan Machinery Co., Ltd. Zhengzhou environmental protection

machinery. <http://en.tybroad.com/lxwm.html>.

164

Tilman, D., Hill, J., & Lehman, C. (2006). Carbon-negative biofuels from low-input high-

diversity grassland biomass. Science, 314(5805), 1598–1600.

Tilman, D., Socolow, R., Foley, J. A., Hill, J., Larson, E., Lynd, L., Pacala, S., & Williams, R.

(2009). Beneficial biofuels—the food, energy, and environment trilemma. Science,

325(5938), 270–271.

Tock, L., & Maréchal, F. (2012). Co-production of hydrogen and electricity from lignocellulosic

biomass: Process design and thermo-economic optimization. Energy, 45(1), 339–349.

Tock, L., Gassner, M., & Maréchal, F. (2010). Thermochemical production liquid fuels from

biomass: Thermo-economic modeling, process design and process integration analysis.

Biomass and Bioenergy, 34(12), 1838–1854.

Tonini, D., & Astrup, T. (2012). LCA of biomass-based energy systems: A case study for

Denmark. Applied Energy, 99(), 234–246.

Trippe, F., Fröhling, M., Schultmann, F., Stahl, R., & Henrich, E. (2011). Techno-economic

assessment of gasification as a process step within biomass-to-liquid (BtL) fuel and

chemicals production. Fuel Processing Technology, 92(11), 2169–2184.

Tunå, P., & Hulteberg, C. (2014). Woody biomass-based transportation fuels–A comparative

techno-economic study. Fuel, 117, 1020–1026.

Tyndall, J. C., Berg, E. J., & Colletti, J. P. (2011). Corn stover as a biofuel feedstock in Iowa’s

bio-economy: An Iowa farmer survey. Biomass and bioenergy, 35(4), 1485–1495.

Uihlein, A., & Schebek, L. (2009). Environmental impacts of a lignocellulose feedstock

biorefinery system: an assessment. Biomass and Bioenergy, 33(5), 793–802.

Ulrich, S., & Pavel, S., (1988). Design and operation of a pervaporation plant for ethanol

dehydration. Journal of Membrane Science, 36, 463–475.

UN Energy (2007). Sustainable bioenergy: A framework for decision makers.

http://esa.un.org/un-energy/pdf/susdev. Biofuels.FAO.pdf.

United Nations Energy Programme (UNEP), (2007). The marrakech process: SCP, industrial

development and energy. <

http://esa.un.org/marrakechprocess/pdf/Issues_Industry_and_Energy.pdf>.

Vadas, P. A., Barnett, K. H., & Undersander, D. J. (2008). Economics and energy of ethanol

production from alfalfa, corn, and switchgrass in the Upper Midwest, USA. BioEnergy

Research, 1(1), 44–55.

165

Van Hoof, V., Van den Abeele, L., Buekenhoudt, A., Dotremont, C., & Leysen, R. (2004).

Economic comparison between azeotropic distillation and different hybrid systems

combining distillation with pervaporation for the dehydration of isopropanol. Separation

and Purification Technology, 37(1), 33–49.

van Kasteren, J. M. N., Dizdarevic, D., van der Waall, W. R., Guo, J., & Verberne, R. (2005).

Bio-ethanol from Syngas. Eindhoven University of Technology (TU/e), I. &. E, Ed.

Vanderghem C, Brostaux Y, Jacquet N, Blecker C, Paquot M: Optimization of formic/acetic acid

delignification of miscanthus × giganteus for enzymatic hydrolysis using response surface

methodology. Industrial Crops and Products, 35(1), 280–286.

Vane, L. M., & Alvarez, F. R. (2008). Membrane‐assisted vapor stripping: energy efficient

hybrid distillation–vapor permeation process for alcohol–water separation. Journal of

Chemical Technology and Biotechnology, 83(9), 1275-1287.

Vega, J. L., Clausen, E. C., Gaddy, J. L., 1990. Design of bioreactors for coal synthesis gas

fermentation. Resources Conservation and Recycling 3 (2–3), 149–160.

Velásquez-Arredondo, H. I., Ruiz-Colorado, A. A., & De Oliveira, S. (2010). Ethanol production

process from banana fruit and its lignocellulosic residues: Energy analysis. Energy, 35(7),

3081–3087.

Vliet, V.O.P.R., Faaij, A.P.C. & Turkenburg, W.C., (2009). Fischer-Tropsch diesel production in

a well-to-wheel perspective: a carbon, energy flow and cost analysis. Energy Conversion

and Management 50(4), 855–876.

von Blottnitz, H., & Curran, M.A. (2007). A review of assessments conducted on bio-ethanol as

a transportation fuel from a net energy, greenhouse gas, and environmental life cycle

perspective. Journal of Cleaner Production, 15(7), 607–619.

Vyn, R. J., Virani, T., & Deen, B. (2012). Examining the economic feasibility of miscanthus in

Ontario: An application to the greenhouse industry. Energy Policy, 50(), 669–676.

Wang, L., Templer, R., Murphy, R. J., (2012). A life cycle assessment (LCA) comparison of

three management options for waste papers: Bioethanol production, recycling and

incineration with energy recovery. Bioresouce Technology, 120, 89–98.

Wang, Y., & Yan, L. (2008). CFD studies on biomass thermochemical conversion. International

journal of molecular sciences, 9(6), 1108–1130.

166

Weber, C., Farwick, A., Benisch, F., Brat, D., Dietz, H., Subtil, T., & Boles, E. (2010). Trends

and challenges in the microbial production of lignocellulosic bioalcohol fuels. Applied

microbiology and biotechnology, 87(4), 1303–1315.

Wei, L, Thomasson J. A., Bricka, R. M., Batchelor, W. D., Columbus, E. P., Wooten, J. (2006).

Experimental study of a downdraft gasifier. Paper No. 066029. St. Joseph, MI: ASAE.

Wei, L., Pordesimo, L. O., Igathinathane, C., & Batchelor, W. D. (2009). Process engineering

evaluation of ethanol production from wood through bioprocessing and chemical catalysis.

