torrefaction and pelletization of different forms of

168
Torrefaction and Pelletization of Different Forms of Biomass of Ontario by Bimal Acharya A Thesis Presented to The University of Guelph In partial fulfillment of requirements for the degree of Master of Applied Science in Engineering © Bimal Acharya, March, 2013

Upload: others

Post on 13-Jan-2022

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Torrefaction and Pelletization of Different Forms of

Torrefaction and Pelletization of Different Forms of Biomass

of Ontario

by

Bimal Acharya

A Thesis

Presented to

The University of Guelph

In partial fulfillment of requirements

for the degree of

Master of Applied Science

in

Engineering

© Bimal Acharya, March, 2013

Page 2: Torrefaction and Pelletization of Different Forms of

ii

ABSTRACT

Torrefaction and Pelletization of Different Forms of Biomass of Ontario

Bimal Acharya Advisor

University of Guelph, 2013 Dr. Animesh Dutta

The purpose of this study is to investigate the torrefaction and pelletization

behavior, hydrophobicity, storage behavior, ash analysis on three different biomasses:

one (willow pellets) from wood products, one (oat pellets) from agricultural products and

one (poultry litter) from the non-lignocellulosic biomass products during the processes.

Four different torrefaction temperatures from 200°C-300°C, at 10-60 minute residence

times, 0%-2.4% oxygen concentration, were considered. Of these, 285°C for willow

pellets, 270°C for oat pellets and 275°C for poultry litter were found to be optimum for

hydrophobicity. Studies of XRD and SEM of biomass ash at 800°C, 900°C and 1000°C

were also carried out. The aforementioned results indicate that torrefaction is a feasible

alternative to improve energy properties of ordinary biomass and prevent moisture re-

absorption during storage.

Page 3: Torrefaction and Pelletization of Different Forms of

iii

ACKNOWLEDGEMENTS

I would firstly like to acknowledge my graduate advisor Dr. Animesh Dutta P.Eng., for

his thoughtful insight and support throughout my graduate career. His mentorship

extended beyond the walls of the University and his support allowed me to obtain several

prestigious awards and compete in Biological Engineering. I would also like to extend

my appreciation towards Dr. Shohel Mohmud who provided insight throughout the

duration of my research.

These acknowledgements would not be complete without naming the following

individuals for their supports: Dr. Mathias Leon: thanks for English correction and

guidance in the lab setup; John Whiteside, Joanne Ryks(School of Engineering): thanks

for providing lab support; and to my colleagues: Idris Sule, Dr Poritosh Roy, Maxime

Moisan, and Mohammad Tushar. I am glad our journeys have crossed paths and I hope

this is just the beginning of a lasting friendship.

I would lastly like to make personal acknowledgements to my family and friends. To my

mother, father, brother and sister who throughout my life always encouraged higher

education and a quest for knowledge. Special gratitude to my employer Nepal Telecom

for granting me study leave to pursue graduate study from a Canadian University.

Finally, to my wife and best friend Sushma Acharya, son Bibhu and daughter Suyasha,

who has always wholeheartedly supported me throughout my scholarly achievements.

Page 4: Torrefaction and Pelletization of Different Forms of

iv

TABLE OF CONTENTS

ABSTRACT ....................................................................................................................... ii

ACKNOWLEDGEMENTS ............................................................................................... iii

TABLE OF CONTENTS .................................................................................................. iv

LIST OF TABLES ............................................................................................................. ix

LIST OF FIGURES .............................................................................................................x

NOMECLATURE .............................................................................................................xv

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

1.1 Introduction ...................................................................................................... 1

1.2 Objectives ......................................................................................................... 4

1.3 Organization of the Thesis ................................................................................ 5

Chapter II: Literature Review ..............................................................................................6

2.1 Overview .......................................................................................................... 6

2.1.1 Fuel Characteristics of Biomass.................................................................... 7

2.1.2 Biomass Components.................................................................................... 8

2.2 Concept of Torrefaction .................................................................................. 11

2.3 Torrefaction Process Methods ........................................................................ 18

2.4 Classification of Reactors ............................................................................... 20

2.5 Commercial Application of Torrefaction in Canada ...................................... 26

2.6 Available Technologies for Torrefaction ....................................................... 29

2.7 Pelletization .................................................................................................... 31

2.8 Gasification ..................................................................................................... 36

Page 5: Torrefaction and Pelletization of Different Forms of

v

2.9 Emission ......................................................................................................... 38

2.10 Storage Behavior ............................................................................................ 38

2.11 Economic Potential ......................................................................................... 40

2.12 Scanning Electron Microscopy (SEM) ........................................................... 43

2.13 X-ray Diffraction (XRD) ................................................................................ 45

2.14 Summary from Literature ............................................................................... 46

Chapter III: Methodology ..................................................................................................49

3.0 Problem Statement .......................................................................................... 49

3.1 Research Scope and Objectives ...................................................................... 50

3.2 Methodology ................................................................................................... 52

Chapter IV: Experiment Setup ...........................................................................................54

4.1 Biomass Characterization ............................................................................... 54

4.1.1 Proximate Analysis ......................................................................................... 54

4.1.2 Ultimate Analysis ........................................................................................... 54

4.1.3 Heating Value ................................................................................................. 54

4.2 Torrefaction .................................................................................................... 55

4.3 Hydrophobicity ............................................................................................... 58

4.4 Pelletization .................................................................................................... 58

4.5 Storage Behaviour .......................................................................................... 60

4.6 Scanning Electron Microscopy (SEM) ........................................................... 60

4.7 X-ray Diffraction (XRD) ................................................................................ 62

Chapter V: Results and Analysis .......................................................................................63

5.1 Poultry Litter Biomass .................................................................................... 63

Page 6: Torrefaction and Pelletization of Different Forms of

vi

5.1.1 Biomass Characterization ........................................................................... 63

5.1.2 Torrefaction................................................................................................. 65

5.1.3 Hydrophobicity ........................................................................................... 68

5.1.4 Storage Behavior ......................................................................................... 69

5.1.5 Optimization by Box Behnken Model ........................................................ 70

5.1.6 Scanning Electron Microscopy (SEM) ....................................................... 74

5.1.8 Summary ..................................................................................................... 77

5.2 Willow Pellets................................................................................................. 78

5.2.1 Biomass Characterization ........................................................................... 78

5.2.2 Torrefaction................................................................................................. 81

5.2.4 Hydrophobicity ........................................................................................... 84

5.2.5 Storage Behavior ......................................................................................... 86

5.2.6 Pelletization................................................................................................. 88

5.2.7 Scanning Electron Microscope (SEM) ....................................................... 91

5.2.9 Summary ..................................................................................................... 95

5.3 Oat Pellets ....................................................................................................... 96

5.3.1 Biomass Characterization ........................................................................... 97

5.3.2 Torrefaction................................................................................................. 99

5.3.4 Hydrophobicity ......................................................................................... 102

5.3.5 Storage Behavior ....................................................................................... 104

Page 7: Torrefaction and Pelletization of Different Forms of

vii

5.3.6 Pelletization............................................................................................... 105

5.3.7 Scanning Electron Microscope (SEM) ..................................................... 107

5.3.9 Summary ................................................................................................... 111

5.4 Comparative Analysis................................................................................... 112

5.4.1 Biomass Characterization ......................................................................... 112

5.4.2 Torrefaction............................................................................................... 116

5.4.3 Hydrophobicity ......................................................................................... 119

5.4.4 Storage Behavior and Moisture Uptake .................................................... 120

5.4.5 Ash Analysis ............................................................................................. 121

5.4.6 Summary ................................................................................................... 123

5.5 Errors and Repeatability Test ....................................................................... 124

Chapter VI: Conclusions and Recommendations ............................................................125

6.1 Conclusions .................................................................................................. 125

6.2 Recommendations ........................................................................................ 127

Chapter VII: References ..................................................................................................129

Chapter VIII: Appendix ...................................................................................................141

Appendix A: Photographs of Willow Pellets .............................................................. 141

Appendix B: Photographs of Oats Pellets ................................................................... 142

Appendix C: Methods and Equipment used in Characterizing Biomass .................... 143

Appendix D: Photograph of Experimental Setup for Torrefaction ............................. 144

Appendix E: Gas Analyzer ......................................................................................... 145

Page 8: Torrefaction and Pelletization of Different Forms of

viii

Appendix F: Ultimate Analysis, Ash Fusion Temp and Ash Elemental Analysis ..... 146

Appendix G: Error and Repeatibility Tests ................................................................ 148

Page 9: Torrefaction and Pelletization of Different Forms of

ix

LIST OF TABLES

Table 5-2 Ultimate Analysis and Heating Value of Raw Biomass 113

Table Page No.

Table 2-1 Summary of Torrefied pellets Properties versus Coal 18

Table 2-2 Comparison of Potential Torrefaction Reactor Technologies 26

Table 2-3 Overview of Torrefaction Projects 30

Table 2-4 Summary of Specifications for Four Different Pelleting

Equipment

33

Table 2-5 Comparison of BO2 Pellet Properties 36

Table 5-1 Proximate Analysis of Raw Biomass 112

Page 10: Torrefaction and Pelletization of Different Forms of

x

LIST OF FIGURES

Figures Page No.

Figure1.1 Energy Density vs. Time and Temperature 3

Figure 2-1 Structure and Pretreatment effect on Biomass 9

Figure 2-2 Carbohydrate in the presence of Carbon Monoxide 10

Figure 2-3 Decomposition regimes of lignocellulosic material during

thermal treatment

13

Figure 2-4 Basic principle of torrefaction Process 20

Figure 2-5 Pictures of raw and pelletized materials 34

Figure 2-6 Schematic Process Flow of the BO2-Technology 35

Figure 2-7 Delivery costs of pelletized biomass. (Numbers indicate

nominal capacity of system dry kilotons of raw biomass feedstock per

year)

40

Figure 2-8 SEM of Wood Ash from Gasifier 45

Figure 3-1 Schematic Block Diagram of Research Procedures 51

Figure 3-2 Flow chart of the methodology 53

Figure 4-1 Experimental Setup for Torrefaction and Weight loss 57

Figure 4-2 Experimental Setup for Pelletization 59

Figure 4-3 SEM Experimental Setup 62

Figure 5-1 Volatile Matter vs. Residence Time for Poultry Litter 64

Figure 5-2 Fixed Carbon vs. Residence Time for Poultry Litter 64

Figure 5-3 Ash Contents vs. Residence Time for Poultry Litter 65

Page 11: Torrefaction and Pelletization of Different Forms of

xi

Figure 5-4 % of Mass Yield and Energy Yield vs. Temperature Poultry

Litter (0% Oxygen with 45 minutes residence time)

66

Figure 5-5 % of Mass Yield and Energy Yield vs. Temperature for

Poultry Litter (2.4% Oxygen with 45 minutes residence time)

67

Figure 5-6 Heating Value vs. Residence Time for Poultry Litter 67

Figure 5-7 Heating Value vs. Residence Time for Poultry Litter 68

Figure 5-8 Hydrophobic behavior of Poultry Litter 69

Figure 5-9 % of Moisture uptake vs. temperature of torrefied biomass 70

Figure 5-10 Cube of Temperature, Moisture and Residence Time Vs.

Mass Yield

71

Figure 5-11 Surface Plots of Temperature, Moisture and Residence

Time Vs. Mass Yield

72

Figure 5-12 Cube of Temperature, Moisture and Residence Time Vs.

Energy Yield

73

Figure 5-13 Surface Plot of Temperature, Moisture and Residence

Time Vs. Energy Yield

73

Figure 5-14(a) SEM for poultry litter ash at 800C 74

Figure 5-14(b) SEM for poultry litter ash at 900C 75

Figure 5-14(c) SEM for poultry litter ash at 1000C 75

Figure 5-15 XRD pattern for poultry litter ash at different temperature 77

Figure 5-16 Volatile Matter vs. Residence Time for Willow 79

Figure 5-17 Fixed Carbon vs. Residence Time for Willow 80

Figure 5-18 Ash Contents vs. Residence Time for Willow 80

Page 12: Torrefaction and Pelletization of Different Forms of

xii

Figure 5-19 % of Mass Yield and Energy Yield vs. Temperature for

Willow(Different Oxygen % and 45 minutes residence time)

82

Figure 5-20 Energy Density variations with Temperature and residence

time

82

Figure 5-21 Heating Value vs. Residence Time for Willow at 0%

Oxygen

83

Figure 5-22 Heating Value vs. Residence Time for Willow with 2.4%

Oxygen

84

Figure 5-23 % of Moisture absorption vs. temperature of torrefied

biomass

86

Figure 5-24 % of Moisture uptake vs. temperature of Willow Pellets 87

Figure 5-25 Pressure for making Pelletization and Force for Breaking

Pellets

90

Figure 5-26 SEM for Willow at 800°C, 900°C and 1000°C Ash 93

Figure 5-27 XRD pattern for Willow ash at different temperature 95

Figure 5-28 Volatile Matters vs. Residence Time for Oats 97

Figure 5-29 Fixed Carbons vs. Residence Time for Oats 98

Figure 5-30 Ash Contents vs. Residence Time for Oats 98

Figure 5-31 % of Mass Yield and Energy Yield vs. Temperature for

Oats (Different Oxygen % and 45 minutes residence time)

100

Figure 5-32 Energy Density variations with Temperature and residence

time

100

Figure 5-33 Heating Value vs. Residence Time for Oats at 0% Oxygen 101

Page 13: Torrefaction and Pelletization of Different Forms of

xiii

Figure 5-34 Heating Value vs. Residence Time for Oats with 2.4%

Oxygen

102

Figure 5-35 % of Moisture absorption vs. temperature of torrefied

biomass

103

Figure 5-36 % of Moisture uptake vs. temperature of Oats Pellets 105

Figure 5-37 Pressure for making Pelletization and Force for Breaking

Pellets

106

Figure 5-38 SEM for Oats at 800°C, 900°C and 1000°C Ash 109

Figure 5-39 XRD pattern for Oats ash at different temperature 110

Figure 5-40 Comparative Study of Ultimate Analysis of different

Biomass

114

Figure 5-41 Comparative Study of Ash Fusion Temperature of different

Biomass

115

Figure 5-42 Comparative Study of Elemental Analysis of different

Biomass

116

Figure 5-43 Comparative Study of Mass Yield of Poultry Litter,

Willow and Oat Pellets at different temp with 45 minutes residence

time

117

Figure 5-44 Comparative Study of Energy Yield of Poultry Litter,

Willow and Oat Pellets at different temp with 45 minutes residence

time

118

Figure 5-45 Comparative Study of Heating Values of different Biomass

with temp

119

Page 14: Torrefaction and Pelletization of Different Forms of

xiv

Figure 5-46 Comparative Study of Hydrophobicity of Poultry Litter,

Willow and Oat at different temp with 45 minutes residence time

120

Figure 5-47 Comparative Study of Storage Behavior of Poultry Litter,

Willow and Oat Pellets at different temp with 45 minutes residence

time

121

Figure 5-48 Comparative Study of Elemental Analysis of Poultry

Litter, Willow and Oat Pellets at different temp with 45 minutes

residence time

122

Figure 5-49 Comparative Study of Ash Fusion Temperature of Poultry

Litter, Willow and Oat Pellets at different temp with 45 minutes

residence time

123

Page 15: Torrefaction and Pelletization of Different Forms of

xv

NOMECLATURE

Mass (daf) of torrefied biomass

Mass (daf) of raw biomass

Higher Heating Value of torrefied biomass

Higher Heating Value of raw biomass

EY Energy Yield

ME Mass Yield

EDR Energy Density Ratio

MC Moisture Contents in %

TT Torrefaction Temperature in ᴼC

RT Residence Time in min

Page 16: Torrefaction and Pelletization of Different Forms of

1

Chapter I: Introduction

1.1 Introduction

Secure supply of sustainable and environmentally friendly energy for future

generations is a global concern these days. “Climate change”, "Global Warming" and

“Ice melting from Artic” are some of the familiar terms we hear on a daily basis. World

events such as the delay in building an oil pipeline between Canada and the United States

of America and the continuing uncertainty in the Middle East remind us of the fragile

supply of energy. Our present world is still so much dependent on fossil fuels,

while looking for other energy options. Processed-biomass has been identified as one of

many options that can significantly contribute to the present global energy requirements.

During photosynthesis, plants convert light (solar radiation) into chemical energy, which

is stored in the form of biomass. Although biomass comes in many forms, wood and

herbaceous biomasses are the main forms of biomass for energy production. Biomass,

traditionally used by mankind in the form of firewood, was the only source of energy that

fulfilled all the energy needs of early man. Of late, global warming concerns and politics

over fossil fuels have again fuelled global interest on our traditional energy source,

“biomass”. New technologies like torrefaction and densification can help in efficient

utilization of biomass, and in reducing the emission of greenhouse gases from burning

fossil fuels. Still, almost 1.5 billion people from the developing world rely on

unprocessed biomass to meet their present energy needs. Processed or treated biomass, in

solid or liquid form, possesses combustion characteristics similar to fossil fuels, and

hence can be used for electricity generation, steam generation or thermal energy

generation by direct combustion, gasification or co-firing. Gasification of biomass

Page 17: Torrefaction and Pelletization of Different Forms of

2

generates high-grade combustible gases such as CO, H2 and methane. Anaerobic

digestion produces biogas from biomass.

Canada has access to abundant forest and agricultural land that serve as great

sources for biomass feedstock to generate sustainable energy production. Biomass can be

converted into heat and electricity by means of thermo-chemical (combustion,

gasification, pyrolysis, and liquefaction) and bio-chemical (anaerobic digestion)

processes (Linghong et al., 2010). The most common forms of conversions are the

thermal processes. Helwig et al.(2002) evaluated the biomass inventory in Canada and

reported that approximately one million oven dry tons per year (odt/y) of cereal straw and

three million odt/y of corn stover are available in eastern Canada and in addition, the

gross energy potentials of these residues is approximately 92 million gigajoules per year.

In Canada, approximately 4.7% of national primary energy for 2006 was derived from

the conversion of renewable biomass and wastes. This fraction is projected to increase to

6–9% over the next 20 years (Douglas, 2009). This projection is low compared to other

industrial nations but it is encouraging.

Despite the tremendous popularity gained by biomass energy in recent

years, the fraction of its utilization in producing energy remains insignificant in the

overall source of energy production in Canada (Douglas, 2009). This can be due to

several factors, including the limitation associated with its properties (Bridgeman et. al.,

2008). The variations in biomass feedstock properties cause several challenges during the

conversion process; these include: low heating value, high moisture content, hygroscopic

nature, excess smoke during combustion, low energy density, low combustion efficiency,

and high ash contents. Torrefaction, a thermal pretreatment process of biomass, has been

Page 18: Torrefaction and Pelletization of Different Forms of

3

proved to improve the combustion properties of biomass (Pimchuai et al., 2010;

Bridgeman et. al., 2008; Bergman, 2005, Deng et al, 2009). Torrefaction is the

thermochemical treatment of biomass at 200 to 300°C, under atmospheric conditions, but

in the absence of oxygen. During the process the biomass partly decomposes, giving off

various types of volatiles. The final product is the remaining solid, which is often referred

to as torrefied biomass, or torrefied product when produced from agricultural or forest

biomass product. Typically, 70% of the mass is retained as a solid product, containing

approximately 90% of the initial energy content. The remaining 30% of the mass is

converted into torrefaction gases, but contain only approximately 10% of the energy

content of the original biomass. Hence a considerable energy densification can be

achieved, typically by a factor of 1.3 on mass basis. This study points out one of the

fundamental advantages of the process, which is the high transition of the chemical

energy from the feedstock to the torrefied product, while concurrently the fuel properties

are improved.

Figure1.1 Mass and Energy and Energy Balance of a Typical Torrefaction Process

(Acharya et al., 2012)

Torrefaction

(200-300°C)

Mild

Pyrolysis

Torrefied Gas as

Loss

30% Mass+

10% Energy

Biomass Feed Stock

input

100% Mass+

100% Energy

Torrefied

Biomass

70% Mass+

90% Energy

Page 19: Torrefaction and Pelletization of Different Forms of

4

1.2 Objectives

It is well known that biomass can be one of the many sources of renewable

energy, and the torrefaction process improves their combustion properties significantly.

Large reduction in the consumption of fossil fuels and development of a sustainable

energy system will require commercial scale production of renewable fuels like torrefied

biomass. However, no commercial operation of torrefaction exists yet. Furthermore only

limited studies are available on agricultural based biomass (Pimchuai et al.,2010;

Tumuluru et al., 2011) and only a few works assessed the commercial applicability of

torrefaction in power plants and the logistics of its impacts on the entire bioenergy supply

chain (Uslu et al., 2008). Hence, this work will expand the torrefaction studies of

different Ontario based biomass feedstock and complete the following objectives by

conducting different experiments and result analysis:

i) Physio-chemical characterization of biomass samples before and after

torrefaction.

ii) Optimization of torrefied conditions based on hydrophobicity

iii) Investigation of pelletization potential before and after torrefaction

iv) Ash analysis of biomass at different combustion temperature.

To achieve the above objectives, the following actions are undertaken: a)

Proximate, Ultimate and Elemental Analysis, and Ash fusion temperature analysis to

characterize the chemical composition of the agricultural based biomass samples with

determination of their respective calorific values prior to and after torrefaction; b) study

of the hydrophobic behavior of torrefied biomass with respect to treatment conditions

Page 20: Torrefaction and Pelletization of Different Forms of

5

(torrefaction temperature and residence time); c) study of the Higher Heating Value of

biomass at different temperature, residence time and Oxygen concentration d) study of

the stress analysis during and after pelletization of torrefied biomass; e) SEM and XRD

analyses of ash formed at different temperatures.

1.3 Organization of the Thesis

The summary of the organization of this thesis is as follows:

Chapter II presents a review of literature on biomass, torrefaction, pelletization, moisture

uptake, ash analysis. It also discusses the available technologies of different torrefiers,

economic and environmental impacts in Canada. At the end of the chapter, it explains in

brief about the scanning electromagnetic microscopy and X-ray diffraction.

Chapter III explains the methodology used to achieve the objectives of the study and

explain in tabular and flow chart form about the flow of experimental procedures.

Chapter IV presents the experimental setup for torrefaction, pelletization, moisture

uptake, hydrophobicity. It also explains the procedures which are taken for proximate,

ultimate, ash analysis. Ash analyses are carried out by SEM and XRD.

Chapter V presents the results and analysis observed during the conduction of different

experiment on torrefaction, pelletization, hydrophobicity, moisture uptake and ash

analysis. Detail comparative study of all the studied results are also presented in this

chapter.

Chapter VI presents conclusions found from this study and recommendation for further

study.

