You are on page 1of 267

Kinetic Study and Phenomenological Modeling of

a Biomass Particle during Fast Pyrolysis Process

Jorge Iván Montoya Arbeláez

Universidad Nacional de Colombia


Facultad de Minas, Departamento de Procesos y Energía
Medellín, Colombia
2016
Kinetic Study and Phenomenological Modeling of
a Biomass Particle during Fast Pyrolysis Process

Jorge Iván Montoya Arbeláez

Thesis submited in partial fulfillments of requeriments for the degree of


Doctor in Philosophy

Advisor:
Ph.D. Farid Chejne Janna
Co-Advisor: Ph.D. Manuel Garcia-Perez

Line of Research:
Energetic Systems
Research Group:
Termodinámica Aplicada y Energías Alternativas TAYEA

Universidad Nacional de Colombia


Facultad de Minas, Departamento de Procesos y Energía
Medellín, Colombia
2016
Dedication
I dedicate this work especially my mother and my brothers who have been a constant support
throughout my professional life.
IV Content

Content

Acknowledgments ................................................................................................................................................... VIII


Summary................................................................................................................................................................... IX
List of Tables ............................................................................................................................................................. XI
List of Figures........................................................................................................................................................... XII
Symbols ............................................................................................................................................................... XVIII
0. General Introduction ........................................................................................................................................................ 22
Dissertation objetive ................................................................................................................................................. 25
Methodology............................................................................................................................................................. 25
Scientific contributions ............................................................................................................................................. 26
Publications .............................................................................................................................................................. 28
References ............................................................................................................................................................... 30
1. Chapter 1. State of the Art ............................................................................................................................................... 32
1.1. Biomass Structure....................................................................................................................................... 33
1.2. Studies of structural parameters in biomass pyrolysis ................................................................................ 34
1.3. Effect of Particle Size .................................................................................................................................. 36
1.4. Effect of mineral matter ............................................................................................................................... 36
1.5. Effect of volatile and solids residence time ................................................................................................. 38
1.6. Effect of reaction temperature ..................................................................................................................... 38
1.7. Effect of Heating Rate ................................................................................................................................. 40
1.8. Models to describe the reaction rate ........................................................................................................... 44
1.8.1. Global lumped models in a single step ............................................................................................... 44
1.8.2. Multi stage-global lumped models ...................................................................................................... 45
1.8.3. Semi-global lumped models ............................................................................................................... 47
1.8.4. Models based in volatile species ........................................................................................................ 48
1.8.5. Distribution activation energy model (DAEM) ..................................................................................... 50
1.9. Secondary reactions ................................................................................................................................... 51
1.10. Existence of intermediate liquid phase (Chronology) .................................................................................. 52
1.11. Modeling intra-particle fast pyrolysis process ............................................................................................. 58
1.12. Final Remarks ............................................................................................................................................. 65
1.13. References.................................................................................................................................................. 66
2. Chapter 2. Identification of the Fractions Responsible for Morphology Conservation in Lignocellulosic Pyrolysis:
Visualization Studies of Sugarcane Bagasse and its Pseudo-Components .............................................................. 77
2.1. Introduction ................................................................................................................................................. 78
2.2. Materials and Methods ................................................................................................................................ 79
2.2.1. Materials and preparation................................................................................................................... 79
2.2.2. Characterization of materials.............................................................................................................. 80
2.2.3. Modified Pyroprobe pyrolysis visualization reactor............................................................................. 82
2.2.4. Temperature calibration ..................................................................................................................... 83
2.2.5. SEM analyzes of solid residues ......................................................................................................... 84
2.3. Results and discussion ............................................................................................................................... 85
2.3.1. Behavior of organosolv lignin during pyrolysis ................................................................................... 85
2.3.2. Behavior of milled wood lignin during pyrolysis .................................................................................. 88
2.3.3. Behavior of xylan (as received) during pyrolysis ................................................................................ 89
2.3.4. Behavior of de-ashed xylan during pyrolysis ...................................................................................... 90
Content V
2.3.5. Behavior of cellulose .......................................................................................................................... 91
2.3.6. Behavior of sugarcane bagasse ......................................................................................................... 93
2.3.7. SEM imaging of solid residue from all materials studied .................................................................... 95
2.4. Conclusions ................................................................................................................................................ 96
2.5. References.................................................................................................................................................. 97
3. Chapter 3. Micro-Explosion of Liquid Intermediates During the Fast Pyrolysis of Sucrose and Lignin................ 101
3.1. Introduction ............................................................................................................................................... 102
3.2. Materials and methods.............................................................................................................................. 103
3.2.1. Materials........................................................................................................................................... 103
3.2.2. Pyroprobe based fast pyrolysis visualization reactor........................................................................ 104
3.2.3. Temperature calibration ................................................................................................................... 105
3.2.4. Processing of visualization data ....................................................................................................... 105
3.2.5. Mathematical model based on global population balance to predict microexplosion intensity and
aerosol ejection intensity ................................................................................................................................... 107
3.3. Results ...................................................................................................................................................... 110
3.3.1. Global behavior of organosolv lignin and sucrose. ........................................................................... 110
3.3.2. Birth rate of bubbles in the liquid phase: Organosolv lignin.............................................................. 112
3.3.3. Birth rate of bubbles in the liquid phase: Sucrose ............................................................................ 114
3.3.4. Bubble growth of organosolv lignin and sucrose .............................................................................. 115
3.3.5. Distribution of the bubbles within the liquid phase of organosolv lignin ............................................ 119
3.3.6. Distribution of the bubbles within the liquid phase to sucrose. ......................................................... 120
3.3.7. Population balance to predict bubble bursting rate, bubbles coalescence, and ejection intensity ... 121
3.4. Conclusions .............................................................................................................................................. 128
3.5. References................................................................................................................................................ 129
4. Chapter 4. Methodology for estimating the distribution and intensity of ejection of aerosol ejected during pyrolysis
of bagasse and models of their components .............................................................................................................. 132
4.1. Introduction ............................................................................................................................................... 133
4.2. Materials and methods.............................................................................................................................. 134
4.2.1. Feedstock characterization .............................................................................................................. 134
4.2.2. Hot plate reactor ............................................................................................................................... 135
4.2.3. Slides visualization by SEM. ............................................................................................................ 136
4.2.4. Aerosol thermal ejection intensity..................................................................................................... 137
4.3. Results ...................................................................................................................................................... 137
4.3.1. Aerosols ejection during the pyrolysis of avicel cellulose ................................................................. 138
4.3.2. Aerosols ejected during the pyrolysis of organosolv lignin ............................................................... 142
4.3.3. Aerosols ejected during the pyrolysis of xylan.................................................................................. 146
4.3.4. Aerosols ejected during the pyrolysis of sugarcane bagasse ........................................................... 149
4.4. Aerosol yield theoretical estimation........................................................................................................... 153
4.5. Final comments......................................................................................................................................... 155
4.6. Conclusions .............................................................................................................................................. 156
4.7. References................................................................................................................................................ 157
5. Chapter 5. Effect of Temperature and Heating Rate on Product Distribution from Pyrolysis of Sugarcane Bagasse
in a New Hot Plate Reactor............................................................................................................................................ 161
5.1. Introduction ............................................................................................................................................... 162
5.2. Materials and methods.............................................................................................................................. 162
5.2.1. Hot plate reactor configuration ......................................................................................................... 162
5.2.2. Plate temperature calibration and temperature distribution .............................................................. 164
5.2.3. Materials........................................................................................................................................... 164
VI Content
5.2.4. Sample preparation .......................................................................................................................... 164
5.2.5. Run sequence .................................................................................................................................. 166
5.2.6. Permanent gases ............................................................................................................................. 166
5.2.7. Char ................................................................................................................................................. 167
5.2.8. Bio-oil ............................................................................................................................................... 167
5.3. Results and discussion ............................................................................................................................. 168
5.3.1. Temperature calibration and distribution .......................................................................................... 168
5.3.2. Impact of holding time and film thickness on char yield ................................................................... 170
5.3.3. Species studied in these experiments .............................................................................................. 171
5.3.4. Effect of peak temperature over volatile species yield identified by GC/MS..................................... 172
5.3.5. Lignin oligomers formation ............................................................................................................... 175
5.3.6. Temperature effect on anhydrosugar yields ..................................................................................... 176
5.3.7. Temperature effect over permanent gases ...................................................................................... 178
5.3.8. Temperature effect on char yield ...................................................................................................... 179
5.3.9. Heating rate effect on main products yield at 500 °C ....................................................................... 180
5.4. Conclusions .............................................................................................................................................. 181
5.5. References................................................................................................................................................ 182
6. Chapter 6. Distribution of Activation Energy Model to Describe Sugarcane Bagasse Thermal Depolymerization
Kinetics ........................................................................................................................................................................... 186
6.1. Introduction ............................................................................................................................................... 187
6.2. Materials and methods.............................................................................................................................. 188
6.2.1. Material ............................................................................................................................................ 188
6.2.2. Sample preparation .......................................................................................................................... 188
6.2.3. Hot plate reactor configuration ......................................................................................................... 188
6.2.4. Operation sequence and experimental plan ..................................................................................... 188
6.2.5. Distribution of activation energy Model............................................................................................. 189
6.2.6. Lignin oligomers decomposition ....................................................................................................... 191
6.2.7. Levoglucosan decomposition ........................................................................................................... 192
6.2.8. Biomass drying ................................................................................................................................. 192
6.3. Results ...................................................................................................................................................... 193
6.3.1. Evaporation and secondary reactions .............................................................................................. 199
6.4. Conclusions .............................................................................................................................................. 202
6.5. References................................................................................................................................................ 203
7. Chapter 7. Single Particle Model for Biomass Pyrolysis: Dynamics of Bubble Formation Inside the Liquid
Intermediate and Their Contribution to Aerosol Formation by Thermal Ejection .................................................... 208
7.1. Introduction ............................................................................................................................................... 209
7.2. Development of the mathematical model .................................................................................................. 210
7.2.1. Physical model ................................................................................................................................. 210
7.2.2. Particle geometry and differential element ....................................................................................... 211
7.2.3. Pyrolysis reactions ........................................................................................................................... 212
7.2.4. Secondary reactions......................................................................................................................... 213
7.2.5. Drying ............................................................................................................................................... 214
7.2.6. Particle shrinking .............................................................................................................................. 214
7.2.7. Mass balances ................................................................................................................................. 214
7.2.8. Population balances in the liquid phase ........................................................................................... 216
7.2.9. Global energy balance ..................................................................................................................... 217
7.2.10. Momentum balance .......................................................................................................................... 218
7.2.11. Aerosol retention efficiency .............................................................................................................. 219
7.2.12. Initial conditions and boundary conditions ........................................................................................ 220
Content VII
7.2.13. Simulation code ................................................................................................................................ 220
7.2.14. Model validation and verification ...................................................................................................... 221
7.3. Simulation results ...................................................................................................................................... 222
7.3.1. Temperature profile .......................................................................................................................... 223
7.3.2. Validation of char and other species yield ........................................................................................ 225
7.3.3. Intra-particle dynamics ..................................................................................................................... 228
7.3.4. Bubbles dynamics ............................................................................................................................ 230
7.3.5. Effect of particle size on product yields ............................................................................................ 235
7.3.6. Effect of final temperature on product yields .................................................................................... 237
7.3.7. Effects of heating rate on product yields .......................................................................................... 240
7.3.8. Sensitivity analysis ........................................................................................................................... 242
7.4. Conclusions .............................................................................................................................................. 243
7.5. References................................................................................................................................................ 243
8. Chapter 8. Conclusions ................................................................................................................................................. 249
8.1. Future works ............................................................................................................................................. 250
8.2. References................................................................................................................................................ 252
9. Appendix ......................................................................................................................................................................... 253
Appendix A. Videos (Chapter 2 – 5) ....................................................................................................................... 253
Appendix B. Chromatograms (Chapter 5 – 6) and Temperature profile at 1000°C/s and Tf=500C. ...................... 253
Appendix C. Auxiliary equations for modeling (Chapter 7)..................................................................................... 258
VIII Content

Acknowledgments

I am particularly grateful to Professor Farid Chejne, my thesis advisor, for his support, dedication and patience in many
discussions that I had made in the development of this thesis. To Professor Manuel Perez-Garcia, for bring me the
opportunity in Pullman, in his facilities. I appreciate his contributions and comments on the review of articles and
chapters of this thesis. I am particularly grateful to Professor Benjamin Rojano, for helping me with Fluorescence and
UV-VIS Test for quantification of oligomers of lignin and sugars. To my friends Javier Fernando de La Cruz, Carlos
Valdes, Carlos Gomez, Gloria Marrugo, for their friendship and strong support in my difficult moments. To engineers
David Roman and Estefania Orrego for their support with laboratory activities, especially those related to testing in the
hot plate reactor. To my friend in USA Brennan Pecha for his valuable friendship and for teaching me many things that
I am presenting in this thesis. To my other friends and doctoral parthners, that without their valuable support it would not
be possible the culmination of this work: Gabe Pecha, Filip Stankovich, Rishi Gohare, Matt Smitt, Jonatnan Lomber,
Jonathan Pulgarin, David Granados, Alejandro Jaramillo, Victor Borda, Diego Carmargo, Jader Allean. To the
undergraduate students: Pedro Arana, Sebastian Hernandez, Marcela Marin, Carolina Londoño, Raiza Manrique and
Daniela Vasquez.

I also thank COLCIENCIAS for their valuable assistance with the financial support for my doctoral studies and my
internship.I wish to thanks at Professor Abel de Jesus Naranjo Agudelo, who bring me financial support for SEM analysis
and some translations belongs to this dissertation.
Content IX

Kinetic Study and Phenomenological Modeling of


a Biomass Particle during Fast Pyrolysis Process

Summary

By Jorge Ivan Montoya Arbelaez


Universidad Nacional de Colombia-Sede Medellin
2016

Advisor: Farid Chejne Jana


Co-Advisor: Manuel Garcia-Perez

In this project, the main physical and chemical phenomena associated with the pyrolysis of sugarcane bagasse were
studied develop a new particle-level model that predicts the evolution of major compounds during the pyrolysis process.
The model incorporates dynamics of bubbling within the liquid intermediate phase and their bubbles contributions to
the ejection of aerosols. The main physical phenomena occurring during pyrolysis of sugarcane bagasse, pseudo-
components, and model compounds (chapters 2-4) were explored. Fof this, new types of reactors (light bulb reactor
and hot plate reactor) and visualization methodologies were employed. Notably, a fast camera was used to visualize
the growth, nucleation, and size of the bubbles within the intermediate liquid phase for two surrogates of lignocellulose
(organosolv lignin, sucrose).

In the hot plate reactor, a new methodology was proposed to identify and classify by size the aerosols ejected during
pyrolysis of sugarcane bagasse and its pseudo-components. The aerosol studies were performed at high and low
heating rates (10 ° C / s and 1200 ° C / s) and under both vacuum (150mbar) and atmospheric pressure (900mbar).
The ejected aerosols were captured on glass and plastic surfaces and visualized by scanning electron microscopy
(SEM). The SEM micrographs were used to estimate the ejection intensity under different conditions. These results
were then used to identify relationships between bubble size and aerosol size as well as the number of bubbles vs.
number of bubbles / number of aerosols used in the mathematical model. The experimental results showed that gas
bubbles sizes and ejected aerosols sizes follow a lognormal distribution for all materials studied (sucrose, lignin, xylan,
cellulose, and bagasse). Under vacuum conditions (150mbar) and high heating rate (1200 ° C / s) conditions, aerosol
formation is promoted by enhancement of oligomer evaporation, bubble bursting, and micro-explosions within the
intermediate liquid phase.

The distribution of primary products and the study of kinetic parameters by using a distribution of activation energy
model were performed using a hot plate reactor with a thin and uniform biomass film (thickness≈60μm). These studies
were carried out under vacuum (150mbar) to minimize secondary reactions. The main compounds released during
devolatilization of sugar cane bagasse were analyzed by mass spectroscopy (GC-MS), UV-fluorescence (UV-
Fluorescence), gas chromatography (GC), and liquid chromatography (HPLC). The results showed that 100 °C/s, no
significant changes are observed in the yields of the products due to the low thermal conductivity of the biomass and
these conditions exhibited the highest yields of sugar and lignin oligomers with the least char. The bio-oil obtained at
X Content
low heating rates, however, was high in water content, organic acids, aldehydes, and ketones with a lower concentration
of oligomers compared with bio-oil obtained at high heating rates.

The temperature profile of the external surface of the biomass was determined via experimentation and model
simulation (Chapter 7). The model and experiments were compared for at different particle thicknesses, temperature,
and heating rates. The reaction rate for each species was obtained by fitting experimental data to a distribution of
activation energy model that describes the evolution of primary products, the main secondary reactions, bubble
formation within the intermediate liquid phase, and aerosol ejection.