Biomass and bioenergy, 33(2), 255–266.

Wilkins, M., (2011). Lignocellulose pretreatment technologies. SunGrantBioWeb,

<http://bioweb.sungrant.org/>.

Williams, E. D., Weber, C. L. and Hawkins, T. R., (2009b). Hybrid framework for managing

uncertainty in life cycle inventories. Journal of Industrial Ecology, 13(6), 928–944.

Williams, P.R., Inman, D., Aden, A., & Heath, G.A. (2009a). Environmental and sustainability

factors associated with next-generation biofuels in the US: what do we really know?.

Environmental Science & Technology, 43(13), 4763–4775.

Wingren, A., Galbe, M., & Zacchi, G. (2008). Techno‐economic evaluation of producing ethanol

from softwood: Comparison of SSF and SHF and identification of bottlenecks.

Biotechnology Progress, 19(4), 1109–1117.

Wooley, R., Ruth, M., Glassner, D., & Sheehan, J. (1999b). Process design and costing of

bioethanol technology: a tool for determining the status and direction of research and

development. Biotechnology Progress, 15(5), 794–803.

Wooley, R., Ruth, M., Sheehan, J., Ibsen, K., Majdeski, H., & Galvez, A. (1999a).

Lignocellulosic biomass to ethanol—process design and economics utilizing co-current

dilute acid prehydrolysis and enzymatic hydrolysis—current and futuristic scenarios,

Report No. TP-580-26157. National Reneawable Energy Laboratory. Golden Colorado,

Colorado USA.

Worden, R.M., Bredwell, M.D., & Grethlein, A.J., (1997). Fuels and Chemicals from Biomass,

ACS Symposium Series No. 666, Washington, DC: pp. 320–336.

Worden, R.M., Grethlein, A.J., Jain, M.K., Datta, R., (1991). Production of butanol and ethanol

from synthesis gas via fermentation. Fuel, 70(5), 615–619.

167

Wright, J. D., Wyman, C. E., & Grohmann, K. (1988). Simultaneous saccharification and

fermentation of lignocellulose. Applied Biochemistry and Biotechnology, 18(1), 75-90.

Wu M, Wang M, & Huo H, (2006). Fuel-cycle assessment of selected bioethanol production

pathways in the United States. Center for Transportation Research, Energy Systems

Division, Argonne National Laboratory, USA.

Wu, M., Wu, Y. & Wang, M. (2008) Mobility chain analysis of technologies for passenger cars

and light-duty vehicles fueled with biofuels: Application of the GREET Model to the role

of biomass in merica’s Energy Future (RB EF) Project Center for Transportation

Research, Energy Systems Division, Argonne National Laboratory,

<http://www.osti.gov/bridge>.

Wyman, C. E. (1994). Ethanol from lignocellulosic biomass: technology, economics, and

opportunities. Bioresource Technology, 50(1), 3–15.

Wyman, C. E. (2007). What is (and is not) vital to advancing cellulosic ethanol. TRENDS in

Biotechnology, 25(4), 153–157.

Xu, J., Chen, H., Kádár, Z., Thomsen, A. B., Schmidt, J. E., & Peng, H. (2011). Optimization of

microwave pretreatment on wheat straw for ethanol production. Biomass and Bioenergy,

35(9), 3859–3864.

Yamamoto, H., Fujino, J., & Yamaji, K. (2001). Evaluation of bioenergy potential with a multi-

regional global-land-use-and-energy model. Biomass and Bioenergy, 21(3), 185–203.

Yang, B., & Wyman, C. E. (2008). Pretreatment: the key to unlocking low‐cost cellulosic

ethanol. Biofuels, Bioproducts and Biorefining, 2(1), 26–40.

Yang, J., Huang, J., Qiu, H., Rozelle, S., & Sombilla, M.A. (2009). Biofuels and the Greater

Mekong Subregion: Assessing the impact on prices, production and trade. Applied Energy,

86, S37–S46.

Yoon, H. H., Wu, Z. W., & Lee, Y. Y. (1995). Ammonia-recycled percolation process for

pretreatment of biomass feedstock. Applied Biochemistry and Biotechnology, 51(1), 5–19.

Yoshida, M., Liu, Y., Uchida, S., Kawarada, K., Ukagami, Y., Ichinose, H. Kaneko, S. &

Fukuda, K. (2008). Effects of cellulose crystallinity, hemicellulose, and lignin on the

enzymatic hydrolysis of Miscanthus sinensis to monosaccharides. Bioscience,

biotechnology, and biochemistry, 72(3), 805–810.

168

Younesi, H., Najafpour, G., & Mohamed, A. R. (2006). Liquid fuel production from synthesis

gas via fermentation process in a continuous tank bioreactor (CSTBR) using Clostridium

ljungdahlii. Iranian J. Biotechnol, 4(1), 45-53.

Younesi, H., Najafpour, G., & Mohamed, A.R. (2005). Ethanol and acetate production from

synthesis gas via fermentation processes using anaerobic bacterium, Clostridium

ljungdahlii. Biochemical Engineering Journal, 27(2), 110–119.

Youngquist, J. A., Krzysik, A. M., English, B. W., Spelter, H. N., & Chow, P. (1996).

Agricultural fibers for use in building components. The Use of Recycled Wood and Paper

in Building Applications, 123–134.

Yu, G., Yano, S., Inoue, H., Inoue, S., Endo, T., & Sawayama, S. (2010). Pretreatment of rice

straw by a hot-compressed water process for enzymatic hydrolysis. Applied Biochemistry

and Biotechnology, 160(2), 539–551.

Yuan, H. K., Ren, J., Ma, X. H., & Xu, Z. L. (2011). Dehydration of ethyl acetate aqueous

solution by pervaporation using PVA/PAN hollow fiber composite membrane.

Desalination, 280(1), 252–258.