At the end, list of the references and appendixes are appended in the thesis.

Page 21: Torrefaction and Pelletization of Different Forms of

6

Chapter II: Literature Review

2.1 Overview

The world economy is slowly but steadily transitioning from non-renewable fossil

fuel energy to renewable and sustainable energy. Energy from biomass can be considered

carbon-neutral because of the carbon cycle. Unlike fossil fuels, biomass is a renewable

source of energy that can be replenished; utilizing biomass for energy production does

not result in net greenhouse gas emission to the atmosphere. A good illustration is wood,

which is obtained from trees. Trees absorb sunlight and CO2 from the atmosphere during

photosynthesis to make cellulose from sugars; consequently, the cellulose, which

contains stored chemical energy, releases this energy as heat when combusted, and the

CO2 liberated as off-gas is equivalent to the amount absorbed during photosynthesis

process. Hence, biomass can be greenhouse gas emission-neutral. Increasing the use of

biomass for energy application thus helps to reduce greenhouse gas (GHG) emission. The

word “biomass” originally meant the total mass of living matter within a specified

environmental area, but more recently, it also describes plants products, vegetation, or

agricultural residues used as an energy source. Tumuluru et al. (2010) also defined

biomass materials as a composite of carbohydrate polymers with a small amount of

inorganic substance and low molecular weight extractable organic elements. Others

defined biomass as a biological or organic material, which can be used as source of

renewable energy after performing certain processes. It can also be classified as carbon-

based material, consisting of organic molecules containing hydrogen, oxygen, nitrogen

and small quantities of atoms containing alkali, alkaline earth and heavy metals (Biomass

Energy Centre, UK). Energy contents of biomass are collected from the sunlight and

Page 22: Torrefaction and Pelletization of Different Forms of

7

stored in the form of chemical energy. Stored energy from the biomass can be converted

in to heat energy at any suitable time with certain transformation processes.

Biomass feedstock can be from agricultural crops, trees or crops grown for energy

production, wood residues, wood wastes, agricultural residues, animal and human wastes,

and municipal wastes.

2.1.1 Fuel Characteristics of Biomass

Fuel characteristics of biomass depend on the origin and types of biomass which

are categorized by their physical and chemical properties. Sizes, shapes, specific

capacity, thermal conductivity, moisture content, bulk density, grindability and porosity

characterize the physical properties of biomass. The chemical properties are determined

by proximate and ultimate analyses and thermal decomposition. The ultimate analysis of

biomass delivers the elemental structures of biomass by weight percentage in the form of

carbon (C), hydrogen (H), Oxygen (O), nitrogen (N), sulfur (S), chlorine (Cl), and ash

elements such as Sodium (Na) and Potassium (K) compounds etc. However, the

proximate analysis provides the percentage weight of fixed carbon (FC), moisture

contents (MC), ash contents (AC), and volatile matter (VM) in biomass. The methods of

performing these analyses are dependent on the standards in ASTM, ISO etc. These

chemical compositions change with oxygen concentration during torrefaction,

temperature and residence time for all types of biomass fuels. Cellulose, hemicelluloses,

lignin, lipids, proteins, simple sugars, starches also influence the combustion process of

biomass. The concentration of each class of compound differs depending on species,

Page 23: Torrefaction and Pelletization of Different Forms of

8

nature of plant tissue, phase of development, and growing environments (Jenkins et al.,

1998).

2.1.2 Biomass Components

The plant cell wall is the strong, usually flexible but sometimes rigid layer that

gives structural support and protection from external mechanical and physical forces. The

major constituents of the primary cell wall are cellulose, carbohydrates, hemicelluloses

and pectin (Tumuluru at el., 2011). Biomass consists of three main polymeric

constituents: Hemicellulose, cellulose, and lignin and generally they cover respectively

20–40, 40–60, and 10-25 wt% for lignocellulosic biomass (McKendry, 2002; Yang et al.,

2005). Pretreatment of biomass changes its physical and chemical structure which makes

more suitable to use for energy application. Figure 2-1 shows the biomass structure and

pretreatment effect on lignocellulosic components.

Cellulose is a linear polymer of biomass that makes up about 45% of the dry weight of

wood, is composed of D-glucose subunits linked together to form long chains (elemental

fibrils), which are further linked together by hydrogen bonds and Van der Waals forces.

The cellular fiber formed by several micro-fibrils coming together can either be

crystalline or amorphous (Pe´rez et al., 2002). Furthermore, cellulose is a high molecular

weight polymer that makes up the fibers in lignocellulosic materials and its degradation

starts anywhere from 240°–350°C because of high resistance of its crystalline structure to

thermal depolymerization owns to its strength (Mohan et al., 2006; Tumuluru et al.,

2010). The waters held in the amorphous regions of the cellulosic wall rupture the

Page 24: Torrefaction and Pelletization of Different Forms of

9

structure when converted into steam as a result of thermal treatment (Tumuluru et al.,

2010).

Figure 2-1: Structure and Pretreatment effect on Biomass (Source: Tumuluru at el.,

2011)

Hemicellulose is a complex carbohydrate polymer with a lower molecular weight

than cellulose and makes up 25–30% of total dry weight of wood. It consists of D-xylose,

D-mannose, D-galactose, D-glucose, L-arabinose, 4-O-methyl-glucuronic, D-

galacturonic and D-glucuronic acids (Pe´rez et al., 2002). The principal component of

hardwood hemicellulose is glucuronoxylan whereas glucomannan is predominant in

softwood (Pe´rez et al., 2002). In contrast to cellulose, hemicelluloses are easily

hydrolysable polymers and do not form aggregates. It consists of shorter polymer chains

with 500–3000 sugar units as compared to the 7,000–15,000 glucose molecules per

polymer seen in cellulose (Tumuluru et al., 2010). Thermal degradation of hemicellulose

occurs between 130° – 260°C, with the majority of weight loss occurring above 180°C

Page 25: Torrefaction and Pelletization of Different Forms of

10

(Mohan et al., 2006; Tumuluru et al., 2010). Hemicellulose produces less tars and char

due to its low degradation temperature range compared to that of the cellulose (Tumuluru

et al., 2010).

Lignin along with cellulose is the most abundant polymer in nature (Pe´rez et al.,

2002). Lignin is an unstructured and highly branched polymer that fills the spaces in the

cell wall between cellulose, hemicellulose, and pectin components (Tumuluru et al.,

2010). It is covalently bonded to hemicellulose and thereby exhibits mechanical strength

on the cell wall. It is relatively hydrophobic and aromatic in nature and decomposes

between 280 and 500°C when subjected to a thermal treatment (Tumuluru et al., 2010;

Mohan et al., 2006). Lignin is difficult to dehydrate and thus converts to more char than

cellulose or hemicelluloses.

Figure 2-2 Carbon Cycle in Biomass (Source: Uslu et al., 2005)

Plants

(Biomass)

Photosynthesis

Sun

Light

Atmo

spher

e

Soil

Biofuel

Energy

Production

(Combustion)

Water +

Minerals

CO2

Processing

Page 26: Torrefaction and Pelletization of Different Forms of

11

2.2 Concept of Torrefaction

Different literatures have defined torrefaction in different terms. Acharya et al.

(2012) has defined torrefaction as a thermal pretreatment process in which isothermal

pyrolysis of biomass takes place at a temperature range of 200°C to 300 °C for a

reasonable residence time, with minimum oxygen concentration. Although almost all the

definitions reveal similarities in terms of processes, the operating temperature range

differs from research to research depending on the types and categories of biomass that

were studied. Sadaka and Negi (2009); Bergman et al.(2005); Rousset et al.(2011) and

Mani (2009) stated the torrefaction temperature range from 200°C - 300°C; Prins et al.

(2006a); Pimchuai et al.(2010) gave temperature between 230°C - 300°C; meanwhile,

Arias et al.(2008) stated temperature range between 220°C - 300°C, and Chen and Kuo

(2011); Zwart et al.(2006) stated temperature range between 225°C - 300°C. Most of the

research has shown that biomass reveal different performance to thermal treatment owing

to their varieties, origin and properties (Bridgeman et al., 2008); hence, the

commencement of biomass decomposition hangs on the biomass types.

According to Bergman et al.(2005), torrefaction is the thermal pretreatment

technique conducted at temperature range of 200°C and 300°C under an inert

environment and reasonably small residence time and slow heating rate less than

50°C/min (Berman et al., 2005, Walton and Van, 2011). The torrefaction process

encompasses the fragmentation of biomass during which various forms of volatiles are

formed and the resulting output is in the form of a solid fuel usually known as torrefied

product or torrefied fuel (Bergman et al., 2005; Pimchuai et al., 2010; Bridgeman et al.,

2008, Tumuluru at el., 2011).

Page 27: Torrefaction and Pelletization of Different Forms of

12

Usually, almost all types of biomass contain hemicellulose, lignin and cellulose,

or lignocellulose in general. The thermal disintegration of biomass during torrefaction

causes various reactions to occur through their lignocellulosic configuration (Mosier et

al., 2005). The disintegration process was well studied by Bergman et al. (2005) as shown

in figure 2-3. At lower torrefaction temperatures, disintegration takes place in the

hemicellulose structure by means of a limited devolatilisation and carbonization;

meanwhile, in the lignin and cellulose structure a minor disintegration is observed. Figure

2-4 shows that hemicellulose undergoes extensive thermal disintegration in the

temperature range of 200°C to 300°C while only partial devolatilisation and

carbonization occurs in the lignin and cellulose structure (Bergman et al., 2005). It can

also be noted that the conversion from one decomposition regime occurs at a narrow

temperature range for hemicellulose while the conversion for lignin and cellulose occurs

over a wider temperature range. Hence, it can be concluded that hemicellulose is the most

reactive polymer constituent of biomass and is attributed to the substantial mass loss in

biomass during torrefaction (Bergman et al., 2005; Chen and Kuo, 2011; Sadaka and

Negi, 2009; Acharjee et al., 2011).

After the maximum moisture loss at temperature 100°C to 120°C, the significance

weight loss of biomass is achieved because of depolymerization and partial

devolatilisation of the hemicellulose during torrefaction process at a temperature range of

200°C to 300°C. However, due to minimum depolymerization and devolatilisation

reactions occurring in lignin and cellulose at this temperature range, maximum energy

content is retained in the torrefied biomass.

Page 28: Torrefaction and Pelletization of Different Forms of

13

Figure 2.3: Decomposition regimes of lignocellulosic material during thermal

treatment (Source: Uslu et al., 2005)

Physical and Chemical properties of biomass are improved significantly after the

thermal treatment torrefaction, resulting in improve combustible properties. The

combustion properties including physical and chemical properties of torrefied biomass

also rely on the biomass properties, torrefaction temperature and residence time.

Following are the major characteristics of the torrefied products for biofuel application:

a. Hydrophobic behavior: Torrefied biomass has hydrophobic properties, i.e.

repels water, and when combined with densification makes bulk storage in

Page 29: Torrefaction and Pelletization of Different Forms of

14

open air feasible. Torrefied Biomass has hydrophobic characteristics owning

to the destruction of is O-H bond structure, hence making it incapable to retain

or absorb moisture. Although standardized test exists yet to come on testing of

hydrophobic strength of Biomass, Pimchuai et al. (2010) demonstrated

hydrophobic test of torrefied biomass in comparison with raw biomass and

confirmed that torrefied biomass is hydrophobic in nature.

b. Elimination of biological activity: All biological activity is stopped, resulting

in total elimination of biological decomposition like rotting and reduced risk

of fire (Kung et al., 2009).

c. Improved grindability: Torrefaction leads to improved grindability of

biomass. This leads to more efficient co-firing in existing coal fired power

stations or entrained-flow gasifier for the production of chemicals and

transportation fuels. TB is more brittle owing to its higher C/H and C/O ratios,

hence provides enhanced pulverization characteristics and requires far less

energy for grinding compared to that of raw biomass (Bergman et al., 2005).

d. Markets for torrefied biomass: Torrefied biomass has added value for different

markets. Biomass in general provides a low-cost, low-risk route to lower CO2-

emissions. When high volumes are needed, torrefaction can make biomass

from distant sources price-competitive as the denser material is easier to

transport and store.

e. Large scale co-firing in coal fired power plants: Torrefied biomass results in

lower handling costs. Torrefied biomass enables higher co-firing rates;

Product can be delivered in a range of LHVs (20 – 25 GJ/ton) and sizes

Page 30: Torrefaction and Pelletization of Different Forms of

15

(briquette, pellet). Co-firing by torrefied biomass with coal leads to reduction

in net power plant emissions.

f. Residential/decentralized heating: Relatively long distance transport of

biomass in the supply chain on wheels makes biomass expensive. Increasing

volumetric energy density decreases cost; Limited storage space increases the

need for higher volumetric density; Moisture content is important as moisture

leads to smoke and smell.

g. Higher heating value: The calorific value of torrefied biomass (TB) increases

with increase in treatment temperature and residence time (Bergman et al.,

2005) and this can be explained by the fact that TB has lost its moisture

content and its oxygen-carbon or hydrogen-carbon ratio reduces with

increasing temperature. The energy density increases with incremental

treatment temperature, which in turn increases the calorific value of TB

(Bergman et al., 2005).

h. Densification: Densification increases the bulk and volumetric density of

biomass. Hence, a combination of torrefaction and pelletization processes

produce torrefied pellets, which through pilot-scale experiments have shown

to have better storage properties than biomass pellets due to their

hydrophobicity. Torrefaction process causes dehydration that initiates and

propagates cracks in the lignocellulosic structure (e.g. wood), as result,

induces porosity and density changes. Increased porosity, due to more particle

voids, decreases particle size but inevitably increases the particle density and

bulk density. Generally, density varies in a different way depending on wood

Page 31: Torrefaction and Pelletization of Different Forms of

16

species during temperature treatment and the changes with respect to

torrefaction might not be very significant (Repellin et al., 2010).

i. Particle Sizes and Distribution of Torrefied Wood: Torrefied biomass produce

more uniform and smooth particle sizes compared to untreated biomass

because of their brittleness, which is similar to that of coal. This behavior is

supported by their lower energy consumption during grinding. In their

experiment to examine the particle size and particle size distribution of a

torrefied pine chips and logging residues, Phanphanich and Mani (2010)

found out that the mean particle size of ground torrefied biomass decreased

with increase in torrefaction temperature. Consequently, torrefaction of

biomass not only decreased the specific energy required for grinding but also

decreased the average particle size of ground biomass. Furthermore, they

concluded that the particle size distribution curves of torrefied biomass

produces smaller particles than that of untreated biomass and their results

were comparable to the studies by Mani et al.(2009),. Cumulative percent

passing curve also showed similar behavior for torrefied biomass.

j. Explosibility of Torrefied Wood: Generally, a process that involves the

handling of dust poses an explosion hazard and the severity increases when

the process operating pressure rises over the atmospheric conditions (Repellin

et al., 2010). Just like pulverized coal, fine particle biomass such as sawdust

generate carbon dust that is combustible; hence, increase in the concentration

of suspended particles that exist in a confined space of a processing equipment

(e.g. boiler) pose a high risk of explosion when exposed to an ignition source.

Page 32: Torrefaction and Pelletization of Different Forms of

17

Consequently, the intensity of explosion of dust particles increases with

increase in the combustible properties of the particles. Hence, torrefied

biomass has a higher dust explosibility than raw biomass and should be

handled with extreme care during plant operation. Although no study exists

that has compared the dust explosibility test of raw biomass to that of torrefied

biomass, it is reasonable to assume that a power plant, for example, should

incorporate dust explosion control measures when designing a plant that

processes pulverized torrefied biomass (wood) for energy supply etc.(Govin et

al., 1988, Ciolkosz and Wallace, 20011)

Additionally, a number of researchers have found that pelletization improves the

biomass bulk and volumetric density. Hence, a combination of torrefaction and

pelletization techniques yield torrefied pellets resulting in easier handling and storage

than raw biomass pellets due to their hydrophobicity and resistance to biodegradability

(Uslu et al., 2008; Kiel et al., 2008). The comparison in fuel characteristics and handling

behaviors of raw wood, wood pellets, torrefied wood pellets, coal and charcoal is

presented in table 2-1.

Page 33: Torrefaction and Pelletization of Different Forms of

18

Table 2-1: Summary of Torrefied pellets Properties versus Coal (Source:

Kleinschmidt, 2011)

Parameters Wood

Wood

pellet

Torrefied

Pellets

Coal

Moisture content (% wt) 30-40 7-10 1-5 10-15

Calorific Value (MJ/kg) 9-12 15-16 20-24 23-28

Volatiles (% db) 70-75 70-75 55-65 15-30

Fixed carbon (% db) 20-25 20-25 28-35 50-55

Bulk density (kg/m3) 200-250 550-750 750-850 800-850

Volumetric energy density

(GJ/m3)

2.0-3.0

7.5-10.4 15.0-18.7 18.4-23.8

Dust explosibility Average Limited Limited Limited

Hydroscopic properties Hydrophilic Hydrophilic Hydrophobic Hydrophobic

Biological degradation Yes Yes No No

Milling requirements Special Special Classic Classic

Handling properties Special Easy Easy Easy

Transport cost High Average Low Low

2.3 Torrefaction Process Methods

Technical steps of the torrefaction can be explained in four steps which are

chopping, drying, mild roasting or pyrolysis at 200ᴼC-300ᴼC and cooling. From the field,

biomass is collected and fed into the chopper, which cuts them into small and more

uniform particles. The chopped biomass then passes via the drying segment to eliminate

the moisture and then into the torrefaction reactor. The output from the torrefaction

Page 34: Torrefaction and Pelletization of Different Forms of

19

reactor is cooled down to room temperature. The moisture gas mixture released during

the drying process is composed of both condensable and non-condensable gases and

volatiles like CO2, CO, H2O, H2 and other organic elements (Kundra and Mujumdar,

2002). The greater the torrefaction temperature, the greater will be the incineration heat

of the volatile gas liberated during the process. The combustion heat of volatiles can be

reused as a supplement for the drying process. The resulting solid product, after the

complete devolatilisation of the biomass, is known as torrefied biomass or char (Bergman

et al., 2005). Careful regulation of torrefaction temperature and residence time is

necessary to the energy density and heating value of the product. The improved

combustion properties of torrefied biomass result in an attractive solid fuel for

combustion and gasification in thermal power plants.

Furthermore, the enhanced grindability of torrefied biomass makes it more

beneficial for pelletization, which enables storage, transportation, and co-combustion of

biomass with coal in existing coal power plants with no additional investment (Bergman

et al., 2005). The output product of the torrefaction process has high energy density and

enhanced overall fuel characteristics. This process involves high temperature treatment

i.e. overall efficiency of the process solely depends on the use of the heat supplied during

the treatment. The overall efficiency can be enhanced by reprocessing the excess heat

liberated from the process.

Page 35: Torrefaction and Pelletization of Different Forms of

20

Figure 2-4: Basic principle of torrefaction Process (Source: Tumuluru et al., 2011)

During the torrefaction process, biomass experiences a series of disintegration

reactions that cause the discharge of gaseous products such as volatile organic

compounds (VOC’s) such as CO, CO2, H2O, H2, and organic volatiles. As a result C, H,

O composition of the biomass is changed, and the ratio of H/C or O/C is decreased as

torrefied biomass loses its hydrogen and oxygen in higher proportion compared to carbon

(Prins et al., 2006; Bergman et al., 2005). The breakdown of biomass polymer structure

throughout torrefaction results in the destruction of its hydroxyl (OH) group and making

it incapable to form hydrogen bond with water and hence, fails in its tendency to absorb

water (Bergman et al. 2005; Sadaka and Negi, 2009).

2.4 Classification of Reactors

Current torrefaction technologies are categorized by their reactor designs. Table

2.2 is a list of some main technology providers with different types of torrefaction

Page 36: Torrefaction and Pelletization of Different Forms of

21

reactors. Each of the torrefaction technologies has its distinct style of heat transfer and

gas-solid or solid-solid mixing pattern in the reactor. These reactors could be generally

categorized into two major classes: a) Directly heated and b) Indirectly heated. In directly

heated reactors, biomass is heated by direct contact with the heating media like hot gas,

hot solids, superheated steam or electromagnetic radiation, either inert or minimum

oxygen environment. In indirectly heated reactors biomass is heated across the wall of the

reactor without contacting the heat transporting medium with the biomass. Therefore, it is

comparatively easier to avoid the presence of oxygen in the reactor during torrefaction.

Based on the mode of gas-solid contacts, the torrefaction reactors can also be

classified into the following four types. a) Plug flow (unidirectional motion of gas and

solids), b) Partial back-mixed (gas is unidirectional but solids are back mixed) c)

Tumbling (solids tumbles or moves around in a rotating drum or cylindrical tunnel) and

d) Entrained (solids are transported by gas). The entrained flow reactor (Topell

technology for example) involves the transport of biomass particles at high velocity (50-

80 m/s) through stationery angled blades at temperatures up to 280ºC which gives a

reactor residence time of less than 5 minutes. In both ‘Belt conveyor’ and ‘Multiple plate’

technologies, biomass moves on surfaces at a defined rate while the heating medium (hot

flue gas, hot nitrogen or superheated steam) sweeps over them, providing heat to the

biomass by convection. The heating is therefore mixed convective type. Heating of

biomass particles is the key part of the torrefaction process. The transfer of heat to the

biomass particles could take place through one of the following means: Gas-particle

convection, Wall particle conduction, Electromagnetic heating, Particle-particle heating

and Liquid-particle heating (FGC, 2010).

Page 37: Torrefaction and Pelletization of Different Forms of

22

Using the above characteristics, torrefaction reactors can be broadly represented

by six generic types though some reactors could use combination of the basic features of

several generic reactors: i) Directly heated: Convective reactor, fluidized bed reactor,

hydrothermal reactor, microwave reactor, and ii) Indirectly heated: Rotating drum type

reactor, screw in stationery shaft type reactor. It may, however, be noted that some

reactors have a combination of these basic features but they are not listed separately.

Commonly used torrefaction reactors are described below:

a) Convective reactor: This is the most common generic type of reactor used for

torrefaction. Here hot gas flows past the biomass particles. The relative velocity between

the particles and gas drives the convective heating of biomass. The hot gas may be

completely inert or with a small amount of oxygen. In a fixed bed the particles remain

stationary while in a moving bed the particles move with respect to the reactor wall

(Marb and Vortmeyer, 1988). The wall of the reactor can be horizontal, vertical or

inclined. The particles may be moved either by gravity or by the force of a mechanical

device like augur. Particle flow through the reactor is unidirectional without back mixing.