The model predictions for aerosols char, light oxygenated compounds, and permanent gases yield were close to
experimental results. The most sensitive variable for aerosols yield was particle size, showing a yield close to 0 with a
2mm particle. Temperature has a positive effect on the aerosol yield by intensifying chemical reactions and bubble
dynamics (nucleation). Heating rate enhances bubble bursting and aerosol ejection but to a lesser degree than particle
size.
Content XI

List of Tables
Page
General Introduction
Table 1. Main biomass Availability and energetic potential in Colombia. Source: UPME [46]. .................................... 25
Chapter 1
Table 1. Main devices used in the study of fast pyrolysis processes. .......................................................................... 42
Table 2. Multistage mechanisms for biomass pyrolysis. .............................................................................................. 45
Table 3. Kinetic schemes proposed by Ranzi, et al. [15] to predict condensable and non-condensable gas yields from
cellulose and hemicellulose pyrolysis . ................................................................................................................ 49
Table 4. Particle-level models ...................................................................................................................................... 58
Chapter 2
Table 1. Composition of sugarcane bagasse as received and after acid wash on a dry ash-free basis. The sums slightly
exceed 100% due to experimental error. The sugarcane bagasse before acid washing has a much higher
extractives content than after acid washing, which contributes to relative changes in the glucan, xylan, arabinose,
and lignin content................................................................................................................................................. 80
Table 2. Physico-chemical characterization of the raw material. Moisture content is on as-received basis. Volatiles,
fixed carbon, ash, glucan, arabinose, lignin, and extractives are reported on a dry basis (w/w). CHN-O data is
reported on a dry, ash-free basis (DAF) [wt. %]................................................................................................... 81
Table 3. Metal species analysis by ICP-MS for sugarcane bagasse and xylan before and after their respective acid
washings. The acid washing reduces the metal content for all samples. Notably, the Na content in xylan is reduced
form 14 mg/g to sub-detection limit concentrations and K to near detection limits (mg/g). .................................. 81
Chapter 3
Table 1. Physico-chemical characterization of the raw material. CHN-O data are reported on a dry basis. .............. 104
Chapter 4
Table 1. Estimated aerosol yield for each material. ................................................................................................... 155
Chapter 5
Table 1. Main Compounds identified by GC-MS, HPLC, UV-Fluorescence, µ-GC. ................................................... 171
Chapter 6
Table 1. Summary of kinetic parameters found by fitting experimental data in distribution of activation energies model.
........................................................................................................................................................................... 193
Table 2. Ultimate analysis of pyrolytic lignin and its solid residue after pyrolysis by TGA. ......................................... 201
Table 3. Elemental analysis of levoglucosan and its solid residue after pyrolysis by TGA. ....................................... 201
Chapter 7
Table 1. Secondary reactions and its kinetic parameters........................................................................................... 213
Table 2. Initial and boundary conditions..................................................................................................................... 220
Table 3. Physicochemical properties for the simulation. ............................................................................................ 221
XII Content

List of Figures
Page
General Introduction
Figure 1. Schematic multi-scale fast pyrolysis biomass............................................................................................... 23
Figure 2. Yields of global products during fluidized bed pyrolysis of different biomass [10] ........................................ 24
Figure 3. Flow chart diagram of main topics studied in this work................................................................................. 26
Chapter 1
Figure 1. Van Krevelen diagram. ................................................................................................................................. 32
Figure 2. General structure of lignocellulosic materials. Adapted from [26] ................................................................. 34
Figure 3. (A)Stages of thermal decomposition for biomass. (B) Temperature profile within a biomass particle. ......... 39
Figure 4. Effect of the heating rate on the mechanism of thermal decomposition of the cellulose. [125] .................... 41
Figure 5. Structural changes of (A) Avicel Cellulose, (B) Alkali Lignin, (C) Xylan, and (D) Tobacco during the pyrolysis
process.Taken from [231] .................................................................................................................................... 54
Figure 6. Photographic record of the evolution of the cellulose during fast pyrolysis. Taken from Dauenhauer et al.
[116]. .................................................................................................................................................................... 55
Figure 7. Images of the evolution of the biomass during the slow pyrolysis process. Figure taken from Thomas et al
[34]. : (A) 260 °C, (B) 299 °C, and (C) 501 °C. Red outlines of xylem cells track individual areas through each
frame. (D) Longitudinally sectioned control poplar showing cell wall faces of ray parenchyma, with white arrows
pointing to examples of cell wall pores (plasmodesma). (E) Longitudinally sectioned hot-stage poplar shows cell
wall faces of ray parenchyma with collected pyrolysis products (black arrows). .................................................. 55
Figure 8. Ejection mechanism of liquid droplets dfrom the intermediate liquid phase originating from the fast pyrolysis
of sucrose (top) with cartoon representations of the observed behavior (bottom). Figure taken from Teixeira et al
[133] ..................................................................................................................................................................... 56
Figure 9. (A).Evolution of the intermediate liquid phase during the fast pyrolysis of hardwood. Figure taken from:
Carlsson et al [238]. (B). Evolution with temperature of a liquid phase during the slow pyrolysis of waste soybeans.
Figure taken from Zhu et al. [239]. ....................................................................................................................... 57
Chapter 2
Figure 1. Visualization setup for studying fast pyrolysis. ............................................................................................. 82
Figure 2. Calibration curves of the various components and the blank disc. ............................................................... 83
Figure 3. Organosolv lignin melting and bubbling behavior during pyrolysis: still frames from video recordings. ........ 87
Figure 4. Milled wood lignin melting and bubbling behavior during pyrolysis: still frames from video recordings. ....... 88
Figure 5. Xylan melting and bubbling behavior during pyrolysis: still frames from video recordings. .......................... 90
Figure 6. De-ashed xylan melting and bubbling behavior during pyrolysis: still frames from video recordings. .......... 91
Figure 7. Cellulose melting and bubbling behavior during pyrolysis: still frames from video recordings. ..................... 92
Figure 8. Sugar cane bagasse melting and bubbling behavior during pyrolysis: still frames from video recordings. .. 93
Figure 9. Acid washed bagasse melting and bubbling behavior during pyrolysis: still frames from video recordings.. 94
Figure 10. SEM micrographs of (A) Organosolv lignin, (B-C) milled wood lignin, (D) xylan, (E-F) de-ashed xylan, (G)
cellulose, (H) sugarcane bagasse, and (I) acid washed sugarcane bagasse chars. Scale bars are indicated on the
figures. ................................................................................................................................................................. 95
Chapter 3
Figure 1. Dynamic life of a bubble on a liquid surface [9]. ......................................................................................... 102
Figure 2. Pyroprobe-based fast pyrolysis visualization reactor. A: Modified heating element and quartz sample disc
with a cellulose sample. B: A view of the setup used for visualization, where the reactor on the right is lit up by a
fiber-optic light source. Nitrogen purges through the reactor, which is open on the end, and a thermocouple can
be fed through a tiny hole in the glass tube to measure temperature on the sample or quartz disc. ................. 104
Content XIII
Figure 3. (A) Positioning the thermocouple and sample calibration curve. (B) Final temperatures on the lignin (~400
°C) and sucrose (~280 °C) are due to the fact that the thermocouple is not touching the quartz. Because char has
a low thermal conductivity, heat is not transferred well through the char, which has room-temperature nitrogen
flowing over the top............................................................................................................................................ 105
Figure 4. Methodology to calculate the control volume using ImageJ using the platinum coil to set the scale. ......... 106
Figure 5. Procedure for counting bubbles in Organosolv Lignin using Matlab’s image analysis software. ................ 106
Figure 6. Scheme of mathematical model diagram with the main phenomena used to describe bubble dynamics .. 108
Figure 7. Sequence of photographs with major changes during devolatilization of sucrose and Organosolv lignin (Disc
Temperature – Sample Temperature). .............................................................................................................. 112
Figure 8. Birth of bubbles in the intermediate stage of lignin pyrolysis. ..................................................................... 113
Figure 9. Experimental and interpolated data for two stages in bubble growth and nucleation of Organosolv lignin. 114
Figure 10. Sequence of frames to describe the birth of bubbles in the liquid phase intermediate formed during the fast
pyrolysis of sucrose. .......................................................................................................................................... 114
Figure 11. Nucleation rate for Sucrose. ..................................................................................................................... 115
Figure 12. Monitoring the growth of a gas bubble in the molten lignin. ...................................................................... 116
Figure 13. Experimental data for bubble size and bubble growth rate of Organosolv lignin versus time with model fitting
to describe the Arrhenius growth rate. ............................................................................................................... 117
Figure 14. Monitoring the growth of a gas bubble into the molten sucrose................................................................ 118
Figure 15. Experimental data for bubble size and bubble growth rate of sucrose versus time with model fitting. ..... 118
Figure 16. Change in the distribution function of the bubbles to the molten phase of Organosolv lignin. .................. 120
Figure 17. Change in the distribution function of the bubbles to the molten phase of sucrose. ................................. 120
Figure 18. Representation of the distribution functions of the net rate of birth and death of bubbles during devolatilization
of Organosolv lignin and Sucrose. A. Net rate of birth and death of bubbles within the liquid phase of the
Organosolv lignin. B. Total net rate of birth and death of bubbles within the liquid phase of the Organosolv lignin. C.
Net rate of birth and death of bubbles within the liquid phase of sucrose. D. Total net rate of birth and death of
bubbles within the liquid phase of sucrose. ....................................................................................................... 122
Figure 19. Sequence of formation and collapse of bubbles in liquid-gas interface during the fast pyrolysis of organosolv
lignin. ................................................................................................................................................................. 123
Figure 20. Sequence formation, breaking bubbles (Bursting) and release of jets, during the fast pyrolysis of sucrose.
........................................................................................................................................................................... 124
Figure 21. Bubble bursting rates for individual bubbles and the sum of bursting rates for Organosolv lignin and
sucrose. A. Bubble bursting rates for Organosolv lignin pyrolysis. B. Total Bubble bursting rate for Organosolv
lignin pyrolysis. C. Bubble bursting rates for sucrose pyrolysis. D. Total bubble bursting rate for sucrose pyrolysis
. .......................................................................................................................................................................... 125
Figure 22. Distribution aerosol function of time and droplet diameter. ...................................................................... 127
Chapter 4
Figure 1. Multi-scale scheme for the fast pyrolysis of biomass. ................................................................................. 134
Figure 2. Impregnation of the samples on the cold-rod stainless steel plate and average film thickness. (A) Finely milled
sample and distilled water. (B) Brush used for impregnation (C). Empty steel plate. (D). Steel plate impregnated
with a water / biomass ≈10% w / w suspension (E). Steel plate impregnated with a water / biomass suspension
after drying in halogen moisture balance at 80 °C. ............................................................................................ 135
Figure 3. Layout of Hot Plate Reactor. ...................................................................................................................... 135
Figure 4. Tone setting in the image to be processed. (A) Original image cellulose 1200 ° C / s vacuum. (B) Image A
converted into binary scale (black-white) using ImajeJ. (C) Manual Processing highlighting aerosols outline (D)
Processed image with outline highlighted. ......................................................................................................... 136
Figure 5. SEM visualization of some pure compounds in bio-oils. ............................................................................. 138
Figure 6. Char yield obtained during avicel cellulose pyrolysis. ................................................................................. 138
Figure 7. SEM images obtained by ejected aerosol during avicel cellulose pyrolysis. .............................................. 139
Figure 8. Aerosols distribution during Avicel cellulose pyrolysis. .............................................................................. 140
XIV Content
Figure 9. Relative aerosol ejection intensity during Avicel cellulose pyrolysis. .......................................................... 141
Figure 10. Char yield obtained in Organosolv lignin pyrolysis. .................................................................................. 142
Figure 11. Images obtained by SEM of ejected aerosols during Organosolv lignin pyrolysis. ................................... 143
Figure 12. Effect of the amount of sample and surface type over aerosol captured geometry during organosolv lignin
pyrolysis at 500 °C, 1200 °C/s and 900 mbar. .................................................................................................. 144
Figure 13. Aerosols distribution during organosolv lignin pyrolysis............................................................................ 144
Figure 14. Aerosols distribution during organosolv lignin pyrolysis............................................................................ 145
Figure 15. Relative aerosols ejection intensity during organosolv lignin pyrolysis. .................................................... 146
Figure 16. Char yield obtained for xylan pyrolysis. .................................................................................................... 147
Figure 17. Images obtained by SEM of ejected aerosols during xylan pyrolysis. ...................................................... 148
Figure 18. Distribution of aerosols from the pyrolysis of xylan. .................................................................................. 149
Figure 19. Char yield obtained for sugarcane bagasse pyrolysis. ............................................................................. 149
Figure 20. Images obtained by SEM of ejected aerosols during sugarcane bagasse pyrolysis. ............................... 150
Figure 21. Raman spectra of a representative bagasse aerosol.(A). Confocal visualization of an aerosol of sugarcane
bagasse. (B). Raman spectrum of sugarcane aerosol and avicel cellulose....................................................... 151
Figure 22. Aerosol distribution during sugarcane bagasse pyrolysis. ........................................................................ 152
Figure 23. Relative aerosols ejection intensity during sugarcane bagasse pyrolysis. ............................................... 153
Figure 24. Geometric representation of volume of a spherical sector as a function of contact angle and diameter. . 153
Figure 25. Contact angles measured experimentally for each model compound of ejected oligomers during pyrolysis.
........................................................................................................................................................................... 154
Figure 26. Importance of aerosol characterization in condenser design.................................................................... 156
Chapter 5
Figure 1. Schematics of the hot plate reactor (1) External housing for cooling (2) Glass window (3) Port for gas phase
thermocouple (4) Pyrometer (5) Valve for sampling gases and pulling vacuum (6) Top cover reactor (7) Reaction
chamber (8) Pressure sensor port (9) Copper electrode (10) Gas tight electrical feed (11) Teflon insulation (12)
Screws for electrodes (13) Steel plate heated by electrodes (14) Fittings for electrode screws (15) Cold Plate (16)
Fittings to adjust the plate and feed cooling or hot fluids (CO2) (17) Copper sinks. (18) Port for carrier gas (Purge).
........................................................................................................................................................................... 163
Figure 2. Power and control scheme for temperature control. ................................................................................... 163
Figure 3. Verification of the temperature distribution in the central area of the steel sheet cold rod. Figure A shows what
the plate looks like after heating, which illustrates that there are distinct heating regions. The pyrometer looks at
point 1 (B) on the plate. The other numbered points indicate placement of the thermocouple for comparison to IR
camera readings. At the end the calibration was verified by comparing melting points of 5 pure metals (indium,
tin, zinc, lead, and aluminum) to the corrected temperature from the pyrometer. .............................................. 164
Figure 4. Preparation of thin film of bagasse on the steel plate (A) Raw ground biomass (B) Steel plate (C) Biomass
impregnated on the steel plate (D) Char after pyrolysis..................................................................................... 165
Figure 5. Viewing the distribution of biomass particles on the steel sheet. (A) 0.64X (B) 1.0X (C) 1.6X (D) 2.5X (E) 4.0X
(F) Increase 20X ................................................................................................................................................ 165
Figure 6. (A) Uncalibrated pyrometer vs centerpoint thermocouple temperature readings used to develop the
calibration curve. (B) Calibrated pyrometer temperature vs melting points of pure metals at atmospheric pressure.
Differences between the melting points and thermocouple were less than 10 °C. ............................................ 168
Figure 7. Comparison of temperature distribution in the steel sheet. (A) Temperature measurements with type K
Thermocouple at 9 points around the central region. Tf = 400 °C, holding time 60s. (B) Frame 240 of sheet
temperature Tf = 400 °C. ................................................................................................................................... 169
Figure 8. Temperature gradients at different times taken with infrared camera when the plate is heated at 50 °C/s and
600 ºC.B. Temperature distribution at different times when plate is heated at 200 °C/s and final temperature of
400 °C. .............................................................................................................................................................. 170
Figure 9. Comparison of records temperature on the surface of the steel sheet and on the biomass surface. ......... 170
Content XV
Figure 10. Film thickness impregnated biomass on the surface of cold steel rod. (A) Visible Light. (B) Fluorescent Light.
........................................................................................................................................................................... 171
Figure 11. A. Char yield versus film thickness for pyrolysis at 500 °C (500 °C/s, 150 mbar pressure, hold time 0s). B.
Char yield versus hold time for film thickness of 60 ± 13 µm at 500 ° C (500 °C/s, 150 mbar pressure). ........ 171
Figure 12. Distribution of light oxygenates analyzed by GC-MS and obtained by sugarcane bagasse pyrolysis in a hot
plate reactor at 1200 °C/s and 5 °C/s and 0 holding time. (A) Water yield (wt.%). (B) Acetoin yield (wt.%). (C)
Formic acid yield (wt.%). (D) Acetic acid yield (wt. %). (E) Furfuryl alcohol yield (wt.%). (F) Glycoaldehyde yield
(wt.%)................................................................................................................................................................. 173
Figure 13. Yields of 5 member ring compounds analyzed by GC-MS obtained by sugarcane bagasse pyrolysis in a hot
plate reactor at 1200 °C/s and 5 °C/s and 0 holding time. (A) Furan yield (wt. %). (B) 5-HMF yield (wt.%)...... 174
Figure 14. Distribution of aromatics products analyzed by GC-MS, and obtained by devolatilization from sugar cane
bagasse (A) Naphtalene (B) Phenol (C) Pyrocatechol (D) Eugenol (E) Vanillin (F) Benzofuran. ...................... 175
Figure 15. A. UV-fluorescence spectra of bio-oil samples obtained by sugarcane bagasse pyrolysis in a hot plate
reactor at 300 °C, 400 °C, 600 °C heating rate of 500 °C/s, 0 holding time. B. Standards of eugenol, vanillin, 1,
2, 3-trimethoxy benzene, and levoglucosan. ..................................................................................................... 176
Figure 16. Effect of temperature and heating rate on pyrolytic lignin yields measured by UV-Fluorescence Spectroscopy
and obtained by devolatilization from sugar cane bagasse in a hot plate reactor at 1200 °C/s and 5 °C/s and 0
holding time. ...................................................................................................................................................... 176
Figure 17. Effect of temperature and heating rate on sugars yields measured by HPLC and obtained by devolatilization
from sugarcane bagasse in a hot plate reactor at 1200°C/s and 5 °C/s and 0 holding time. The sugars yield
represents cellobiosan, cellotriosan, and 1,4:3,6 dianhydroglucopyranose (DGP). ........................................... 177
Figure 18. Distribution of permanent gases at different temperatures and heating rates. ......................................... 178
Figure 19. Distribution of Char depending on the temperature and heating rate. ...................................................... 179
Figure 20. SEM visualization of morphological changes of sugarcane bagasse chars. ............................................ 180
Figure 21. Effect of heating rate on product distribution during the pyrolysis of sugarcane bagasse in a hot plate reactor
at 500 °C (773K)................................................................................................................................................ 181
Chapter 6
Figure 1. Humidity comparison measured by GC-MS vs. Karl Fisher. The red dashed line represents the best fit line
for the parity plot. ............................................................................................................................................... 193
Figure 2. Mathematical fitting of specified light oxygenated compounds released during Sugarcane bagasse pyrolysis
in a hot plate reactor at different temperatures and Heating rates at 150 mbar. Experimental data at 1200 °C/s
(o); Experimental data at 500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o);
Experimental data at 50 °C/s (o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at
1200 °C/s (--); Fitting at 500 °C/s (--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting
at 10 °C/s (--); Fitting at 5 °C/s (--). ................................................................................................................... 195
Figure 3. Mathematical fitting of specified oligomers compounds and water released during Sugarcane bagasse
pyrolysis in a hot plate reactor at different temperatures and Heating rates at 150 mbar. Experimental data at 1200
°C/s (o); Experimental data at 500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s
(o); Experimental data at 50 °C/s (o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting
at 1200 °C/s (--); Fitting at 500 °C/s (--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--);
Fitting at 10 °C/s (--); Fitting at 5 °C/s (--). ........................................................................................................ 196
Figure 4. Mathematical fitting of permanent gases released during Sugarcane bagasse pyrolysis in a hot plate reactor
at differen temperatures and heating rates at150 mbar. Experimental data at 1200 °C/s (o); Experimental data at
500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o); Experimental data at 50 °C/s
(o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at 1200 °C/s (--); Fitting at 500 °C/s
(--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting at 10 °C/s (--); Fitting at 5 °C/s
(--). ..................................................................................................................................................................... 198
Figure 5. Solid residue conversion at different Heating rates and temperatures. Experimental data at 1200 °C/s (o);
Experimental data at 500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o);
XVI Content
Experimental data at 50 °C/s (o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at
1200 °C/s (--); Fitting at 500 °C/s (--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting
at 10 °C/s (--); Fitting at 5 °C/s (--). ................................................................................................................... 199
Figure 6. Evaporation and secondary reactions of lignin oligomers and levoglucosan. (A).Sugarcane bagasse drying
by halogen balance. (B). Evaporation of 5 mg of levoglucosan at different heating rate in Thermobalance (C).
Thermogravimetric behavior of Pyrolytic lignin (lignin oligomers) at different heating rates. (D). Thermogravimetric
behavior of 20 mg of levoglucosan at different heating rates............................................................................. 200
Chapter 7
Figure 1. Scheme for the differential element of a particle of biomass in a slab.(A).Representation of biomass structure
(views and cuts). (B).Abstract Representation of biomass structure in order to represent the mathematical model.
........................................................................................................................................................................... 210
Figure 2. Dimensions of sugarcane bagasse fibbers. ................................................................................................ 212
Figure 3. Differential element for global mass and energy balance. .......................................................................... 212
Figure 4. Biomass film configuration for external temperature measurement.(A).Frontal view of temperature
measurement on biomass surface to model validation.(B). Top view of set-up for temperature measurement. 222
Figure 5. Mesh independency for simulation verify with last node temperature of biomass particle.......................... 223
Figure 6. Comparison of experimental data vs. simulation for temperature profiles of the sugarcane bagasse film
surface at different thicknesses, heating rates and final reaction temperature. (A) Temperatures profiles
(Simulated and experimental) on biomass surface for a 0.5mm particle thickness at different heating rates (50°C/s
and 500°C/s) and temperatures (400°C-600°C). (B) Temperatures profiles (Simulated and experimental) on
biomass surface for a 1mm particle thickness at different heating rates (50°C/s and 500°C/s) and temperatures
(400°C-600°C). (C) Temperatures profiles (Simulated and experimental) on biomass surface for a 1.5mm particle
thickness at different heating rates (50°C/s and 500°C/s) and temperatures (400°C-600°C). (D) Temperatures
profiles (Simulated and experimental) on biomass surface at 600°C and 500°C/s and particle thickness between
0.5mm-1.5mm.................................................................................................................................................... 224
Figure 7. Verification of proposed model results vs. experimental data reported by Sepman et al [80] for 1mm biomass
particle pyrolysis in a Hot Plate Reactor at 600°C final temperature. ................................................................ 225
Figure 8. Comparison of the experimental and theoretical values of chemical compounds released during sugarcane
bagasse pyrolysis at 500 K/s using a 10 s of hold time and different particle thickness and temperatures.(A).Acetic
acid yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (B).Furfuryl
alcohol yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (C).5-
HMF yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (D).Phenol
yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (E). Eugenol
yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (F).Vanillin yield
at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (G).Carbon monoxide
yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. (H). Carbon
dioxide yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at 500°C/s. ...... 226
Figure 9. Comparison of the experimental and theoretical values for char, levoglucosan, water and aerosols released
during sugarcane bagasse pyrolysis at 500K/s with a 10 s hold time, and different particle thicknesses and
temperatures.(A).Char yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at
500°C/s. (B).Water yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at
500°C/s. (C).Levoglucosan yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm)
at 500°C/s. (D).Aerosol yield at different temperatures (400°C500°C) and particle thickness (0.5mm-1.5mm) at
500°C/s.............................................................................................................................................................. 227
Figure 10. Verification of the solid residue yield predicted by the model using the experimental results reported by
Drummond et al. [83] and Aiman et al. [84] for 0.4 mm sugarcane bagasse particles pyrolyzed at 1000 °C/s, using
1 s and 30 s hold times. ..................................................................................................................................... 228
Figure 11. Intraparticle dynamics. (A). Pressure and pressure gradient profiles at z=0 mm, z=0.5 mm and z=1 mm. (B).
Char density profile at z=0 mm, z=0.5 mm and z=1 mm. with volume shrinkage shown on the second axis. (C)
Content XVII
CO2 density profile at z=0 mm, z=0.5 mm and z=1 mm. (D) Levoglucosan density profile at z=0 mm, z=0.5mm
and z=1 mm. (E) Bubble density profile at z=0 mm, z=0.5 mm and z=1 mm and cumulative bubbles density
(second axis). (F). Bubble size distribution at z=0 mm, z=0.5 mm and z=1 mm................................................ 229
Figure 12. Bubbles dynamics at different heating rates and final temperatures. (A) Bubbles distribution at z=0.5mm
and heating rates 1000°C/s-10°C/s and 600°C-400°C. (B). Bubbles distribution at z=0.5mm and different reaction
times (0s, 1.5s, 5s). (C). Cumulative total bubbles density (Bubbles/m3) at different heating rates and final
temperature. ...................................................................................................................................................... 231
Figure 13. Deconvolution of main bubble formation dynamics, during devolatilization of a 1mm sugarcane bagasse
particle at 1000°C/s and 500°C. ........................................................................................................................ 233
Figure 14. Effect of particle size on product distribution during fast pyrolysis of bagasse at 1000°C/s and 500°C. Using
particles of 0.5 mm, 1.0 mm and 2.0 mm........................................................................................................... 236
Figure 15. Products global yield predicted by the model during pyrolysis with different particle thickness at 1000°C/s
and 500°C. ........................................................................................................................................................ 237
Figure 16. Effect of temperature on the product distribution obtained from the fast pyrolysis of a 1 mm thick particle of
sugarcane bagasse at heating rates of 1000 °C/s. ............................................................................................ 239
Figure 17. Global product yields predicted by the proposed model during pyrolysis at different reaction temperatures,
with a heating rate of 1000 °C/s and 1 mm particle thickness. .......................................................................... 239
Figure 18. The effect of heating rate on the product distribution obtained from the fast pyrolysis of 1 mm sugarcane
bagasse at 500°C. ............................................................................................................................................. 241
Figure 19. Global products yields as predicted by the proposed model from the pyrolysis of a 1 mm sugarcane bagasse
particle at 500°C using different heating rates................................................................................................... 241
XVIII Content