Zabihi, S., Alinia, R., Esmaeilzadeh, F., & Kalajahi, J.F., (2010). Pretreatment of wheat straw

using steam, steam/acetic acid and steam/ethanol and its enzymatic hydrolysis for sugar

production. Biosystems Engineering, 105(3), 288–297.

Zacchi, G., & Axelsson, A. (1989). Economic evaluation of preconcentration in production of

ethanol from dilute sugar solutions. Biotechnology and bioengineering, 34(2), 223–233.

Zafeiriou, E., Arabatzis, G., Tampakis, S., & Soutsas, K. (2014). The impact of energy prices on

the volatility of ethanol prices and the role of gasoline emissions. Renewable and

Sustainable Energy Reviews, 33, 87–95.

Zaldivar, J., Nielsen, J., & Olsson, L. (2001). Fuel ethanol production from lignocellulose: a

challenge for metabolic engineering and process integration. Applied Microbiology and

Biotechnology, 56(1), 17–34.

Zhang, T., Wyman, C. E., Jakob, K., & Yang, B. (2012). Rapid selection and identification of

Miscanthus genotypes with enhanced glucan and xylan yields from hydrothermal

pretreatment followed by enzymatic hydrolysis. Biotechnology for biofuels, 5(1), 56.

Zhang, Y. (2008). Rethinking thermochemical conversion of biomass into biofuel. Resource,

15(2), 6 & 27.

169

Zhu, J. Y., & Pan, X. J. (2010). Woody biomass pretreatment for cellulosic ethanol production:

technology and energy consumption evaluation. Bioresource Technology, 101(13), 4992–

5002.

Zhu, J. Y., Pan, X. J., Wang, G. S., & Gleisner, R. (2009b). Sulfite pretreatment (SPORL) for

robust enzymatic saccharification of spruce and red pine. Bioresource Technology, 100(8),

2411–2418.

Zhu, J. Y., Wang, G. S., Pan, X. J., & Gleisner, R. (2009a). Specific surface to evaluate the

efficiencies of milling and pretreatment of wood for enzymatic saccharification. Chemical

Engineering Science, 64(3), 474–485.

170

Appendices

A-2-1 The schematic diagram of chemical looping gasification (CLG) system

Source: Acharya, 2011

171

A-2-2 Brief summary of microorganisms identified and used for syngas fermentation

Species Optimum

Temp. C pH Time, h Products Reference

Mesophilic bacteria

Clostridium ljungdahlii 37 6.0 3.8 Acetate, ethanol Tanner et al., 1993

Clostridium autoethanogenum 37 5.8-6.0 - Acetate, ethanol Abrini et al., 1994

Clostridium carboxidivorans 38 6.2 6.2

Acetate, ethanol,

butyrate, butanol Liou et al., 2005

Oxobacter pfennigii 36–38 7.3 13.9 Acetate, n-butyrate Krumholz et al., 1985

Acetobacterium woodii 30 6.8 13 Acetate Sharak et al., 1987

Eubacterium limosum 38–39 7.0–7.2 7 Acetate Sharak et al.,1982, 1987

Rhodospirillum rubrum 30 6. 8 8.4 H2 Kerby et al., 1995

Rubrivivax gelatinosus 34 6.7–6.9 6.7 H2

Dashekvicz & Uffen,

1979; Uffen, 1976

Rhodopseudomonas palustris P4 30 nr 23 H2 Jung et al., 1999

Mesophilic archaea

Methanosarcina barkeri 37 7.4 65 CH4 O’Brien et al , 984

Methanosarcina acetivorans strain C2A 37 7 24

Acetate, formate,

CH4 Rother & Metcalf, 2004

Thermophilic bacteria

Moorella thermoacetica 55 6.5–6.8 10 Acetate Daniel et al., 1990

Moorella thermoautotrophica 58 6.1 7 Acetate Savage et al., 1987

172

Moorella strain AMP

60–65

6.9

-

H2

Jiang, 2006

Carboxydothermus hydrogenoformans 70–72 6.8–7.0 2 H2 Svetlitchnyi et al., 2001

Carboxydibrachium pacificus 70 6.8–7.1 7.1 H2 Sokolova et al., 2001

Carboxydocella sporoproducens 60 6.8 1 H2 Slepova et al., 2006

Carboxydocella thermoautotrophica 58 7 1.1 H2 Sokolova et al., 2002

Thermincola carboxydiphila 55 8 1.3 H2 Sokolova et al., 2005

Thermosinus carboxydivorans 60 6.8–7.0 1.2 H2 Sokolova et al., 2004

Thermophilic archaea

Methanothermobacter

thermoautotrophicus 65 7.4 140 CH4 Daniels et al., 1977

Thermococcus strain AM4 82 6.8 - H2 Sokolova et al., 2004

Source: Henstra et al., 2007.

Jung GY, Kim JR, Park JY, Park S: Hydrogen production by a new chemoheterotrophic bacterium Citrobacter sp. Y19. Int J

Hydrogen Energy 2002, 27:601-610.

Svetlitchnyi V, Peschel C, Acker G, Meyer O: Two membraneassociated NiFeS-carbon monoxide dehydrogenases from the anaerobic

carbon-monoxide-utilizing eubacterium Carboxydothermus hydrogenoformans. J Bacteriol 2001, 183:5134-5144.

Dashekvicz MP, Uffen RL: Identification of a carbon monoxide metabolizing bacterium as a strain of Rhodopseudomonas gelatinosa

(Molisch) Vanniel. Int J Syst Bacteriol 1979, 29:145-148.

Uffen RL: Anaerobic growth of a Rhodopseudomonas species in the dark with carbon monoxide as sole carbon and energy substrate.

Proc Natl Acad Sci USA 1976, 73:3298-3302.

A-2-2 Continued.

173

Jung GY, Jung HO, Kim JR, Ahn Y, Park S: Isolation and characterization of Rhodopseudomonas palustris P4 which utilizes CO with

the production of H2. Biotechnol Lett 1999, 21:525-529.

Jiang, B., 2006. PhD thesis, Wageningen University, 2006.