The heat transfer is primarily through gas solid convection. Bergman et al. (2005)

estimates the biomass-to-gas heat transfer to be high in the range of 200 W/m2.K. Their

estimation of heating time (1 minute to heat 10x30x50 mm3 wood chips to 280ᴼC),

however, is much less than the 20 minutes measured in the present work for a wood

cylinder (25 diax75mm to 280ᴼC) in a convective reactor. Some convective type directly

heated torrefier uses a rotating drum where the biomass is heated directly by hot gases

passing through the rotating drum. In this case, the drum simply serves as a mixing

device while heat transfer takes place through gas-particle convection. Bergman et al.

Page 38: Torrefaction and Pelletization of Different Forms of

23

(2005) estimates heat transfer coefficient in such reactors to be low, in the range of 40

W/m2K.

b) Fluidized Bed: In this type of torrefier, inert gas is blown through a bed of

granular heat carrier solids in a way that the solids behave like a fluid. These hot

particles, in vigorously mixed and agitated state, easily heat up any fresh biomass

particles dropped amongst them (Basu P., 2006, Pipatmanomai S., 2011). The biomass

particles undergo torrefaction in a well-mixed state with uniform temperature

distribution. This system, therefore, ensures uniform product quality that is a problem

with many other reactors. Separation of heat carrier solids from torrefaction product and

entrainment of fine particles are some of the limitations of this technology. Here the

dominant mode of heat transfer is particle-to-particle heat transfer. Because of the high

degree of particle-particle mixing, the heat transfer in this reactor is very high. This work

measured a heating period of less than14 minutes for heating a wood cylinder (25 mm dia

x 75 mm long) to 280ᴼC. Though this type is not in common use, it can provide very

uniform quality of the torrefaction product. Rapid heating to the torrefaction temperature

could potentially increase the throughput of the reactor without affecting product quality.

c) Hydrothermal Reactor: Here the biomass is subjected to heating in high-

pressure water and thus it obviates the need for drying (Yan et al., 2009). The dominant

mode of heat transfer in hydrothermal reactor is that between hot water (fluid) and solid.

While this process has several potential advantages, the energy required for

pressurization and movement of large volume of biomass across pressure barrier poses

major practical difficulties. No commercial application of this torrefaction is in use at the

moment.

Page 39: Torrefaction and Pelletization of Different Forms of

24

d) Microwave: Microwave irradiation involves electromagnetic waves of

frequency in the range of 300 MHz to 300 GHz. Typical microwave ovens or microwave

reactors usually work at 2.45 GHz. The microwave irradiation produces efficient internal

heating by direct coupling of microwave energy with the molecules of biomass. The

electric component of electromagnetic microwave radiation causes heating by two main

mechanisms: dipolar polarization and ionic conduction (Leoneli and Mason, 2010). The

heating depends on the ability of the materials being heated to absorb microwaves and

convert it into heat. Metals for example reflect microwave, while biomass absorbs it

(Miura et al., 2004). This type of reactor is thus different from other directly heated

reactors, where biomass particles are heated externally, which means heat from the heat

carrier (gas, solid, liquid or reactor wall) first arrives at the surface of the biomass particle

and then it is conducted into the interior of the biomass. In a microwave reactor, the

biomass particles are heated from within (Salem and Ani, 2011). Other reactors heat

biomass through conductive heating by an external source, where heat received on the

biomass surface is conducted inside (De la Hoz et al., 2005). Biomass being poor thermal

conductor such heating is less efficient. In microwave processing, heating is internal. The

heating is, thus, volumetric rather than surface heating. Here no other heat transfer

medium, hot wall, particle or gas is needed.

e) Entrained flow reactor: Here ground biomass particles are entrained in high

velocity jet of hot gases. The Torbed process is an example of entrained reactor. Torbed

reactor heats the biomass particles to a temperature of up to 350ᴼC within a relatively

very short time of 90 sec (Michel et al., 2011) that greatly increases the throughput of the

reactor. A major characteristic of such reactor is high heat and mass transfer rates and the

Page 40: Torrefaction and Pelletization of Different Forms of

25

absence of moving parts in the reactor. This process is similar to low temperature flash

pyrolysis. Fast heating and short residence time increases the volatile yield and reduces

char production during heating in the absence of oxygen (Michel et al., 2011).

Torrefaction process attempts to maximize char production while minimizing volatiles to

retain higher fraction of the biomass energy in the torrefied solids. From this standpoint

this type of reactor may have lower energy yield. This inference however cannot be

verified due to the absence of any research paper in the open literature with relevant

details on this process.

f) Rotating drum: The indirectly heated rotating drum tumbles the biomass in an

environment of inert gaseous medium. Here, heat is transferred from the hotter drum wall

to the biomass particles. This type of reactor has two major advantages: first the heating

medium does not have to be oxygen free and then the volatiles released are not diluted by

the gas passing through it, so it can be combusted to supplement the thermal load of the

reactor. Heat transfer from the wall to the biomass particles is the controlling factor in

such reactors.

g) Screw or stationary shaft: Here the torrefaction reactor (circular or rectangular

cross-section) is stationary, and it could be vertical, horizontal or inclined. A rotating

screw churns and moves the biomass through the reactor to enhance the heat transfer

between the wall and the bulk of the biomass and at the same time moving the biomass

along its length (Shu-de et al., 1996). To avoid direct contact with oxygen carrying hot

gases, the biomass is heated indirectly by the outer heated wall. Some designs may,

however, have holes for the products of torrefaction to escape. Biomass may also be

heated by the screw heated from inside the reactor (Waje et al., 2007).

Page 41: Torrefaction and Pelletization of Different Forms of

26

Table 2.2: Comparison of Potential Torrefaction Reactor Technologies

(Kleinschmidt CP, 2011)

Torrefiers

Technology

Mode of

Heading

Status Criteria

Rotary drum

reactor

Direct Proven Technology, Minimum Heat

Transfer, High heating rate, medium

temperature control, good residence time

control, excellent heating integration,

enhanced mixing, large size tolerance, high

moving parts, good fouling,, little scaling

problem

Fluidized Bed

Reactor

Proven Technology, Enhanced Heat

Transfer, High heating rate, medium

temperature control, medium residence

time control, excellent scalability, excellent

heating integration, excellent uniform

heating materials, enhanced mixing

Moving Bed

Reactor

Direct Under development, Enhanced Heat and

Transfer, High heating rate, medium

temperature control, good residence time

control, excellent heating integration,

enhanced mixing, Good fouling

Screw conveyor Direct Indirect Proven technology, Enhanced Heat and

Transfer, High heating rate, medium

temperature control, good residence time

control, excellent heating integration,

enhanced mixing, large size tolerance, high

moving parts, best fouling and scaling.

Microwave Direct Indirect Under R&D, Enhanced Heat and Transfer,

High heating rate, Good temperature

control, good residence time control

Multiple Hearths

Furnace

Direct Proven Technology, Enhanced Heat and

Transfer, High heating rate, medium

temperature control, good residence time

control, excellent heating integration,

enhanced mixing, large size tolerance, high

moving parts, perfect scaling and best

scalability

2.5 Commercial Application of Torrefaction in Canada

The first and only commercially operated torrefaction plant was built in the 1980s

by the French company, Pechiney, to yield 12000-ton/acre of torrefied wood as a coke

Page 42: Torrefaction and Pelletization of Different Forms of

27

substitute for the metallurgical industry. The plant was dismantled in the 1990s for

economic reasons (Bergman et al., 2005). Although the Pechiney demonstration plant

was considered state-of-the-art technology, Bergman et al.(2005) questioned its

feasibility for large-scale production; hence, they recognized its commercial failure partly

to its small scale pilot plant and large residence time during torrefaction, leading to loss

of energy efficiency. As a result, it drove the operating costs to an unmaintainable level.

However, the failure of Pechiney plant still played a positive role in thoughtfully

evaluating the technologies that will deliver an optimized torrefaction process at

minimum costs, while fulfilling the feedstock variability and the end user quality

requirements.

Several technology developers from the EU and the United States have invested

significantly in torrefaction technology. Canada still lags behind in its efforts and

commitment in bringing the technology to the market. The ECN (Energy Centre of the

Netherlands) is currently working on assessing the viability of commercial application of

torrefaction by up-scaling the concept of the Pechiney practice to enhance the efficiency

and sustainability of the operation (ECN, 2010). In June 2010, ECN commissioned a pilot

installation of a torrefaction plant in 2008 with a capacity of 50 kg of biomass per hour.

The plant reportedly achieved a stable process operation for 100 hours. This indicates a

prospect in the installation of a commercial scale torrefaction. Currently, ECN is

developing a demonstration installation with a capacity of nearly 5 tons of biomass per

hour.

Canada has approximately 979.1 million hectares of land. Out of the 397.2

million hectares of treed land, forested land covers 347.7 million hectares with 7.8

Page 43: Torrefaction and Pelletization of Different Forms of

28

million and 41.8 million hectares classified as “other land with tree cover” and “other

wooded land” respectively. In 2004, Canadian sawmills produced 83.5 million cubic

meter of lumber, 47% of which was from British Columbia followed by 24% from

Quebec, and 10% from Ontario. Although Lumber production has slowed down in recent

years due to the US-led recession (Douglas, 2009), two main sources of bioenergy

feedstock in Canada are agriculture residues and wood-based products. Wood waste from

mills, residual biomass after harvest, or from stands grown specifically for biomass

production are the sources for wood biomass (Wood and Layzell, 2003). The wood

waste from mills is widely used among the three categories for pellet production and

other biofuel applications.

BIOCAP Canada published results of the analysis of mainly agriculture and

forestry based resources to evaluate the availability of bio-resources obtainable in Canada

(Wood and Layzell, 2003). On the basis of bioenergy stock, 245 million hectares of

timber forest in Canada have a biomass carbon stock of about 15.8 billion tons of carbon,

totaling 566 exajoules or 69 times Canada’s annual energy demand met by fossil fuel; on

the basis of annual harvest, the annual energy content of the biomass crop in Canada

amounts to 5.1 exajoules, which is 62% of the energy recovered from fossil fuel ignition;

and on the basis of biomass residue, about 60 million tons of carbon streams may be

considered “available” feedstock for a bio-based economy and this is conservatively

expected between 1.5-2.2 exajoules of energy contents per year, which is 18 – 27% of the

energy that Canada received from fossil fuels in 2000 (Wood and Layzell, 2003).

Canada retains substantial benefits in bioenergy from its arable land and forested

areas. About 1,866 megawatts of biomass power capacity is currently available in Canada

Page 44: Torrefaction and Pelletization of Different Forms of

29

(Center for Energy, 2011). In 2007, 11.856 million tons of Municipal solid waste (MSW)

amounting to about 211-187 MW electricity generating capacity was available in Canada

without considering the thousands of MW of potential energy lost in the sewage. Biomass

feedstock extends from forest and agricultural residues to poultry litters and MSW. So, an

effective implementation commercially operable torrefaction plants can revolutionize the

green energy business potential leading the economy towards sustainable energy systems

and helping in minimizing the utilization of fossil fuels. Processed or unprocessed

biomass can be used for the replacement of coal in future.

2.6 Available Technologies for Torrefaction

Many researchers from the universities and government sectors are involved in

the torrefaction research, but a commercial scale torrefier has still not been developed. In

parallel, a number of private institutions are also involved in the commercial

development of the technology. In fact, majority have either given up their efforts or

delayed the development due to a lack of investment, while others are far behind their

forecasted operation date. On the other hand, the investors are not enthusiastic to invest

into torrefaction technology for fear of financial risk due to technological uncertainty.

Table 2-3 illustrates a number of torrefaction developments, including the companies,

suppliers and projected date of execution. Further studies have shown that a majority of

these projects is either being delayed, cancelled, or status not known.

Page 45: Torrefaction and Pelletization of Different Forms of

30

Table 2-3: Overview of Torrefaction Projects (Source: Kleinschmidt CP, 2011)

Company Demo

Technology Supplier Locations

Prod.

Capacit

y (t/a)

ESD1 of

Operation

3RAgrocarbon,

Hungary

Rotary Kiln

(3R Pyrolysis

Biochar)

Unknown Unknown Unknow

n Unknown

4Energy Invest.

(BE) Unknown

Stramproy

Green Tech.

(NL)

Amel (BE) 40,000 Q4 2010

Agri-Tech

Producers LLC

(US/SC)

Belt

Conveyor

Kuster

Zima

Corporation

(US/SC)

Unknown Unknow

n 2010

Andritz

(Austria) Unknown Unknown Unknown 50,000 Unknown

Atmosclear

(CH) Rotary Drum CDS (UK)

Latvia,

New

Zealand,

US

50,000 Q4 2010

BioEnergy

Development

(SWE)

Rotary Drum

Unknown

Ö-vik

(SWE)

25,000 –

30,000

2011/2012

Biogreen

Energy (FR)

Screw

Conveyor ETIA (FR) Unknown

Unknow

n Unknown

Biolake BV

(NL)

Screw

Conveyor Unknown

Eastern

Europe

5,000 –

10,000 Q4 2010

CDS (UK) Rotary Kiln Unknown Unknown Unknow

n Unknown

CMI (NESA)

Multiple

Heath

Furnace

Unknown Unknown Unknow

n Unknown

EBES AG (AT) Rotary Drum Andritz (T) Frohnleite

n (AU) 10,000 2011

ECN (NL) Moving Bed Unknown Unknown Unknow

n Unknown

FoxCoal B.V.

(NL)

Screw

Conveyor Unknown

Winschote

n (NL) 35,000 2012

Integro Earth

Fuels, LLC

(US/NC)

TurboDryer Wyssmont

(US/NC)

Roxboro,

NC 50,000 2010

New Earth

Renewable

Energy Fuels,

Fixed

Bed/Pyrovac

Pyrovac

Group

(CA/QU)

Unknown Unknow

n Unknown

Page 46: Torrefaction and Pelletization of Different Forms of

31

Inc. (US/WA)

Rotawave Ltd.

(UK)

Microwave

Heating

Group’s

Vikoma

Terrace ,

B.C,

Canada

110,000 Q4 2011

Stramproy

Green

Investment B.V.

(NL)

Oscillating

belt Conveyor

Stramproy

Green Tech.

(NL)

Sreenwijk

(NL) 45,000 Q3 2010

Thermya (FR) Moving Bed Lantec

Group (SP)

San

Sebastian

(SP)

20, 000 2011

Topell Energy

B.V. NL Torbed

Torftech Inc

(UK)

Duiven

(NL) 60,000 Q4 2010

Torr-Coal B.V. Rotary Drum Unknown

Dilsen-

Stokkem

(BE)

35,000 Q3 2010

Torrefaction

System Inc.

(US)

Unknown

Bepex

Internationa

l (US/MN)

Unknown Unknow

n 2013

Vattenfall

(SWE) Moving Bed Unknown Unknown

Unknow

n Unknown

WPAC (CA) Unknown Unknown Unknown 35, 000 2011

Zilkha Biomass

Energy (US) Unknown Unknown

Crockett,

Texas

(US)

40,000 Q4 2010

2.7 Pelletization

Pelletization is a densification process of biomass that increases the energy

density and bulk density, minimizes the moisture content, resulting in significant savings

in transportation costs (Holley, 1983; Mani et al., 2003; Obernberger et al., 2004;

McMullen et al., 2005). Kumar et al. (2003) showed a detailed study in Western Canada

on the cost to produce biomass electricity by direct combustion and determined that

transportation was the second most important factor that influences the net cost of

operation. This can be addressed by densifying the biomass in the form of pellets or

briquettes or cubes (Kaliyan and Vance, 2009). . Because of uniform shape and sizes,

densified products can be easily handled using the standard handling and storage

Page 47: Torrefaction and Pelletization of Different Forms of

32

equipment, and can be easily implemented in direct-combustion or co-firing (Kaliyan and

Vance, 2009).

There are a number of pellet producers around the world, producing tons of

pellets for domestic as well as commercial applications. Pellet plant consists of the

following parts: a) chopper: to break down the feedstock into small and uniform particles;

b) crusher/hammer mill: to convert the small particles into fine powder with diameter less

than 3 mm; d) dryer: to decrease the moisture content to less than 15%; and e) pellet mill:

to densify the material into pellets. The capacity of a pellet mill is generally in the range

of 0.30 tons/h to 10 tons/h.

A German-based company AMANDUS KAHL is one of the leading producers of

pellet equipment from small to industrial scale. KAHL pelleting plants have been applied

successfully for compacting organic products of different particle sizes, moisture

contents, and bulk densities. Their pelleting presses are designed for array of feedstock

characteristics as seen in figure 2-5 below. Available pelleting presses consist of a drive

power of 3 kW to 500 kW and a throughput between 0.3 tons/h and 8 tons/h. KAHL

recently developed pellet press equipment with 15 to 20 tons/h capacity. Table 2-4 below

shows the summary of specifications of different pelletizing equipment from four

different manufacturers.

Page 48: Torrefaction and Pelletization of Different Forms of

33

Table 2-4: Summary of Specifications for Four Different Pelleting Equipment

Company La Meccanica NOVA Pellet Kerry Die Amandus Kahl

Model CLM 800 P

LG

N-Plus B-Mass 800 60-1250

Roller Quantity 2 Unknown 6 4 – 5

Drive Power (KW) Up to 280 160 450 3 - 500

Energy Consumption Unknown Unknown Unknown 40 – 60 KWh/t

Capacity (T/H) 2.3 – 3 Up to 2.5 10 15 – 20

Operation Mode Continuous Continuous Continuous Continuous

Weight (kg) 10800 7500 Unknown 9370

Roll Diameter (mm) Unknown 245 250 450

Motor speed (rpm) 750 Unknown 1490 Unknown

Roller Speed (m/s) 6.5 – 7.5 Variable variable 2.5

Die Diameter (mm) Unknown 580 840 175 – 1250

Input Density Unknown Unknown Unknown 150

Output Density (kg/m3) Unknown Unknown Unknown 550 – 650

Feedstock Moisture Unknown 8 – 12% Unknown 12 – 15 wt%

Feedstock Size Unknown 0.5 – 1.5 mm Unknown 4 mm

Pellet Moisture 9 – 12 wt% Unknown Unknown 12 wt%

Pellet diameter (mm) 6 6 8 2 – 30

Page 49: Torrefaction and Pelletization of Different Forms of

34

Figure 2-5: Pictures of raw and pelletized materials (Source:

http://www.akahl.de/akahl/files/Prospekte/Prospekte_englisch/1322_Strohpell_10e.pdf)

More recently, research has found a biomass treatment process that combines the

densification (pelletization) and torrefaction to increase the bulk density and the calorific

value of biomass. Kiel et al., (2008) presented BO2-technology under the umbrella of

Energy-research Centre of The Netherlands (ECN) for biomass improvement into

commodity fuel; a technology that combines torrefaction and pelletization methods to

develop products named as torrefied pellets (BO2 pellets ). BO2 pellets possess the

Page 50: Torrefaction and Pelletization of Different Forms of

35

benefits of both process but with higher bulk density and calorific values developed from

wide range of biomass such as woodchips, agricultural residues and different residues

from the food and feed processing industry (Kiel et al., 2008). A schematic course flow

of the BO2-technology is shown figure 2-6 below and table 2-5 displays the comparison

of “BO2 Pellets ” characteristics from raw wood chips, wood pellets, and torrefied

woods.

Figure 2-6: Schematic Process Flow of the BO2-Technology (Kiel et al., 2008)

Page 51: Torrefaction and Pelletization of Different Forms of

36

Table 2-5: Comparison of BO2 Pellet Properties (Source: Kiel et al., 2008)

Properties

(Typical Values) Wood Chips

Torrefied

Wood Wood Pellets

BO2

Pellet

Moisture wt% 35 0 10 3

LHV KJ/Kg

Dry

As received

17.7

10.5

20.4

20.4

17.7

15.6

20.4

19.9

Bulk Density

Kg/m3

MJ/m3

475

5.0

230

4.7

650

10.1

750

14.9

2.8 Gasification

Gasification is a process that converts biomass into carbon monoxide, hydrogen

and carbon dioxide. This is achieved by reacting the material at high temperatures

(>700°C), without combustion, with a controlled amount of oxygen and/or steam. The

resulting gas mixture is called syngas (from synthesis gas or synthetic gas) or producer

gas and is itself a fuel. The power derived from gasification of biomass and combustion

of the resultant gas is considered to be a source of renewable energy; the gasification of

fossil fuel derived materials such as plastic is not considered to be renewable energy.

The main application of torrefied biomass (wood) is as a renewable fuel for

combustion or gasification. Prins et al. (2006b) studied the possibility of more efficient

biomass gasification via torrefaction in different systems; air-blown circulating fluidized

bed gasification of wood, wood torrefaction and circulating fluidized bed gasification of

torrefied wood, and wood torrefaction integrated with entrained flow gasification of

torrefied wood (Dangtran et al., 2000, Svoboda et al., 2009).

The advantage of gasification is that using the syngas is potentially more

efficient than direct combustion of the original biomass because it can be combusted at

Page 52: Torrefaction and Pelletization of Different Forms of

37

higher temperatures or even in fuel cells, so that the thermodynamic upper limit to the

efficiency defined by Carnot's rule is higher or not applicable. Syngas may be burned

directly in gas engines, used to produce methanol and hydrogen, or converted via the

Fischer-Tropsch process into synthetic fuel. Gasification can also begin with material

which would otherwise have been disposed of such as biodegradable waste. In addition,

the high-temperature process refines out corrosive ash elements such as chloride and

potassium, allowing clean gas production from otherwise problematic fuels. Gasification

of fossil fuels is currently widely used in industrial scale to generate electricity.

Gasification of biomass, which in many ways is a more efficient use of the

feedstock, is nowadays an interesting alternative to combustion for many industries but is

still limited. Tar in product gas is a major drawback of wood gasification in any

conventional gasifier. This can be addressed by using torrefied biomass instead of un-

torrefied biomass. Other disadvantages of un-torrefied biomass are their relatively low

energy content and hydroscopic character. Prins et al.(2006) have shown that higher

gasification efficiency can be achieved by fuels with lower O/C ratio by thermo-chemical

process. Torrefaction is a process that effectively lowers the O/C ratio of biomass in a

simple way and lowers the energy consumption during milling and transportation. The

output product in the form of powder greatly enhances the feeding properties (Kiel et al.,

2007). Although extensive studies have been carried out on the solid product and its

application in gasification, only limited publications have been made on the utilization of

torrefied product in existing thermochemical processes.

Page 53: Torrefaction and Pelletization of Different Forms of

38

2.9 Emission

Burning biomass is known to be carbon neutral; net carbon emissions would be

zero, which is helpful in the fight against global warming. Many industrialized countries

are planning to replace coal as a fuel in power plants with biomass to minimize the

greenhouse gas emission. Torrefied biomass will be an environmentally friendly fuel

compared to fossil fuels.