Symbols
Chapter 4
AI Distribution of number of aerosols No Bubbles·kg-1
mo Initial biomass mass kg
mf Final biomass mass kg
mslide−after Slide mass after experiment kg
mslide−before Slide mass before experiment kg
Aslide Slide área m2
Aimage Visualized slide area m2
D Aerosol diameter m
F(D) Function distribution of number of aerosols No Bubbles·kg-1
Atotal Total number of captured aerosols m2
𝑉 Volume m3
𝜃 Contact angle Degrees
𝑎 Droplet radius m
𝜌 Aerosol density kg·m-3
F(V) Aerosol distribution function No Bubbles·m3

Chapter 6
Pafter experiment Pressure after experiment mbar
Pbefore experiment Pressure before experiment mbar
Vreactor Reactor volume l
Troom Room temperature K
R Ideal gas constant mbar·l·K-1·mol-1
Mwi Molecular weight of species i kg·kmol-1
%moli Molar fraction of species i
mbiomass Biomass mass k
Mwmixture Molecular weight of mixture kg·kmol-1
Ci Mass fraction
mg Gas mass kg
mo Initial mass kg
dm(E) Amount of differential reacted mass kg
ko Frequency factor s-1
E Activation energy kJ·mol-1
T Temperature K
m(E) Total reacting species mass g
f(E) Reaction activation energy distribution function mol·kJ-1
σ Standard deviation kJ·mol-1
E0 Average activation energy kJ·mol-1
V∗ Maximum volatile mass of species i kg
V(T) Volatile mass of species i kg
K DAEM Kinetic coefficient from DAEM s-1
Y∗ Final yield of species i %w/w
Content XIX
Y(T) Yield of species i %w/w
Levevap Levoglucosan mass kg
K vap Evaporation kinetic coefficient s-1
𝑚 Biomass weight kg
𝑚0 Initial biomass weight kg
𝑚∞ Biomass weight at equilibrium kg
X(T) Solid residue conversion % Wt.

Chapter 7
ri Reaction rate kg·m-3·s-1
mi Mass of species i pero volume kg·m-3
K DAEM−i Kinetic coefficient s-1
Yield∗i Final yield species i per volume kg·m-3
ρbiomass Biomass density kg·m-3
Yi∗ Final mass yield of species i %w/w
ρi Density of species i kg·m-3
εj Volumetric fraction of phase j m3j·m-3particle
ṙ sec−i Reaction rate of secondary reaction kg·m-3·s-1
δij Stoichiometric coefficient mass basis of species i Dimensionless
koi Frequency factor s-1
Eai Activation energy kJ·mol-1
R Ideal gas constant mbar·l·K-1·mol-1
T Temperature K
εk Volumetric fraction, phase where secondary reaction occurs m k·m-3particle
3

V Particle volume m3
Vo Initial particle volume m3
Vf Final particle volume m3
ρs Solid phase density kg·m-3
ρl Liquid phase density kg·m-3
ρ0 Particle density kg·m-3
ρcf Final char density kg·m-3
εg Gas phase volumetric fraction m3g·m-3particle
εgs Gas fraction in solid phase m3g·m-3s
εgl Gas fraction in liquid phase m3g·m-3l
𝑣𝑔 Gas velocity in solid phase pores m·s-1
𝑣𝑏 Average velocity in liquid phase m·s-1
Deffv Vapor effective diffusivity m2·s-1
ρv Vapor density kg·m-3
DeffN2 Nitrogen effective diffusivity m2·s-1
ρN2 Nitrogen density kg·m-3
εl Liquid phase volumetric fraction m ·m-3particle
3

K lvg Levoglucosan evaporation rate s-1


Ylvg Levoglucosan yield %w/w
İaerosol Aerosol ejection intensity kg·m-3·s-1
XX Content
ρlvg−v Levogrucosan density in gas phase kg·m-3
Defflvg Levoglucosan effective diffusivity m2·s-1
Ḃb Bubble bursting rate No Bubbles·m-3·s-1
Naerosol Number of aerosols No Aerosols
Nb Number of bubbles No Bubbles
R aerosol Aerosol radius m
F(t, r, x) Number of bubbles in node x No Bubbles·m-3
Ḋ Bubble growth rate m·s-1
J̇nuc Nucleation rate No Bubbles·m-3·s-1
δ Dirac delta function Dimensionless
rb Bubble radius m
r∗ Minimum gas cluster radius to form a bubble m
Mn Momentum n
rn Length scale (Radius) m
φ Probability of two adjacent bubbles to grow enough to coalesce See Apendix C
α Factor of the size of a new bubble formed by coalescence See Apendix C
σ Standard deviation N Bubbles-1·m3·s1
o

tb Bubble bursting time s


No Formation frequency factor of new bubbles No Bubbles·m-3·s-1
∆Ecri Minimum free energy to form the first gas cluster kJ
Kb Boltzmann constant kJ·K-1
̅̅̅̅
Cpi Heat capacity of liquid phase kJ·kg-1·K-1
K eff Thermal conductivity W·m-1·K-1
Cpg Heat capacity of gas phase kJ·kg-1·K-1
ṙ water Drying rate of water kg·m-3·s-1
∆Hvap Enthalpy of vaporization of water kJ·kg-1
∆Hrxnj Enthalpy of reaction kJ·kg-1
K lvg Evaporation kinetic coefficient s-1
∆Hvaplvg Enthalpy of vaporization of levoglucosan kJ·kg-1
K rad Radiation contribution W·m-1·K-1
Ug Gas phase permeability m2
μg Gas phase viscosity Pa·s
yi Molar fraction of species i
ϑr Aerosols retention efficiency Dimensionless
ω Degree of compaction of porous medium Dimensionless
a BaggaseTotal internal área respect to outer area of the particle m2
̂dporo Average pore diameter m
d̂aero Average aerosol diameter m
ABET Specific surface area from BET isotherm m2·g-1
ho Biomass particle thickness m
K mass Mass transfer coefficient s-1
yi∞ Accumulation of gas species around biomass
Ni Mass flux kg·m-3·s-1
Nbubbles Number of bubbles No Bubbles
Content XXI
Naerosol Number of aerosols No Aerosols
Dbubbles Diameter of bubbles m
Daerosol Diameter of aerosols m
22 General Introduction

0. General Introduction
Interest in the development of biofuels derived from thermochemical processes from biomass has grown in recent years
as a promising alternative in mild and long term to supply partially rising energy consumption [1]. Thermochemical
conversion processes use mostly agro-industrial wastes as main source of raw materials, so they do not threaten food
security, not required complex pre-treatments and reaction times are short; which are striking advantages over
biological transformation processes for the same purposes [2]. The main thermochemical processes implemented in
the industry for the production of fuels and energy from biomass are gasification, combustion, pyrolysis and torrefaction
[3–5]. Pyrolysis consist of thermal decomposition of the lignocellulosic material in an inert atmosphere or with oxygen
shortage (less than stoichiometric), at temperatures between 500°C - 600°C, with particle sizes between 0.5mm - 2mm,
with the aim of produce the highest yield of a liquid fuel with a high oxygen content, called bio-oil, a solid residue ‘’bio-
carbon’’ or char and non-condensable gases with low calorific value (rich in CO, CO2 CH4, H2) which can be recirculated
in the reactor to supply heating power [6]. Depending on the operating conditions of the reactor, pyrolysis can be
classified as: slow pyrolysis and fast pyrolysis [7].

The first case is typical of processes at low temperatures, with large particles (> 2mm), normally in fixed bed reactors.
The main products are char (50-75% yield), bio-oil (15-20% yield) and non-condensable gases (20-30% yield) [8,9].
Contrary to this, in fast pyrolysis processes the main product is bio-oil (50-70% yield), whose yield is limited to the type
of biomass, operating temperature, residence time of the gases into the reactor, heating rate, particle size, content and
type of mineral material [10–13]. To achieve high selectivity towards the production of bio-oil high heating rates of the
particle are required, higher than 1000 °C/s, particle sizes between 1mm - 5mm, reactor temperatures between 400 °C
- 600 °C and residence time of gases under 1 s. The chemical composition and quality of the final bio-oil is related to
many factors, such as the nature of the raw material, heating rate, particle size, reaction temperature, among many
others [14–17]. Bio-oil recovered is a viscous liquid, containing a mixture of compounds with high oxygen content
(aldehydes, furans, oligomers lignin and sugars, organic acids, alcohols, phenols), limiting its direct use as a fuel [18,19].

Although various technologies have been developed for the production of bio-oil via fast pyrolysis with some of them
already working [11,20,21], some operational problems have emerged associated with short operating times (hours),
bio-oil low yields, lack of reproducibility in testing, low thermal stability of bio-oil, among others, that do not allowed its
mass industrial scale [6]. The main reason associated with these problems is the lack of some vital level phenomena
that affect the final particle distribution and product performance [22,23].

The phenomenological models proposed at particle and reactor scale are not accurate in predicting the chemical
species released during the devolatilization, this happens because of the lack of knowledge in the main chemical and
physical behavior of lignocellulosic materials during pyrolysis. A lot of effort is required to develop particle and reactor
scale model because of the rigor of heat and mass transfer phenomena inside the particle (convection, diffusion) and
morphological changes (swelling, shrinking, and pore structure). However the chemical description are poorly
developed, generally as lumped models. Despite of the plenty information available in the literature, few models
describe the dynamics of main species (primary products) at different reaction conditions (temperatures and heating
rates).

In order to obtain more accurate models that allow to track primary products and to develop ‘’chemical tunning’’ for a
desired product, it is necessary to know the kinetics parameters of primary products released during the desvolatilization
stage. The coupling of kinetics models fitted in ‘’kinetics control’’ regime with important intra-particle physical
phenomena bring information to develop new technologies for reactors and separations systems (condensers), to
improve the quality of bio-oil.
General Introduction 23

Among these physical phenomena is the formation of the intermediate liquid phase generated in the process of
depolymerization of the biomass, the dynamics of bubble formation within this phase, affecting the mechanisms of
formation and ejection of aerosols inside and outside of the particles [23–28]. Many researchers have begun to apply
the dynamics of this phase as the main step in the whole pyrolysis process [26,27,29,30]; i.e. the understanding of the
formation and consumption of this stage can allow the explanation of the main products formed during devolatilization,
including char. The importance in description and understanding of these phenomena is an open door towards the
development of new product recovery processes such as fractional condensation, the use of catalysts or foaming agents
that guide reactions towards high value - added products, such as glucose and lignin oligomers or other products of
interest. It will also allow the development of new mathematical models with better predictive capacity.

Considering everything above, in particle level, the pyrolysis of biomass is a multistage process (Figure 1) involving
primary reactions in solid phase and liquid, heterogeneous reactions between volatile and char generated and / or the
liquid phase. In recent years, it has been demonstrated by visualization techniques that during devolatilization of
biomass, a liquid phase is formed, generally composed of glucose oligomers and / or pyrolytic lignin [26,31–34],
resulting in a multiphase reaction system. The existence of this liquid phase has been important to explain the existence
of high molecular weight compounds such as anhydro-sugars DP> 2 and pyrolytic lignin in the bio-oil by aerosol ejection
mechanisms [26,28,35]. Within this liquid phase occur simultaneously many physiochemical phenomena like solid /
liquid reactions, resistance to mass transfer and energy in the liquid-solid-gas interface, nucleation mechanisms,
coalescence, death and growth of bubbles, among others, which have not yet been clarified.

Figure 1. Schematic multi-scale fast pyrolysis biomass.

There are many restrictions from the instrumental and reactors point of view to fully understand and study these
chemical (kinetics) and physical phenomena, mainly due to short reaction times and heating (<1s), and limitations with
sampling systems and analysis of the reaction products in real time [22]. Despite these, first steps have begun to take
using visualization techniques, coupled reactors with low heating rates, all aimed to elucidate structural and
morphological changes before and after devolatilization.

Some of these reported studies focus on the use of SEM, TEM [36–39] and other optical techniques [37,40] to elucidate
the changes in the internal structure of particles after pyrolysis. This information is often correlated with surface area
analysis [41,42], to provide quantitative information on changes of the porous structure before and after
pyrolysis. Further analysis by FT-IR, Raman Spectroscopy, XRD, H-MNR, have also been used to describe structural
changes from the chemical point of view (functional groups, types of links, etc.). Despite all this, due to the complexity
24 General Introduction
of the technical and technological limitations, these efforts have been reduced to comparative studies of solid structures
before and after pyrolysis [38,43,44]; in the best case, studies in real time using commercial hot-stages reactors, are
limited to low heating rates (max 100 ° C / min) [40,45].

The important contribution of the development of this thesis is a first step in the exploration of the chemical and physical
phenomena that occur during the fast pyrolysis of biomass, which have not been described yet, with the help of new
methodologies, and reactors design that allow in-situ viewing of the biomasses during devolatilization. Some of these
phenomena are the description of kinetics of main primary products released during the devolatilization stage at different
heating rates and temperatures, and also the description of physical phenomena, like bubble formation and dynamics
for model compounds present in biomass and the distribution of aerosols formed during pyrolysis of biomass and its
model compounds (cellulose, hemicellulose, and lignin). The above studies are the link for finally develop a new single
particle model that describes the evolution of main species released during the devolatilization stage, the bubbles
dynamics inside the melted phase (intermediate liquid phase) and the aerosol ejection intensity. For this goal, a new
prototypes of reactors that allows to set controlled reaction conditions and to capture and visualize the released products
during devolatilization are developed, together with new methodologies as thin film to reach as closer possible the
‘kinetics control’. Sugarcane bagasse was chosen for this thesis after a project of investigation between ECOPETROL-
Universidad Nacional de Colombia (308010010868-Producción de bioaceite por pirólisis rápida– ECOPETROL)
identified four promising biomasses for energy production in Colombia that are available in high quantities. These
species were sugarcane bagasse, palm rachis, banana rachis, and pine sawdust. The criteria for choosing these four
were availability in Colombia, pre-processing requirements before pyrolysis, and bio-oil yield.