Kerby RL, Ludden PW, Roberts GP: Carbon monoxidedependent growth of Rhodospirillum rubrum. J Bacteriol 1995, 177:2241-

2244.

Sharak Genthner BR, Bryant MP: Growth of Eubacterium limosum with carbon monoxide as the energy source. Appl Environ

Microbiol 1982, 43:70-74.

Sharak Genthner BR, Bryant MP: Additional characteristics of one-carbon-compound utilization by Eubacterium limosum and

Acetobacterium woodii. Appl Environ Microbiol 1987, 53:471-476.

Krumholz LR, Bryant MP: Clostridium pfennigii sp. nov. uses methoxyl groups of monobenzenoids and produces butyrate. Int J Syst

Bacteriol 1985, 35:454-456.

O’Brien J , Wolkin RH, oench TT, organ JB, Zeikus JG, Lupton FS, Conrad R: ssociation of hydrogen metabolism with

unitrophic or mixotrophic growth of Methanosarcina barkeri on carbon monoxide. J Bacteriol 1984, 158:373-375.

Rother M, Metcalf WW: Anaerobic growth of Methanosarcina acetivorans C2A on carbon monoxide: an unusual way of life for a

methanogenic archaeon. Proc Natl Acad Sci USA 2004, 101:16929-16934.

Sokolova TG, Gonzalez JM, Kostrikina NA, Chernyh NA, Tourova TP, Kato C, Bonch Osmolovskaya EA, Robb FT:

Carboxydobrachium pacificum gen. nov., sp nov., a new anaerobic, thermophilic, CO-utilizing marine bacterium from Okinawa

Trough. Int J Syst Evol Microbiol 2001, 51:141-149.

Sokolova TG, Kostrikina NA, Chernyh NA, Tourova TP, Kolganova TV, Bonch-Osmolovskaya EA: Carboxydocella

thermautotrophica gen. nov., sp. nov., a novel anaerobic, CO-utilizing thermophile from a Kamchatkan hot spring. Int J Syst

Evol Microbiol 2002, 52:1961-1967.

174

Sokolova TG, Kostrikina NA, Chernyh NA, Kolganova TV, Tourova TP, Bonch-Osmolovskaya EA: Thermincola carboxydiphila gen.

nov., sp. nov., a novel anaerobic, carboxydotrophic, hydrogenogenic bacterium from a hot spring of the Kae Baikal area. Int J

Syst Evol Microbiol 2005, 55:2069-2073.

Sokolova TG, Gonzalez JM, Kostrikina NA, Chernyh NA, Slepova TV, Bonch-Osmolovskaya EA, Robb FT: Thermosinus

carboxydivorans gen. nov., sp. nov., a new anaerobic, thermophilic, carbon monoxide-oxidizing, hydrogenogenic bacterium

from a hot pool of Yellowstone National Park. Int J Syst Evol Microbiol 2004, 54:2353-2359.

Daniels L, Fuchs G, Thauer RK, Zeikus JG: Carbon monoxide oxidation by methanogenic bacteria. J Bacteriol 1977, 132:118-126.

Sokolova TG, Jeanthon C, Kostrikina NA, Chernyh NA, Lebedinsky AV, Stackebrandt E, Bonch-Osmolovskaya EA: The first

evidence of anaerobic CO oxidation coupled with H2 production by a hyperthermophilic archaeon isolated from a deep-sea

hydrothermal vent. Extremophiles 2004, 8:317-323.

Savage MD, Wu ZG, Daniel SL, Lundie LL, Drake HL: Carbon monoxide-dependent chemolithotrophic growth of Clostridium

thermoautotrophicum. Appl Environ Microbiol 1987, 53:1902-1906.

Daniel SL, Hsu T, Dean SI, Drake HL: Characterization of the hydrogen- and carbon monoxide-dependent chemolithotrophic

potentials of the acetogens Clostridium thermoaceticum and Acetogenium kivui. J Bacteriol 1990, 172:4464-4471.

Savage MD, Wu ZG, Daniel SL, Lundie LL, Drake HL: Carbon monoxide-dependent chemolithotrophic growth of Clostridium

thermoautotrophicum. Appl Environ Microbiol 1987, 53:1902-1906.

Slepova TV, Sokolova TG, Lysenko AM, Tourova TP, Kolganova TV, Kamzolkina OV, Karpov GA, Bonch- Osmolovskaya EA:

Carboxydocella sporoproducens sp. nov., a novel anaerobic CO-utilizing/H2-producing thermophilic bacterium from a

Kamchatka hot spring. Int J Syst Evol Microbiol 2006, 56:797.

175

A-2-3 Syngas fermentation parameters and ethanol yield

Experimental conditions Yield Reference

Microorganism Cell

density

Media Reactor Feedstock Temp. 0C

pH Agitation,

rpm

C. Ljungdahlii,

ATCC#55383

10 g/L ATCC1754 PETC Two stage: 1L

CSTR & 4-L

bubble column

60% CO, 35% H2, & 5%

CO2

37 5.5

(stage 1)

4.4-4.8

(stage 2)

200 0.37 g/(L·h). Richter et

al., 2013

Clostridium

ragsdalei

0.95 g/L - Batch 20%CO, 5%/H2,

15%/CO2, & 60% /N2

32 6 - 1.89 g/L ethanol,

1.45 g/L acetate

Kundiyana

et al., 2011

C. Ljungdahlii,

ATCC#55383

ATCC1754 PETC Hollow fiber

membrane reactor

(125 mL)

50% CO, 30% H2, and

20% CO2

35 - - 6 g/L ethanol to

acetate ratio 2.6

Lee, 2010

C. Ljungdahlii

(growth phase)

Modified nitrogen

limitation media

Batch (80 ml) - 37 6.8 - 0.02 g/L ethanol

> 0.06 g/L acetate

Cotter et

al., 2009

C. Ljungdahlii

(resting phase)