However, from the torrefaction process, the output product contains gaseous

volatiles, organic acids and primary tars. After capturing the gaseous and liquid products

of the process, the remaining emissions consists only CO2, H2O, NOx and SOx. NOx

emissions can be negligible due to low temperature and SOx emissions can also be

considered as zero due to negligible sulfur content in lignocellulosic biomass. Condensed

tars are a major concern in the application of torrefied biomass. As the temperature

increases during torrefaction, the tar formation also increases exponentially. This issue

needs to be addressed very carefully. According to Kleinschmidt C.P. (2011), test results

have shown that even after combustion, the flue gas contains some organic compounds

like hydrogen fluorides, sulfides and nitrates that need to be removed before collecting

the flue gas. Bag filters and ceramic filters with an absorbent are suggested to minimize

these emissions. The emission from biomass torrefaction is not a major technical

challenge, but the ash, chlorine, sulfur and alkaline production should be minimized.

2.10 Storage Behavior

There is always chance of off-gassing and self-heating in any kinds of solid

biomass because of presence of moisture and pores by chemical oxidation and

microbiological contamination. Throughout storing of such biomass, there is always

Page 54: Torrefaction and Pelletization of Different Forms of

39

chance of occurring chemical-microbial reactions. According to Tumuluru et al.(2011),

high storage temperatures of about 50°C may generate high CO and CO2 emissions, and

the concentrations of these off-gases can range up to 1.7 and 6% during the storage

period of sixty days. Emitted products are also sensitive to relative humidity and moisture

content. Torrefied biomass either pelletized or raw products show superior performance

than the raw biomass or raw pellets because of its hydrophobic nature and low moisture

uptake even under severe storage circumstances. Off gassing and self-heating are at

minimum level in torrefied biomass as most of the solid, liquid, and gaseous products,

because chemically and microbiologically active components are eliminated in the

torrefaction procedure. Studied at University of British Columbia, Vancouver, Canada on

off gassing from torrefied wood chips showed minimum emission of CO and CO2; nearly

one third of the emissions from regular wood chips (Van der at el., 2011). Delivery cost

of pelletized biomass including shipping, trucking, storage, pre-treatment, chopping and

others is presented in figure 2-7 (Van der at el., 2011 and Zwart R. et al., 2006).

Page 55: Torrefaction and Pelletization of Different Forms of

40

Fig. 2-7 Delivery costs of pelletized biomass. (Numbers indicate nominal

capacity of system in dry kilotons of raw biomass feedstock per year) (Source: Van

der et al, 2011)

2.11 Economic Potential

To analyze the details of net profit of torrefaction, the impact of the process on the

all steps of the value chain is to be discussed. The segments of benefits are transport,

storage, carbon neutral and production. Higher energy density, condensation,

pelletization and dried mass of the torrefied products make economic benefit on the

transportation. Hydrophobic behavior of torrefied biomass can be successfully stored

outdoors, thus obviating the need for an enclosed storage bin or building but further

studied is required on this issue. However, it should be noted that, in dry climates, wood

chips have been successfully stored in large outdoor piles. The relative fuel losses

(shrinkage) during storage are not well known, but can be expected to be higher for

Page 56: Torrefaction and Pelletization of Different Forms of

41

outdoor storage. Comparisons of shrinkage losses of torrefied vs. raw biomass are needed

for different storage conditions and climates. Utilization benefits are related to the higher

energy content, lower oxygen content, and (probable) lower moisture content, relative to

unprocessed biomass. Torrefied biomass is expected to perform as well or better than raw

biomass for many bioenergy applications, including combustion, gasification, and fuel

production applications (Svoboda et al., 2009). Enhanced conversion and utilization,

when compared to the other steps in the supply chain, probably provide the most

significant opportunity for cost savings (followed by transport costs). Torrefied biomass

is believed to be a superior solid fuel for combustion, especially when co-fired with coal

due to its higher energy density and coal-like handling properties. Torrefied biomass is

also expected to provide advantages as a fuel for thermochemical processing, due to the

removal of acids and oxygen. Gasification using torrefied biomass allows for improved

flow properties of the feedstock, increased levels of H2 and CO in the resulting syngas,

and improved overall process efficiencies (Svoboda et al., 2009, IEA, 2010). Torrefaction

combined with pelletization provides a lower cost fuel for power or fuel production when

compared to pelletizing alone, with cost savings ranging from 4% to 16%, depending on

the end use of the biomass. Fig 2-7 shows supply chain costs for several scales and

processing options for biomass, indicating that pelletizing of torrefied biomass

significantly reduces costs, that larger-scale operations are more cost efficient, and that

integrated torrefaction and pelletizing is less costly than pelletizing alone. Zwart et

al.(2006) conclude that, while torrefaction is one of the most cost-effective options for

supply of overseas biomass, modifications to the supply chain, such as the centralized

processing of raw feedstock, can result in similar reductions in overall costs.

Page 57: Torrefaction and Pelletization of Different Forms of

42

According to Van der Stelt et al.(2011), the torrefaction step represents an

additional unit operation in the biomass utilization chain. The attendant capital and

operating costs, as well as conversion losses are, however, offset by savings elsewhere.

Recent cost estimates for the ECN torrefaction technology indicate that the total capital

investment of a standalone 75 kton/a plant will be in the range 6.1 to 7.3 MV. The

assumed feedstock is wet softwood chips. The plant consists of a conventional rotary

drum for drying the biomass, ECN torrefaction technology and conventional grinding

equipment and pellet mill. No feedstock preparation (e.g. chipping) before drying was

included. At 75 ktons/a production rate (design), the total production costs are calculated

at 37 V/ton product (2.0 V/GJ), produced from a feedstock with 35% moisture content.

At 50% and 25% moisture content this is 50 V/ton (2.6 V/GJ) and 34 V/ton (1.9 V/GJ) of

product, respectively. The moisture content is one of the most influential parameters of

the torrefaction process as it predominantly determines the energy input of the process.

These data represent the added cost for the torrefaction process without preprocessing of

pre-drying process of biomass.

There is always a cost for any processing so torrefaction also adds some cost in

the processing than raw biomass (Magalhaes et al., 2009). Because of low moisture

contents, it also reduces cost of transportation and grinding. Torrefied biomass acts as

good fuel for any gasification process. Gasification with torrefied biomass enhances flow

properties of the feedstock, improves contents of H2 and CO in the syngas, and enriched

process efficiency resulting with cost savings from 4% to 16%. Zwart et al.(2006)

concluded that torrefaction is one of the best cost-effective alternatives for delivery at

overseas destination with optimum overall costs.

Page 58: Torrefaction and Pelletization of Different Forms of

43

ECN, Netherland estimated that torrefaction technology capital investment of a

standalone 75 kton/a plant will be in the range 6.1 to 7.3 M€. The plant comprises of

drying unit (rotary drum), torrefier unit, grinding unit and pellet mill. No feedstock

preparation (e.g. chipping) before drying was counted in for the estimation. For 75

ktons/year manufacturing of torrefied products, the total production costs are estimated as

37 €/ton, with 35% moisture content, 50 €/ton with 50% moisture contents and 34 €/ton

with 25% moisture content.

2.12 Scanning Electron Microscopy (SEM)

A scanning electron microscope (SEM) is used to project the images of a sample

by scanning over it with a high energy focused electron beam. The electrons interact with

electrons in the sample, generating secondary electrons, back-scattered electrons, and

characteristic x-rays that can be detected and that comprise data about the sample's

superficial structure and configuration. The electron beam is generally scanned in a raster

scan pattern, and the beam's position is combined with the detected signal to produce an

image. The electron beam can be concentrated to a spot approximately 1 nanometer in

diameter, and microscopes are able to resolve details ranging from 1–20 nm in size. To

avoid accumulating charge on the surface, samples must be electronically conductive;

non-conducting samples are often coated with an ultrathin coating of metal. Conventional

SEM requires samples to be imaged under vacuum, but methods have been developed

that allow imaging biological samples (Suramya DFM, 2012).

The large content of Silica, Potassium and Chlorine in biomass significantly

increases the deposit formation and corrosion of the thermal power plants, compared to

boilers firing with coal (Basu, 2006; Beatrice et al., 2007). It has been verified that the

Page 59: Torrefaction and Pelletization of Different Forms of

44

alkali species which cause the bed agglomeration originate from the biomass ash. So,

study on the behaviors of alkali species in biomass ash is crucial for understanding the

mechanisms of agglomerates and coating layers formation. The ash block can maintain a

fixed shape and presents a little compression strength due to the sintering. Skrifvars et al.

(1997) characterized the sintering tendency of ten biomass ashes and classified the ash

components into three groups: simple alkali salts, silicates and the rest non-melt. We can

see that the fusible compounds melt and coat the surface of ash particles during

combustion. According to the experimental data and plenty of study results by other

investigators (Skrifvars et al., 1997), these melts should be the alkali silicates. Based on

the K2O–CaO–SiO2 phase diagram and the Na2O–SiO2 and K2O–SiO2 phase systems

(Jenkins et al.,1995) the melting point of these alkali silicates should be in the range 800–

1200°C. At the same time, these silicon dioxide grains are coated by the molten alkali

silicates which can freely flow through the gaps between the silicon dioxide grains. From

above findings we see that the alkali metals such as K and Na are mainly found in the

outer layer of biomass ash. The alkali silicates formed during combustion can melt and

coat the surface of ash particles at high temperature. Further, it can be imagined that the

sizes of ash particles must vary in a wider range due to the fragmentation and attrition in

the bed. When the small-sized ash particles collide with bed particles, the molten alkali

species on them will be transferred to surfaces of bed particles. In addition, the large-

sized ash particles may increase the amount of melts on local surface of bed particle and

may act as the necks for the agglomerate formation. So, the large- and small-sized ash

particles may play the different roles in the bed agglomeration. There is no article found

Page 60: Torrefaction and Pelletization of Different Forms of

45

in the open literature on SEM analysis of poultry litter ash and this kind of analysis is of

its first kind. The SEM of wood mixture from gasifier is shown in figure 2-8.

2-8. SEM of Wood Ash from Gasifier (Beatrice et al., 2007)

2.13 X-ray Diffraction (XRD)

According to online definition, the scattering of x-rays by crystal atoms,

producing a diffraction pattern that yields information about the structure of the crystal is

defined as XRD. The wavelengths of X-rays are of the same order of magnitude as the

distances between atoms or ions in a molecule or crystal having less than Nano meter in

size. A crystal diffracts an X-ray beam passing through it to produce beams at specific

angles depending on the X-ray wavelength, the crystal orientation, and the structure of

the crystal. X-rays are mainly diffracted by electron density and analysis of the

diffraction angles creates an electron density map of any types of crystal (Cheng et al.,

2011). XRD instrument consists of an X-ray generator, a goniometer and sample holder,

and an X-ray detector such as photographic film or a movable proportional counter. X-

Page 61: Torrefaction and Pelletization of Different Forms of

46

ray tubes generate X-rays by bombarding a metal target with high-energy (10 - 100 keV)

electrons that knock out core electrons. An electron in an outer shell fills the hole in the

inner shell and emits an X-ray photon. Two common targets are Mg and Cu, which have

strong K(alpha) X-ray emission at 0.71073 and 1.5418, respectively. X-rays can also be

generated by decelerating electrons in a target or a synchrotron ring. These sources

produce a continuous spectrum of X-rays and require a crystal monochomator to select a

single wavelength.

Powder X-ray Diffraction (XRD) is one of the primary techniques used by

mineralogists and solid state chemists to examine the physic-chemical make-up of

unknown solids. This technique provides information that cannot be obtained any other

way. The information obtained includes types and nature of crystalline phases

present, structural make-up of phases, degree of crystal formation, amount of

amorphous content, micro-strain & size and orientation of crystallites.

2.14 Summary from Literature

From the analysis of above literature review, following conclusions can be made:

a) Biomass can play important role as a carbon neutral fuel in the future.

b) Torrefaction process includes chopping, drying, roasting and cooling of any kinds

of biomass samples which leads polymerization and carbonization phenomenon.

Study on torrefaction gets more and more important for making solid fuels from

biomass even though research on it is still at the primitive stage of commercial

development. Torrefaction process depends on many parameters which are still

Page 62: Torrefaction and Pelletization of Different Forms of

47

needs to be studied. Although, researches on the technology and science of

torrefaction have been carried out, further investigations are still required.

c) Torrefaction results are different for different biomasses so it is important to

investigate the different biomass sample separately. None of the published

materials shows the studies on torrefaction characterization of lignocellulosic and

non-lignocellulosic biomasses from Ontario.

d) Different types of torrefaction reactors are now available in the market but still

further research is essential to improve the reactor efficiency.

e) Torrefaction improves the hydrophobicity, grindability, non-degradability,

storability. However, the findings of the optimization of torrefaction temperature

and residence time are still under the investigation which best fit the long storage

capability without bio-degradation.

f) Pelletization is a compressing techniques which reduces the volume of the

biomass drastically and improves the energy density. It will reduce transportation

cost of biomass pellets. Pelletization techniques for raw and torrefied biomass

needs further investigation as limited studies are only available.

g) Safety process during the torrefaction techniques are still under study like dust

explosibility, emissions, health hazard, environmental impact on the surroundings

etc.

h) Major researches on torrefaction are basically based on the lignocellulosic

biomass having cellulose, hemicellulose and lignin but there is lack of in-depth

study on the non-lignocellulosic biomasses like animal wastes, municipal waste

and waste from agro-based plants.

Page 63: Torrefaction and Pelletization of Different Forms of

48

i) Torrefied bio-fuels are the potential input for the electrical or thermal power

plants. Biomass creates slagging and fouling at the tubes of boilers. Biomass ash

management is very important issues for designing future bio-fuel boilers. None

of the papers are available on the detail ash analysis of lignocellulosic and non-

lignocellulosic biomass

Hence, this study tries to fill the small gap on torrefaction characterization,

pelletization techniques, optimum temperature for hydrophobicity, storage behaviors and

ash analysis using SEM and XRD. Finally, comparative studies on above mentioned

studies are presented.

Page 64: Torrefaction and Pelletization of Different Forms of

49

Chapter III: Methodology

3.0 Problem Statement

The Ontario government is on track to phase out coal-fired electricity plants by

2014. According to Ontario Ministry of Energy, the power demand will increase by

about 15% between 2010 and 2030. Ontario should take timely initiative for the

production of the clean energy as per its’ future demand. The development of clean and

economically viable biomass to bioenergy conversion technologies for a domestic market

is thus imperative to promote the local utilization of biomass in Canada. Certain portion

of the energy demand can be fulfilled by the biomass energy after the successful up

gradation of its quality. One of the methods to upgrade the quality of biomass is the

torrefaction, densification and gasification.

Hence, more interest for research in the area of torrefaction and pelletization has

grown. There is potential to improve the quality of agricultural and woody biomass and

its residues as energy fuel which can be used for many energy applications. Agricultural

products, woody biomass and non-lignocellulosic bio-residues are attractive sources of

energy that can be greenhouse gas emission neutral and can provide sustainable energy

sources to meet the energy economy. Although biomass is abundant and renewable, its

properties pose several challenges during thermal conversion process; hence, limits its

applications in power plant operations even though it has been successfully used as an

upgraded solid fuel in electric power plants and gasification plants. Many studies are

available on torrefaction process; up to date however, majority of them have focused

mainly on lab-scale analysis of biomass compositions and characterization of torrefied

biomass in terms of grindability and energy value while only few have examined its

Page 65: Torrefaction and Pelletization of Different Forms of

50

feasibility on commercial applications. These torrefied fuels can be used for high quality

smokeless solid fuels for industrial, commercial and domestic applications, solid fuel for

cofiring directly with pulverized coal at electric power plants, an upgraded feed stock for

fuel pellets, briquettes and densified other biofuels. Not all aspects of torrefaction and its

influence on other processing operations have been explored (Tumuluru et al., 2011;

Khodier, 2011).

There exists several gaps in the development of torrefaction technologies and its

maturities and there is need for continued research and development to characterize and

optimize this promising option for bioenergy feedstock processing for the application of

next generation fuel prior to the depletion of fossil fuels. Governments, private parties

and universities are investing a lot in the field of biomass applications. Several

achievements are still under the scope of laboratory. The most challenging is to see the

laboratory experiment in the commercial applications. For this, in depth study on

composition and application of tar, char, ash has yet to be established, in part due to the

complex chemical nature of the feedstock. Environmental effect, storage behavior of

torrefied biomass, energy analysis of the torrefied products, temperature effect, heating

values due to different temperature, practical reactors, residue management, effective

transportation possibilities are few areas of research on the torrefaction process.

3.1 Research Scope and Objectives

Current research has wide range of scope in the era of global warming reduction

by maximizing the utilization of green fuels. There are number of challenges to minimize

the dependency on fossil fuels and accomplish a sustainable, renewable energy source.

Energy from biomass can be produced from different thermochemical combustion,

Page 66: Torrefaction and Pelletization of Different Forms of

51

gasification, and pyrolysis, where direct combustion can provide an energy solution. The

increasing attention in the use of biomass is considered as a solid fuel. This consists of

combustion to produce steam for electrical power and commercial plants. Still, the use of

either producer gas or gas turbines, or to produce higher value chemicals and fuels, is

limited due to biomass feedstock preparation, storage, transportation, logistics, ash

management and economics. This study will add new scope on these aspects so that

biomass can be used as bio fuel in the future.

Figure 3.1 Schematic Block Diagram of Research Procedures

Hence, this work will expand the torrefaction studies on the different lignocellulosic

and non-lignocellulosic biomass of Ontario and complete the following objectives by

conducting different experiments and result analysis:

i) Physio-chemical characterization of biomass samples before and after

torrefaction.

ii) Optimization of torrefied conditions based on hydrophobicity

iii) Investigation of pelletization potential before and after torrefaction

Biomass Product

Input: Willow,

Oats and Poultry

Litter

Energy Input

Torrefier,

Reactor,

Pelletizer, Bomb

Calorie meter Pelletization

Strength,

Heating values

and Ash analysis

Output products

analysis,

Torrefaction,

Hydrophobicity

Page 67: Torrefaction and Pelletization of Different Forms of

52

iv) Ash analysis of biomass at different combustion temperature.

3.2 Methodology

The proposed research will focus biomass analyses before and after torrefaction

and will be represented on a dry-ash-free basis. A laboratory setup is prepared and

intensive lab test will be carried out to find the performance on selected biomass samples,

Poultry litter, willow and oat pallet samples to characterize their properties in terms of

moisture content, volatile matter, ash content, and fixed carbon. The proximate analysis,

elemental analysis, ash analysis will be performed on all samples.

The methodology will be based on the following flow chart as mentioned in figure 3.2.

Page 68: Torrefaction and Pelletization of Different Forms of

53

Figure 3.2: Flow Chart of Methodology

Biomass Sample: Willow, Oats and

Poultry Litter

Characterization:

Proximate, Ultimate,

Energy Density and

HHV

Torrefaction

-Varied O2

-Temperature (200°C -300ᴼC)

-Residence Time (15-60min)

Comparative Study of

Torrefied and Raw

Biomass:

Characterizations,

Hydrophobicity

Pelletization, Ash

Ash Analysis at

800°C, 900°C,

1000ᴼC

-SEM

-XRD

- Fusion Temp

Hydrophobicity

-Optimum

Torrefaction

Temperature

-Moisture Uptake

Investigation of

Pelletization Potential

-Making Force

-Breaking Force

Page 69: Torrefaction and Pelletization of Different Forms of

54

Chapter IV: Experiment Setup

4.1 Biomass Characterization

Biomass samples of poultry litter, willow pellets and oats pellets were collected

from the different locations of Ontario. Drying process of all biomass samples were

conducted at 105°C according to ASTM standard D1762-84 procedure. Muffle furnace

was used to determine the moisture content of the biomass. Different samples of biomass

weighing from 1-2 gram were placed in the muffle furnace for two hours at 105°C and

allowed them to cool at desiccator and weight were taken. The experiments were repeated

till the constant weight reached. The change in weight of biomass was considered as the

total moisture present in biomass and its percentage determined.

4.1.1 Proximate Analysis

Proximate analysis was conducted to determine the moisture, volatile matter, ash

and fixed carbon contents according to the procedure as specified on a modified ASTM D

5142-04 method on a muffle furnace of Thermo Scientific.

4.1.2 Ultimate Analysis

Ultimate analysis of the samples was carried out according to ASTM D 5373-08

method. All samples were dried at 105°C for 24 hours prior to the experiment for

Ultimate analysis. During the experiment, the combustion was carried out at 925°C under

Helium atmosphere while the reduction was carried out at 650°C.

4.1.3 Heating Value

After the specified residence time, crucible were removed from the reactor and

placed for in a vacuum desiccator to cool. The samples were grinded in a coffee grinder

to prepare sample for Bomb Calorimeter (Model IKA C-200) by IKA. The complete

Page 70: Torrefaction and Pelletization of Different Forms of

55

basses of calculation for the calorific value were based on ASTM D 240 and ASTM D

5865. Combustion was carried out in a calorimeter in the presence of Oxygen. The

decomposition vessel was filled with a fuel sample of less than 1 gram and pure Oxygen

of maximum of 30 bar, the fuel sample was ignited and the temperature increase in the

calorimeter system measured. The heat quantity required to raise the temperature of the

calorimeter system by one Kelvin was used to determine the C-Value of the system. The

specific calorific values of the sample were calculated as follows:

Where

m = weight of fuel of the sample

C = Heat capacity (C-value) of the calorimeter system.

DT =Calculated temperature increase of water in inner vessel of measuring cell

Correction value for the heat energy generated by the cotton thread as

ignition aid

Correction value for the heat energy generated from other burning aid

4.2 Torrefaction

A novel reactor (Fig. 4.1), similar to Quartz Wool matrix (QWM) reactor (Bashu

P., 2010), was designed, developed and fabricated in the machine lab at University of

Guelph for the purpose of continuous torrefaction to produce a torrefied product, which

is. The reactor consists of a Stainless Steel (SS) tube heated by four electric heaters of

1.25KW capacity in close contact with the reactor wall and separately controlled by two

PID controllers. The SS tube has an inner diameter of 75 mm and height of 600 mm. The

Page 71: Torrefaction and Pelletization of Different Forms of

56

percentages of different composition of gaseous particles inside the reactor were

observed using gas analyzer (Testo-350). Experiment were conducted for different

samples at different temperature 210°C, 250°C, 275°C and 300°C and at different

residence time up to 60 minutes. This reactor allows simulation of any gas-solid relative

velocity; gas composition and gas temperature in a reactor while a precision electronic

balance continuously measures the mass change of a reaction. Such reactor could thus

accurately simulate conditions one would expect in a fixed, moving or entrained flow

reactor. Before starting an experiment, the reactor was heated until an equilibrium

temperature or steady state was attained. A stream of inter gas (N2) of flow rate of 1-16

liters per minutes was flushed through flow meter (FMA 5400/5500, Mass Flow

Controller, Omega, USA) to maintain an inert environment inside the reactor.