Figure 2. Yields of global products during fluidized bed pyrolysis of different biomass [10]

Pine sawdust has a high volatile content and produces close to 75 %w/w in bio-oil (Figure 2). However it is only available
in remote places, so it was disregarded as a practical option. Banana rachis has a high content of moisture and ash,
which forms slags in a fluidized bed. In addition, the high moisture content of this material impedes to use it directly for
fast pyrolysis. Although it is possible to add pre-drying steps, more work is still needed to find the best conditions for
making drying economically attractive for high moisture content materials. Also the bio-oil yield is low and with high
water content. For these reasons it was also discarded.

The final decision was between sugarcane bagasse and palm rachis. Both produce high bio-oil yields (Figure 2) and
low char yields. However, palm rachis requires significant pre-processing before feeding it into reactors. Due to the
fibrous nature of palm rachis, 4 milling steps are required; but the milling is expensive and likely not feasible for industrial
applications. Table 1 shows the availability and energy content of the types of biomass discussed herein. Based on the
availability in Colombia, energy potential, bio-oil yield, and low pre-processing steps before pyrolysis, sugarcane
bagasse was selected as the strategic feedstock to produce bio-oil for future use in oil refineries. Other studies
General Introduction 25
developed in faculty of mines facilities, shows the potential of sugarcane bagasse to produce bio-oil by fast pyrolysis
processes [47].

Table 1. Main biomass availability and energetic potential in Colombia. Source: UPME [46].
Biomass Availability (Millions of t/year) Energetic potential (TJ/year)
Sugarcane bagasse 12.68 139086
Palm rachis 0.924 6607
Bannana rachis 5.1 2231
Pine sawdust Not available Not available

Dissertation objetive
This project aims at advancing the current state of knowledge of biomass fast pyrolysis with the helps of new
methodologies and rector configurations. In order to understand and develop mathematical models of pyrolysis process
it is needed information about primary products and main physical phenomena that contribute to explain the products
distribution. These phenomena can be studied by employing hot plate reactors and light bulb reactors as tools, where
conditions can be easily controlled, and it may be possible to visualize and to recover products for post analysis. This
thesis aims to study physical and chemical phenomena that still non-clarified and then to bring new information about
the actual behavior of biomass during pyrolysis process. The following specific objectives are proposed:

 To propose a new methodology to study the behaviour of biomass and model compounds during fast pyrolysis
process, specially the formation of molten phase, and it´s contribution to bubbles dynamics and aerosol
ejection.
 To propose new methodologies, to quantify total the aerosol distribution and intensity during biomass and
surrogates during fast pyrolysis in different conditions.
 To investigate the product distribution and kinetics fitting of main products released during sugarcane bagasse
fast pyrolysis in a hot plate reactor operrated in vaccum to assure kinetics control and analysis of primary
products.
 To couple for the first time the physical phenomena like bubbles distributions and aerosol ejection, also the
kinetics parameters fitted experimentally in a single particle model that describes the behaviour of biomass
during pyrolysis process.
 To investigate the behaviour of sugarcane baggase during pyrolysis in a fluidized bed reactor wich is typical
in industrial processes.

Methodology
In order to bring clarity about the sequence in this thesis, the flow char, in figure 3, presents the main topics to study
and the experimental tools used for each task. The thesis is divide into has 5 sections.

First section (Chapter 1): Literature review, focusing in parametric study that affect the product yields, kinetics
schemes, evidence of intermediate liquid phase, and particle modelling.
26 General Introduction
Second section (Chapter 2-4): These chapters, study important physical phenomena that happen during pyrolysis,
such as the formation of intermediate liquid phase, the bubbles dynamics within liquid phase, and characterization of
aerosols collected during pyrolysis of biomass and its constituents. In chapters 2 and 3, it is described a new reactor
that allows the visualization of the main events during pyrolysis of biomass using a fast speed camera. These chapters
also describe the morphology behaviour of model compounds during pyrolysis process. Using image-processing tools
the bubbles dynamics inside the melted phase in lignin and cellulose surrogate (sucrose) were described. In chapter 4
it is proposed a new methodology to study aerosols ejection intensity and distribution by SEM characterization.

Third Section (Chapter 5-6): Study of chemical phenomena such as products evolution during the pyrolysis of thin
layer of sugarcane bagasse in a new vacuum-hot plate reactor. In this section, a new methodology for pyrolysis studies
that allows tracking the main species released during the devolatilization (including heavy oligomers) is presented. Also,
were fitted experimental data to DAEM (Distribution of Activation Energy Model) to obtain kinetics parameters useful
for particle modelling.

Fourth Section (Chapter 7): In this chapter a new single particle model for biomass pyrolysis that couple physical
phenomena such as bubble dynamics and aerosol ejection, together with kinetics of main species released during the
pyrolysis is presented.

Figure 3. Flow chart diagram of main topics studied in this work.

Scientific contributions
1. The work develops a new pyrolysis reactors (light bulb reactor, hot plate reactor) and applies new
methodologies (thin film) to study pyrolysis process that allows the study of chemical species released by
primary reactions.
2. The in-line visualization of the main morphological changes of Lignocellulosic materials and it’s components
during pyrolysis process has beed conducted for the first time.
3. The development of methodologies to study the bubbling dynamics within intermediate liquid phase, formed
during lignocellulosic materials and surrogates has beed conducted for the first time.
General Introduction 27
4. The development of a new methodology to characterize the aerosols ejected during bubbles bursting within
intermediate liquid phase formed by pyrolysis of lignocellulosic materials and model compounds has not
previously been reported in the literature. In this work, the effect of reaction pressure and heating rate over
aerosol intensity is presented for the first time.
5. A new methodology combining thin film impregnation of biomass, vacuum operation to study primary products
released during sugarcane bagasse at low and high heating rates.
6. The work present for the first time a single particle model that couples main physical phenomena such as liquid
phase formation, bubbles dynamics, and aerosol ejection together with chemical phenomena such as kinetics
reactions of principal species released during the devolatilization. The model also includes secondary
reactions of heavy oligomers (Pyrolytic lignin) as an important way to explain the secondary char formation.
The model predict product yields including the aerosol intensity and the bubbles behavior within liquid phase
as well.
28 General Introduction

Publications
Books and book chapter
 Pyrolysis rápida de biomasas. ECOPETROL-Universidad Nacional de Colombia. Medellin 2014. ISBN: 978-
958-761-774-0.
 Book Chapter: Pecha B, Smith M, Montoya J, Garcia-Perez M: Pyrolyse rapide de la biomasse
lignocellulosique. In: Chemie pour la transformation durable de la resource lignocellulosique. Edited by
Tatjana Stevanovic. 2016.

Papers published
 Montoya Jorge, Chejne Janna, Farid, Garcia-Pérez, Manuel. Fast pyrolysis of biomass: A review of relevant
aspects. Part I: Parametric study. Dyna. 2015. V 82 issue 192. Pg 239-248. DOI:10.15446/dyna.v82n192.44701.
 Jorge Montoya, Brennan Pecha, David Roman, Farid Chejne Janna, Manuel Garcia-Pérez. Effect of
Temperature and Heating Rate on Product Distribution from Pyrolysis of Sugarcane Bagasse in a Hot Plate
Reactor. in Journal of analytical and applied pyrolysis. 2016.
 Montoya, J.I.Valdés, C.Chejne, F.Gómez, C.A.Blanco, A.Marrugo, G.Osorio, J.Castillo, Aristóbulo, J.Acero, J.
Bio-oil production from Colombian bagasse by fast pyrolysis in a fluidized bed: An experimental study. Journal
of analytical and applied pyrolysis.2015.V 112. Pg 379-387.DOI : 10.1016/j.jaap.2014.11.007.
 Montoya .J, B. Pecha, F. Chejne Janna, M. Garcia-Perez. Micro-Explosion of Liquid Intermediates during the
Fast Pyrolysis of Sucrose and Organosolv Lignin. Journal of analytical and applied pyrolysis. 2016. In press.
DOI:10.1016/j.jaap.2016.10.010.
 Carlos F. Valdés, Gloria Marrugo, Farid Chejne, Juan David Román, Jorge I. Montoya. Effect of atmosphere
reaction and heating rate on the devolatilization of a Colombian sub-bituminous coal. Journal of analytical and
applied pyrolysis. 2016. V 121. September 2016, Pages 93–101. DOI: 10.1016/j.jaap.2016.07.007.
 Carlos F. Valdés Farid Chejne, Gloria MarrugoR, Robert J. Macias,Carlos A. Gómez, Jorge I. Montoya, Carlos
A. Londoño, Javier De La Cruz, Erika Arenas. Co-gasification of sub-bituminous coal with palm kernel shell in
fluidized bed coupled to a ceramic industry process. Applied Thermal Engineering. 2016. V 107. Issue 25. Pg
1201-1209.DOI: 10.1016/j.applthermaleng.2016.07.086.
 Macías, R. J., Chejne, F., Montoya, J. I. y Blanco, A. (2014). Gasificación de bagazo de caña y carbón en planta
piloto. Revista Mutis 4(1); pag. 24-32.
 Jorge Montoya, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez. Identification of the Fractions
Responsible for Morphology Conservation in Lignocellulosic Pyrolysis: Visualization Studies of Sugarcane
Bagasse and its Pseudo-Components. Journal of analytical and applied pyrolysis. 2016.
 Jorge Montoya, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez. Single Particle Model for
Biomass Pyrolysis with Bubble Formation Dynamics inside the Liquid Intermediate and Its Contribution to
Aerosol Formation by Thermal Ejection. Journal of analytical and applied pyrolysis 2016.

Papers Submitted

 Jorge Montoya, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez. Methodology for Estimation of
Thermal Ejection Particle Size Distribution and Intensity during the Pyrolysis of Sugarcane Bagasse and its
Model Fractions. To be submitet to Journal of analytical and applied pyrolysis 2017.
General Introduction 29

Papers in preparation
 Jorge Montoya, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez. Kinetics of Pyrolysis of a
Sugarcane Bagasse-Thin Film in a Hot Plate Reactor. To be submitet to Journal of analytical and applied
pyrolysis 2017.
 Jorge Montoya, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez.Coupling single particle model
for Sugarcane bagasse slabs with DAEM kinetics model. To be submitet to Journal of analytical and applied
pyrolysis 2017.
 Jorge Montoya, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez. Effect of Pressure and Heating
rate on sucrose and organosolv lignin micro-explosions. To be submitet to Journal of analytical and applied
pyrolysis 2017.
 Jorge Montoya, David Roman, Brennan Pecha, Farid Chejne Janna, Manuel García-Pérez. Comparisson of
product yields obtained by rice Straw and Sugarcane bagasse at high Heating rates in a hot plate reactor. To
be submitet to Journal of analytical and applied pyrolysis 2017.
30 General Introduction

References
[1] T. Damartzis, A. Zabaniotou, Thermochemical conversion of biomass to second generation biofuels through
integrated process design—A review, Renew. Sustain. Energy Rev. 15 (2011) 366–378.
[2] R.M. Baldwin, B.S. Donohoe, M.R. Nimlos, K.A. Magrini-Bair, J.E. Hensley, P. Pepiot, S.D. Phillips, Current
research on thermochemical conversion of biomass at the National Renewable Energy Laboratory, Appl. Catal.
B Environ. 115–116 (2012) 320–329.
[3] P. Basu, Biomass gasification and pyrolysis practical design and theory, Academic Press, Burlington, MA, 2010.
[4] P. Basu, Biomass Gasification and Pyrolysis, © 2010 Elsevier Inc., 2010.
[5] C.F. Valdés, F. Chejne, G. Marrugo, R.J. Macias, C.A. Gómez, J.I. Montoya, C.A. Londoño, J. De La Cruz, E.
Arenas, Co-gasification of sub-bituminous coal with palm kernel shell in fluidized bed coupled to a ceramic
industry process, Appl. Therm. Eng. 107 (2016) 1201–1209.
[6] A. Bridgwater, Fast Pyrolysis of Biomass: A Handbook, 3rd ed., PyNe, 2008.
[7] J. Lédé, Biomass Pyrolysis: Comments on Some Sources of Confusions in the Definitions of Temperatures and
Heating Rates, Energies. 3 (2010) 886–898.
[8] O. Onay, O. Mete Koçkar, Fixed-bed pyrolysis of rapeseed (Brassica napus L.), Biomass and Bioenergy. 26
(2004) 289–299.
[9] Ö.M. Koçkar, Ö. Onay, S.H. Beis, Fixed-bed pyrolysis of safflower seed: influence of pyrolysis parameters on
product yields and compositions, Renew. Energy. 26 (2002) 21–32.
[10] J.I. Montoya, C. Valdés, F. Chejne, C.A. Gómez, A. Blanco, G. Marrugo, J. Osorio, E. Castillo, J. Aristóbulo, J.
Acero, Bio-oil production from Colombian bagasse by fast pyrolysis in a fluidized bed: An experimental study,
J. Anal. Appl. Pyrolysis. 112 (2015) 379–387.
[11] R.H. Venderbosch, W. Prins, Fast pyrolysis of biomass for energy and chemicals: Technologies at various
scales, in: Sustain. Dev. Process Ind. Cases Impact, Harmsen J., Jhon Wiley & Sons, Inc, 2010: p. 270.
[12] J.I. Montoya Arbeláez, F. Chejne Janna, M. Garcia-Pérez, Fast pyrolysis of biomass: A review of relevant
aspects. Part I: Parametric study, Dyna. 82 (2015) 239–248.
[13] A.V. Bridgwater, Review of fast pyrolysis of biomass and product upgrading, Biomass and Bioenergy. 38 (2012)
68–94.
[14] C. Branca, C. Di Blasi, Multistep mechanism for the devolatilization of biomass fast pyrolysis oils, Ind. Eng.
Chem. Res. 45 (2006) 5891–5899.
[15] C.A. Mullen, A.A. Boateng, Characterization of water insoluble solids isolated from various biomass fast
pyrolysis oils, J. Anal. Appl. Pyrolysis. 90 (2011) 197–203.
[16] F. Broust, J. Lédé, F.-T. Ndiaye, M. Ferrer, Properties of bio-oils produced by biomass fast pyrolysis in a cyclone
reactor, Fuel. 86 (2007) 1800–1810.
[17] P.K. Kanaujia, Y.K. Sharma, M.O. Garg, D. Tripathi, R. Singh, Review of analytical strategies in the production
and upgrading of bio-oils derived from lignocellulosic biomass, J. Anal. Appl. Pyrolysis. 105 (2014) 55–74.
[18] T. Cheng, Y. Han, Y. Zhang, C. Xu, Molecular composition of oxygenated compounds in fast pyrolysis bio-oil
and its supercritical fluid extracts, Fuel. 172 (2016) 49–57.
[19] M. Garcia-Perez, A. Chaala, H. Pakdel, D. Kretschmer, C. Roy, Characterization of bio-oils in chemical families,
Biomass and Bioenergy. 31 (2007) 222–242.
[20] M.-K.K. Bahng, C. Mukarakate, D.J. Robichaud, M.R. Nimlos, Current technologies for analysis of biomass
thermochemical processing: A review, Anal. Chim. Acta. 651 (2009) 117–138.
[21] R.H. Venderbosch, W. Prins, Fast pyrolysis technology development, Biofuels, Bioprod. Biorefining. 4 (2010)
178–208.
[22] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[23] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials, Curr. Opin.
Biotechnol. 24 (2013) 414–20.
[24] A.R. Teixeira, R.J. Hermann, J.S. Kruger, W.J. Suszynski, L.D. Schmidt, D.P. Schmidt, P.J. Dauenhauer,
Microexplosions in the upgrading of biomass-derived pyrolysis oils and the effects of simple fuel processing,
ACS Sustain. Chem. Eng. 1 (2013) 341–348.
[25] A.R. Teixeira, Transport Limitations in Zeolites and Biomass Pyrolysis, University of Massachusetts - Amherst,
2015. Doctoral thesis.
General Introduction 31
[26] A. Texeira, K. Mooney, J. Kruger, L. Williams, W. Suzinsky, L. Shmidt, D. Schmidt, P. Dauenhauer, Aerosol
generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011) 4306.
[27] S. Zhou, M. Garcia-Perez, B. Pecha, A.G. McDonald, R.J.M. Westerhof, Effect of particle size on the
composition of lignin derived oligomers obtained by fast pyrolysis of beech wood, Fuel. 125 (2014) 15–19.
[28] L.D.S. Paul J Dauenhauer, Joshua L Colby, Christine M Balonek, Wieslaw J Suszynski, Reactive boiling of
cellulose for integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[29] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[30] M. Zhou, S., Pecha, B., van Kuppevelt, M., McDonald, A.G. and Garcia-Perez, Slow and fast pyrolysis of
Douglas- Fir Lignin:Importance of liquid intermediate formation on distribution of products, Biomass and
Bioenergy. (2014) 398–409.
[31] J. Lédé, Cellulose pyrolysis kinetics: An historical review on the existence and role of intermediate active
cellulose, J. Anal. Appl. Pyrolysis. 94 (2012) 17–32.
[32] C. Di Blasi, Heat transfer mechanisms and multi-step kinetics in the ablative pyrolysis of cellulose, Chem. Eng.
Sci. 51 (1996) 2211–2220.
[33] S. Nordin, J. Nyren, E. Back, An Indication of Molten Cellulose Produced in a Laser Beam, Text. Res. J. 44
(1974) 915–917.
[34] D.A.I. Goring, Thermal Softening of Lignin, Hemicellulose and Cellulose, Pulp Pap. Mag. Canada. 64 (1963)
T517--T527.
[35] S. Saka, M. Asmadi, H. Kawamoto, Gas- and solid/liquid-phase reactions during pyrolysis of softwood and
hardwood lignins, J. Anal. Appl. Pyrolysis. 92 (2011) 417–425.
[36] E. Bolat, K. Açıkalın, F. Karaca, Pyrolysis of pistachio shell: Effects of pyrolysis conditions and analysis of
products, Fuel. 95 (2012) 169–177.
[37] T.J. Haas, M.R. Nimlos, B.S. Donohoe, Real-Time and Post-reaction Microscopic Structural Analysis of
Biomass Undergoing Pyrolysis, Energy & Fuels. 23 (2009) 3810–3817.
[38] R. Bassilakis, R.. M. Carangelo, M.. A. Wójtowicz, TG-FTIR analysis of biomass pyrolysis, Fuel. 80 (2001)
1765–1786.
[39] G.A. Zickler, W. Wagermaier, S.S. Funari, M. Burghammer, O. Paris, In situ X-ray diffraction investigation of
thermal decomposition of wood cellulose, J. Anal. Appl. Pyrolysis. 80 (2007) 134–140.
[40] G. Zhu, X. Zhu, Z. Xiao, F. Yi, Study of cellulose pyrolysis using an in situ visualization technique and
thermogravimetric analyzer, J. Anal. Appl. Pyrolysis. 94 (2012) 126–130.
[41] W. Li, K. Yang, J. Peng, L. Zhang, S. Guo, H. Xia, Effects of carbonization temperatures on characteristics of
porosity in coconut shell chars and activated carbons derived from carbonized coconut shell chars, Ind. Crops
Prod. 28 (2008) 190–198.
[42] L. Sun, P. Fu, S. Hu, S. Su, J. Wang, J. Xiang, Evaluation of the porous structure development of chars from
pyrolysis of rice straw: Effects of pyrolysis temperature and heating rate, J. Anal. Appl. Pyrolysis. 98 (2012)
177–183.
[43] Z. a. Mayer, A. Apfelbacher, A. Hornung, Effect of sample preparation on the thermal degradation of metal-
added biomass, J. Anal. Appl. Pyrolysis. 94 (2012) 170–176.
[44] V. Agarwal, G.W. Huber, W.C. Conner, S.M. Auerbach, Simulating infrared spectra and hydrogen bonding in
cellulose Iβ at elevated temperatures., J. Chem. Phys. 135 (2011) 134506.
[45] F. Yi, G. Zhu, X. Zhu, R. Zhou, Z. Xiao, Pyrolysis characteristics of bean dregs and in situ visualization of
pyrolysis transformation, Waste Manag. (2012).
[46] O.M. Hernandez, E. Humberto, Orduz, P. janneth, Zapata, L. Henry, Cardona, R. Maria Cecilia. Duarte, Atlas
del Potencial Energético de la Biomasa Residual en Colombia, 2013.
www1.upme.gov.co/sites/default/files/article/1768/files/Atlas de Biomasa Residual Colombia__.pdf.
[47] G. Marrugo, C.F. Valdés, F. Chejne, Characterization of Colombian Agroindustrial Biomass Residues as Energy
Resources, Energy & Fuels. 30 (2016) 8386–8398.
32 Chapter 1