Modified nitrogen

limitation media

Batch (80 ml) - 37 4.5 0.02 g/L ethanol

> 0.06 g/L acetate

Cotter et

al., 2009

C.

autoethanogenum

DSMZ 640 Batch (250 ml) 20% CO, 10% H2, 20%

CO2, and 50% N2

(flow rare: 10 mL/min)

37 6.0 - 0.66 g/L ethanol

1.40 g/L acetate

Cotter et

al., 2008

C. Ljungdahlii

(growth phase)

Modified

reinforced

Clostridial

Batch (250 ml) 20% CO, 10% H2, 20%

CO2, and 50% N2

(flow rare: 7.5 mL/min)

37 6.8 - 0.23 g/L ethanol

2.10 g/L acetate

Cotter et

al., 2008

Clostridium

aceticum

0.8 g/L Batch 78% CO, 4% H2, 18%

Ar

30 8.5 - 2.27 g/L acetate Sim et al.,

2007

C. Ljungdahlii 1.2 g/L ATCC1754 PET Batch (125 ml) 55% CO, 20% H2, 10%

CO2, and 15% Ar

37 4.5 - 0.60 g/L ethanol

1.11 g/L acetate Younesi et

al., 2005

176

C. Ljungdahlii ATCC1754 PETC Batch (125 mL) 55% CO, 20% H2, 10%

CO2, and 15% Ar

37 - - 0.012 g/L ethanol

0.024 g/L acetate

Najafpour &

Younesi,

2006

C.Lljungdahlii ATCC1754 PETC CSTR (20L) 55% CO, 20% H2, 10%

CO2, and 15% Ar

37 4.5 11 g/L ethanol and

acetate

Younesi et

al., 2006

Eubacterium

limosum

0.75 g/L - Gas lift reactor 100% CO 37 6.8 - Butyrate, acetate, &

38 mmol ethanol

Chang et

al., 1998

C.Ljungdahlii ATCC1754 PETC STR, with cell

recycle (13.5 L)

55% CO, 20% H2, 10%

CO2, 15% Ar

37 4.5 300-500 1.5 g/L ethanol

3.5 g/L acetate

Phillips et

al., 1994

Designed media

based on E. Coli

13.5 L STR, with

cell recycle

55% CO, 20% H2, 10%

CO2, and 15% Ar

37 4.5 300-500 23 g/L ethanol

27 g/L acetate

Phillips et

al., 1994

C.Ljungdahlii 4 g/L Cell recycled 55% CO, 20% H2, 10%

CO2, 15%Ar

36 4.5 300-500 48 g/L ethanol Phillips et

al., 1993

Note: Partly from Lee, P.H., 2010. Syngas fermentation to ethanol using innovative hollow fiber membrane. PhD Thesis. Iowa State University, USA.

Phillips, J. R., Klasson, K. T., Clausen, E. C., & Gaddy, J. L. (1993). Biological production of ethanol from coal synthesis gas. Applied

biochemistry and biotechnology, 39(1), 559-571.

Sim, J. H., Kamaruddin, A. H., Long, W. S., & Najafpour, G. (2007). Clostridium aceticum—A potential organism in catalyzing carbon

monoxide to acetic acid: Application of response surface methodology. Enzyme and Microbial Technology, 40(5), 1234-1243.

A-2-2 Continued.

177

A-6-1 Land classification in Ontario

Classification Descriptions

Class 1 Well developed and has no significant limitations in use for crops.

Class 2 Has moderate limitations that restrict the range of crops or requires moderate conservation practices.

Class 3 Has moderately severe limitations that restrict the range of crops or requires special conservation practices.

Class 4

Has severe limitations that restrict the choices of crops or requires special conservation and management practices, or

both.

Class 5

Has very severe limitations that restrict their capability in producing perennial forage crops, and improvement practices

are feasible.

Class 6 Unsuitable for cultivation, and can be used for unimproved permanent pasture.

Class 7 Has no capacity of arable culture or permanent pasture.

Source: AAFC, 2008; OMAF, 2013

178

A-6-2 On-farm inputs for miscanthus cultivation in different regions

Items/Parameter Other

input values

Region

Western Ontario Southern Ontario Eastern Ontario Central Ontario

Land class

1-2 3 3 4-5

Soil type

Silt Loam-Clay

Loam

Clay Clay Loam

Displace crop rotation

Corn-soy

rotation

Continuous

soybean rotation

Corn-Corn-

Forage- Forage-

Forage rotation

Long term

pasture

Production

Rhizome yield (production) (tonne/ha) 10 10

Rhizome required (tonne/ha) 0.8 0.8

Distance rhizome producer to farm (km) 100 100 100 100 200

Moisture content at harvest (%) 15 15

Miscanthus crop lifespan (years) 20 20 20 15 15

Crop residue left on soil at harvest (% of peak

biomass produced) 30 30

Below ground biomass contribution to total

mass (%) 35 35

Spring yield years 3 to 20 (dry tonne/ha) 11.10 11.10 11.10 10.03 8.90

Stand failure year 2 (% of area) 10 10 20 40 40

N application rate year 2-20 (kg/ha) 60 60 80 80 60

P application rate year 2-20 (kg/ha) 12 12 11 11 9

K application rate year 2-20 (kg/ha) 105 105 105 95 79

Percent of N requirements fulfilled with urea 100%

Total herbicide application Year 1 (kg active

ingredient /ha) 10 10 10 10 11.8

Herbicide/pesticide application rate Year 2-20

(kg active ingredient /ha) 4.00 4

179

Herbicide application rate at termination (kg

active ingredient /ha) 4.00 4

Fuel use for Miscanthus planting (L/ha) 69.06 69.06

Production efficiency 1 1 1 1 0.85

Displaced crop C pools

Harvested material yield (dry t/ha) 2.3 2.3 1.1 3.7 2.4

Residues yield (dry t/ha) 2.3 2.3 1.1 1.9 1.6

Total above ground biomass (dry t/ha) 4.7 4.7 2.1 5.6 4.0

Belowground biomass (dry t/ha) 2.2 2.2 0.5 3.1 3.0

Soil C (dry t/ha) 80.9 80.9 67.8 76.2 89.1

Total C (dry t C/ha) 87.8 87.8 72.6 90.4 100.0

Miscanthus C pools (average of life expectancy of stand)