Temperatures were measured at two different locations one from upper mid portion of

reactor and another from lower mid portion of the reactor by two separate thermocouples

through the temperature controller (CNi16D, Temperature and Process Controller,

Omega, USA). Then the sample of biomass of known mass and moisture content were

placed into the reactor. The residence time was recorded from that instant. During

torrefaction, temperatures of the gas passing through the biomass are continuously

recorded. The electronic balance (Model: MS204S, Mettler Toledo, Switzerland)

continuously measured the mass of the biomass. After a specified time, the biomass is

taken out. Thereafter, the sample is cooled down in desiccator and weighed. Its

composition is analyzed by proximate analysis (ASTM E870-82) and energy density by a

bomb calorimeter (model C-200 by IKA). Similar experiments are repeated for separate

temperature and residence time for individual biomass samples.

Page 72: Torrefaction and Pelletization of Different Forms of

57

Percentage of mass yield (MY), percentage of energy yield (EY) and energy

density radio (EDR) are determined by using following formula:

Where,

= Mass (daf) of torrefied biomass

= Mass (daf) of raw biomass

= Higher Heating Value of torrefied biomass

= Higher Heating Value of raw biomass

Figure 4.1 Experimental Setup for Torrefaction and Weight loss

To chimney

Electric tubular

furnace

Analytical balance

N2

SS porous

basket

Gas

analyzer

Temperature

Controller

Pre-heater

Data

Logger

Reactor tube

Mass flow controller

Gas

sample

Thermocouple tube

Page 73: Torrefaction and Pelletization of Different Forms of

58

4.3 Hydrophobicity

Very few papers have stated the experimental setup of the hydrophobicity test for

torrefied samples. Pimchaui et al. (2010) demonstrated hydrophobic test by immersing

raw and torrefied biomass in water for two hours, allowed to dry and determined the

weight change as a measure of moisture absorption. This approach was similar to Felfli et

al. (2005) in investigating the hydrophobic characteristics of briquettes where briquettes

were immersed for 70 minutes in water.

Here, Raw and Torrefied Samples were immersed in water for 2 hours at room

temperature of 22°C-25°C and relative humidity of 40%-50% and the hydrophobic

characteristics of torrefied and raw samples were investigated. The contents of the

moisture is determined after drying the immersed samples at the interval of 1 hours till

the nearly constant weight of the sample were achieved. In addition, hydrophobic test

were conducted in controlled environment pressure by using humidifier. The humidity

level was from 88%.

4.4 Pelletization

According to Wolfgang et al. (2011), the pellets were prepared using a single

2812 pellet press from Parr which was modified at the workshop of University of Guelph

as shown in figure 4-2, such that it can connect with Omega LC1001-500 and

CNiS8DH33 to measure the force in Newton maximum of 2224N. The press consisted of

a cylindrical die 6.4 mm in diameter, made of hardened steel. The end of the die was

closed using a removable backstop. Pressure was applied and the force could be

measured using an Omega Process Gauge controller (CNiS6DH33), USA. The die was

rinsed with acetone, and wiped clean using a paper towel before each use, and when

Page 74: Torrefaction and Pelletization of Different Forms of

59

changing raw materials. The pressure was released after five seconds, the piston

removed, and more biomass was loaded and compressed until the pellet had a length of

about 16 mm. This results in a layered structure, similar to pellets obtained by

commercial units, although there are some differences. The most significant difference is

that the lower part of the pellet is pressed repeatedly, and the upper layers are pressed

fewer times, with the top layer being pressed only once. For determination of pelletizing

pressure in the press channel of the pellet mill, FN, the pellets were removed from the die

by removing the backstop and pushing out the pellet at a rate of 2mm/s. The applied

maximum force was logged and FN was calculated based on the pellet surface area.

Figure 4-2 Experimental Setup for Pelletization

Modified Parr Pellet Press 2812 pellet press

Strength Meter Omega LC1001-500

Handle of Pellet Press

Heating Tape over die

Temperature Controller CNiS8DH33

Omega Process Gauge controller (CNiS6DH33)

Page 75: Torrefaction and Pelletization of Different Forms of

60

The internal strength of the manufactured pellets was analyzed by compression

testing and determined as the force at break. Pellets 15-17 mm in length and between 6-7

mm in diameter were produced in the single pellet press, stored at a relative humidity of

50% and 20°C for three weeks, and tested under the same conditions. The pellets were

placed on their side in the same material tester as was used for pellet preparation.

Breaking test was also carried out in similar setup. The average force at break was

calculated based on 3 replications per test.

4.5 Storage Behaviour

Here, real environmental effects during the storage of the torrefied products were

monitored. Raw and torrefied biomass samples were kept in particular controlled

environmental conditions. Indoor storage behaviors was monitored by placing samples in

a room and see the weight loss or biodegradable behavior for 48 hours. Humidity of test

room was maintained by humidifier and measured the percentage of moisture uptake by

each biomass.

4.6 Scanning Electron Microscopy (SEM)

In this study, ash samples of 800°C, 900°C and 1000°C were prepared and SEM

images analyses were performed on the ash samples. Morphology of the biomass samples

Page 76: Torrefaction and Pelletization of Different Forms of

61

were analyzed using SEM. Samples were coated with gold (20 nm) with a sputter coater

(Model K550; Emitech, Ashford, Kent, England) prior to SEM analysis. SEM

micrographs were taken by using a model S-570 (Hitachi High 115, Technologies Corp.,

Tokyo, Japan) at 10 kV accelerating voltage. Many images were captured by extracting

electrons from a sharp tungsten tip, and formed into a fine beam by a series of

electromagnetic lenses. The beam was directed over the surface of the sample, and the

signals were collected point by point to produce a grayscale image. Resolutions of up to

one nanometer could be produced using the SEM as shown in figure 4.3.

(a) Photograph of SEM Setup

Page 77: Torrefaction and Pelletization of Different Forms of

62

(b) Schematic Diagram of Setup

Figure 4.3: SEM Experimental Setup (Hitachi S-570)

4.7 X-ray Diffraction (XRD)

XRD patterns were measured in Bragg-Mrentano geometry in a STOE two

circle goniometer using Cu Kalfa characteristic radiation (Lambda = 1.54178 A)

produced by an ENRAF NONIUS FR 571 rotating anode generator. The x-rays were

detected with a MOXTEK energy sensitive Si detector connected to a single channel

analyzer that was set to accept only the Cu Kalfa photons.

The samples were grounded in an agate mortar and the loose powder was packed into an

aluminum sample holder and then X-Ray diffraction tests were carried out on the ash

samples of the biomass. The XRD test gave the structure of constituent elements,

minerals and ores present in the ash. The patterns were measured in the 2θ interval from

5° to 70°, with a step size of 0.02° and 24 s counting time per step.

Page 78: Torrefaction and Pelletization of Different Forms of

63

Chapter V: Results and Analysis

5.1 Poultry Litter Biomass

For the study of non-lignocellulose, poultry litter from a poultry farm located

inside the Ontario was collected. Collected poultry litter was dry, mixture of pellets and

powder form. It also contains small feathers of poultry and bed materials of the poultry

farm. Collected samples were dried as per ASTM standard D1762-84 until samples were

dried on the Thermo Scientific muffle furnace with 2416 controller (model: F48055-60)

for two hours at 105°C and conducted different types of experiments and the following

results were obtained.

5.1.1 Biomass Characterization

Proximate analysis of biomass was carried out in the muffle furnace manufactured

by Thermo Scientific muffle furnace with 2416 controller (model: F48055-60) on raw

and torrefied samples and result obtained were listed in figure 5-1 to 5-3. From the

results, volatile matter decreases as the residence time and torrefaction temperature

increases whereas ash contents and amount of fixed carbon slightly increases as the

residence time and torrefaction temperature increases. The cause of variation is due to the

loss of moisture contents of the biomass as the temperature and residence time increases.

The moisture content of the raw sample was 20.1% and the heating value of it was 10.088

MJ/Kg. Other details of biomass characteristics are listed in section 5.4 of this report.

Variation of volatile matter from 55% to 38%, carbon from 3% to 22% and ash contents

from 22% to 38% is observed as the temperature and residence time increases.

Page 79: Torrefaction and Pelletization of Different Forms of

64

Figure 5-1 Volatile Matter vs. Residence Time for Poultry Litter

Figure 5-2 Fixed Carbon vs. Residence Time for Poultry Litter

Page 80: Torrefaction and Pelletization of Different Forms of

65

Figure 5-3 Ash Contents vs. Residence Time for Poultry Litter

5.1.2 Torrefaction

The torrefaction of biomass samples at 210°C, 250°C, 275°C and 300°C resulted

in four products of very light brown, light brown, dark brown and black color as shown in

appendix. The color change is mainly attributed to chemical changes of the biomass (Lam

et al., 2011). The loss of dry matter through volatilization during torrefaction was within

5% at 210°C, about 20% at 250°C and above 30% at 280°C and more than 50% at 300

°C. Mass yield and energy yield at 210°C, 250°C, 275°C and 300°C are shown in figure

5-4 to 5-7. From the result, it is observed that lower oxygen concentration has more

energy yield and more mass loss than the torrefaction environment with higher

concentration of oxygen. Higher Heating Value of the torrefied samples with lower

oxygen concentration has higher than the torrefied sample of higher oxygen

Page 81: Torrefaction and Pelletization of Different Forms of

66

concentration. HHV varies from 10.00 to 12.40 MJ/Kg with 0-0.6% Oxygen

concentration whereas 10.00-11.90 MJ/Kg with 2.4% oxygen concentration. During the

torrefaction process, sample flue gases were measured by the gas analyzer which shows

percentage contents of different gases like CO, NO, SO2. At temperature 275°C, the

highest CO of 101 ppm and 29 ppm of SO2 were observed. Mass yield and energy yield

figures also displayed error bars with 3.5-4.5% standard deviation.

Figure 5-4 Mass Yield and Energy Yield (%) vs. Temp. (45min residence time)

Page 82: Torrefaction and Pelletization of Different Forms of

67

Figure 5-5 % of Mass Yield and Energy Yield vs. Temperature for Poultry Litter

(2.4% Oxygen with 45 minutes residence time)

Figure 5-6 Heating Value vs. Residence Time for Poultry Litter (0% O2)

Page 83: Torrefaction and Pelletization of Different Forms of

68

Figure 5-7 Heating Value vs. Residence Time for Poultry Litter (2.4% O2)

5.1.3 Hydrophobicity

Figure 5-8 shows that percentage of moisture absorption decreased by multiple

folds after torrefaction and displayed error bars with 2.5-4.0% standard deviation. The

more the torrefaction temperature and residence time the more the decrease in the

percentage of moisture absorption. Similarly, the moisture increases as the torrefaction

parameters are increased which is caused by the emission of volatiles. The emission of

volatiles becomes more intensive as torrefied temperature and residence time is raised

which lead to increase the porosity and hygroscopic characteristics of biomass. For

testing the hydrophobicity behavior, torrefaction at 275°C for any residence time from 30

to 60 minutes shows the best result among tested results.

Page 84: Torrefaction and Pelletization of Different Forms of

69

Figure 5-8 Hydrophobic behavior of Poultry Litter

5.1.4 Storage Behavior

Figure 5-9 shows the moisture uptake behavior of the torrefied poultry

litter at different temperature of 30 minutes residence time at 88% relative humidity of a

storage room and displayed error bars with 3.75-4.75% standard deviation. It shows that

maximum absorption of moisture from the humid environment of the dried raw biomass

was about 16% maximum whereas minimum absorption of moisture from the atmosphere

was performed by the torrefied poultry litter at 250°C and 275°C. But, the moisture

absorption increased when the temperature of torrefaction increases to 300°C because of

high porosity and hygroscopic characteristics of biomass at higher temperature. Hence,

Page 85: Torrefaction and Pelletization of Different Forms of

70

for storing purpose, torrefied biomass at temperature of 275°C was found as optimum

option for poultry litter.

Figure 5-9 % of Moisture uptake vs. temperature of torrefied biomass

5.1.5 Optimization by Box Behnken Model

A Box Behnken design was used to analyze the data and model the torrefaction process

in term of mass and energy yields and evaluate the significance of temperature, residence

time and moisture content on mass yield and energy yield were observed. From the result

using response surface model, the optimization of Mass Yield and Energy Yield were

found as:

Where

MC =Moisture Contents in %

TT =Torrefaction Temperature in C

RT =Residence Time in Minutes

Page 86: Torrefaction and Pelletization of Different Forms of

71

The strength of effect of three significant process parameters on mass yield was better

revealed by surface plot as shown in figure 5-10 and 5-11 which showed the result of

mass yield varies from 55% to 97% as the moisture varies from 3% to 30% and

temperature varies from 250ᴼC to 300ᴼC according to the residence time variation from

15 minutes to 45 minutes.

Figure 5-10 Cubes of Temperature, Moisture and Residence Time Vs. Mass Yield

Page 87: Torrefaction and Pelletization of Different Forms of

72

Figure 5-11 Surface Plots of Temperature, Moisture and Residence Time Vs. Mass

Yield

The strength of effect of three significant process parameters on energy yield was

also better revealed by surface plot as shown in figure 5-11 and 5-12 which showed the

result of mass yield varied from 70% to 98% as the moisture varies from 3% to 30% and

temperature varied from 250°C to 300°C according to the residence time variation from

15 minutes to 45 minutes.

Page 88: Torrefaction and Pelletization of Different Forms of

73

Figure 5-12 Cubes of Temperature, Moisture and Residence Time Vs. Energy Yield

Figure 5-13 Surface Plots of Temperature, Moisture and Residence Time Vs.

Energy Yield

Page 89: Torrefaction and Pelletization of Different Forms of

74

5.1.6 Scanning Electron Microscopy (SEM)

Figure 5-14 (a), (b) and (c) show the images of the poultry litter ash samples at

800°C, 900°C and 1000°C respectively. At low temperature, the sample contains some

wool like structures which gradually disappears with the increase in combustion

temperature. This may be due to the presence of combustible particles. This wool like

structures are not combustible at low temperature and can only be combusted at high

temperature. Also at high temperature, particles become more brittle as the minerals

matters in the form of crystal increases. Hence the particles become smaller and many

cracks start to show up as the combustion temperature increases. The increase in smooth

surfaces in the samples with the increase in combustion temperature also supports the

relative increase in mineral matters in the samples which is obtained in XRD analysis.

a) 800ᴼC

Page 90: Torrefaction and Pelletization of Different Forms of

75

b) 900ᴼC

c) 1000ᴼC

Figure 5-14: SEM for poultry litter ash at 800°C, 900°C and 1000ᴼ C

Page 91: Torrefaction and Pelletization of Different Forms of

76

5.1.7 X-Ray Diffraction (XRD)

Figure 5-15 shows the XRD patterns for poultry litter ash prepared at

temperatures 800°C, 900°C and 1000°C. As it is well known that calcium is an important

diet for the poultry and hence it is expected to have some kind of crystalline structure of

calcium compound in the litter samples. The XRD analysis confirm that the basic forms

in poultry litter ash were CaCO3, SiO2 and K2Ca(CO3)2, which determined the nature of

the alkaline extracts in water. Poultry litter ash could be considered as an attractive

material for neutralizing acidic soil and could be a good source of material for cement

production because of the availability of high calcium components. It was also observed

that with the increased in temperature, the intensity of peaks increased.

Page 92: Torrefaction and Pelletization of Different Forms of

77

Figure 5-15: XRD pattern for poultry litter ash at different temperature

5.1.8 Summary

The non-lignocellulosic biomass of poultry litter showed that the decrease in the

moisture contents increases in the energy yield and mass yield as the temperature and

residence time increases. Increase in temperature and residence time also makes

hydrophobic to the non-lignocellulosic biomass of poultry litter. From the HHV and

hydrophobicity point of view, torrefaction at 260°C-280°C for residence time of 20-40

minutes was found as the optimum region for poultry litter. Dried and torrefied particles

do not bind well which can be solved by adding small quantities of water. The severity of

INT

EN

SIT

Y

Page 93: Torrefaction and Pelletization of Different Forms of

78

torrefaction improves the degree of hydrophobicity. Torrefied poultry litter at 275°C

performs considerable resistance to moisture admission.

SEM confirms the brittleness and smoothness of the surface of the ash at 1000°C

than 800°C and 900°C. The XRD analysis confirmed that the basic forms in poultry litter

ash were CaCO3, SiO2 and K2Ca(CO3)2, which determine the nature of the alkaline

extracts in water and poultry litter ash could be an attractive material for neutralizing

acidic soil and be a good source of material for cement production because of the

availability of high calcium components.

5.2 Willow Pellets

For the study of wood family of lignocellulosic biomass, willow samples, in the

form of mixture of powder and pellets from Ontario, were collected. Collected samples

were dried as per ASTM standard D1762-84 until samples were dried on the Thermo

Scientific muffle furnace with 2416 controller (model: F48055-60) for two hours at

105°C and conducted different types of experiments and the following results were

obtained.

5.2.1 Biomass Characterization

From the results as shown in figures 5-16 to 5-18, volatile matter decreases as the

residence time and torrefaction temperature increases whereas ash contents and amount

of fixed carbon slightly increases as the residence time and torrefaction temperature

increases. The cause of variation is due to the loss of moisture contents of the biomass as

the temperature and residence time increases. The moisture content of the raw sample

was 5.15% and the heating value of it was 16.121 MJ/Kg. Other details of biomass

Page 94: Torrefaction and Pelletization of Different Forms of

79

characteristics are listed in section 5.4 of this report. It was also observed that an increase

in the temperature resulted to decrease in solid product and an increase in the volatile

portion. During the torrefaction, mostly water is produced and the energy content of the

volatiles is mainly preserved in the lipids and organics. Torrefaction testing condition,

environment and properties of samples have a substantial effect on the amount of solid

residue and the volatile and gaseous outputs generated during the process of torrefaction.

Figure 5-16 Volatile Matter vs. Residence Time for Willow

Page 95: Torrefaction and Pelletization of Different Forms of

80

Figure 5-17 Fixed Carbon vs. Residence Time for Willow

Figure 5-18 Ash Contents vs. Residence Time for Willow

Page 96: Torrefaction and Pelletization of Different Forms of

81

5.2.2 Torrefaction

The torrefaction of biomass samples at 210°C, 250°C, 280°C and 300°C resulted

in four products of very light brown, light brown, dark brown and black color. The color

change is mainly attributed to chemical changes of the lignin, i.e. the formation of

chromaphoric groups, mainly the increase of carbonyl groups. The loss of dry matter

(anhydrous weight loss) through volatilization during torrefaction was within 5% at

210°C, about 25% at 250°C and above 35% at 280°C and more than 50% at 300°C. Mass

yield, energy yield and energy density at 210°C, 250°C, 275°C and 300°C are shown in

figure 19 and 20. From the result, it is observed that less Oxygen concentration has more

energy yield and more mass loss than the torrefaction environment with higher

concentration of oxygen. During an inert environment at 250°C, energy yield was near

90% and the mass yield was near 75% which is matching with other finding (Bergman et

al., 2005 and Pimchuai et al., 2010) . At 250°C, maximum energy yield was observed.

Temperature range from 250°C to 285°C is found the optimum temperature range for

perfect torrefaction. Figure 5-20 shows that the energy density continues to increase with

the increase in temperature and residence time. The residence time has insignificant

impact up to 250°C. However, more influence of residence time was seen above 250°C to

300°C. The largest increment ratio of energy density is found to be 1.31 at 300°C for 45

minutes residence time and the lowest ratio is found as 1.02 at 210°C of 15 minutes

residence time. Mainly, moisture and hemicellulose are lost during torrefaction process,

which results in a significant mass loss of raw feedstock without compromising much of

its energy value. Mass yield and energy yield figures also displayed error bars with 4.0-

5.0% standard deviation.

Page 97: Torrefaction and Pelletization of Different Forms of

82

Figure 5-19 % of Mass Yield and Energy Yield vs. Temperature for Willow

(Different Oxygen % and 45 minutes residence time)

Figure 5-20 Energy Density variations with Temperature and residence time

Page 98: Torrefaction and Pelletization of Different Forms of

83

5.2.3 Higher Heating Value

After the specified residence time, samples were removed from the reactor and

placed in a vacuum desiccator to cool. The samples were grinded in a coffee grinder to

prepare sample for Bomb Calorimeter (Model IKA C-200) by IKA, USA. The basis of

complete calculation for the calorific value was based on ASTM D 240 and ASTM D

5865. From the figure 5-21 and 5-22, there is no significance difference on heating value

at an inert environment and 2.4% oxygen concentration. HHV varies from 16 to 21

MJ/Kg with an inert environment which is increase of about 31% more than raw biomass

whereas increase of about 29% observed with 2.4% oxygen concentration. Results show

that torrefaction temperature has more impact on the increase of HHV than residence

time. Moisture reduction and increase in the carbon concentration could be the main

factor contributing to the increase in the heating value.

Figure 5-21 Heating Value vs. Residence Time for Willow at 0% Oxygen

Page 99: Torrefaction and Pelletization of Different Forms of

84

Figure 5-22 Heating Value vs. Residence Time for Willow with 2.4% Oxygen

5.2.4 Hydrophobicity

Figure 5-23 shows that percentage of moisture absorption decreased by multiple

folds after torrefaction and displayed error bars with 2.0 - 3.5% of standard deviation.

The more the torrefaction temperature from 200°C to 300°C and residence time up to 60

minutes, the more the decrease in the percentage of moisture absorption. The lower

moisture could be the result of the tar condensation inside the pores, obstructing the

passage of moist air through the solid, and then avoiding the condensation of water vapor.

Another reason for this could be the polar character of condensed tar on the solid, also

preventing the condensation of water vapor inside the pores. Similarly, the moisture

increases as the torrefaction parameters are increased which is caused by the emission of

volatiles. The emission of volatiles becomes more intensive as torrefied temperature and

residence time is raised which lead to increase the porosity and hygroscopic

Page 100: Torrefaction and Pelletization of Different Forms of

85

characteristics of biomass. For the hydrophobicity behavior, torrefaction at 275°C to 285

°C for any residence time from 30 to 60 minutes shows the best result among tested

results at 0 to 2.4% of oxygen concentration. During torrefaction, depolymerization of the

polymers occurs. The hemicellulose is largely destroyed, disabling the greatest moisture

absorption capacity. According to Verhoeff F.(2011), many oxygen groups such as

hydroxyl, carbonyl and carboxyl are removed from the cell wall polymers during

torrefaction, making room for furan-aromatic, aliphatic structures. With this change in

structure, the hydrophilic groups are replaced by hydrophobic groups, so water is rather

rejected from than attracted to the torrefied biomass. Slight increment of moisture uptake

can be observed above 285°C because of high porosity and hygroscopic characteristics of

biomass at higher temperature. From the experiment, it is seen that torrefied willow at

275°C-285°C acts as good storage behavior and most water repellent so can be good for

storage. It becomes quickly dry even if it is immersed in water. The optimum temperature

for the hydrophobicity is found at 275°C.