1. Chapter 1. State of the Art


The total worldwide energy consumption in recent years is estimated to be approximately 515 EJ/year (10^18 J/year),
80% of which is supplied by petroleum fuels [1,2]. Global energy consumption has shown exponentially increasing
trends over time due to phenomena associated with population growth and increased per capita demands from
emerging countries such as Brazil, Russia, India and China (BRIC) [3,4]. Biomass, solar radiation, wind, water and
geothermal are all potential alternative energy sources that can be used in the near future to offset fossil energy
consumption. Of these sustainable sources, biomass is the only source of carbon for renewable fuels production that
can be quickly inserted into the liquid fuels markets [5]. Colombia has a total biomass availability of about 72 million
t/year when residuals from agricultural activities are considered, including: sugarcane bagasse, banana, corn, coffee,
rice, and oil palm [6]. The energy potential of these residuals is close to 332 million GJ/year, equivalent to 7.4% of the
total energy consumption in the country [6,7]. These numbers show that there exist a huge opportunity to use waste
biomass for biofuels production in the country. Fast pyrolysis processes offer an attractive potential technology for the
rapid thermal processing of agro-industrial wastes for bio-oil production. The development of this technology is relevant
and appropriate with benefits over current biofuel productions methods, In particular: (1) it does not compete with food
security, (2) unlike the transesterification and fermentation processes, commonly used to produce ethanol and biodiesel,
and (3) can be easily incorporated into existing petroleum refining infrastructure, achieving a reduction in operating
costs [8–12].

Pyrolysis processes require heating of biomass at moderate temperatures between 400-600°C, in an inert or oxygen-
less atmosphere to produce bio-oil, biochar and non-condensable gases with high calorific value (rich in CO, CH4, H2).
As the pyrolysis process progresses, volatiles are released as both condensable compounds (bio-oil) and non-
condensable gases (pyrolysis gas or Py-gas). The remaining solid product (biochar) from this process is high in carbon.
The pyrolysis process produces a solid residue (biochar) that is high in carbon with a low oxygen/carbon ratio, similar
to high rank coal, as shown in the Van Krevelen diagram in Figure. 1. Bio-oil can be integrated into refineries as a raw
material while biochar can be integrated into the process to supply reaction energy and process heat. Low ash biochar
is also attractive for metallurgical applications as a reducing agent or as a gas/water filtration medium. Finally, pyrolysis
gas can be combusted and recirculated into the reactor to provide heat, or can be used in subsequent processes such
as catalytic methane reforming or Fisher Tropsch process [13,14].

Figure 1. Van Krevelen diagram.


Chapter 1 33
Biomass contains a mixture of polymers with different physicochemical structures resulting in a thermal decomposition
that progresses in stages. Initially, hemicellulose, breaks down at temperatures between 200-300 °C. This is followed
by the degradation of cellulose, which decomposes between 300-450 °C and, finally, lignin decomposition occurs
across the full temperature range of 250-500 °C, but is the last polymer to fully degrade. Lignin is the major contributor
to char production at the end of pyrolysis [15–18]. Depending on operating conditions, specifically heating rate, the
pyrolysis process is classified typically as slow pyrolysis, intermediate pyrolysis and fast pyrolysis. Slow pyrolysis is
used primarily for char production (≈30-35% yield), bio-oil (≈30% yield) and non-condensable gases ( 30% w/w) [19],
intermediate pyrolysis produce close to 20%w/w in char,30%w/w in permanent gases and close to 50%w/w in bio-oil
[12,20]. In contrast, in the fast pyrolysis processes, the main product is bio-oil (50-80% yield) with minor yields of biochar
(10-20%w/w) and gas (≈10%w/w) [12,21]. In each case, yields are strongly influenced by the type of biomass, operating
temperature, gas residence time in the reactor, heating rate, particle size and mineral content. High heating rates are
required, for achieving high selectivity towards the production of bio-oil. A particle size between 1-5 mm is required to
achieve high heating rates throughout the entire particle. Reactor temperatures between 400-600 °C and gas residence
time less than 1s are also necessary to maximize bio-oil yields. Although various technologies have been developed
for producing bio-oils via fast pyrolysis, some of which put into commercial operation with developed technologies
(Ensyn, BTG, Dynamotive, GTI, KIT, VTT, PyTec), there have been problems associated with short operating times
(hours), low bio-oils yields, non-repeatability in testing, and low thermal stability for oils that prevents direct use in
turbines or boilers [22,23].

Several poorly understood particle-level phenomena are associated with the operational challenges that affect final
products distribution, including: formation of an intermediate liquid phase, bubble formation and rupture within the liquid
phase, and aerosol ejection. The composition and reactivity of the biomass affects the formation of the intermediate
liquid phase during the thermal depolymerization process of biomass; thus affecting the mechanisms of gas transport
within the particle. Information about the existence and form of this intermediate liquid phase during the pyrolysis of
lignocellulosic material is very limited. To date there are no studies that provide an in-depth examination of the bubble
formation dynamics inside the liquid phase or the contribution of bubbles within this phase to aerosols ejection. In this
chapter, the principals parameters associated with biomass fast pyrolysis, the focus of this thesis, are reviewed. Current
studies related to the identification of the intermediate liquid phase and particle modelling are also examined.

1.1. Biomass Structure


Biomass can be defined as an organic material composed primarily of a heterogeneous mixture of polymers and a
small fraction of inert materials. The organic fraction consists of polymers containing three major macromolecules:
cellulose, hemicellulose, and lignin as shown in figure 2. In addition to these primary polymers, some biomasses contain
a small quantity of lipids, pectin, and extractives, which do not exceed 10% w/w [24,25].
34 Chapter 1

Figure 2. General structure of lignocellulosic materials. Adapted from [26]

Cellulose corresponds to about 40 – 60% of the total biomass weight, hemicellulose to 15 - 25% and lignin (15 - 25%)
[26–28]. Cellulose is a linear, mostly crystalline, polymer composed of glucose structural units linked by β-D type
glycosidic bonds. These monomeric units link together to form chains of 10,000 or more structural units (degree of
polymerization). Cellulose polymers form fibers inside biomass, and are responsible for the fibrous nature of cell walls.

Hemicellulose is an amorphous, branched polymeric structure, with a substantially lower degree of polymerization than
cellulose (approximately 100). It is primarily composed of sugars with 5 and 6 carbons per unit (xylose, mannose,
arabinose, galactose and glucose).

Lignin is a complex, branched, non-crystalline macromolecule, composed of three primary aromatic monomeric
constituents: sinapyl, coniferyl, and coumaryl alcohols (see figure 2). linked by β-ether linkages and other variety of C-
C and C-O-C bonds exist between monomer units of lignin as β-O-4, α-O-4, β-5, 5-5, 4-O-5, β-1, and β-β bonds [28,29].
Both lignin and hemicellulose are responsible for binding cellulose fibers together, maintaining the structure of the cell
walls [29,30].

Mineral matter content depends on the type of biomass, but comprises 2 - 25% of the total solid weight and commonly
consist of minerals such as Na, K, Ca, Mg, Mn, Co, Zn, and Cu. These minerals are present as like oxides and salts
such as chlorides, carbonates, phosphates and sulphates [31–33]. Mineral matter plays an important role in thermal
decomposition of biomass, acting as catalyst for critical reactions, such as dehydration and catalytic cracking of the
volatiles.

1.2. Studies of structural parameters in biomass pyrolysis


Final products and yields obtained from the pyrolysis of lignocellulosic material [34–37] and model compounds directly
related to cellulose crystallinity [38–40], degree of polymerization and bond orientation in cellulose structure [41,42],
type and structure of lignin molecules [43,44], and interactions between pseudo-components inside biomass solid matrix
[17,45–47] have been widely reported. Ponder et al. [41] studied the effects of the orientation and position of glucan
links (cellulose constituent units) in the production of volatiles. These studies show that the position and orientation of
linkages do not have a significant effect on levoglucosan yields (the main compound obtained from the devolatilization
of cellulose). It is important to note that this study is only an approximation, since only low and medium molecular weight
glycosides (150 Dalton) with behavior that differs from cellulose molecules (polymerization degree > 1000) were used.
Mettler et al. [42] examined the degradation of pure cellulose, mixtures of cellulose, and other constituents of biomass
Chapter 1 35
to understand whether or not a relationship between the bonds orientation and the formation of volatiles exists for
macromolecules more consistent with whole biomass. The work demonstrated for cellulose that at high degrees of
polymerization the production of levoglucosan increases and the formation of char decreases. The points under
maximum stress, and therefore the most feasible to fracture during pyrolysis, are precisely those that are located at the
interface of the amorphous and crystalline structures [48]. It has been established [49] that the maximum temperature
required to achieve cellulose decomposition is higher for crystalline cellulose structures than for amorphous cellulose.
This is due to the extra energy required to destroy the crystal lattice structure formed by inter-chain hydrogen bonds. In
addition, the ring structure of the glucose units is more easily maintained in crystal structures due to the network of
hydrogen bonds. The crystalline lattice acts as a thermal energy sink; therefore, the fragmentation pattern is not as
wide as in amorphous cellulose, from which an array of compounds with a low degree of polymerization are obtained
(glucose, levoglucosan, cellobiosan, cellotriosan, etc.).

Due to the complex structure of lignin within biomass, the final structure of separated lignins shows significant
differences in molecular weight and aromatic structures distribution based on the separation methods employed and
biomass used [50,51]. Product yields have been found to vary substantially depending on the type of lignin used for
pyrolysis [43,52]. Saka et al. [43] reported that lignins from softwoods tend to pyrolyze at lower temperatures and
produce more volatiles and less char than lignins derived from hardwoods. This observation is attributed to the
molecular structure of hardwoods lignins, which are more complex, and have a high molecular weight than softwood
lignins, providing greater resistance to thermal decomposition. In the following chapters, it will be shown that during
organosolv lignin decomposition, an intermediate liquid phase is formed that has a lower viscosity than the one formed
from extracted ball milled lignin. This promotes gas bubble formation in the liquid phase and aerosol ejection by bubbles
bursting [53]. Unfortunately, there are very few studies regarding the effect of the type and nature of lignins on the
product distribution during pyrolysis.
36 Chapter 1

1.3. Effect of Particle Size


Biomass from any source is a poor conductor of heat. Because of this it readily develops temperature gradients within
particles, when subjected to rapid heating. This gradient is positively correlated to particle size, meaning that the size
and shape of the biomass particle affect the residence time of volatiles inside the particle [54–56].

Increasing vapor residence time favors cracking reactions that reduce the yield of condensable vapors [24,27,57].
Therefore, secondary reactions between char and volatiles are more prominent in larger particles, which increases the
yield of char and gases at the expense of boil-oil production [58].Recent studies have shown that as particle size is
decreased by milling, pore walls within the biomass are destroyed, favoring the rapid escape of volatiles and high
molecular weight compounds during pyrolysis [59,60]. Recently, Zhou et al. [61] showed the effect of particle size on
the formation of char and lignin oligomers during beech wood pyrolysis. The authors suggest two distinct pyrolysis
regimes based on particle size to explain the oligomer yields obtained. In the first regime, for particles with diameters
between 0.3 mm – 3 mm, lignin oligomer yield decreased when particle size was increased. For the second regime, for
particles > 3mm, no changes were observed in oligomer yield when particle size changed. The authors hypothezed that
lignin oligomers come from aerosol ejection from the intermediate liquid phase, and these are retained within the particle
solid matrix, with the retention efficiency increasing with particle size. In contrast, char yield increased with increasing
particle size. A reasonable explanation for this behavior is the increase in char formation by crosslinking reactions of
aerosols trapped inside the particle matrix. In general, particle size is directly related to the heating rate; therefore, for
fast pyrolysis processes the formation of bio-oils is favored with the high heating rates obtained with small particle sizes
(< 1 mm ) [22,24,32,58,62–64].

It should be clarified that, in industrial or pilot scale processes, the costs associated with grinding to reduce the biomass
particle size are high, and this may limit the profitability of the process. Bridgwater et al. [64] presented specifications
for particle sizes to be used with each of the major technologies available. For example: particle sizes of less than 200
mm are recommendable for jet rotary cone technology, less than 6 mm for fluidized bed reactor and circulating bed
designs, and less than 10 cm are common for ablative rotary disc processes.

Shena. J et al. [65] found that bio-oil yields obtained from mallee (Australian Eucalyptus) in a fluidized bed reactor
decreases as the particle diameter increases from 0.3 to 1.5 mm, and thereafter bio-oil yield is insensitive to particle
size increases. Onay et al. [66] reported that the maximum bio-oil yield obtained from fixed-bed pyrolysis of safflower
is reached with a 0.42 mm particle size. Onay et al. [36] and Nurul et al. [67], have reported ideal particle sizes of
between 0.85 to 200 mm for fixed-bed pyrolysis of rapeseed, with bio-oil yields of up to 60% (w/w) obtained. Laboratory
scale investigations of primary biomass pyrolysis reactions have shown that rapid removal of products from within the
solid matrix is critical.This removal is facilitated by the use of small particle sizes (< 100 μm) and vacuum conditions
[68,69]. Other researchers have suggested that to successfully study primary reactions impregnation surface
techniques are required to obtain uniform biomass or model compounds films of less than 100 μm, in which internal
resistance to energy transfer is negligible, and the exhaust of reaction products is nearly instantaneous [70].

1.4. Effect of mineral matter


Biomass contains traces of mineral material (K, Na, Ca, Zn, Cu, P, Mg, Mn), usually in the form of oxides or salts such
as chlorides, sulphates, carbonates and phosphates [31,33,71–73]. The proportions of these minerals is dependent on
the type of biomass, as well as collection, drying and grinding conditions. The content of inorganic material affects
product yields and is an important parameter for the study of side reactions; generally, as the mineral material content
Chapter 1 37
increases, bio-oil yields (and levoglucosan) decrease and gas and char production increase [15,23,24,26,32,63,64,74].
Changes in product yield are largely due to the catalytic effect of the mineral matter on several reaction pathways
including the dehydration and decarboxylation reactions [75]. Alkalis tend strongly interact with functional groups such
as –COOH and -OH, forming alkali-oxygen clusters that promote cracking reactions [73]. Additional effects on reaction
pathways have also been observed; for example, sodium restricts cellulose and hemicellulose trans glycosylation
reactions, and facilitates demethoxylation, demethylation, and dehydration of lignin [76]. The presence of silicon has
also been found to decrease the production of volatiles [77]. In contrast to the effects of alkali and alkaline metals,
adding anions (chloride, nitrates, sulfates) has been found to increase levoglucosan production and decreases char
yields [78]. In recent years, numerous works have been published that examine the effect of mineral matter on final
product distribution, [31,33,72,79–81]. However, despite these efforts, accurate theoretical models for the prediction of
the effects of inorganic material on the product distribution of fast pyrolysis processes are not currently available.

Raveendran et al. [72] evaluated the effect of mineral content on the yield of volatiles and char, for 20 different biomass
samples, which were studied with and without a demineralization step (acid washing). These tests demonstrated that
higher char yields are obtained for the biomass without washing, than those that were acid washed prior to pyrolysis.
From these results, up to 57% decrease in the char yield and increases of 21% in the production of volatile compounds
was observed for corn cobs pyrolysis. A separate study undertaken by Patwardah et al. [75] evaluated the effect of
mineral matter on the yield of char and volatiles obtained during pyrolysis of Avicel microcrystalline cellulose. In these
studies, the authors used particles of 50 microns to minimize the effects of heat and mass transfer, and conducted the
pyrolysis tests in a Frontier free fall pyrolysis micro reactor coupled to a gas equipment, to ensure high heating rates
(5000 °C/s). In the study, the authors impregnated cellulose with solutions of 0.5 to 5% w/w of salts such as NaCl, KCl,
MgCl2.6H2O, CaCl2, Ca(OH)2, Ca(NO3), CaCO3, CaHPO4, to evaluate the effects of these minerals exert a greater
influence on the global production of volatiles products, specifically: levoglucosan, acetol, furaldehyde, 5HMF (Hydroxy-
methyl-furfural), glycol aldehyde and formic acid. They determined that low concentrations, e.g. 0.5% w/w, increase
char yields by up to 10% and also increase the yield of low molecular weight species such as acetol, glycol aldehyde
and formic acid while levoglucosan yield decreases. The minerals with the greatest impact on volatile yields were, in
order: K+, Na+, Ca+2, Mg+2, Cl-, NO-, OH-, CO2-, PO3-.. In another study, Yang H.et al. [33] Studied the effect of KCl,
K2CO3, Na2CO3, CaMg (CO3)2, Fe2O3 and Al2O3 on the yield of char produced in the slow pyrolysis of cellulose,
hemicellulose and pure lignin using a thermogravimetric analyzer. Inorganic and the organic ratios of 1:10 w/w were
used for each salt and biomass combination (cellulose, hemicellulose, and lignin). According to the authors, none of
the minerals studied, except K2CO3, had any appreciable effect on the yield of char, in contrast with results reported by
Shimomura K et al. [82], Patwardah P et al. [31], Richards G et al. [77], Nik-Azar et al. [71], Zsuzsa A et al. [78].

To study the effects of mineral matter on the volatile and char yields, samples of biomass (natural or synthetic) are
subjected to washing with deionized water and with acidic solutions (HCl, H2SO4, HNO3, H3PO4), or the addition of
minerals in solution (Na2 (NO3), K2(NO3), Ca(NO3), etc.) [83,84]. Each of these pre-treatment methods has limitations,
which make it difficult to fully explain the results based solely on the effects of mineral matter. Acid washing for example,
dissolves hemicellulose and cellulose (hydrolysis) almost entirely, [78,84]; therefore, the results obtained are due to a
combination of effects, creating uncertainty regarding which effects produce the greatest influence on yields.