Harvested material yield (dry t C/ha) 5.2 5.19 5.19 4.65 4.13

Residues yield (dry t C/ha) 2.2 2.23 2.23 1.99 1.77

Total above ground biomass (dry C t/ha) 7.4 7.42 7.42 6.65 5.90

Belowground biomass (dry t C/ha) 4.0 4.00 4.00 3.58 3.18

Loss soil C at conversion (%) 2.5 0.0 2.5

Initial soil C after conversion (dry t C/ha) -

assume loss at conversion 78.9 78.90 66.14 74.27 86.89

Total C returned to land (dry t C/ha)-above only 2.30

Total C returned to land (dry t C/ha) - above and

below 6.3

ICBM output - final soil C (dry t C/ha) 85.3 85.27 85.27 84.95 84.95

Total final C (dry t C/ha) 96.7 96.7 96.7 95.2 94.0

Total average C (dry t C/ha) 93.5 93.5 87.1 89.8 95.0

Annual system C capture (t C/ha*yr) 0.45 0.45 1.21 0.32 - 0.40

Annual system C capture (t CO2/ha*yr) 1.63 1.63 4.42 1.17 -1.46

Source: Sanscartier et al., 2013

A-6-2 Continued.

180

A-6-3 On-farm energy and other inputs for miscanthus cultivation

Items/Parameters

Inputs and

other values

Energy use (diesel)

Diesel use (L/ha) L/t dry baled miscanthus

Year one - Stand establishment

Rhizome production

12.23 1.18

Rhizome required (tonne/ha) 0.80

Rhizome delivery

3.29 0.32

Rhizome planting

69.06 6.65

Pesticide production and distribution (input = application rate in kg/ha) 10.00

1.28

Pesticides application

0.88 0.08

Mowing

6.39 0.62

Crop residue contribution to total biomass produced 0.30

Total crop residue (above ground and below ground) returned to soil

(dry tonne/ha) 6.56

Year two

Stand failure (%) 10%

Rhizome production and delivery

1.55 0.15

Rhizome planting (by hand, assumption: 5% of energy required for

machine planting)

0.35 0.03

Fertilizer production and distribution

N (input = application rate in kg/ha) 60.00

0.08

P (input = application rate in kg/ha) 12.00

0.20

K (input = application rate in kg/ha) 105.00

0.07

Application of fertilizers

2.26 0.22

Pesticide production and distribution (input = application rate in kg/ha) 4.00

0.51

181

Pesticides application

0.88 0.08

Harvest

40.25 3.88

Transport to on-farm storage

1.61

Crop residue contribution to total biomass produced 0.30

Total crop residue (above ground and below ground) returned to soil

(dry tonne/ha) 13.13

Year 3 to 20

Harvest

40.25 3.88

Transport to on-farm storage

1.61

Fertilizer production and distribution

N (input = application rate in kg/ha) 60.00

0.08

P (input = application rate in kg/ha) 12.00

0.20

K (input = application rate in kg/ha) 105.00

0.07

Fertilizer application

2.26 0.22

Pesticide production and distribution (input = application rate in kg/ha) 4.00

0.51

Application of pesticides

0.88 0.08

Crop residue contribution to total biomass 0.30

Total crop residue (above ground and below ground) returned to soil

(dry tonne/ha) 13.13

Termination

Herbicide production and distribution (input = application rate in

kg/ha) 4.00

0.51

Application of pesticides

0.88 0.08

Ploughing

11.61 1.12

Total on-farm per year (allocating Years 1 and 2 to 18 years after

establishment of crop) 51.69 7.68

A-6-3 Continued.

Source: Sanscartier et al., 2013

182

A-6-4 Estimated emission from farm input and carbon sequestration

Parameters CO2e (g/ODT miscanthus delivered at farmgate)

Location

Western

Ontario

Southern

Ontario

Eastern

Ontario

Central

Ontario

Land class 1-2 3 3 4-5

Soil type Silt loam - clay

loam Clay Clay Loam

Displace crop rotation Corn-

soy

rotation

Continuous

soybean

rotation

Corn-corn-forage-

forage

rotation

Long

term

pasture

Emission from farm input 128858.17 149714.94 166284.58 155710.04

Change in C content in all pools -157434.66 -237784.33 1185.43 11957.97

Net emission -28576.49 -88069.39 167470.01 167668.00

Raw data source: Sanscartier et al., 2013

183

A-6-5 Calculation of energy consumption and material cost of enzyme production

Energy consumption and material cost were calculated based on the enzyme production

cost reported by Wooley et al., 1999 (enzyme loading: 15 FPU/g-cellulose≈ 9263 FPU/L).

Electricity price was 4.2¢/kWh in 1997 in USA (EIA, 2010). Based on these information, and

enzyme loading rate of this study material and energy cost (based on the electricity price in

Canada in 2012) for enzyme were estimated.

Reference

EIA (Energy Information Administration), 2010. Official Energy Statistics from the US Govt., <

http://www.eia.doe.gov/cneaf/electricity/epa/epat7p4.html>

Wooley, R., Ruth, M., Sheehan, J., Ibsen, K., Majdeski H., Galvez, A., 1999. Lignocellulosic

biomass to ethanol—process design and economics utilizing co-current dilute acid

prehydrolysis and enzymatic hydrolysis—current and futuristic scenarios. Report No. TP-

580-26157. National Renewable Energy Laboratory. Golden Colorado USA.