Page 101: Torrefaction and Pelletization of Different Forms of

86

Figure 5-23 % of Moisture absorption vs. temperature of torrefied biomass

5.2.5 Storage Behavior

One of the major concerns of raw biomass handling is the storage and

biodegradability by absorbing water from the atmosphere. It is hydrophilic and easily

decomposed with damp during storage. This can be improved by torrefaction converting

hydrophilic characteristic to hydrophobic character. Figure 5-24 shows the moisture

uptake behavior of the torrefied willow with 3.0-4.5% standard deviation of error bars at

different temperature of 30 minutes residence time at 88% relative humidity of a storage

Page 102: Torrefaction and Pelletization of Different Forms of

87

room. It shows that maximum absorption of moisture from the humid environment of the

dried raw biomass is about 4% whereas minimum uptake of moisture from the

atmosphere is performed by the torrefied Willow at 275°C and 300°C. Slight increment

of moisture uptake can be observed above 285°C because of high porosity and

hygroscopic characteristics of biomass at higher temperature. Hence, for indoor or under

the shade storing purpose, torrefied biomass at temperature of 275°C -285°C is found as

optimum temperature for Willow. It absorbs minimum moisture even if it exposes to the

higher atmospheric/surrounding humidity. This is because of hydrophobic nature of

torrefied willow. The samples were stored for one week to observe the biodegradability

of the biomass by placing them in separate plastic bags. After three days, raw biomass is

seen with some biodegradation whereas torrefied at 210°C for 45 minutes was affected

after seven days whereas torrefied biomass above 250°C has not any sign of

biodegradation even after 7 days of storage.

Figure 5-24 % of Moisture uptake vs. temperature of Willow Pellets

Page 103: Torrefaction and Pelletization of Different Forms of

88

5.2.6 Pelletization

Torrefaction could have a strong effect on the mechanical stability and

combustion behavior of any types of biomass (Prins et al., 2006). During torrefaction, the

hydrogen bonding hydroxyl groups are removed which causes reduction in the moisture

uptake of the torrefied biomass as the temperature of torrefaction increases. The

differences in composition and water content have also a strong effect on the pelletizing

properties of biomass. The pelletizing pressure in the press is a crucial parameter in

pelletizing processes in terms of process energy consumption and pellet quality (Gilbert

et al., 2009). Pellet pressure increases drastically when comparing raw biomass and

torrefied biomass. This increase is mostly likely attributed to the lack of water and low

hemicelluloses content in the torrefied biomass. Water acts as a plasticizer, lowering the

softening temperature of biomass. At raw stage, the hemicellulose binds lignin and

cellulose and provides flexibility in the plant cell wall. Their degradation in binding

forces makes easier to break into small particles (Arias et at., 2008). The degradation of

hemicellulose, lignin and cellulose can affect pelletizing properties like friction

coefficient and Poisson ratio (Gilbert et al., 2009). During the pelletization, cumulative

pressure in MN/m2

is calculated and plotted in figure 5-25. This shows pressure to

prepare pellets of 6.4mm size of length of 16mm from raw Willow is about 14 MN/m2

which is much less than the pellet prepared from torrefied willow at 275°C for 30

minutes residence time of 81MN/m2. As the torrefaction temperature increases, the

pressure increases sharply.

During the testing of breaking force of the pellets, it is observed that a rapid

decrease in the pellet compression strength of torrefied pellet as the torrefied temperature

Page 104: Torrefaction and Pelletization of Different Forms of

89

increased. Minimum force is required to break the torrefied pellets of 275°C and 300°C.

Wolfgang et al.(2011) have tested the compression strength and observed similar

behavior to this result for spruce. They found that the breaking force of their samples

decreased both with treatment temperature and residence time and concluded that

strength loss degradation is connected with the degradation of hemicelluloses for the

biomass. For the Willow, decrease in the water contents and above similar reasons could

be the result for the degradation of strength loss. Both plots (b) and (c) also displayed

error bars with 3.5-5.5% of standard deviation.

(a) 3D plot of Pelletization Pressure (MN/m2) and Breaking Force (N)

Page 105: Torrefaction and Pelletization of Different Forms of

90

(b) Pelletization Pressure (MN/m2) vs. Torrefaction Temp at 45 min residence time

(c) Breaking Force (N) vs. Torrefaction Temperature at 45 minutes residence time

Figure 5-25 Pressure for making Pelletization and Force for Breaking Pellets

Page 106: Torrefaction and Pelletization of Different Forms of

91

5.2.7 Scanning Electron Microscope (SEM)

The large content of potassium and chlorine in biomass greatly increases the

deposit formation and corrosion of the thermal power plants, compared to coal fired

boilers. It has been verified that the alkali species which cause the bed agglomeration

come from the biomass ash. So, study on the behaviors of alkali species in biomass ash is

crucial for understanding the formation mechanisms of agglomerates and coating layers.

The ash block can maintain a fixed shape and presents a little compression strength due

to the sintering. Figure 5-26(a), (b) and (c) show the images of the Willow ash samples at

800°C, 900°C and 1000°C respectively. At low temperature, the sample contains some

wool like structures which gradually disappears with the increase in combustion

temperature. This may be due to the presence of combustible particles. This wool like

structures are not combustible at low temperature and can only be combusted at high

temperature. The surface has irregular structure which signifies the process has not been

enough for complete coalescence of the ash into spherical particles. Some of the particles

were long, fiber like agglomerates still resembling the original wood fiber structure. The

surfaces of the particles were usually formed in small sizes which had often sintered

together forming chain-like agglomerate structures. The primary particles were mainly

cotton shaped but also cornered ones were seen like crystalline structure. This signifies,

the ash particles does not contain significant amount of unburned samples. Also at high

temperature, particles become more brittle as the minerals matters in the form of crystal.

Hence the particles become smaller and many cracks start to show up as the combustion

temperature increases. Because of the presence of high Silicon dioxide, Magnesium

dioxide and other alkalis, the structure contains smooth surface which represents the

Page 107: Torrefaction and Pelletization of Different Forms of

92

presence of metallic components in the ash sample. The increase in smooth surfaces in

the samples with the increase in combustion temperature also supports the relative

increase in mineral matters in the samples which is also obtained in XRD analysis and

ash analysis.

(a) SEM for Willow ash at 800C

Page 108: Torrefaction and Pelletization of Different Forms of

93

(b) SEM for Willow ash at 900C

(C) SEM for Willow ash at 1000C

Figure 5-26 SEM for Willow at 800°C, 900°C and 1000°C Ash

Page 109: Torrefaction and Pelletization of Different Forms of

94

5.2.8 X-ray Diffraction (XRD)

X-Ray diffraction tests were carried out on the ash samples of willow as

mentioned in 5.1. The major components of the Biomass are quartz (SiO2), anhydrite

(CaSO4), iron sulfite (FeSO3), Potassium Aluminum Silicate (KAlSiO4), Calcium

Aluminum Silicate (CaAl2SiO6) and hematite (Fe2O3) (Suramya DFM, 2012) . The ashes

were analyzed with X-Ray Diffraction to determine the crystalline components. Figure 5-

27 shows the XRD patterns for willow ash prepared at temperatures 800, 900 and

1000°C. The peaks at 21, 28, 32 and 36 of 2θ represent the crystalline structure of quartz,

which shows the height of the peak increases with the increase in the temperature. This

can be attributing to the increase in relative amount of crystalline structure in the sample

with the increase in combustion temperature. The XRD analysis confirm that the basic

forms in Willow ash are SiO2, CaSO4 and K2Ca(CO3)2, which determine the nature of the

alkaline extracts in water and Willow ash. The ash fusion temperature of willow is about

1100°C -1300°C which is caused due to the presence of alkalis. This obviously limits the

temperature of the practical furnace of a boiler.

Page 110: Torrefaction and Pelletization of Different Forms of

95

Figure 5-27: XRD pattern for Willow ash at different temperature

5.2.9 Summary

The optimum torrefaction treatment temperature was found at 275°C for

hydrophobic characteristics and storage behavior of willow. Because of depolymerization

of the polymers during torrefaction, the hemicellulose is largely ruined leading towards

disabling the largest moisture absorption capacity. The results showed that torrefaction at

higher temperature and residence time had a positive effect on the hydrophobic behavior

by showing smaller amount of water assimilation by the torrefied willow. Pellet pressure

increased by 3-7 folds to prepare pellet from the torrefied products above 250°C. It is not

Page 111: Torrefaction and Pelletization of Different Forms of

96

preferable to have pelletization of willow after the treatment at 285°C. Torrefied willow

pellets are brittle in nature which reduces the grinding energy and make efficient burning

during the combustion process at the thermal plant. Pellets made from torrefied material

have a lower density than pellets made from raw samples. It takes more energy to make

pellets from torrefied than from raw willow. Dried and torrefied particles do not bind

well which can be solved by adding small quantities of water or binders. The severity of

torrefaction increases the degree of hydrophobicity. Torrefied willow at 250°C-285°C

performs considerable resistance to moisture admission. It was observed that HHV

increased by about 31% as the temperature and residence time increases whereas there is

insignificance effect on HHV observed by increasing Oxygen concentration by 2.4%.

SEM confirmed the brittleness and smoothness of the surface of the ash at 1000°C than

800°C and 900°C. The ash fusion started from 1115°C temperature for willow because of

presence of alkalis. The XRD analysis confirmed that the basic forms in willow ash were

SiO2, CaCO4 and K2Ca(CO3)2, which determined the nature of the alkaline extracts in

water and willow ash could be an attractive material for neutralizing acidic for soil and

could be a good source of material for farming and cement production because of the

availability of high silica and magnesium components.

5.3 Oat Pellets

For the study of agricultural lignocellulosic biomass, oat samples from Ontario

were collected. Collected oat was dry and in the form of mixture of powder and raw

pellets form. Collected samples were dried as per ASTM standard and conducted

different types of experiments and the following results were obtained.

Page 112: Torrefaction and Pelletization of Different Forms of

97

5.3.1 Biomass Characterization

From figures 5-28 to 5-30, the moisture content of the raw sample was 7.74%. It

was also observed that an increase in the temperature resulted to decrease in carbon

products and decrease in volatile portion. The product of volatile matter was found from

68-74% which may be due to the limited devolatization and carbonization of

hemicellulose component till the temperature reaches 250°C whereas as the temperature

above 250°C started the decomposition of lignin and cellulose so decrease in volatile

matter was seen between 250°C-300°C. Volatile matter variation was in the range of 1-

10% with the variation of temperature and residence time. This was also the reason for

the poor moisture contents above 250°C. These results are in agreement with other

researchers (Rosillo-Cakke F, 2007; Rousset et al.,2011; Chen and Kuo, 2011; Tumuluru

et al., 2011).

Figure 5-28 Volatile Matters vs. Residence Time for Oats

Page 113: Torrefaction and Pelletization of Different Forms of

98

Figure 5-29 Fixed Carbons vs. Residence Time for Oats

Figure 5-30 Ash Contents vs. Residence Time for Oats

Page 114: Torrefaction and Pelletization of Different Forms of

99

5.3.2 Torrefaction

The torrefaction of biomass samples at 210°C, 250°C, 275°C and 300°C resulted

in four products of very light brown, light brown, dark brown and black color as shown in

Appendix B. The color change may be mainly attributed to chemical changes of the

lignin, i.e. the formation of chromaphoric groups, mainly the increase of carbonyl groups.

The weight loss of dry matter through volatilization during torrefaction was within 3% at

210°C, about 40% at 250°C and above 55% at 275°C and more than 55% at 300°C. Mass

yield, energy yield and energy density at 210°C, 250°C, 275°C and 300°C are shown in

figure 5-31 and 5-32. From the result, it was observed that less Oxygen concentration had

more energy yield and less mass loss than the torrefaction environment with higher

concentration of oxygen. During an inert environment at 250°C, energy yield was near

87% and the mass yield was near 71% at 30 minutes residence time which is matching

with results of other researchers (Woolf et al., 2010; Aries et al., 2008; Rousset et al.,

2006). At 250°C, maximum energy yield was observed. Temperature range from 250°C

to 275°C is found the optimum temperature range for perfect torrefaction. Figure 5-32

showed that the energy density continued to increase with the increase in temperature and

residence time and error bars of positive standard deviation up to 4.5%. The residence

time has insignificant impact up to 250°C. However, more influence of residence time

was seen above 250°C to 300°C. The largest increment ratio of energy density was found

to be 1.44 at 300°C for 45 minutes residence time and the lowest ratio was found as 1.00

at 210°C of 15 minutes residence time. Mainly, moisture and hemicellulose are lost

during torrefaction process, which results in a significant mass loss of raw feedstock

without compromising much of its energy value.

Page 115: Torrefaction and Pelletization of Different Forms of

100

Figure 5-31 % of Mass Yield and Energy Yield vs. Temperature for Oats

(Different Oxygen % and 45 minutes residence time)

Figure 5-32 Energy Density variations with Temperature and residence time

Page 116: Torrefaction and Pelletization of Different Forms of

101

5.3.3 Higher Heating Value (HHV)

From the figure 5-33 and 5-34, there was no significance difference on heating

value at an inert environment and 2.4% oxygen concentration. HHV varies from 16 to 24

MJ/Kg with an inert environment which was increase of about 44% more than raw

biomass whereas increase of about 42% observed with 2.4% oxygen concentration.

Results showed that torrefaction temperature had more impact on the increase of HHV

than residence time. Moisture reduction and increase in the carbon concentration could be

the main factor contributing to the increase in the heating value.

Figure 5-33 Heating Value vs. Residence Time for Oats at 0% Oxygen

Page 117: Torrefaction and Pelletization of Different Forms of

102

Figure 5-34 Heating Value vs. Residence Time for Oats with 2.4% Oxygen

5.3.4 Hydrophobicity

Figure 5-35 showed that percentage of moisture absorption decreased by multiple

folds after torrefaction and displayed error bars with 2.75-4.0% standard deviation. The

more the torrefaction temperature from 200°C to 300°C and residence time up to 60

minutes, the more the decrease in the percentage of moisture absorption. The lower

moisture could be the result of the tar condensation inside the pores, obstructing the

passage of moist air through the solid, and then avoiding the condensation of water vapor.

Similarly, it was observed that the moisture absorption increases as the torrefaction

parameters were increased. For the hydrophobicity behavior, torrefaction at 270°C to 285

°C for any residence time from 30 to 60 minutes showed the optimum result among

tested data. The optimum temperature for the hydrophobic behavior for oats was found at

270°C. Raw oats absorbed more than 70% of moisture whereas torrefied oats at 270°C

Page 118: Torrefaction and Pelletization of Different Forms of

103

absorbs only about 2% of moisture. According to Verhoeff F.(2011), many oxygen

groups such as hydroxyl, carbonyl and carboxyl are removed from the cell wall polymers

during torrefaction, making room for furan-aromatic, aliphatic structures. With this

change in structure, the hydrophilic groups are replaced by hydrophobic groups, so water

is rather rejected from than attracted to the torrefied biomass. Slight increment of

moisture uptake could be observed above 285°C because of high porosity and

hygroscopic characteristics of biomass at higher temperature. From the experiment, it is

seen that torrefied oats at 270°C-285°C acts as optimum storage behavior and most water

repellent so can be good for storage. It became quickly dry even if it was immersed in

water.

Figure 5-35 Moisture absorption vs. temperature of torrefied biomass

Page 119: Torrefaction and Pelletization of Different Forms of

104

5.3.5 Storage Behavior

One of the major concerns of raw biomass handling is the storage and

biodegradability by absorbing water from the atmosphere. It is hydrophilic and easily

decomposed with damp during storage. This can be improved by torrefaction converting

hydrophilic characteristic to hydrophobic character. Figure 5-36 showed the moisture

uptake behavior of the torrefied oats with 2.5-5.0% of standard deviation of error values

at different temperature of 30 minutes residence time at 88% relative humidity of a

storage room. It showed that maximum absorption of moisture from the humid

environment of the dried raw biomass was about 10% whereas minimum about 2%

uptake of moisture from the atmosphere was performed by the torrefied oats at 275°C and

300°C. Slight increment of moisture uptake could be observed above 285°C because of

high porosity and hygroscopic characteristics of biomass at higher temperature. Hence,

for indoor or under the shade storing purpose, torrefied biomass at temperature of 275°C

-285°C was found as optimum temperature for oats. It absorbed minimum moisture even

if it was exposed to the higher atmospheric/surrounding humidity. This may be because

of hydrophobic nature of torrefied products. The samples were stored for one week to

observe the biodegradability of the biomass by placing them in separate plastic bags.

After three days, raw biomass was seen with some biodegradation whereas torrefied at

210°C for 45 minutes was affected after seven days whereas torrefied biomass above

250°C had not any sign of biodegradation even after 7 days of storage.

Page 120: Torrefaction and Pelletization of Different Forms of

105

Figure 5-36 % of Moisture uptake vs. temperature of Oats Pellets

5.3.6 Pelletization

During the pelletization, cumulative pressure in MN/m2

was calculated and

plotted in figure 5-37. This showed pressure to prepare pellets of 6.4 mm size of length of

8-16 mm from raw oat was about 9 MN/m2

which was much less than the pellet prepared

from torrefied oats at 275°C for 45 minutes residence time of 65 MN/m2.

During the testing of breaking strength of the pellets, it was observed that a rapid

decrease in the pellet compression strength of torrefied pellet as the torrefied temperature

increased. Minimum force 51N and 7N were required to break the torrefied pellets of

275°C and 300°C respectively whereas 1101N was required to break the raw pellets. For

the oats, decrease in the water contents and degradation of lignin and hemicellulose could

be the reasons for getting such result for the degradation of strength loss. Both figures

displayed error bars up to 5% standard deviation.

Page 121: Torrefaction and Pelletization of Different Forms of

106

(a) Pelletization Pressure vs. Torrefaction Temperature at 45 min residence time

(b) Breaking Force vs. Torrefaction Temperature at 45 minutes residence time

Figure 5-37 Pressure for making Pelletization and Force for Breaking Pellets

Page 122: Torrefaction and Pelletization of Different Forms of

107

5.3.7 Scanning Electron Microscope (SEM)

Fig. 5-38 showed a close-up view of ash sample. The ash block can maintain a

fixed shape and presents a little compression strength due to the sintering. Researcher

(Skrifvars et al., 1997) characterized the sintering tendency of ten biomass ashes and

classified the ash components into three groups: simple alkali salts, silicates and the rest

non-melt. We could see that the fusible compounds melt and coat the surface of ash

particles during combustion. The spot analyses indicated that these melts were rich in K

or Na. From the result of ash fusion analysis, the ash fusion temperature of oats started

from 1279°C. The alkali silicates formed during combustion can melt and coat the

surface of ash particles at high temperature. In addition, the large-sized ash particles may

increase the amount of melts on local surface of bed particle and may act as the necks for

the agglomerate formation. So, the large- and small-sized ash particles may play the

different roles in the bed agglomeration.

Figure 5-38 (a), (b) and (c) showed the images of the oats ash samples at 800°C,

900°C and 1000°C respectively. At low temperature, the sample contained some wool

like structures which gradually disappears with the increase in combustion temperature.

This may be due to the presence of combustible particles. This broken wood like

structures was not combustible at low temperature and can only be combusted at high

temperature. The surface had irregular structure which signified the process had not been

enough for complete coalescence of the ash into spherical particles. Some of the particles

were long, fiber like agglomerates still resembling the original oats fiber structure. At

900°C, the ash metallic components were formed and hole like spots were observed. The

primary particles were mainly cotton shaped but also cornered ones were seen like

Page 123: Torrefaction and Pelletization of Different Forms of

108

crystalline structure. This signified, the ash particles did not contain significant amount of

unburned samples. Also at high temperature, particles became more brittle as the

minerals matters in the form of crystal. Hence the particles became smaller and many

cracks start to show up as the combustion temperature increases. Because of the presence

of high Silicon dioxide, Magnesium dioxide and other alkalis, the structure contained

smooth surface which represented the presence of metallic components in the ash sample.

The increase in smooth surfaces in the samples with the increase in combustion

temperature supported the relative increase in mineral matters in the samples which was

also obtained from ash analysis.

(a) SEM for Oats ash at 800°C

Page 124: Torrefaction and Pelletization of Different Forms of

109

(b) SEM for Oats ash at 900°C

(c) SEM for Oats ash at 1000°C

Figure 5-38 SEM for Oats at 800°C, 900°C and 1000°C Ash

Page 125: Torrefaction and Pelletization of Different Forms of

110

5.3.8 X-ray Diffraction (XRD)

Figure 5-39 shows the XRD patterns for oats ash prepared at temperatures 800°C,

900°C and 1000°C. The peaks at 25 and 45 of 2θ represent the crystalline structure of

quartz, which shows the highest peak is observed at 900°C and then followed by 800°C

and 1000°C. The XRD analysis confirm that the basic forms in oats ash are SiO2, CaSO4

and K2Ca(CO3)2, which determine the nature of the alkaline extracts in water and oats

ash. The ash fusion temperature of oats is about 1250°C -1350°C which is caused due to

the presence of alkalis. This obviously limits the temperature of the practical furnace of a

boiler.