Recently Oudenhoven et al. [85,86], evaluated the effect of washing 5 biomasses with acetic acid on the yields of
monomers, sugars oligomers, and phenolics. The authors report that washing with an acid solution (pH 2) at 90 ° C,
does not alter the native structure of cellulose and lignin of biomass; however, it removed a substantial fraction of
mineral material, resulting in inorganic fractions below 90 mg/kg. After acid washing, dehydration and fragmentation
reactions and were minimized, resulting in a single phase bio-oil. Between 30 - 40% of sugars contained in native
38 Chapter 1
biomass (cellulose, hemicellulose) were recovered in the bio-oil after acid washing compared with only 5 - 10% recovery
from untreated biomass.

The presence of mineral matter has also been found to affect morphological changes during biomass pyrolysis. Iwasaki
et al. [87] showed that a possible cause of particles agglomeration inside fluidized bed reactors is the formation of an
intermediate liquid phase during devolatilization, that covers all particles inside the fluidized bed and promotes
agglomeration. The formation of this liquid phase is favored for biomasses that have been pretreated with acids.
Visualization by SEM of inert material particles from fluidized beds shows a coating which acts as glue for other particles.
The formation of this film is most evident for biomasses that were previously pre-treated with organic acids [88].
Oudenhoven et al. [88] have shown that part of the solid residue left by biomasses treated with acid has lost its original
structure, forming new structures similar to a melted polymer. Here cavities can be seen that are formed when bubbles
explode inside the liquid phase. Biomasses that have not been acid pre-treated tend to maintain their original fibrous
structure after pyrolysis.

As it will be shown in Chapter 2, the behavior of bagasse and Xylan during fast pyrolysis are altered by the presence
of mineral matter. Xylan pre-treated with organic acids rapidly forms a liquid phase, losing its initial morphology. In
contrast, Xylan without pre-treatments maintains its original structure during devolatilization.

1.5. Effect of volatile and solids residence time


It is accepted that short volatile residence time increases the bio-oil yield [7,32,59,89,90]. For the majority of pyrolysis
reactor designs, the residence time of volatiles and solids are different. High solid biomass residence times are needed
to ensure complete devolatilization, while volatile time should be short to minimize secondary reactions [32,59,91].
These times are directly related to the operating conditions (such as carrier gas flow for fluidized beds) and the
technology used in the pyrolysis process. For instance, the residence time of solids in fixed bed reactors is higher (t →
∞), while in fluid bed reactors the residence time for both, solids and volatile, tends to be very short (a few seconds)
depending on the height of the reactor and the gas flow. Scott et al. [92] have observed that for the pyrolysis of sorghum
bagasse at 550 °C, the bio-oil yield drastically reduces from 75% to 57%, while char yields increase as the volatile
residence time increases. An increase in the organic vapor residence time promotes recondensation reactions of
biomass constituents (cellulose, lignin, and hemicellulose) due to long time available for reactions. Some authors
suggest that short volatile residence times (< 1s) can be counterproductive, because the depolymerization reactions
are incomplete, resulting in increased oligomer production [19]. The volatile residence time in a fluidized bed reactor is
manipulated by changing the carrier gas flow [7,93,94], however it can also be altered by gas phase pressure. At high
pressures, the volatile residence time within particles is high, promoting secondary cracking reactions [54,95,96]. In
reactivity studies using lab scale reactors, in order to minimize secondary reactions, it is preferable to have short
residence time of volatiles inside the particle. In wire mesh and hot plate reactors this condition is commonly achieved
by operating the reactor under vacuum (< 200mbar) [68,97,98]. Under these conditions, evaporation of high molecular
weight oligomers from cellulose decomposition is favored at low temperatures. Aerosol ejection from the intermediate
liquid phase is also more intense under these conditions due to bubble bursting [99,100].

1.6. Effect of reaction temperature


The devolatilization process is sequential and depends on the temperature reached by the particle (see Figure 3A). The
first stage occurs between 20 -120 °C and corresponds to water evaporation from the biomass. The second stage
between 120 - 300 °C shows no appreciable biomass weight loss, though small quantities of some permanent gases,
Chapter 1 39
such as CO, CO2 and steam, are released, primarily from dehydration reactions and decarboxylation of R-CO-R’ groups
belonging to hemicellulose and lignin. [28,58,101]. The third stage, for cellulose, occurs between 200 and 350 °C
[63,102–104] and relates to the depolymerization reactions that give rise to oligomers and anhydrous sugars
(levoglucosan, cellobiosan, cellotriosan, glucose, etc.) that form an intermediate liquid phase known by many
researchers as ''active cellulose” and ''molten cellulose¨ or simply ''intermediary liquid compound.’’ [105–107].The final
stage occurs between 300 and 400 °C and results in the maximum weight loss (about 80%), this corresponds with the
maximum rate of volatiles released. In this stage, random fragmentation of glucosidic bonds in cellulose, hemicellulose
and their oligomers generate volatile compounds with high oxygen content, leaving a carbonaceous residue known as
char or biochar [21,57,108]. At temperatures above 400 °C, CO, CO2 are released by depolymerization reactions from
the lignin-rich aromatic carbonaceous matrix.

The effect of temperature on bio-oil production is well understood and has widely been explored in the literature. It has
been shown that here is a parabolic behavior for bio-oil production with temperature the maximum bio-oil yield is
obtained between 400 - 550°C, depending on the type of biomass [7,11,31,59,87,109,110]. As temperature is increased
initial bio-oil yields increase because the rate of biomass devolatilization increases (this is reflected in char reduction).
At higher temperatures, volatile cracking reactions are promoted, which reduce the yield of bio-oil and increase the
production of non-condensable gases. [65,111–113]. The effect of temperature on the production of bio-oil by cellulose
(Avicel) and maple fast pyrolysis in a fluidized bed and entrained flow reactor was studied by Scott et al. [114]. In this
study, particle size and volatile residence time were held constant at 100 µm, and 500 ms respectively, while the
process temperature was varied between 400 - 700 °C. The authors found that maximum production of bio-oil was
achieved in the fluidized bed reactor at 450 °C, with yields of 80% and 90% for maple and cellulose, respectively. For
entrained flow reactors, yields were 70 and 60%, respectively, at 600 °C. Kalgo et al. [115] pyrolyzed sawdust with
particle diameters of 0.7 mm in a fluid bed reactor to estimate, the effect of temperature on bio-oil yields. They found
that, the highest bio-oil yields are achieved at 450 °C (60 % w/w). Similar results are also reported by Hajaligol et al.
[116,117] and Dauenhauer et al. [116,117] .

A. Thermal decomposition of biomass B. Temperature profile

Figure 3. (A)Stages of thermal decomposition for biomass. (B) Temperature profile within a biomass particle.

In addition the bio-oil chemical composition is altered. [118]. At low temperatures (< 300°C), most volatiles belongs to
hemicellulose degradation and the siccion of cellulose weaky linkages, being the main products levoglucosan,
levoglucosenone, hydroxymethyl furfural, acetic acid, acetone, Guiacyl acetone, glyoxal, methanol and formic acid. Fu
et al. [119] demonstrated the formation of primarily short chain vapors and gases, including: formic acid, methanol,
ethane, ethylene, formaldehyde, acetone, hydrogen cyanide, carbon monoxide and carbon dioxide, during pyrolysis of
cotton stalk, corn and rice, at a temperature range of 300 - 400 °C. At temperatures above 400 °C, intact rings such as
benzene, carboxylic acids, phenols, p-cresol, furans, furfural, were found in important amounts (> 1% w/w) [31]. As the
40 Chapter 1
temperature increases further, the yield of these volatile compounds decreases and the yield of more thermally stable
polyhydroxyl aromatic compounds (PHA) increases. For instance, the yields of aldehydes, alcohols, paraffin, and
acetone gradually react, at temperatures between 580 - 800°C, givin new aromatic compounds such as benzene,
naphthalene, cresol, toluene [32,101,120].

Recently, Garcia-Perez et al. [121], studied the effect of reaction temperature on lignin oligomer yield during malle wood
pyrolysis in a fluidized bed reactor. The results show that lignin oligomer yields increase with temperature until 500 °C
due to the increased rates of depolymerization reactions with temperature. At temperatures higher than 500 °C,
decomposition reactions for aerosols, which contain these oligomers, become more intense, generating secondary
lower molecular weight volatiles. Westerhof et al. [59], reported similar results for pine pyrolysis in a fluidized bed
reactor. Between 380 - 480 °C, the authors reported an increase in lignin oligomer yields, attributing this behavior to
increased biomass conversion. At higher temperatures, oligomer yields decreased, presumably due to cracking
reactions.

Despite the wide assortment of studies conducted to evaluate the effect of temperature on the yield of char, bio-oil and
non-condensable gases, there is ambiguity in the definition of the optimum temperature of pyrolysis. Unfortunately, in
both laboratory scale and industrial scale pyrolysis equipment, it is not possible to directly measure the temperature of
the biomass particles, and so it is generally assumed to be equal to the reactor temperature [58,103] or in some cases
is estimated by mathematical modeling [104,122]. For slow pyrolysis conditions utilizing small particle sizes (less than
1mm), the assumption that the biomass particle temperature is equal to the reactor temperature is valid due to the
higher heating times, which allows thermal equilibrium between the biomass particles and surroundings to be achieved.
[105,112]. However, these assumptions are often far from reality, especially for fast pyrolysis conditions, resulting in
mismatches (thermal lag) of temperatures as high as 100 °C or more. In general, temperature lags are greater when
heating is undertaken at high temperatures or high heating rates. This thermal lag causes errors in estimating the actual
particle temperature, which can result in over-estimates of kinetic parameters such as the activation energy and the
coefficient of the reaction rate [112,113].

1.7. Effect of Heating Rate


The heating rate of biomass particles is the main parameter that allows us to obtain a difference in products and yields
between slow pyrolysis and fast pyrolysis processes. The particle heating rate is direcltly related to the time it takes the
biomass particle to reach reaction temperature. It has been reported that heating rates of the order 1 °C/min – 100
°C/min are typical for slow pyrolysis processes [24,32], while fast pyrolysis processes require heating rates higher than
1000 °C/min [21,123]. High heating rates promote cellulose and hemicellulose depolymerization reactions and minimize
volatile residence time inside the particle, and limits secondary reactions. Because of this, condensable gas release
occurs quickly [124], achieving high yields of bio-oil and the lowest production of char. Chaiwat et al. [125] has proposed
a set of competitive reactions between cellulose depolymerization and dehydration, as shown in fig 4. These results
are supported by Agarwal et al. [126] who, through computer simulations, studied the reactions of cellulose during fast
pyrolysis process. These authors found that heating rate affects hydrogen-type bonding in cellulose structures. At low
temperatures and low heating rates, strong intra-chain hydrogen bonds of cellulose functional groups, increasing the
probability of collisions that produce dehydration reactions. At high heating rates, intra-chain hydrogen bonds weaken,
resulting in greater separation between the cellulose chains and thus decreasing the possibility of collisions that facilitate
dehydration reactions.
Chapter 1 41

Figure 4. Effect of the heating rate on the mechanism of thermal decomposition of the cellulose. [125]

There are many reports in the literature that point out the importance of heating rate effects on bio-oil and char yields.
Ozlem et al. [127] studied the fast pyrolysis of safflower seeds in a fixed-bed reactor, showing the heating rate effect
on bio-oil and biochar yield by evaluating the pyrolysis of 3 grams of this sample (particle size 1.25 mm) at heating rates
of 100, 300 and 800 °C/min. The results showed that for heating rates above 300 °C/min, maximum yield of 55 % bio-
oil and minimum yield of 17 % for char were achieved. Thangalazhy et al. [128] conducted a study at the micro-scale,
using a pyroprobe reactor (Pyroprobe model 5200, CDS Analytical Inc., Oxford, PA), to evaluate the effect of
temperature and heating rate on yield and the distribution of some compounds in the bio-oil. The experiments were
carried out with pine and grass fodder, and heating rates of the pyrolyzer filament of 50, 100, 500, 1000, 2000 °C/s. An
interesting result of this study was that at a constant temperature (550 °C), regardless of the filament heating rate, the
final product distribution was the same (30% yield for pine wood bio-oil and 17% for forage grass). This result is
attributed to the slow heat transfer through biomass (Poor thermal conductivity) resulting in a heating rate that is lower
than that of the filament. Although these experiments were carefully prepared, yields are far lower than usually reported
by other researchers, [127,129,130]. The heating rate also plays an important role on the quality of bio-oil and char
structure obtained. Bio-oils can be obtained with lower moisture content at higher heating rates, primarily due to the
inhibition of secondary reactions, such as volatile dehydration and cracking [131]. Higher heating rates also reduce the
formation of the water-soluble fractions (formic acid, methanol, acetic acid) and aromatic fractions (rich in phenol and
its derivatives). In contrast, CO and CO2 yield increase with higher heating rates [64]. Ketones, levoglucosan, phenol
and toluene yields also increase with heating rate for the fast pyrolysis of pine wood and grasses when heating rates
exceed 50 °C/s. Heating rate can generate significant pressure gradients between the inside and outside of the particle,
because the volatiles are produced quickly and do not have the ability to be instantly evacuated. This causes some
internal structures to crack, increasing the quantity of both macropores and micropores. In addition, when the biomass
passes through a melting phase (metaplastic), especially at high heating rates, the internal structures can warp and
block pores [118,119].
42 Chapter 1

An interesting study was conducted by Zhou, et al. [132] in which they showed the importance of the intermediate liquid
phase formation during slow and fast pyrolysis of douglas fir lignin. The authors showed that for both slow pyrolysis (2
K/s) and fast pyrolysis (1000 K/s), lignin decomposes to form an intermediate liquid phase, which swells, evaporates
and shrinks to form a droplet. They observed no further details of the bubbles inside this intermediate liquid phase. At
high heating rates, an intense foaming of oligomeric compounds in the liquid phase occurs, which may be ejected from
the surface as vapors or aerosols through a reactive ejection mechanism [133]. At low heating rates, the intermediate
liquid provides an ionic medium in which cross-linking reactions can occur, promoting char formation and decreasing
high molecular weight oligomers yield [124]. No work has been found that examines the effect of heating rate on aerosol
ejection intensity or bubble dynamics inside the intermediate liquid phase. However, analogy can be made to other
substances, better studied liquids where it has been proven that during boiling of a liquid at high heating rates, an
intense bubbling is produced due to violent expansion of gases in over very short periods of time, which facilitates fluid
collapse by micro-explosions [100,134,135]. These micro-explosions release small liquid droplets as aerosols into the
gas stream.

Although heating rate is a critical factor to define the conditions of fast pyrolysis, it cannot be measured experimentally;
its estimation is limited only to theoretical models. Often, these parameters are subjective, and depend strongly on the
pyrolysis system used (see table 1.), measurement equipment, model assumptions, and experimental conditions (type
of biomass, carrier gas flow, particle size), making it difficult to compare experimental results obtained for global yields
and kinetic parameters published in scientific papers. There is ambiguity in determining the correct heating rate from
values reported in publications. In most cases, heating rate is reported as being the programmed rate for heating
elements (electrical resistance, radiation lamp, hot plate), rather than a particle heating rate. Particle heating is
commonly assumed to follow a linear heating, mathematically represented as: T=βt+To with β, t and To as the heating
rate, time and initial temperature of the particle respectively [103,136,137]. This simplification can be valid under certain
experimental conditions, such as those used in thermogravimetric analysis, or pyrolysis subjected to kinetic control
conditions (small particle sizes < 100 um). Because of the challenges in identifying the correct biomass heating rate, it
would be necessary to fully recreate the experimental conditions to obtain comparable biomass heating rates for direct
comparison of experimental results.

Table 1. Main devices used in the study of fast pyrolysis processes.


Chapter 1 43

Reactor T °C HR (°C/s) Time (s) P (atm) Diam Remarks

Drop Tube High heating rates. It is not possible to measure particle temperatures. Moderate gas
1500 104 - 105 0.1- 4 Vac - 69 Powder
residence time. Errors of 5% in closing mass balances.
Wire Mesh Uncertainty in particle temperature measurement. Short residence times of volatiles. Low
1500 1000 1 - 3600 Vac - 69 Powder
uncertainty in the mass balances.
TGA Low heating rates and temperatures. High volatile residence time. Good estimation of sample
1200 1 h Vac - 2 < 2mm
temperature. Online monitoring sample weight.
Radiation
2000 104 – 106 ms 1 Powder High heating rates. Inaccurate temperature measurement.
The temperature history of the system is well established. In shock tube reaction, times longer
Shock Tubes
700 - 2200 106 ms Vac Fine than 3.0 ms cannot be achieved. Also, sampling of the reaction zone for gas chromatographic
analysis is time consuming.
Py-MS Direct measurement of volatile and light gases. Indirect measurement of the sample
700 20000 10 - 3600 Vac - 1 Fine
temperature.
Fluid bed 700 - 1200 1000 -10000 1-3600 High uncertainties in the closure of mass balances. Volatile quantifying uncertain. It is not
1 - 100 1 - 5 mm
possible to accurately determine particle temperature and residence time.
Plate reactor Accurate temperature measurement of the particle. Uniformity on solid and temperature
300 - 700 102 – 106 s 1-2 µm - film
distribution. Allow visualization of solid sample
Ablative Volatile cracking reactions. Solids temperature measurement is not possible. Difficult volatile
600 ----- 1 1 cm - m
monitoring.
AUGER. Bio-oil contamination of sand. Difficult estimation of solids temperature. Higher volatile
600 ------ min 1 Chips
residence times.
The temperature ceases to rise when the Curie-point of the metal has been reached; that is
Curie point the exact reproducible temperature at which the ferromagnetic material loses its magnetism.
300 - 900 ----- s 1 µm - mm
However, the choice of different pyrolysis temperatures is limited since they are determined
by the Curie-points of available materials
Microwave Power 500- It favors the formation of hot spots in the sample (uneven heating). It is not possible to make
------ min 1 Powder.
2000W precise sample temperature control. Selective heating of polar groups.

HR: Heating rate


Vac: Vacuum
P: Pressure
44 Chapter 1
Fig 3b, shows that for large particles (> 1mm), temperature gradients exist and heating rates at each point are different,
modifying reactivity patterns at each point within the biomass particle (the reaction rate depends on the heating rate
and reaction temperature). This consideration has been ignored in many studies [74,102,104], where normally the same
reaction rate expression is used to describe the reactivity of the whole particle. Obviously, this can cause problems
when predicting the final product distribution, because each point inside the particle has different kinetic parameters
due to changes in the heating rates and temperatures. For fast pyrolysis studies, there are many techniques and
devices, which can ensure conditions of high heating rates, such as: Py-MBMS (Pyrolysis-Molecular Beam Mass
Spectrometry), TGA (Thermo gravimetric analysis), flash radiation lamps, wire mesh reactors, discharge tubes, and
free fall reactors (Drop tube reactor). Table 1 summarizes the main characteristics of these devices. For kinetic studies,
the most commonly used devices are free fall reactors, Py-MS (Pyrolysis-Mass Spectrometry) reactors, and wire mesh
(hot wire) reactors [69,136,138–140]. Within the academic community, wire mesh reactors have gained acceptance
over other technologies for their versatility (they can be used to study almost any kind of biomass and coal without
many modifications to the reactor), directly measurable temperature (this data is important for kinetic parameters
estimation), short volatile residence time (< 1s) which minimizes side reactions, and the allowance for investigation of
kinetically control regimes.