22.1 cm (100%); 0.30 $/galCapital Material Process electricity Fixed cost

7.5 cm

33.94%

3.8 cm

17.19%

9.4 cm

42.53%1.4 cm

6.33%

Enzyme production

Assumed electricity price: 4.2 ¢/kWh

Energy consumption= 0.3$×100×0.4253/4.2/3.785=0.802 kWh/L

Material cost= 0.3$×100×0.1719/3.785= 1.362¢/L

184

A-8-1 Membrane separator

A-8-2 Membrane support

185

A-8-3 List of materials/accessories for the developed bioreactor

Name of items Quantity Material type Model No. Manufacturer Cost, $ Remarks

Bioreactor

1

Plexiglass(R)

VH-100 Acrylic

Resin, & PVC-

9002-86-2 - Developed - -

Swagelok 12 Stainless steel 316Z77 - - -

Swagelok 1 Stainless steel - - -

Swagelok 1 Plastic - - -

pH meter 1 - MV-RS232 Omega 148.5 -

pH probe 1 - HHWT-SD1-ATC Omega 63.0 -

Temperature probe 1 - PHE-1411 Omega 40.5 -

Pressure gauge 1 - DPG110/120 Omega Digital

Aeration tube 1 - - - 11.0 Rubber

Support for aeration tube 1 Stainless steel - Developed - -

Micro pump 1 GF-F155001 Gilson 1295.0

Pump head 1 - GF-F117800 Gilson 550.8 Two channel head

Peristatic Tubing 10 Polypropylene GF-F1825121 Gilson 92.65 -

Connection tube 1 PVC 06422-05 Cole-Parmer 35.0 25 feet

Connector 1 Plastic F1179931 Mandel - Set of 10

Connector 1 Plastic

Membrane separator 1 - - Developed - Materials from Dr. Sheng

Membrane support 1 Stainless steel - Developed -

Jars (for media &

effluent) 6 Glass - - -

-

Reactor lead opener 2 Steel - - - -

Teflon ferrule 1 Teflon - - - Set of 10

186

A-8-4 List of materials/accessories for anaerobic gas chamber

Name of items Quantity Material type Model No. Manufacturer Cost, $ Remarks

Anaerobic gas

chamber

1

PVC, 58092421,

Bayer Material

Science LLC - Developed -

Valve 2 PVC - - - -

Glove (pair) 1 Rubber - - -

-

Pipe fittings

2 PVC 9002-86-2

National Pipe &

Plastics, Inc.

USA -

-

Hinge 2 Steel - -

-

Handle 3 Steel - - -

-

A-8-5 List of chemicals and their amount used for broth media

Peptone…………………………………… 0 0 g

Beef E tract………………………………… 0 0 g

Yeast E tract……………………………… 3 0 g

De trose…………………………………… 5 0 g

NaCl…………………………………………5 0 g

Soluble Starch…………………………… 0 g

L-Cysteine HCl…………………………… 0 5 g

Sodium cetate…………………………… 3 0 g

Resa urin (0 025%)……………………… 4 ml

DI Water…………………………………… 000 ml

Add all ingredients except L-Cysteine HCl. Bring media to boil to drive off oxygen. Cool

down media while bubbling with oxygen free gas. Add L-Cysteine HCl and adjust pH to 6.8.

Dispense under same gas phase and autoclave at 121ºC.

187

A-8-6 Photograph of the incubator

A-8-7 Calibration curve of the pump

yin = 0.08044x - 0.01799

(R² = 0.9989; n=36)

yout = 0.08044x - 0.01470

(R² = 0.9987; n=36)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 5 10 15 20

Flo

w r

ate,

mL

/min

Pump speed, rpm

In Out

Heratherm IGS60

188

A-8-8 Photographs of overall experimental setup

189

A-9-1 Experimental setup of torrefaction process adopted in this study

A-9-2 Composition of flue gas from biomass torrefaction process

Feedstock

Torrefaction

time, min

Composition of flue gas, ppm

CO2 CO O2 SO2 NO2 NOx

Miscanthus

0 19.85 0.00 0.25 0-1.0 0.00 0.00

15 19.88 0-1.0 0.12 0-1.0 0.00 0.00

30 19.89 0-1.0 0.11 0-1.0 0.00 0.00

45 19.88 0-1.0 0.12 0-1.0

0.00 0.00

Wheat straw

0 16.15 1.00 3.98 1.00 0.00 0.00

15 15.60 0-1.0 4.62 1.00 0.00 0.00

30 15.66 0-1.0 4.52 0-1.0 0.00 0-0.1

45 15.68 0-1.0 4.40 0-1.0 0.00 0.00

Sawdust (pine) 0 18.11 1.00 2.09 0.00

0.00 0.00

15 18.87 0.00 1.25 0.00

0.00 0.00

30 18.85 0.00 1.20 0.00

0.00 0.00

45 19.17 0-1.0 0.89 0-1.0

0-0.1 0.00

Source: Acharya, 2013

190

A-9-3 Energy consumption in the torrefaction of biomass (for 45 min)

Specific heat of biomass has been calculated based on the following formula (Ravi et al., 2004):

= 0.1031 + 0.003867T kJ/kg-K . . . . . (Eq. A-1)

+ (0.02355T − 1.32M − 6.191)M kJ/kg-K . (Eq. A-2)

where, T= absolute temperature, K

M= the mass fraction of moisture in wood

The estimated specific heat (Cp) of wood = 1.522 kJ/kg-k (moisture is assumed to be 15%).

Energy consumption in torrefaction process = mCpt=0.41 MJ/kg.

Reference: Ravi, M. R., Jhalani, A., Sinha, S., & Ray, A. (2004). Development of a semi-

empirical model for pyrolysis of an annular sawdust bed. Journal of analytical and applied

pyrolysis, 71(1), 353-374.