Figure 5-39: XRD pattern for Oats ash at different temperature

Page 126: Torrefaction and Pelletization of Different Forms of

111

5.3.9 Summary

Torrefied products produced from oat were characterized in the view of

combustion applications. The optimum treatment temperature was found at 270°C for

hydrophobic characteristics and storage behavior of oats. Because of depolymerization of

the polymers during torrefaction, the hemicellulose was largely ruined leading towards

disabling the largest moisture absorption capacity. The results showed that torrefaction at

higher temperature and residence time had a positive effect on the hydrophobic behavior

by showing smaller amount of water assimilation by the torrefied oats. Pellet pressure

increased by 4-7 folds to prepare pellet from the torrefied products above 250°C. It is not

preferable to have pelletization of biomass after the treatment at 285°C. Torrefied pellets

were brittle in nature which could reduce the grinding energy and make efficient burning

during the combustion process at the thermal plant. Pellets made from torrefied material

had a lower density than pellets made from raw samples. It took more energy to make

pellets from torrefied than from raw oats. Dried and torrefied particles did not bind well

which could be solved by adding small quantities of water or binders. The severity of

torrefaction increases the degree of hydrophobicity. Fuels made from torrefied oats at

250°C-285°C perform considerable resistance to moisture admission. It was observed

that HHV increased by about 43% as the temperature and residence time increases

whereas there was insignificant effect on HHV observed by increasing oxygen

concentration by 2.4%. SEM confirmed the brittleness and smoothness of the surface of

the ash at 1000°C than 800°C and 900°C. The ash fusion started from 1279°C

temperature for oats because of presence of alkalis. The XRD analysis confirmed that the

basic forms in oats ash are SiO2, Al2O3 and K2O which determine the nature of the

Page 127: Torrefaction and Pelletization of Different Forms of

112

alkaline extracts in water and oats ash could also be an attractive material for neutralizing

acidic for soil and could be a good source of material for farming and cement production

because of the availability of high silica and magnesium components like other

lignocellulosic biomass.

5.4 Comparative Analysis

5.4.1 Biomass Characterization

a) Proximate Analysis of Raw Biomass

The results of the proximate analysis are listed in the following Table 5-1 which indicated

that poultry litter had the highest moisture contents and willow pellets had the lowest.

The ash contents on the poultry litter had more than 20% with minimum carbon contents

which consequently made the lowest heating value of it.

Table 5-1 Proximate Analysis of Raw Biomass

Sample

Moisture

(%)

Volatile Matter

(%) Ash Contents (%) Fixed Carbon (%)

Poultry

litter 20.1 54.29 22.28 3.33

Willow

Pallets 5.15 73.18 11.45 10.22

Oat Pallets 7.74 74.28 5.64 12.34

Page 128: Torrefaction and Pelletization of Different Forms of

113

b) Ultimate Analysis:

Results of the ultimate analysis of the raw biomass were given in the Table 5-2.

All of raw biomass contained more than 40% of carbon, more than 5% of hydrogen but

poultry litter had the lowest oxygen less than 6% and the highest sulfur with more than

1%.

Table 5-2 Ultimate Analysis and Heating Value of Raw Biomass

Components

Poultry

Litter Willow Pallets Oat Pallets

Carbon 43.30% 50.65% 52.23%

Hydrogen 6.62% 5.86% 6.59%

Nitrogen 5.72% 0.52% 0.62%

Sulphur 1.15% 0.44% 0.29%

Oxygen 5.95% 24.07% 33.98%

Heating Value

MJ/Kg 10.88 16.12 16.41

From the figure 5-40, the highest components in all biomass were observed as

carbon with more than 40% whereas minimum component was Sulfur with less than 1%.

Lignocellulosic biomass contained more carbon than non-lignocellulosic which could be

because of photosynthesis (carbon capture nature) of biomass. Poultry litter has more

sulfur and nitrogen whereas it contained minimum oxygen. Oats contained maximum

carbon and oxygen among all. All biomass has almost similar percentage of hydrogen.

Page 129: Torrefaction and Pelletization of Different Forms of

114

Figure 5-40. Comparative Study of Ultimate Analysis of different Biomass

c) Ash Fusion Temperature

From figure 5-41, Poultry litter had highest initial deformation temperature of

about 1400°C whereas willow had the lowest at 1150°C. Oats had about 1300°C. This

could be because of higher alkali contents in the lignocellulosic biomass and high amount

of calcium components in the poultry litter.

Page 130: Torrefaction and Pelletization of Different Forms of

115

Figure 5-41. Comparative Study of Ash Fusion Temperature of different Biomass

d) Elemental Analysis

From figure 5-42, it was observed that Poultry Litter contained CaO of about 65%

(daf) whereas willow and oats contained more than 60% (daf) of SiO2. This may be due

to high calcium contents in the food of poultry and high silica contents in the soil of

Ontario for the lignocellulosic biomass. In the second components, poultry litter contains

more than 18 % of P2O5, Willow contains 20% of MgO and 15% of AlO3 which

symbolized that woody biomass had more magnesium components and agricultural

biomass contained more aluminum. Similarly, all biomass contained alkalis K2O and

Na2O.

Page 131: Torrefaction and Pelletization of Different Forms of

116

Figure 5-42. Comparative Study of Elemental Analysis of different Biomass

5.4.2 Torrefaction

a) Mass Yield

From figure 5-43, poultry litter lost less weight than other two lignocellulosic

biomasses with the increase in temperature and residence time and error bars showed 3.0-

4.5% variation on standard deviation. This was due to moisture loss of all the biomasses

and fast devolatilisation of hemicellulose of lignocellulosic of biomasses. Most severe

was the agricultural residue oats because of the loose bonding of the hemicellulose

components than willow.

Page 132: Torrefaction and Pelletization of Different Forms of

117

Figure 5-43. Comparative Study of Mass Yield of Poultry Litter, Willow and Oat

Pellets at different temp with 45 minutes residence time

b) Energy Yield

From figure 5-44, maximum energy yield was seen on the sample of poultry litter

with more than 80% because of its high mass contents and low heating values whereas

the oats had the lowest energy yield with about 60%. Mass decomposition of agricultural

residue was found the fastest than willow and poultry litter. This is due to lower contents

of calcium and silica in oats than other biomass. Oats has the weak bonding of

hemicellulose than willow so faster evaporation and devolatilisation occurs than other

biomasses. Error bars showed 3.5-5.5% of standard deviation.

Page 133: Torrefaction and Pelletization of Different Forms of

118

Figure 5-44. Comparative Study of Energy Yield of Poultry Litter, Willow and Oat

Pellets at different temp with 45 minutes residence time

c) Higher Heating Value

From figure 5-45, oats had highest heating value of about 24 MJ/Kg and poultry

litter had the lowest with 12 MJ/Kg while willow had about 22 MJ/Kg. It was seen that

higher heating values increased with temperature. The highest heating value of all the

samples were achieved at 300°C because of the removal of moisture and devolatization

of hemicellulose of woody biomass. However, no significant variation of heating values

was observed after the 275°C. This signifies higher heating value with torrefaction

temperature at 250°C -275°C can be justified. Error bars showed 3.0-5.0% of standard

deviation.

Page 134: Torrefaction and Pelletization of Different Forms of

119

Figure 5-45. Comparative Study of Heating Values of different Biomass with temp

5.4.3 Hydrophobicity

From figure 5-46, it was observed that oats had the highest moisture absorption

capacity than other two samples. Torrefied oats at 250°C absorbed about 35% whereas

other two samples absorbed only about 7%. All samples performed the minimum

absorption at torrefaction temperature 275°C out of which poultry litter showed the best

performance of hydrophobicity. With the increase of temperature beyond 285°C,

devolatization of hemicellulose and lignin occurred which consequently increased the

porosity of the torrefied biomass and resulted little more absorption of moisture. Hence,

270-285°C was found optimum temperature range for torrefaction for poultry litter, oats

and willow. It also displayed error bars with 2.5-4.5% of standard deviation.

Page 135: Torrefaction and Pelletization of Different Forms of

120

Figure 5-46. Comparative Study of Hydrophobicity of Poultry Litter, Willow and

Oat at different temp with 45 minutes residence time

5.4.4 Storage Behavior and Moisture Uptake

From the figure 5-47, the optimum moisture uptake was observed on all of the

samples with 4.25-5.0% standard deviation on error bars. Willow and oats performed the

optimum moisture uptake at 275°C. Storing of torrefied biomass at 275°C was found as

optimum. Torrefaction decomposes the hydro-oxy components from the biomass and

make more hydrophobic in nature. This effect could be more severe in the lignocellulosic

biomass than non-lignocellulosic biomass.

Page 136: Torrefaction and Pelletization of Different Forms of

121

Figure 5-47. Comparative Study of Storage Behavior of Poultry Litter, Willow and

Oat Pellets at different temp with 45 minutes residence time

5.4.5 Ash Analysis

Elemental ash analysis was done based on the ASTM standard and find from

figure 5-48 that Poultry litter contained excessive amounts 65.17% of MgO and 17.46%

of P2O5 whereas only 9.4% and 5.59% of MgO and 1.74% and 4.54% of P2O5 in willow

and oats respectively. Lignocellulosic biomass willow and oats contains more than 60%

of SiO2. Willow contains 20.7% of MgO and 9.47% of Al2O3 whereas oats contains

15.62% of Al2O3 and 9.85% of K2O. The major compounds in poultry litter identified in

the ash are CaO (65%), P2O5 (17.5%) and K2O (6.4%), which falls into the category of

biomass ash with rich in calcium, phosphorous. The concentration of alkali metals

sodium oxide/Na2O, and potassium oxide/K2O were higher in poultry litter than in

willow and oats, indicating that poultry litter is a more challenging fuel than

lignocellulosic biomass for combustion applications. High alkali content, especially in

Page 137: Torrefaction and Pelletization of Different Forms of

122

conjunction high chloride levels, results in a high potential for slagging, fouling,

particulate emissions, and corrosion. Silica salts formed by K and Na show strong

tendencies to become sticky and form slag on the hot surfaces of the combustion

equipment and the boiler. Maintaining low combustion temperatures will also help in

controlling alkali-related slagging and fouling problems.

Figure 5-48. Comparative Study of Elemental Analysis of Poultry Litter, Willow

and Oat Pellets at different temp with 45 minutes residence time

From figure 5-49, it is observed that out of tested three samples, only poultry litter

contained very small amount of sulfur. Lignocellulosic biomass samples of willow and

oats had lower ash initial deformation temperature 1115°C and 1279°C respectively than

non-lignocellulosic biomass poultry litter at 1421.67°C. This may be due to higher

contents of alkalis in the lignocellulosic biomass than non lignocellulosic biomass. The

Page 138: Torrefaction and Pelletization of Different Forms of

123

high ash content in poultry litter may require high-volume ash-handling equipment and

more attention to particulate removal, slagging, and fouling while used in a

combustor/boiler.

Figure 5-49. Comparative Study of Ash Fusion Temperature of Poultry Litter,

Willow and Oat Pellets at different temp with 45 minutes residence time

5.4.6 Summary

After torrefaction, the maximum heating value of 24 MJ/Kg for oats, 22 MJ/Kg

for willow and 12 MJ/Kg for poultry litter were found. Mass yield varied from 42%-91%

whereas energy yield varied from 61%-89% with operating temperature and residence

time. Oats showed the fastest mass and energy yield whereas poultry litter showed the

least. For hydrophobicity and moisture uptake, the optimum temperature were found at

285°C for willow, 270°C for oat and 275°C for poultry litter at 45 minutes of residence

time. It was observed that all torrefied products showed hydrophobic character and

Page 139: Torrefaction and Pelletization of Different Forms of

124

remain unaffected from biodegradation when they were immersed in water after

torrefaction.

5.5 Errors and Repeatability Test

Each test was carried out at least for three times in each experiment. During the

torrefaction, the oxygen inert environment was measured by gas analyzer. The accuracy

of the analyzer was 0.5% of full scale volume and the resolution was 0.01% of volume.

During the control of the oxygen concentration of the torrefaction environment, the

nitrogen gas was released using flow meter which had the accuracy of 0.5% in the flow

volume. The measurement of repeatability was carried out during the 2.4% oxygen two

times and measured 2.4%±0.0069 and 2.39%±0.018 at two different dates of test. This

may be due the environmental condition, human activities (opening and closing moment)

and technical variation of the equipment itself.

During the hydrophobicity and moisture uptake, two similar experiments were

performed for different sample in similar condition and the results are presented in the

appendix G.

Page 140: Torrefaction and Pelletization of Different Forms of

125

Chapter VI: Conclusions and Recommendations

6.1 Conclusions

Poultry litter (non-lignocellulosic biomass), oat and willow wood (lignocellulosic

biomass) samples from Ontario were collected and torrefied in a locally designed and

fabricated lab-scale reactor. A number of torrefaction experiments were conducted to

characterize different samples at different torrefaction temperatures, ranging from 200°C

to 300°C, residence time ranging from 15 minutes to 60 minutes, and different oxygen

concentrations (from about 0% up to 2.4%) in the torrefaction chamber. Results indicated

that the effect of oxygen concentration on the mass loss, energy loss and higher heating

value were not significant.

During torrefaction, both lignocellulosic and non-lignocellulosic biomass showed

a decrease in moisture content, increase in HHV and decrease in weight as the

temperature and residence time increased. Increase in temperature and residence time

also improved the hydrophobic characteristics of both biomasses. From the HHV and

hydrophobicity point of view, torrefaction at a temperature of 260°C-285°C for a

residence time of 20-40 minutes is found to be the optimum condition for all three

biomasses tested.

After torrefaction, a maximum higher heating value of 24 MJ/Kg for oats, 22

MJ/Kg for willow and 12 MJ/Kg for poultry litter were obtained. Mass yield varied from

42- 91% while energy yield varied from 61- 89% at different operating temperatures and

residence times. Oats showed the fastest mass and energy yields whereas poultry litter

showed the least. For hydrophobicity and moisture uptake, the optimum temperatures

found were 285°C for willow, 270°C for oat and 275°C for poultry litter at 45 minutes of

Page 141: Torrefaction and Pelletization of Different Forms of

126

residence time. It was observed that all torrefied products showed hydrophobic

characteristics and remained resistant to moisture re-absorption when they were

immersed in water after torrefaction.

Due to the depolymerization of the polymers during torrefaction, the

hemicellulose was largely removed from the biomass, causing the loss of moisture

absorption capacity. The results showed that torrefaction at higher temperatures and

residence times had a positive effect on the hydrophobic behavior by showing smaller

amount of water assimilation by the torrefied lignocellulosic products. Pelletizing

pressure increased by multiple folds to make pellets from the torrefied willow and oats at

about 250°C. Hence, it took more energy to make pellets from torrefied lignocellulosic

biomass than from raw biomass. It was not preferable to have pelletization of willow and

oats after the treatment at 285°C. Dried and torrefied particles do not bind well, which

can be solved by adding small quantities of water or binders. Torrefied pellets are brittle

in nature which reduces the grinding energy and makes efficient burning during the

combustion process at the thermal plant. Pellets made from torrefied material had a lower

density than pellets made from the raw samples. The level of hydrophobicity was found

to increase with the degree of torrefaction.

SEM analysis of the biomass ash confirmed the brittleness and smoothness of the

surface of the ash at 1000°C rather than at 800°C or 900°C. The XRD analysis indicated

that the basic constituents in poultry litter ash were CaCO3, SiO2 and K2Ca(CO3)2,

whereas SiO2, Al2O3 and K2O in oats and willow ash. These determine the nature of the

alkaline extracts in water. Biomass ash is an attractive material for neutralizing acidic soil

and could be a good source of material for cement production because of the availability

Page 142: Torrefaction and Pelletization of Different Forms of

127

of high calcium components. The ash fusion starts from 1115°C temperature for willow,

1279°C for oats, because of the presence of alkalis, and 1421°C for poultry litter.

6.2 Recommendations

Torrefaction has been identified as a key biomass pretreatment/upgrading process

that can offer significant advantages in terms of improvements in storage stability,

hydrophobicity and chemical properties important for thermochemical processes such as

co-firing with coals and gasification. However, more advanced design of commercial

torrefiers are yet to be developed. Multipurpose torrefier and gasifier appears to offer

promise for optimized processing of biomass. From the literature review, followings are

the gaps still available in the torrefaction research:

1) Optimization on the degradation of hemicellulose, lignin and cellulose during the

process of torrefaction.

2) Molecular level understanding during the process of torrefaction like molecular

breaking energy.

3) Study on the changing of colors during torrefaction using advanced device like

colorimeter.

4) Design and development of integrated processes of torrefier, combustion and

gasification for potential applications.

5) Study of the storage behavior of torrefied products in terms of off-gassing and

spontaneous combustion at different temperature.

6) Life cycle assessment of the torrefied biomass in the application of electrical

power generation.

Page 143: Torrefaction and Pelletization of Different Forms of

128

7) Study on the explosibility of torrefied biomass and its measures to rectify it.

8) Study of torrefaction and pelletization integrated process

9) Design of economical commercial torrefier and its challenges

10) Economic analysis on the use of torrefied biomass for co-firing with coal or

replacing coal from the electrical power plants.

Page 144: Torrefaction and Pelletization of Different Forms of

129

Chapter VII: References

1. Acharjee T.C., Coronella C. J., and Vasquez V.R., (2011) “Effect of thermal

pretreatment on equilibrium moisture content of lignocellulosic biomass,”

Bioresource Technology Journal, Vol. 102, No. 7, pp. 4849 – 4854

2. Acharya B., Sule I. & Dutta A., (2012) “A review on advances of torrefaction

technologies for biomass processing”, Biomass Conversion and Biorefinery,

December, Volume 2, Issue 4, pp 349-369.

3. Arias, B., Pedida, C., Fermoso, J., Plaza, M. G., Rubiera, F., and Pis, J. J, (2008)

“Influence of torrefaction on the grindability and reactivity of woody biomass,”

Fuel Process Technology Journal, Vol. 89, No. 2, pp. 169–175.

4. Basu, P., (2006) “Combustion and Gasification in Fluidized Beds,” Taylor &

Francis, USA

5. Basu, P., (2010) “Biomass Gasification and Pyrolysis - Practical Design and

Theory,” Academic Press, New York, pp- 92-96.

6. Bergman PCA, Boersma AR, Kiel JHA, Prins MJ, Ptasinski KJ, and Janssen

FGGJ (2005) “Torrefied biomass for entrained-flow gasification of biomass”,

Report ECN-C--05-026, ECN.

7. Bergman PCA, Boersma AR, Zwart RWH, and Kiel JHA (2005) “Torrefaction

for biomass co-firing in existing coal-fired power stations”, Report ECN-C--05-

013, ECN.

8. Bergman, P. C.A, and Kiel, J.H.A., (2005) ‘Torrefaction for biomass upgrading”,

Published at 14th European Biomass Conference & Exhibition, Paris, France, 17-

21 October.

Page 145: Torrefaction and Pelletization of Different Forms of

130

9. Bioenergy Update (2000), Vol. 2 No. 4 https://www.bioenergyupdate.com/

magazine/security/NL0400/bioenergy_update_april_2000.htm (Cited on January

15, 2012)

10. Bridgeman T.G., Jones J.M., Williams A., and Waldron D.J., (2010) “An

investigation of the grindability of two torrefied energy crops,” Fuel journal, Vol.

89, No. 12, pp. 3911 – 3918

11. Bridgeman T.G., Jones, J.M., Shield, I., and Williams P.T., (2008) “Torrefaction

of reed canary grass, wheat straw and willow to enhance solid fuel qualities and

combustion properties”, Journal of Fuels, Vol. 87, pp. 844 – 856.

12. Beatrice Coda, Mariusz K. Cieplik, Paul J. de Wild, and Jacob H. A. Kiel (2007)

“Slagging Behavior of Wood Ash under Entrained-Flow Gasification

Conditions”, Energy & Fuels 21, 3644–3652

13. Chen W. and Kuo P., (2011) “Torrefaction and co-torrefaction characterization of

hemicellulose, cellulose and lignin as well as torrefaction of some basic

constituents in biomass,” Energy Journal, Vol. 36, No. 2, pp. 803 – 811

14. Chen W-H, Cheng W-Y, Lu K-M, Huang YPA.,(2011) “In evaluation on

improvement of pulverized biomass property for solid fuel through torrefaction.

Applied Energy 88:3636–44.

15. Chen W-H, Hsu H-C, Lu K-M, Lee W-J, Lin T-C. (2011) “Thermal pretreatment

of woo(Lauan) block by torrefaction and its influence on the properties of the

biomass”. Energy; 36: 3012–21.

16. Cheng G., Varanasi P., Li C., Liu H., Melnichenko YB, Simmons BA, Kent MS

and Singh S., (2011) Transition of Cellulose Crystalline Structure and Surface

Page 146: Torrefaction and Pelletization of Different Forms of

131

Morphology of Biomass as a Function of Ionic Liquid Pretreatment and its

Relation to Enzymatic Hydrolysis, ACS Publications,

dx.doi.org/10.1021/bm1012.40z/Biomacromolecules. (Cited on July 17, 2012)

17. Chew, J.J, Doshi V.,(2011) “Recent advances in biomass pretreatment –

Torrefaction fundamentals and technology”, Renewable and Sustainable Energy

Reviews 15 (2011) 4212–4222

18. Ciolkosz D. and Wallace R.,(2011) “A review of torrefaction for bioenergy

feedstock production”, Biofuels, Bioproducts and Biorefining, Vol. 5, No. 3, pp.

317–329.

19. Dangtran K.Y., Mullen J.F., and Mayrose D.T.,(2000) “A Comparison of Fluid

Bed and Multiple Hearth Biosolids Incineration”, The 14th Annual Residuals &

Sludge Management Conference.

20. De la Hoz A., Dı´az-Ortiz A. and Moreno A.,(2005) “Microwaves in organic

synthesis. Thermal and non-thermal microwave effects”, Chemical Society

Reviews, Vol. 34, pp 164–178

21. Deng J, Wang G-j, Kuang J-h, Zhang Y-l, Luo Y., (2009) “Pretreatment of

agricultural residues for co-gasification via torrefaction”. Journal of Analytical

and Applied Pyrolysis; 86:331–7.

22. Douglas, B., (2009) “Canada Report on Bioenergy,” (Online Material) URL:

http://www.canbio.ca/documents/publications/canadacountryreport2009.pdf

(Cited on March 13, 2012)

23. Felfli F.F., Luengo C.A., Soler P.B., and Rocha J.D.,(2004) “Mathematical

Modelling of Woods and Briquettes Torrefaction, Proceedings of the 5th

Page 147: Torrefaction and Pelletization of Different Forms of

132

Encontro de Energia no Meio Rural, Campinas, Spain, October 19–2

http://www.feagri.unicamp.br/energia/agre2004/Fscommand/PDF/Agrener/Trabal

ho%205.pdf (Cited on October 11, 2011)

24. Felfli FF, Luengo CA, Suárez JA, Beatón PA.,(2005) “Wood briquette

torrefaction. Energy for Sustainable Development ;9:19–22.

25. FGC Group, (2010) URL: http://www.fgcgroupllc.com/ multiple hearth_

furnaces.html (Cited on October 09, 2011)

26. Gilbert P, Ryu C, Sharifi V, Swithenbank J. (2009) “Effect of process parameters

on pelletization of herbaceous crops. Fuel; 88:1491–7.