1.8. Models to describe the reaction rate


Due to the complex physicochemical nature of biomass, there are a wide array of reaction mechanisms and products
associated with thermal decomposition. The search for these reaction mechanisms is a developing science; every day
new results are published, with sophisticated experimental equipment, new analytical techniques, and computational
tools (Monte Carlo simulations, Density Functional Theory [DFT] investigations) to elucidate the controlling steps and
the evolution of species during devolatilization processes. The reaction mechanisms and rate expressions are the heart
of the analysis of pyrolysis processes; these factors have become the primary challenges for researchers in
understanding the controlling phenomena of the devolatilization stage, which is critical to the efficient design and
optimization of reactors [21,141]. Most reaction rate models are focused on the study of global products obtained from
biomass (total volatiles, total light gases, char) and simultaneous multi-stage and/or competitive reactions. These
models can be classified under the following categories: global models in a single step, global models in multiple steps,
semi-global models and models focused on the formation of volatile species.

1.8.1. Global lumped models in a single step

In single step, global lumped models, it is assumed that all the products are directly released from the biomass in a
single reaction stage [142,143], see equation 1:

Biomass → Char + Volatiles (1)

The expressions for the reaction rate are represented as a function of temperature and biomass conversion (see
equations 2 and 3) according to the mathematical expression:

dα (2)
= 𝑘(T)f(α)
d𝑡

Where:
Chapter 1 45
𝑚(t) − 𝑚0 (3)
α=
𝑚∞ − 𝑚0

𝑑𝛼
Where 𝑑𝑡
represents the rate of decomposition of biomass, t represents the reaction time, α represents the biomass
conversion, T represents the temperature of the sample, f(α) represents the expression for the conversion function,
m(t) represents the weight of biomass at a given time t (sub index 0 and ∞, indicate the weight of the biomass at t = 0
s and at equilibrium, respectively).

For non-isothermal systems such as thermogravimetric processes, the mathematical expression of the reaction rate
𝑑𝑇
may be expressed as a function of temperature by taking into account that the heating rate = 𝑑𝑡
, thus the mathematical
expression for reaction rate can be expressed by equation 4:

dα dα dT 1 (4)
= = 𝑘(T)f(α)
dt dT dt β

Gronli [143], OÂrfao [142], Varhegyi [144,145] and Antal [146], among others, have extensively used this reaction
scheme. The authors agree fundamentally that it is possible to achieve good representations of the thermogravimetric
curves using a simple scheme that does not require sophisticated technical data fitting.

Because of its simplicity, this type of model does not allow prediction of the individual species generated in
devolatilization. The reaction rate is monitored by following the change in the weight of the solid residue with time, which
is usually done using a thermobalance, which limits most experimental information found in the literature to slow heating
conditions that are not useful representations of fast pyrolysis processes, as demonstrated by Milosavljevic et al [147].
Another weakness of this reaction scheme is that it does not include the formation of the intermediate liquid phase,
which is important in the processes of pyrolysis, especially at high heating [10,62,103,107,124,141,148,149]. The
presence of this phase alters mass and energy transport mechanisms, and further explains phenomena such as aerosol
formation that are largely responsible for the presence of oligosaccharides in the volatile stream [141].

1.8.2. Multi stage-global lumped models

Many researchers have proposed a variety of complex reaction schemes with multiple stages that monitor nearly all
species involved in the overall reaction (condensable volatile tar, light gases, char, and biomass) [150]. Initially, these
reaction schemes were developed to explain the reactivity of cellulose, and then extrapolated to whole biomass models
[21,150,151]. Table 2 summarizes the main schemes proposed for the decomposition of biomass in multiple stages.
Reaction schemes 1, 2, 4, 5, 6, 7, 9 and 10 model two competing reactions in the first stage. One of these reactions is
associated with dehydration of the cellulose to form anhydro-cellulose and water, which can then, according to the
kinetic schemes (1, 2, 4, 5, and 10) react again to produce char and non-condensable gases. The second reaction
involves the depolymerization of cellulose (breaking of glycoside bonds) to produce anhydrous tar that is rich sugars
(mainly levoglucosan). With these models, it is not possible to explain the effect of the heating rate on the overall
distribution of products. It has been proposed that dehydration reactions are favored at low heating rates and
temperatures below 300°C, increasing the production of char and light gases, [21,147,152] whereas at higher
temperatures (300 - 500°C) and high heating rates depolymerization reactions are promoted that increase the yield of
condensable volatiles.

Table 2. Multistage mechanisms for biomass pyrolysis.


46 Chapter 1

1-Kilzer and Broido [153] 2-Modified Kilzer and Broido [154]

3-Broido-Shafizadeh [155] 4-Diebold [156]

5-Shafizadeh [157] 6-Koufopanos [158]

7-Banyaz [159] 8-Agarwal[160]

9-Bradbury [161] 10-Varhegyi [162]

Models 3, 4, 5 and 10 have the novel inclusion of an intermediate phase (many authors call this phase active cellulose
or intermediate liquid molten biomass [107]) which is formed by random breakage of glucosidic bonds in cellulose and
hemicellulose. The presence of this phase explains the “Lag” in biomass mass loss at temperatures below 320°C when
evaluating the experimental results obtained by thermobalance [163,164]. These schemes have been widely used by
many researchers studying biomass slow and fast pyrolysis. For example, Boutin et al. [62] studied the fast pyrolysis
of cellulose pellets using a radiation heated reactor (Xenon lamp) by varying the heating rate (changing the incident
radiation power). An interesting result of this work was the experimental tracking (measured weight) of oligomers
belonging to the intermediate liquid phase formed during devolatilization. In addition, more conventional products such
as, the total char weight and the total volatiles, which are recovered by condensation of products released to the gas
stream released during devolatilization, were also monitored. The experimental data collected satisfactorily fit the
Broido-Shafizadeh scheme (model 3), with activation energies of around 242 and 198 kJ/mol. These results are similar
to those reported by the same authors [152,165] for biomass and cellulose pyrolysis under heating from solar radiation
and infrared radiation (xenon lamp).

Alexander [74] in his doctoral thesis compared five different reaction schemes (decomposition in one step, models as
8, and two multi component decomposition schemes) to explain the slow and fast pyrolysis of Loblolly pine in a hot
plate reactor (wire mesh reactor). The author found that the kinetic parameters obtained using low and high heating
Chapter 1 47
rates are distinct and not interchangeable, i.e. you cannot use the parameter obtained at low heating rates (5-20 K/
min) to explain behavior at high rates heating (400 °C/s). These results corroborate the results presented by
Milosavljevict [147]. He found that the decomposition scheme in a single step fits well with experimental data only for
temperatures below 320 °C (before the depolymerization of cellulose begins [124]), confirming that this scheme is not
accurate for biomass fast pyrolysis study at high heating rates.

Varhegyi et al. [155] has made strong critiques of the classic Broido-Shafizadeh reaction model (model 3). They
questioned the existence of the intermediate (activated cellulose), citing that the experiments performed by Broido and
Weinstein [163] were not performed under kinetic control, since the size and quantity of the sample used favored a
delay in the decomposition of cellulose due to heat transfer resistance. They claimed that this experimental data could
be well represented by a simple model in a single step. Di Blasi [164] presented the results of a comparative study of
most models presented in Table 1 (models 1, 3, 4, 6, 8) for cellulose and biomass slow pyrolysis at heating rates of
between 20 and 50 K/min at a temperature range between 550 and 1000 K. The primary findings of this study were:

 For cellulose, all schemes, from a qualitative point of view, have the same trend: volatile species and char
increase and non-condensable gases decrease as the reaction conditions become more severe (higher
heating rates).
 Models 1 and 3 predict similar volatile yields, while model 4 under predicts volatile yields.
 For biomass, from the qualitative point of view, there are large differences in data trends, especially in the yield
of non-condensable gases.
 Models 6 and 8 predict that char yield remained constant as the heating rate increases from 40 to 160 K/min.
The authors suggest that at high temperatures, the reaction is thermally controlled and no kinetic control exists.
The primary reactions responsible for the formation of condensable volatile compounds are strongly
endothermic, with part of the energy supplied to the reaction used to evaporate the tar condensate (liquid
phase within the biomass), cooling the particle as the vapors escape.

All kinetic models obtained for both cellulose and biomass, were developed using data collected under different
operating conditions (different particle sizes and temperature ranges) and process equipment (fluidized bed, fixed bed
thermobalance, Pyroprobe). This of course means that there is subjectivity in each model since, in some, kinetic control
could exist and not others, or the experimental conditions are more tightly controlled by some teams, minimizing the
risk of uncontrolled secondary effects.

In summary, the study of the reaction rate expressions within these models are based on monitoring global yields such
as, the total weight of the solid residue char (char + unprocessed biomass + liquid phase), total volatiles (obtained by
condensing the total gas), and light, non-condensable gases (gas chromatography). These models do not predict the
evolution of species released during devolatilization, and many of them contain high degrees of freedom for the estimate
the kinetic parameters (more variables than experimental data), so that there may be difficulties with adjustment and
subjectivity in the analysis of experimental data. Other kinetics schemes can be consulted in [63,166].

1.8.3. Semi-global lumped models

Semi-global models describe the reaction rate of biomass as a linear combination of reaction rates for each of its
components (cellulose, hemicellulose and lignin), see equation 5. For each constituent, models have been adopted that
are similar to the reactions outlined in table 2.

rbiomass = xcellulose ∗ rcellulose + xhemicellulose ∗ rhemicellulose + xlignin ∗ rlignin (5)


48 Chapter 1

Where 𝑟𝑖 and 𝑥𝑖 are the reaction rate and the weight fraction of species i (cellulose, hemicellulose, lignin), respectively.
This type of models has generated substantial controversy. While many authors claim that there is no interaction
between biomass components, so it is valid to model the overall biomass pyrolysis behavior [158,167–169], others
claim the opposite, arguing that it is virtually impossible to achieve the same structure and pattern of actual biomass
crossovers with simple mixtures of components. [17,45,170,171].

Yang et al. [167] investigated the role of cellulose, hemicellulose and lignin during biomass slow pyrolysis at a heating
rate of 10 K/min. They found that synthetic biomass behavior can be explained as the sum of the decomposition of
each component separately and that this behavior is similar to that of raw biomass. Rao et al. [169] have proposed a
kinetic model of order n, to assess the behavior of cellulose, hazel, wood, rice husks and olive husk under slow pyrolysis
conditions (20 K/min), using predefined kinetics parameters for cellulose, hemicellulose and lignin. In this brief paper,
the authors found good correlation between the results predicted by the model and those found experimentally for each
biomass. Olajire et al. [172] adopted a two-stage model (model 7 table 1) to explain the behavior of each biomass
component during pyrolysis, and used these results to explain the behavior of bamboo during devolatilization between
5 - 30 K/min. The results showed that this model represented the experimental data well. The activation energy obtained
varied from 33 to 144 kJ/mol for the formation of volatiles and 60 - 91 kJ/mol for the formation of char for each
component.

Hosoya et al. [45] investigated the interactions of cellulose, hemicellulose and lignin in biomass pyrolysis at 800 °C. To
study these interactions the authors made binary mixtures of components at various proportions (cellulose-lignin,
cellulose-hemicellulose, lignin-hemicellulose) as well as a ternary mixture. In this study the authors, rather than seek a
kinetic scheme to represent the overall behavior of a biomass from its components, evaluated how each component
affects the distribution of species in the mixture of volatile and non-condensable gases formed during pyrolysis. They
found that cellulose-hemicellulose interactions are negligible and do not alter the distribution of product species and
global product yields, but found the interactions between cellulose and lignin to be important. For the lignin-cellulose
mixtures (2:1) a increases in the formation of light gases such as CO, CO2 and decreases the formation of char were
found. The effect of cellulose (cellulose-lignin 2:1) reduces the formation of char, and increases the production of
guaiacol, 4-methyl-guaiacol and 4-vinyl-guaiacol. Despite these revealing results, the authors do not report the
theoretical arguments to explain these findings.

Couhert et al [17], obtained similar results to those reported by Hosoya et al. [45], concluding that it is not possible to
represent the behavior of a biomass by an additive law of its components. The authors suggest that the type of contact
between constituents is very important the in raw biomass, since there are components that are inter-linked by both
inter-and intra-molecular forces that hold them together. Such unions are not present when doing a simple mixing
components, but may affect degradation patterns during thermal devolatilization.

Meanwhile Alexander [74] concluded that a two components (celullose and lignin), two stage model (model 7 table 2)
is not a suitable representation for loblolly pine pyrolysis. As mentioned, there is ambiguity regarding the use of additive
laws to represent the behavior of whole biomass from its components, because of arguments regarding the
arrangement and potential interaction of each component in raw biomass. In most of these models, only global
components (total volatiles, char, total light gases), and used to describe the evolution of species, or the evolution of
the condensed intermediate phase, and are usually are restricted to the used of a thermobalance to measure the
experimental conditions.

1.8.4. Models based in volatile species


Chapter 1 49

These models allow for the species present in the volatile and non-condensable gases to be determined as a function
of the operating parameters (temperature, heating rate). This type of model is limited in the literature [15], withmany of
those that do exist simply extensions of the models described in table 2. Only a few works have been reported that
study the evolution of species from cellulose and glucose derivatives from a micro kinetic view (reaction mechanism)
[173]. Ranzi et al. [15,174], proposed a multicomponent kinetic model that takes into account changes in the primary
gas in the condensable and non-condensable gas fractions, such as CO, CO2, H2, CH4, H2O, HMF, levoglucosan,
phenol, ethanol and methanol. The model was validated using results from various biomass pyrolysis tests conducted
in a fluidized bed reactor. The model, despite being very descriptive, does not accurately predict the behavior of
levoglucosan, HMF, ethanol and other species, especially at elevated temperatures. In addition Ranzi et al. do not
report the experimental conditions under which the model was developed, and claim that it is only a first approximation,
with more work needed to improve reproducibility.

Table 3. Kinetic schemes proposed by Ranzi, et al. [15] to predict condensable and non-condensable gas yields from
cellulose and hemicellulose pyrolysis .

Lumped kinetic scheme for cellulose pyrolysis


reactions rate units
𝐂𝐄𝐋𝐋 → 𝐂𝐄𝐋𝐋𝐀 −𝟒𝟔𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]
𝐊=𝟖∗ 𝟏𝟎𝟏𝟑 𝐞 𝐑𝐓

𝐂𝐄𝐋𝐋𝐀 → 𝟎. 𝟗𝟓𝐇𝐀𝐀 + 𝟎. 𝟐𝟓𝐆𝐥𝐲𝐨𝐱𝐚𝐥 + 𝟎. 𝟐𝟎𝐂𝐇𝟑 𝐂𝐇𝐎 −𝟑𝟎𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]


𝐊 = 𝟏 ∗ 𝟏𝟎𝟗 𝐞 𝐑𝐓
+ 𝟎. 𝟐𝟎𝐂𝟑 𝐇𝟔 𝐎 + 𝟎. 𝟐𝟓𝐇𝐌𝐅𝐔 + 𝟎. 𝟐𝟎𝐂𝐎𝟐
+ 𝟎. 𝟏𝟓𝐂𝐎 + 𝟎. 𝟏𝐂𝐇𝟒 + 𝟎. 𝟗𝐇𝟐 𝐎 + 𝟎. 𝟔𝟐𝐂𝐇𝐀𝐑

𝐂𝐄𝐋𝐋𝐀 → 𝐋𝐕𝐆 −𝟏𝟎𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]


𝐊 = 𝟒𝐓𝐞 𝐑𝐓

𝐂𝐄𝐋𝐋 → 𝟓𝐇𝟐 𝐎 + 𝟔𝐂𝐇𝐀𝐑 −𝟑𝟐𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]


𝐊 = 𝟖 ∗ 𝟏𝟎𝟏𝟑 𝐞 𝐑𝐓

Lumped kinetic scheme for hemicellulose pyrolysis


𝐇𝐂𝐄 → 𝟎. 𝟒𝐇𝐂𝐄𝟏 + 𝟎. 𝟔𝐇𝐂𝐄𝟐 −𝟑𝟏𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]
𝐊 = 𝟏 ∗ 𝟏𝟎𝟏𝟎 𝐞 𝐑𝐓

𝐇𝐂𝐄𝟏 → 𝟐. 𝟓𝐇𝟐 + 𝟎. 𝟏𝟐𝟓𝐇𝟐 𝐎 + 𝐂𝐎 + 𝐂𝐎𝟐 + 𝟎. 𝟓𝐂𝐇𝟐 𝐎 −𝟐𝟕𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]


𝐊 = 𝟑 ∗ 𝟏𝟎𝟗 𝐞 𝐑𝐓
+ 𝟎. 𝟐𝟓𝐂𝐇𝟑 𝐎𝐇 + 𝟎. 𝟏𝟐𝟓𝐂𝟐 𝐇𝟓 𝐎𝐇 + 𝟐𝐂𝐇𝐀𝐑
𝐇𝐂𝐄𝟏 → 𝐗𝐘𝐋𝐎𝐒𝐄 −𝟏𝟏𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]
𝐊 = 𝟑𝐓𝐞 𝐑𝐓

𝐇𝐂𝐄𝟐 → 𝟏. 𝟓𝐇𝟐 + 𝟎. 𝟏𝟐𝟓𝐇𝟐 𝐎 + 𝟎. 𝟐𝐂𝐎𝟐 + 𝟎. 𝟕𝐂𝐇𝟐 𝐎 −𝟑𝟑𝟎𝟎𝟎⁄ [𝐬 −𝟏 ]