A-9-4 Flash 200 CHNS-O, Organic Elemental Analyzer

Flash 200 CHNS-O, Organic Elemental Analyzer,

191

A-9-5 Photograph of thermo gravimetric analyzer (TGA)

SDT-Q600

A-9-6 Photograph of Fourier transform infrared spectroscopy (FT-IR)

Thermo Scientific Nicolet 6700

192

A-9-7 TGA/FT-IR experimental parameters

Pretreatment Feedstock Sample size, mg Air flow rate, mL/min

Miscanthus 10.18 13.00

Raw Wheat straw 8.00 10.00

Sawdust (pine) 10.13 14.00

Miscanthus 8.83 13.00

PT Wheat straw 9.52 13.00

Sawdust (pine) 9.51 14.00

Miscanthus 19.86 13.00

Raw-CaO Wheat straw 19.98 13.00

Sawdust (pine) 20.30 14.00

Miscanthus 18.00 13.00

PT-CaO Wheat straw 18.00 13.00

Sawdust (pine) 19.15 14.00

Note: Thermal degradation temperature: 900 °C; Heating rate: 20 °C/min; Pretreatment (PT): 275 °C@45

min, ER: 0.3; Ratio of CaO and biomass is 1.

A- 9-8 Comparison among various raw biomasses

-5

0

5

10

15

20

0

20

40

60

80

100

0 100 200 300 400 500 600 700 800 900

Wei

ght

loss

, %

Temperature, °C

MS-Wt

SD-Wt

WS-Wt

MS-HF

SD-HF

WS-HF

Hea

t fl

ow

, W

/g

(MS: miscanthus; SD: sawdust ; WS: wheat straw; Wt: weight loss; HF: heat flow)

193

A- 9-9 Comparison among various torrefied biomasses

A- 9-10 Comparison among various raw biomasses degraded with CaO

-5

0

5

10

15

20

25

30

35

40

0

20

40

60

80

100

0 100 200 300 400 500 600 700 800 900

Wei

ght

loss

, %

Temperature, °C

MS-Wt

SD-Wt

WS-Wt

MS-HF

SD-HF

WS-HF

Hea

t fl

ow

, W

/g

-4

-2

0

2

4

6

8

10

12

14

0

20

40

60

80

100

0 100 200 300 400 500 600 700 800 900

Wei

gh

t lo

ss, %

Temperature, °C

MS-Wt

SD-Wt

WS-Wt

MS-HF

SD-HF

WS-HF

Hea

t fl

ow

, W

/g

(MS: miscanthus; SD: sawdust ; WS: wheat straw; Wt: weight loss; HF: heat flow)

(MS: miscanthus; SD: sawdust ; WS: wheat straw; Wt: weight loss; HF: heat flow)

194

A- 9-11 Comparison among various torrefied biomasses degraded with CaO

A- 9-12 Comparison among raw and torrefied with or without CaO (miscanthus)

-5

0

5

10

15

20

0

10

20

30

40

50

60

70

80

90

100

0 100 200 300 400 500 600 700 800 900

Wei

ght

loss

, %

Temperature, °C

MS-Wt

SD-Wt

WS-Wt

MS-HF

SD-HF

WS-HF

Hea

t fl

ow

, W

/g

-5

0

5

10

15

20

25

30

35

0

20

40

60

80

100

0 100 200 300 400 500 600 700 800 900

Wei

ght

loss

, %

Temperature, °C

R-Wt

R+CaO-Wt

T-Wt

T+CaO-Wt

R-HF

R+CaO-HF

T-HF

T+CaO-HF

Hea

t fl

ow

, W

/g

(MS: miscanthus; SD: sawdust ; WS: wheat straw; Wt: weight loss; HF: heat flow)

(T: torrefied; R: raw; weight loss; HF: heat flow)

195

A-9-13 Cold gas efficiency (CGE) calculation for steam gasification (Gai & Dong, 2012).

. . . . . (Eq. A-3)

where, = volume of product gas from the gasification (Nm3/kg-fuel)

The lower heating value (LHV) product gas is calculated from the following equation (Alamo et

al., 2009).

LHV (MJ/Nm-3

) = ∑ × . . . . . . (Eq. A-4)

where, X = mole fraction of the corresponding gas

= lower heating value of corresponding gas

Reference:

Gai, C., & Dong, Y. (2012). Experimental study on non-woody biomass gasification in a

downdraft gasifier. International Journal of Hydrogen Energy, 37(6), 4935-4944.

del Alamo, G., Hart, A., Grimshaw, A., & Lundstrøm, P. (2012). Characterization of syngas

produced from MSW gasification at commercial-scale ENERGOS Plants. Waste

management, 32(10), 1835-1842.

A-9-14 Summary of ASPEN simulation parameters (Feed stream input conditions for CLG

simulation)

Feed Stream

Input Conditions

Temperature

(°C)

Pressure (atm) Flowrate

(kmol h-1

) Component

BIOMASS 25 1 1a Biomass

H2O-FEED 25 1 1 H2O (Conventional)

CAO-FEED 25 1 6b CaO (Conventional

Solid)

STEAM 400 1 85b H2O (Conventional)

a Input as mass flowrate (kg h

-1) using biomass molecular weight.

b Fed in excess of required stoichiometric amount.

196

A-9-15 Summary of ASPEN simulation parameters and CLG block diagram.

Fig. CLG block diagram (Stefan, 2014)

197

A-9-16 CLG simulation flowsheet (Stefan, 2014)

198

A-9-17 Product gas compositions (simulated) and CGE

Feedstock Conditions Gas composition, %

Nm3/kg-

FS

LHV,

MJ/N

m3

CGE,

%

CO H2 CO2 CH4

Raw 40.76 51.44 6.04 1.76

1.653 11.317 78.830

Miscanthus Torrefied 40.57 51.67 5.99 1.78

1.774 11.324 78.652

Raw, CLG 8.24 89.52 1.05 1.19

1.602 11.118 79.055

Torrefied,

CLG 8.21 89.56 1.04 1.19

1.721 11.091 78.668

Wheat straw Raw 41.28 50.81 6.20 1.72

1.658 11.300 74.829

Raw, CLG 8.33 89.42 1.07 1.18

1.604 11.115 83.357

Sawdust Raw 40.36 51.92 5.93 1.79

1.704 11.331 71.652

Raw, CLG 8.33 89.42 1.07 1.18

1.655 11.119 79.436

Note: Volume of CO2 excluded while calculation the CGE.


Recommended