27. Govin A., Repellin V., Rolland M., and Duplan J., (1988) “Effect of torrefaction

on grinding energy requirement for thin wood particle production,” (online

materials), URL: http://hal.archives-ouvertes.fr/docs/00/46/23/39/PDF/AG-

SFGP09.pdf (Cited on October 09, 2011)

28. Helwig T., Jannasch R., Samson R., DeMaio A., and Caumartin D.,(2010)

“Agricultural Biomass Residue Inventories and Conversion Systems for Energy

Production in Eastern Canada”, Prepared for Natural Resources – Canada (online

material) URL: http://www.reap Canada.com/online_library/feedstock

biomass/7Agricultural%20Biomass%20Residue%20Inventories%20and%20Conv

ersion..Samson%20et%20al.%202002.pdf (Cited on October 10, 2011)

29. Holley C.A.,(1983) “The densification of biomass by roll briquetting,”

Proceedings of the Institute for Briquetting and Agglomeration (IBA), Vol. 18,

No., pp 95–102.

Page 148: Torrefaction and Pelletization of Different Forms of

133

30. http://www.ecotechenergygroup.com/index.php/alternative-energy; (Cited on

November 03, 2012)

31. IEA. Bioenergy (2010) “Technology roadmaps”, International Energy Agency

(IEA); (Cited on December 21, 2011)

32. Jenkins B.M., Baxter L.L., and Miles Jr. T.R.,(1998) “Combustion properties of

biomass”, Fuel Processing Technology, Vol. 54, pp 17 – 46.

33. Kaliyan N, Vance Morey R., (2009) “Factors affecting strength and durability of

densified biomass products”, Biomass and Bioenergy;33:337–59.

34. Khodier AHM., (2011) “Co-firing fossil fuels and biomass: combustion,

deposition and modeling”, School of applied sciences, PhD Thesis, Academic

January 2011, CRANFIELD UNIVERSITY

35. Kiel J.H.A., Verhoeff F, Gerhauser H and Meuleman B., (2008) “BO2

Technology for Biomass Upgrading into solid fuel-Pilot Scale Testing and Market

Implementation”, Energy research Centre of the Netherlands (ECN), P.O. Box

1755 ZG Petten, the Netherlands.

36. Kiel JHA (2007) “Torrefaction for biomass upgrading into commodity fuels”, In:

Proceedings of the IEA Bioenergy Task 32 Workshop on Fuel Storage, Handling

and Preparation and System Analysis for Biomass Combustion Technologies.

Berlin, Germany.

37. Kleinschmidt C.P.,(2011) “Overview of international developments in

torrefaction,” Bio-energy Trade, Torrefaction workshop, January, online material

(Cited on January 26, 2012) URL:

http://www.bioenergytrade.org/downloads/grazkleinschmidtpaper2011.pdf

Page 149: Torrefaction and Pelletization of Different Forms of

134

38. Kuang X, Tumuluru JS, Bi XT, Lim CJ, Sokhansanj S, and Melin S. (2009) “Rate

and peak concentrations of off-gas emissions in stored wood pellets —

sensitivities to temperature, relative humidity, and headspace volume”, Ann

Occup Hyg 53(8), 789–796.

39. Kudra T., and Mujumdar A.S., (2002) “Advanced Drying Technologies,” Marcel

Dekker, Inc. New York, pp. 69-70, 81 – 83, and 335 – 336

40. Kumar A., Cameron J.B., and Flynn P.C., (2003) “Biomass power cost and

optimum plant size in western Canada,” Biomass and Bioenergy Journal, Vol. 24,

No. 6, pp. 445 – 464

41. Lam PK, Sokhansanj S, Bi T, and Lim CJ.,(2011) “Quantifying color of steam

treated pellets made from western Douglas Fir (Pseudotsuga Menziesii, L.). Trans

ASABE.

42. Leonelli C. and Mason T.J.,(2010) “Microwave and ultrasonic processing: Now a

realistic option for industry”, Chemical Engineering and Processing Journal, Vol.

49, No. 9, pp 885 – 900

43. Linghong, Z., Chunbao, C.X., and Pascale, C.,(2010) “Overview of recent

advances in thermo-chemical conversion of biomass”, Energy, 51, 969 – 982.

44. Magalhaes A, Petrovic D, Rodriguez A, Putra Z and Thielemans G,(2009)

“Techno economic assessment of biomass pre-conversion processes as a part of

biomass-to-liquids line-up. Biofuels, Bioprod Bioref 3:584–600.

45. Mani S., (2009) “Integrating biomass torrefaction with thermo-chemical

conversion processes,” Proceedings of the Annual Meeting of AIChE, Nashville,

TN. Nov 8 – 13, 2009, Paper # 160229, (online materials), URL:

Page 150: Torrefaction and Pelletization of Different Forms of

135

http://www.aicheproceedings.org/2009/Fall/data/papers/Paper160229.pdf (Cited

on November 23, 2011)

46. Marb C. M. and Vortmeyer D.,(1988) “Multiple Steady States of a Cross flow

Moving Bed Reactor”, Chemical Engineering Science, Vol. 43 (4), pp 811 – 819.

47. McKendry P. (2002) “Energy production from biomass (part1): Overview of

biomass”, Biores Technology, Vol. 83, pp 37–46.

48. McMullen J., Fasina O.O., Wood C.W., and Feng Y.,(2005) “Storage and

handling characteristics of pellets from poultry litter,” Applied Engineering in

Agriculture, Vol. 21, No. 4, pp 645–651.

49. Measey G.J., Silva J.B., and Di-Bernardo M., (2002) “Testing for Repeatability in

Measurements of Length and Mass in Chthonerpeton indistinctum (Amphibia:

Gymnophiona), Including a NovelMethod of Calculating Total Length of Live

Caecilians, Herpetological Review, Society for the Study of Amphibians and

Reptiles, (Cited on December 31, 2012)

URL:http://john.measey.com/pdf/Measey.techs.pdf

50. Michel, J.B., Mahmed, C., Ropp, J., Richard, J., and Sattler, M., (2011)

“Combustion evaluation of torrefied wood pellets on a 50 kWth boiler,”

http://www.sib.heigvd.ch/contact/Documents (Cited on May 15, 2012).

51. Miura M., Kaga H., Sakurai A., Kakuchi T., and Takahashi K.,(2004) “Rapid

pyrolysis of wood block by microwave heating”, J. Anal. Appl. Pyrolysis, Vol.

71, pp 187–199

52. Mohan D., Pittman C.U., and Steele P.H., (2006) “Pyrolysis of Wood/Biomass for

Bio‐oil: A Critical Review”, Energy & Fuels, Vol. 20:3, pp. 848–889.

Page 151: Torrefaction and Pelletization of Different Forms of

136

53. Mosier N, Wyman C, Dale B, Elander R, Lee YY, Holtzapple M.,(2005)

“Features of promising technologies for pretreatment of lignocellulosic biomass”,

Bioresource Technology; 96:673–86.

54. Obernberger I. and Thek G.,(2004) “Physical characterization and chemical

composition of densified biomass fuels with regard to their combustion behavior,”

Biomass and Bioenergy, Vol 27, No. 6, pp. 653–69

55. Pe´rez J., Mun˜oz-Dorado J., Rubia T., and Martı´nez J.,(2002) “Biodegradation

and biological treatments of cellulose, hemicellulose and lignin: an overview,”

International Microbiology, Vol. 5, No. 2, pp. 53 – 63

56. Phanphanich M. and Mani S.,(2011) “Impact of torrefaction on the grindability

and fuel characteristics of forest biomass,” Bioresource Technology Journal, Vol.

102, No. 2, pp. 1246 – 1253

57. Pimchuai A., Dutta A., and Basu P,(2010) “Torrefaction of Agricultural Residue

to Enhance Combustible Properties,” Energy Fuel Journal, Vol. 24, No. 9, pp.

4638–4645.

58. Pipatmanomai S. (2011) “Overview and experiences of biomass fluidized bed

gasification in Thailand". J Sustainable Energy Environ Special Issue, 29-11.

59. Prins MJ, Ptasinski KJ, Janssen FJJG., (2006a) “Torrefaction of wood: Part 1.

Weight loss kinetics”. Journal of Analytical and Applied Pyrolysis; 77:28–34.

60. Prins, M. J., Ptasinski, K. J., and Janssen, F. J. J. G., (2006b) “More efficient

biomass gasification via torrefaction”, Energy Fuels Journal, Vol. 31, No. 15, pp.

3458–3470.

Page 152: Torrefaction and Pelletization of Different Forms of

137

61. Repellin V., Govin A., Rolland M., and Guyonne R.,(2010) “Energy requirement

for fine grinding of torrefied wood,” Biomass and Bioenergy Journal, Vol. 34,

No. 7, pp. 923 – 930

62. Rosillo-Calle F.(2007) “Biomass assessment handbook – bioenergy for a

sustainable environment. first ed. London, UK: Earthscan.

63. Rousset P, Turner I, Donnot A, Perré P. (2006) “The choice of a low temperature

pyrolysis model at the microscopic level for use in amacroscopic formulation.

Annals of Forest Science;63:213–29.

64. Rousset P., Davrieux F., Macedo L., and Perre´ P., (2011) “Characterizations of

the torrefaction of beech wood using NIRS: Combined effects of temperature and

duration”, Biomass and Bioenergy Journal, Vol. 35, No. 3, pp. 1219 – 1226

65. Sadaka, S., and Negi, S.,(2009) “Improvements of biomass physical and

thermochemical characteristics via torrefaction process,” Environmental Progress

and Sustainable Energy, AlChE Journal, Vol. 28, No. 3, pp. 427–434.

66. Salem A.A. and Ani F.N., (2011) “Microwave Induced Pyrolysis of Oil Palm

Biomass,” Biosource Technology Journal, Vol. 102, No. 3, pp 3388 – 3395

67. Shu-de Q., Fang-Zhen G. and Dong-li Z.,(1996) “The study on performance of

twin screw conveyor, Drying Technology Vol. 14 (7and 8), pp 1859–1870

68. Skrifvars, B.-J., SÞris, G., Backman, R., Widegren-Dafgasrd, K. and Hupa, M.

(1997) “Ash behavior in a CFB boiler during combustion of Salix”, Energy Fuels

11, 843Ð848.

Page 153: Torrefaction and Pelletization of Different Forms of

138

69. Suramya DFM., (2012) “Investigations of Heat Seal Parameters and Oxygen

Detection in Flexible Packages”, PhD Dissertation, University of Guelph, May

2012.

70. Svoboda K. Pohořelý, M., Hartman M., and Martinec J., (2009) “Pretreatment and

feeding of biomass for pressurized entrained flow gasification”, Fuel Processing

Technology Journal, Vol 90, No. 5, pp. 629 – 635.

71. Tumuluru J.S., Sokhansanj S, Hess J.R., Wright C.T., and Boardman R.D, (2011)

“A review on biomass torrefaction process and product properties for energy

applications, Industrial Biotechnology, October.

72. Tumuluru J.S., Sokhansanj S., Wright J.C.T, and Boardman R.D, (2010)

“Biomass Torrefaction Process Review and Moving Bed Torrefaction System

Model Development”, Oak Ridge National Laboratory, INT/EXT-10019569 and

INL/CON-10-18636

73. United Nations (UN). Kyoto protocol to the United Nations framework

convention on climate change [Online]. Available at:

http://unfccc.int/resource/docs/ (Cited on March 27, 2012)

74. Uslu A., Faaij A.P.C., and Bergman P.C.A., (2008) “Pre-treatment technologies,

and their effect on international bioenergy supply chain logistics. Techno-

economic evaluation of torrefaction, fast pyrolysis and pelletization”, Elsevier

Energy Vol. 33, pp 1206–1223

75. Van der Stelt M.J.C, Gerhauser H., Kiel J.H.A. and Ptasinski K.J., (2011)

“Biomass upgrading by torrefaction for the production of biofuels: A review”,

biomass and bio energy 35 ; 3748 e3762

Page 154: Torrefaction and Pelletization of Different Forms of

139

76. Verhoeff F, Arnuelos AA, Boersma AR,Pels JR, Lensselink J, Kiel JHA and

Schukken H.,(2011) “Torrefaction Technology for the production of solid

bioenergy carriers from biomass and waste”, ECN-E--11-039

77. Waje S. S., Patel A. K., Thorat B. N. and Mujumdar, A. S.,(2007) “Study of

Residence Time Distribution in a Pilot-Scale Screw Conveyor Dryer”, Drying

Technology, Vol. 25, No.1, pp. 249 – 259

78. Walton R.A. and Van Bommel I.B.G.,(2011) “A complete and comprehensive

overview of torrefaction technologies”,

79. Wang G, Luo Y, Deng J, Kuang J, Zhang Y. (2011) “Pretreatment of biomass by

torrefaction”, Chinese Science Bulletin ;56:1442–8.

80. Wood S.M. and Layzell D. B., (2003) “A Canadian Biomass Inventory:

Feedstocks for a Bio-based Economy BIOCAP Canada Foundation, Queen’s

University, 156 Barrie Street, Kingston, Ontario, Canada K7L 3N6,

URL:http://www.biocap.ca/images/pdfs/BIOCAP_Biomass_Inventory.pdf (cited

January, 2012)

81. Wolfgang S, Craig C, Jens K. H, Anand R. S, Jesper A, Lei S, Ulrik B. H., (2011)

“Pelletizing properties of torrefied spruce”, Biomass and bio energy (35) 4690-

4698

82. Yan W, Acharjee TC, Coronella CJ, Vásquez VR..(2009) “Thermal pretreatment

of lignocellulosic biomass”, Environmental Progress and Sustainable Energy;

28:435–40.

Page 155: Torrefaction and Pelletization of Different Forms of

140

83. Yang H, Yan R, Chen H, Zheng C, Lee DH, Liang DT., (2005) “In-depth

investigation of biomass pyrolysis based on three major components:

hemicellulose, cellulose and lignin. Energy and Fuels”;20:388–93.

84. Zwart, R. W. J., Boerrigter, H., and Drift, A. V. D.,(2006) “The impact of

biomass pretreatment on the feasibility of overseas biomass conversion to

Fischer-Tropsch products,” Energy Fuels Journal, Vol. 20, No. 5,pp. 2192–2197.

85. http://www.biomassenergycentre.org.uk/portal/page?_pageid=76,15049&_dad=p

ortal&_schema=PORTAL (Cited on December 02, 2012)

Page 156: Torrefaction and Pelletization of Different Forms of

141

Chapter VIII: Appendix

Appendix A: Photographs of Willow Pellets

Raw Pellets:

Pellets at 200°C Pellets at 250°C

Pellets at 275°C

Pellets at 300°C

Page 157: Torrefaction and Pelletization of Different Forms of

142

Appendix B: Photographs of Oats Pellets

Raw Pellets:

Pellets at 200°C

Pellets at 250°C

Pellets at 275°C

Pellets at 300°C

Biodegradable Raw Oats

Page 158: Torrefaction and Pelletization of Different Forms of

Appendix C: Methods and Equipment used in Characterizing Biomass

Proximate Analysis Ultimate Analysis

Higher Heating Value

(HHV) and Ash analysis

Parameter MC AC VM FC C, H, N S O HHV Ash Analysis

Method ASTM

E871

ASTM

E1755

ASTM

D3175

ASTM

E872

ASTM

D5375

ASTM

D4239

ASTM

E870

ASTM

E711

ASTM

D6349

Equipment

Used

Muffle

furnace

Muffle

Furnace

Muffle

Furnace

calculate

d value

Leco

CHN-

1000

Elemental

Analyser

Leco SC-

432

Elemental

Analyser

calculated

value

C-200

Bomb

Calorimeter

Muffle

furnace,

SEM (S-570-

Hitachi High

115) and

XRD (Bragg-

Mrentano

geometry )

Results from

experiment

C=52.2%

H=6.59%

N=0.62%

0.29% 33.98% See: Table 2

Page 159: Torrefaction and Pelletization of Different Forms of

Appendix D: Photograph of Experimental Setup for Torrefaction

Page 160: Torrefaction and Pelletization of Different Forms of

145

Appendix E: Gas Analyzer

Page 161: Torrefaction and Pelletization of Different Forms of

146

Appendix F: Ultimate Analysis, Ash Fusion Temp and Ash Elemental Analysis

Table F.1: Ultimate Analysis of different biomass samples

Components Poultry Litter Willow Pallets Oat Pallets

Carbon 43.30% 50.65% 52.23%

Hydrogen 6.62% 5.86% 6.59%

Nitrogen 5.72% 0.52% 0.62%

Sulphur 1.15% 0.44% 0.29%

Oxygen 5.95% 24.07% 33.98%

Table F.2: Chlorine, Ash Content and Ash Fusion temperatures of different

Biomasses in dry basis

Particular

Poultry

Litter Willow Pellets Oat Pellets

Chlorine (%) 0.07 0 0

Ash Fusion Temperature (°C)

Initial Deformation (IT) (°C) 1421.67 1115 1279

Softening Temperature (ST)

(°C) 1433.89 1171 1303

Hemispherical Temperature

(HT) (°C) 1437.78 1258 1338

Fluid Temperature (FT) (°C) 1440 1292 1354

Page 162: Torrefaction and Pelletization of Different Forms of

147

Table F.3. Elemental Analysis of Poultry Litter in % Ash Basis

Components Poultry Litter Willow Pellets Oat Pellets

SiO2 2.69 67.77 60.68

TiO2 0.02 0.44 0.61

Al2O3 0.31 9.47 15.62

Fe2O3 0.57 2.89 0.4

MnO 0.33 0.11 0.09

MgO 4.53 20.7 2.06

CaO 65.17 9.4 5.59

K2O 6.36 3.53 9.85

Na2O 2.48 2.44 0.47

P2O5 17.46 1.74 4.54

Cr2O3 0.03 0.03 0.07

Page 163: Torrefaction and Pelletization of Different Forms of

148

Appendix G: Error and Repeatibility Tests

G.1 Error

Errors observed during the testing of characterization of biomasses are listed

which are due to the use of high precision instruments.

Table G-1 Systematic errors

S/N Name of the instrument Accuracy from

manufacturer

Unit

1 Weighing Machine 0.0001 g

2 Temperature and Process Controller ±0.5 ᴼC

3 Flue Gas Analyzer for Oxygen 0.01 Vol %

4 Mass Flow Controller ±1.0 %

5 Water volumetric flask 1 ml

6 Humidity meter ±0.75 %

For the systematic error calculation, procedures with following equations were

taken from http://teacher.pas.rochester.edu/PHY_LABS/AppendixB/AppendixB.html

link and used for the error calculations.

(G.1)

(G.2)

(G.3)

(G.4)

(G.5)

(G.6)

Page 164: Torrefaction and Pelletization of Different Forms of

149

(G.7)

Table G-2 Maximum Systematic error

S/N Particular Systematic Error Unit

1 Mass Yield 0.15 %

2 Energy Yield 0.10 %

3 Fixed Carbon 0.1 %

4 Volatile Matter 0.05 %

5 Ash 2.0 %

G.2 Repeatability Test

The accuracy of any lab tests are basically depended on the variability of the data

recorded equipment setup and measuring instruments. The characterizations,

hydrophobicity and moisture uptake experiments involve lots of uncertainty because of

the manual operation for running the experiment and taking the readings using weighing

machine. Here three tests were repeated and presented in the figure G1-G3 for

hydrophobicity and G4-G6 for moisture uptake. All graphs in each categories showed

that results are consistent with insignificant variation from the test results. Repeatability

tests were carried out by using practical tests repetitions rather than some statistical

software.

Page 165: Torrefaction and Pelletization of Different Forms of

150

Table G-3 Proximate Analysis:

Repeated Twice for Proximate Analysis

Sample A1 A2 Error A1 A2 Error A1 A2 Error

Volatile Matter (%) Ash Contents (%) Fixed Carbon (%)

I 54.91 53.98 0.93 19.54 21.01 1.47 3.12 3.96 0.84

II 73.88 74.25 0.37 10.91 10.36 0.55 11.05 10.57 0.48

III 73.97 75.01 1.04 4.77 5.22 0.45 13.12 12.58 0.54

Because of the accuracy of the bomb calorimeter, the results are pretty close with the

repetition of the experiment on three different samples.

Table G-4 Heating Value:

a) Repeatability test on Hydrophobicity

Some variations were found during 2.4% oxygen sample which are acceptable

because of the dependability of the results in the environmental conditions.

Repeated Twice for Heating Values

Sample A1 A2 Error A1 A2 Error A1 A2 Error

Heating Value of Raw

(MJ/Kg)

Heating Value of

Torrefied at 250C for 30

min (MJ/Kg)

Heating Value of

Torrefied at 275C for 30

min (MJ/Kg)

I 9.89 10.55 0.66 11.34 12.85 1.51 11.85 12.86 1.01

II 73.88 74.25 0.37 16.86 18.57 1.71 19.89 20.58 0.69

III 73.97 75.01 1.04 20.01 21.22 1.21 22.48 23.97 1.49

Page 166: Torrefaction and Pelletization of Different Forms of

151

Figure G- 1: Repeatability test on the Hydrophobicity of a Sample with 0% O2

(30min)

Figure G- 2: Repeatability test on the Hydrophobicity of a Sample with 2.4% O2

(30min)

Page 167: Torrefaction and Pelletization of Different Forms of

152

b) Repeatability test on Moisture Uptake

Satisfactory results were obtained during the experimental repetitions.

Figure G- 3: Repeatability test on the Moisture Uptake for a sample 0% O2

Figure G- 4: Repeatability test on the Moisture Uptake for a sample 2.4% O2

In order to test the significance variation on the data of these experiments, ANOVA, a

statistical model was used to see the hydrophobicity and moisture uptake with inert and

0

0.5

1

1.5

2

2.5

3

3.5

0 5 10 15 20 25 30

Mo

istu

re U

pta

ke (

%)

Exposure Time in hours

275C with 30 min 275C with 30min

Page 168: Torrefaction and Pelletization of Different Forms of

153

2.4% oxygen concentration. The null hypothesis states that there is no difference on the

results collected with the oxygen variations.

MS Excel based ANOVA was run with 95% confidence level and the result

obtained for the biomass torrefied at 275ᴼC for 30 minutes are presented in the table G-5.

The P-value obtained was greater than 0.03, which means that the hydrophobicity tested

at different oxygen concentration has no significance variations on the results.

G-5: ANOVA between the hydrophobicity tests of Willow for 30 min residence time

Mean Square

(MS)

Temperature SS df MS F P-Value

Between ground 275ᴼC 3.6112 1 3.6112 42.0746 0.02377

Within Group 0.1716 2 0.0858

Total 3.7828 3

According to Measey et al (2002),

G-8

Where, n= numbers of repeated measurements (2)

Once above data are substituted in the equation G-8, which shows the good

repeatability because ri within range of 70% to 90% are considered as very high

repeatability (Measey et al, 2002).