𝐊 = 𝟏 ∗ 𝟏𝟎𝟏𝟎 𝐞 𝐑𝐓
+ 𝟎. 𝟐𝟓𝐂𝐇𝟑 𝐎𝐇 + 𝟎. 𝟏𝟐𝟓𝐂𝟐 𝐇𝟓 𝐎𝐇 + 𝟎. 𝟖𝐆{𝐂𝐎𝟐 }
+ 𝟐𝐂𝐇𝐀𝐑
Lumped kinetic scheme for hemicellulose pyrolysis
𝐋𝐈𝐆 − 𝐂 → 𝟎. 𝟑𝟓𝐋𝐈𝐆𝐂𝐂 + 𝟎. 𝟏𝐩𝐂𝐎𝐔𝐑𝐌𝐀𝐑𝐘𝐋 + 𝟎. 𝟎𝟖𝐏𝐇𝐄𝐍𝐎𝐋 𝐊 = 𝟒 ∗ 𝟏𝟎𝟏𝟓 𝐞
−𝟒𝟖𝟓𝟎𝟎⁄
𝐑𝐓 [𝐬 −𝟏 ]
+ 𝟏. 𝟒𝟗𝐇𝟐 + 𝐇𝟐 𝐎 + 𝟏. 𝟑𝟐𝐆{𝐂𝐎𝐇𝟐 }
+ 𝟕. 𝟎𝟓𝐂𝐇𝐀𝐑
𝐋𝐈𝐆 − 𝐇 → 𝐋𝐈𝐆𝐎𝐇 + 𝐂𝟑 𝐇𝟔 𝐎 𝐊 = 𝟐 ∗ 𝟏𝟎𝟏𝟑 𝐞
−𝟑𝟕𝟓𝟎𝟎⁄
𝐑𝐓 [𝐬 −𝟏 ]
𝐋𝐈𝐆 − 𝐎 → 𝐋𝐈𝐆𝐎𝐇 + 𝐂𝐎𝟐 𝐊 = 𝟏 ∗ 𝟏𝟎𝟏𝟎 𝐞
−𝟐𝟐𝟓𝟎𝟎⁄
𝐑𝐓 [𝐬 −𝟏 ]
𝐋𝐈𝐆𝐂𝐂 → 𝟎. 𝟑𝐩𝐂𝐎𝐔𝐌𝐀𝐑𝐘𝐋 + 𝟎. 𝟐𝐏𝐇𝐄𝐍𝐎𝐋 + 𝟎. 𝟑𝟓𝐂𝟑 𝐇𝟒 𝐎𝟐 𝐊 = 𝟓 ∗ 𝟏𝟎𝟔 𝐞
−𝟑𝟏𝟓𝟎𝟎⁄
𝐑𝐓 [𝐬 −𝟏 ]
+ 𝟏. 𝟐𝐇𝟐 + 𝟎. 𝟕𝐇𝟐 𝐎 + 𝟎. 𝟐𝟓𝐂𝐇𝟒 + 𝟏. 𝟑𝐆{𝐂𝐎𝐇𝟐 }
+ 𝟎. 𝟓𝐆{𝐂𝐎} + 𝟕. 𝟓𝐂𝐇𝐀𝐑
𝐋𝐈𝐆 → 𝐂𝟏𝟏 𝐇𝟏𝟐 𝐎𝟒 𝐊 = 𝟏 ∗ 𝟏𝟎𝟏𝟑 𝐞
−𝟒𝟗𝟓𝟎𝟎⁄
𝐑𝐓 [𝐬 −𝟏 ]
50 Chapter 1
𝐋𝐈𝐆 → 𝟎. 𝟕𝐇𝟐 + 𝐇𝟐 𝐎 + 𝟎. 𝟐𝐂𝐇𝟐 𝐎 + 𝟎. 𝟓𝐂𝐎 + 𝟎. 𝟐𝐂𝐇𝟐 𝐎 𝐊 = 𝟏 ∗ 𝟏𝟎𝟓 𝐞
−𝟐𝟎𝟓𝟎𝟎⁄
𝐑𝐓 [𝐬 −𝟏 ]
+ 𝟎. 𝟒𝐂𝐇𝟑 𝐎𝐇 + 𝟎. 𝟐𝐂𝐇𝟑 𝐂𝐇𝐎 + 𝟎. 𝟐𝐂𝟑 𝐇𝟔 𝐎𝟐
+ 𝟒𝐂𝐇𝟒 + 𝟎. 𝟓𝐂𝟐 𝐇𝟒 + 𝐆{𝐂𝐎} + 𝟎. 𝟓𝐆{𝐂𝐎𝐇𝟐 }
+ 𝟔𝐂𝐇𝐀𝐑

Liao et al. [175] studied cellulose fast pyrolysis, using the experimental data to a create a modified Shafizadeh Broido
scheme. This model takes into account the evolution of: levoglucosan, acetol, 5-HMF, HAA, CO, CO2, H2, H2O, and
furfural. Although the model includes the formation of an intermediate phase (activated cellulose), this phase is not
directly monitored, but is instead determined by fitting experimental data. Other similar schemes may be found in
[176,177]. Vinu et al. [173] for first time developed a micro kinetic model for glucose-based carbohydrates pyrolysis.
The scheme is fairly rigorous, involving 41 elementary reactions, including: depolymerization, cracking, drying, aldol
condensation, and tautomerization reactions among others. This model describes the step by step the evolution of the
primary volatile condensable and non-condensable gases, as well as the evolution of char and intermediate
anhydrosugars. Despite being a rigorous scheme, successful implementation is doubtful because the kinetic
parameters of each elementary reaction come from different sources (some are theoretically estimated by computer
simulation, other measured experimentally, and others are taken from the literature, under varying experimental
conditions). Other theoretical models have also been developed to explain the evolution of species during biomass
devolatilization, however; because they are commercial packages, these models remain black boxes, with details as to
the theoretical development lacking within the literature. These models are: Prime Kinetics Model (Reaction Network)
[178], Bio-Flashing [179], Bio-FG-VDC [180] (Functional Group-Depolymerization-volatilization-Crosslinking).

1.8.5. Distribution activation energy model (DAEM)

Another kind of kinetic model that has been developed for coal and is being widely used to explain both coal and
biomass pyrolysis is the distribution of activation energies model (DAEM) [181–188]. These are multi-stage and
multicomponent models, which involve the formation of a species from a very large number of independent parallel
reactions that reflect variations in bond cleavage and reactant structures. These models have been used to explain
data obtained under slow pyrolysis using a thermobalance [181,186,187,189–193], but so far few applications have
been reported under fast pyrolysis conditions for biomass.

Usually, in the study of the effect of heating rate on the overall yield of volatiles, char and non-condensable gases, there
are very few models that extend to the study of individual species of both condensable and non-condensable gases
[194]. To achieve this level of detail in the analysis of product species, integral kinetic modes, such as the one given in
equation 6 are employed. The main advantage of this model is that it covers a wide experimental conditions, including
temperatures and heating rates, and is able to represent slow and fast heating kinetics in one kinetics scheme. The
activation energy distribution can be represented by continuous functions as normal, lognormal and Weibull
distributions.

∞ t
V∗ − V 1 −E (E − E0 )2
= ∫ exp [−A ∫ exp ( ) dt] [ ] dE (6)
V∗ σ√2π RT(t) 2σ2
−∞ 0

Here V* and V represent the maximum amount of volatile or individual species in the gas or vapor phase at a given
time t, A is the rate constant, E is the activation energy, σ is the variance and E0 is the mean of the statistical distribution
of activation energies. The kinetic parameters can be obtained by numerical fitting using numerical methods to solve
Chapter 1 51
the exponential integral [187,191,195] or with differential methods as proposed by Miura, et al. [196,197]. In peak
temperature studies, the kinetics parameters must be fitted, because there are not enough degrees of freedom to use
differential methods. Despite this difficulty, the literature widely recognizes that these models are the most rigorous and
accurate methods available to correlate the experimental data from biomass pyrolysis. These models avoid
underestimating the activation energy compared to single stage reaction schemes, but the most striking feature of this
scheme is that a mathematical model is derived that is applicable at various heating rates, unlike either single or multi
stage that only apply to a single heating rate [198]. The difficulty with this technique is that the right side integral has no
analytic solution, so theoretical approaches must be employed, or tedious numerical methods must be used to obtain
a solution.

1.9. Secondary reactions


Depending on the reaction conditions (particle size, temperature, mineral material, heating rate), the residence time of
condensable volatiles inside the particle can be high, promoting their reaction with char [124,127,146].

The reactions of primary products either on the inside or the outside of the particle that produce new compounds are
called secondary reactions. Several types of secondary reactions have been described, including: reactions of
condensable volatiles at the gas/liquid interface [43], reactions of condensable volatile in the liquid phase [38,132,199]
(polymerization, crosslinking, dehydration) and reactions of gases and condensable volatiles with the solid structure
(char) catalyzed by mineral matter [200,201]. From this set of reactions the most studied reactions are decomposition
reactions that occur in the gas phase (cracking) and solid-gas phase reactions (reactions gasification, combustion, and
oxidation) [96,202–204]. Cracking reactions are important at high temperatures (T > 500°C) and are more prominent
when long residence time in the hot zones exist [91,203,205]. Numerous studies related to solid-gas secondary
reactions have also been reported [204,206–209]. These reactions have been studied at an experimental level, and
have been measured by increases in char production, and decreases in volatiles caused by an increase in the time of
solid-gas contact.

Several studies conducted at both the experimental and theoretical level, have reported the reactions responsible for
the thermal decomposition of volatiles (cracking). Norinaga et al. [203] reports a rigorous theoretical and experimental
study volatile cracking reactions released from biomass fast pyrolysis. The authors utilized two reactors connected in
series, the first is for biomass pyrolysis and the second examines the cracking of the volatiles generated in the first
stage. Volatile composition in the first and second reaction stage were monitored by gas chromatography. These results
were compared with data obtained from the theoretical model simulated using the DETCHEM modeling package for
volatile cracking. The model predicted with good accuracy the behavior of light gases (CO, CO2, H2, CH4, C2H4), but
provided less accurate predictions for furan compounds as well as benzene, toluene, acetic acid, aldehyde, methanol.
Shin et al. [210] proposed a 3 stage kinetic model for the decomposition of volatiles produced from cellulose,
levoglucosan and 5-HMF pyrolysis. The compounds were monitored online by the MBMS technique (Molecular Beam
Mass Spectrometry). Good fits were obtained from experimental data. They found that in 0.6 seconds and 750 °C 80
% of the volatiles generated in during pyrolysis decompose. Further experimental studies at fast pyrolysis conditions
were carried out by Morf et al. [91], Graham et al. [211] and Zhang et al. [212]. Each evaluated char, total volatiles,
condensable species and light gases yields as a function of temperature and gas residence time in two stage reactors.
In these studies, it was found that the maximum decomposition rate was achieved at a temperature range of between
600 - 700°C. The poly-hydroxy-aromatic (PHA) formation is another interesting secondary reaction that results primary
from the volatile derivatives of lignin (aromatic). Investigations have shown that these reactions are favored at very high
temperatures (> 800°C) and long residence times (> 10s) [212,213], most typical of operational conditions for
gasification and combustion processes.
52 Chapter 1

Another set of solid-gas and gas phase secondary reactions are related to char gasification with steam or CO2. These
reactions include: char oxidation, the Boudouard reaction, reforming reactions, water gas-shift reactions and
methanation, which have been widely explored, in biomass and coal gasification and combustion processes
[77,214,215].

The liquid-gas and liquid-liquid phase reactions, are largely unexplored with no theoretical models to report., This is
primarily due to uncertainty regarding the components present in the liquid phase and the difficulties associated with
monitoring these species during pyrolysis. Understanding of the liquid phase reaction is one of the main challenges to
be addressed in achieving a better understanding of the reactivity of biomass in the ast pyrolysis process [141]. Despite
this lack of knowledge, it is know that reactions in the liquid phase are the primary contributors to the production of
secondary char, light oxygenated compounds and permanent gases [124,199].

Recent studies focused on levoglucosan reactivity as a surrogate for the intermediate liquid phase formed during
cellulose pyrolysis [216–218]. Paulsen [216], studied the reactivity of mixtures of levoglucosan/fructose to simulate the
environment and interactions inside the intermediate liquid phase formed during cellulose fast pyrolysis. Cyclization
and elimination reactions were found to be responsible for the production of light oxygenated compounds and char.
Both reactions are more intense in the pyrolysis of powders than for thin films. Furthermore, the author suggested that
catalysts, such as palladium, promote decarbonylation reactions in the liquid phase, reducing aldehydic furans and
improving the bio-oil quality (less oxygen content in bio-oil). Bai, et al [217], suggests a competitive mechanism between
levoglucosan evaporation and polymerization during cellulose pyrolysis in TGA. The intensity of both reactions depends
on the quantity ofsample, the carrier gas flow rate, and the heating rate. They show that low sample quantities, and
high carrier gas flow promote evaporation (lower mass transfer resistance). Similar mechanisms were proposed by
Hosoya, et al [219] for levoglucosan pyrolysis in an ampoule reactor. Other important reactions in the liquid phase are
related to the decomposition of lignin oligomers (Pyrolytic lignin) to produce monomers, char and permanent gases.
However, there are few studies focused on the pyrolysis of lignin oligomers, and none relating experimental results
using kinetic parameters. It has been suggested that these compounds are generated by thermal ejection caused by
intense bubbling of the liquid intermediates [61]. These aerosols can be trapped inside the solid matrix after ejection,
with the collision probability increase with particle size [65,220].

1.10. Existence of intermediate liquid phase (Chronology)


The possible existence of an intermediate liquid phase during biomass pyrolysis has sparked much debate for several
decades. In the mid - 60s and early 70s, researchers took the first steps toward a description of this phase. Goring et
al. [221] observed that for cellulose, lignin and hemicellulose a softening occurred when samples were immersed in a
hot oil bath. The authors attributed this behavior to the formation of an intermediate liquid phase (thermal degradation),
rather than a phase change associated with the glass transition point. The softening point was estimated as the
temperature at which the solid structure collapsed (227- 253 °C for cellulose), which is higher than the glass transition
point of the cellulose (145 - 175 °C).

In 1974, Nordin et al. [222] carried out an interesting experiment in which visual evidence of a condensed phase during
flash pyrolysis was presented (heating laser in less than 0.1 ms) for various types of leaf paper. Once samples reach
the pyrolysis temperature, volatiles were collected, rapidly cooled and analyzed. Based on these results the authors
argued that the structure of cellulose was clearly broken and the crystallinity index had decreased by 35%. These
results were attributed to a physical phenomenon (fusion of cellulose) instead of thermal degradation. The above studies
argue that cellulose and biomass, when they are subjected to high heating rates, pass through a condensed phase,
Chapter 1 53
although it could not be determined if this behavior was due to thermal degradation (chemical process) or a merger
(physical phenomenon). In both studies, the melt phase was not analyzed by chromatography methods. Over the course
of this decade, the first theoretical models to explain cellulose and biomass pyrolysis were developed. These schemes
include an intermediate liquid phase (activated cellulose).

Kiler and Broido [154] rigorously analyzed the data reported in the literature related to cellulose pyrolysis obtained using
a thermobalance. They proposed a two-step competitive model (model 1 table 2). In this model he describes the
transformation of cellulose into an unstable intermediate compound, which in most cases is associated with
levoglucosan (the principal active product obtained from cellulose pyrolysis).

Between 1971 -1975, Broido and Nelson [163], Shafizadeh [157] and Bardbury [161] proposed similar reaction
schemes. Patai et al. [223] and Golova [224] studied changes in degree of polymerization of cellulose at a temperature
range of 170 - 230 °C, where the mass loss only reaches approximately 5%. These researchers associated the
reduction of the degree of polymerization with random breaking of cellulose glycosidic bonds, giving rise to anhydrous
glucose derived sugars (levoglucosan, cellobiosan, cellotriosan), which are the main precursors of the liquid
intermediate phase.

In October 1980, during the Specialists' Workshop on Biomass Fast Pyrolysis held at Copper Mountain (USA), Diebold
[225] carried out, in front of all attendees, an experiment in which he demonstrated a hot thread (nichrome wire) passing
through a piece of biomass, as if it were butter, without a trace of char. A detailed analysis of the biomass edges after
cutting showed virgin biomass structures covered by a beige layer, like a varnish. The explanation for this observation
was the formation of a molten phase, which is liquid at pyrolysis temperatures and solid at room temperature.

Excited by Diebold’s demonstration, in 1985 Ledé et al. [226] presented the results of a biomass pyrolysis experiment
in a spinning disk reactor. In this experiment large cylindrical wood pieces were pressed at high pressure (0.1 - 3.5
MPa) against a hot disk (500 - 900 °C) that rotates at constant velocity, thereby degrading biomass. They observed
that, during the reaction on the heated disc, a liquid film formed in the wake of the biomass, like a lubricant, direct
evidence of the formation of an intermediate liquid phase that faded away with time. This was further evidence to
corroborate the results presented by Diebold [225].

In 1989, Pouwels et al. [227] carried out a number of interesting experiments for the study of the fast pyrolysis in-situ
and ex-situ of cellulose, using a mass spectrometer DCI-MS (Desorption Chemical Ionization Mass Spectrometry), DC
& MS (Desorption Ionization Mass Spectrometry) and curie-point-DCI-MS. The authors identified anhydrous sugars
containing up to 6 degrees of polymerization (celohexosas) at high concentrations, thereby indicating that the first step
of the devolatilization process is depolymerization of lower molecular weight compounds with high melting points (> 250
°C ) and boiling points around 339 - 550 °C (levoglucosan and celobiosan) to form an intermediate liquid phase.

In 1998 Boutin et al. [228] studied cellulose fast pyrolysis in a reactor heated by radiation (Xenon lamp - 5 kW, 0.2 s
exposure). With observations made using a microscope, they again confirmed the formation of an off-yellow film, on
the cellulose surface. This film was soluble in water and polar solvents, indicating that this was a new compound and
not molten cellulose, which has a hydrophobic behavior. Piskorz et al. [229] states that, considering the condensed
phase as a simple fusion of cellulose is questionable, as the product is soluble in water, unlike cellulose, and at the
reaction temperatures used (466 °C), it is very difficult to reach the glass transition point without undergoing any thermal
degradation. Piskorz et al. [230] presented experimental evidence, showing that it is possible to depolymerize cellulose
to produce oligosaccharides with varying degrees of polymerization. They proved by SEM (Scanning Electron
Microscopy) analysis, that the cellulose structure was deformed, and showed behavior similar to a swollen polymer.
54 Chapter 1
Qualitative analysis by HPLC of the methanol soluble solid (char) fraction showed a variety of anhydrosugars with
degrees of polymerization of 1-7, and a heavier fraction which could not be identified. These results again corroborate
that cellulose passes through an intermediate liquid-state product during depolymerization.

In 2003 Baliga et al. [231] studied the char obtained from pyrolysis of tobacco leaves and its components (lignin,
cellulose, hemicellulose). Using HTEM characterization, they demonstrated that snuff char, as well as its components,
pass through a melt phase at pyrolysis temperatures between 450 - 550 °C (see Figure 5). This phase is identified by
the formation of small vesicles. The authors did not assess the chemical composition of melt phase. The structural
behavior of the cellulose and hemicellulose presented in these images is similar to that found for the tobacco leaves,
and shown by Dufour et al [232].

Figure 5. Structural changes of (A) Avicel Cellulose, (B) Alkali Lignin, (C) Xylan, and (D) Tobacco during the pyrolysis
process.Taken from [231]

In 2008, Liu et al. [233] carried out experiments in an oven heated by radiation (under fast pyrolysis conditions) to
estimate the evolution of the condensed phase formed during cellulose pyrolysis. The authors confirmed results that
had been obtained in previous studies [62,228,234]. They demonstrated that up to 68 % (w/w) of the initial cellulose
formed a water soluble compound of yellow color, consisting primarily of oligomers of glucose. They proved that under
lower severity heating conditions lower yields of the condensed phase were obtained, with maximum yields dropping
to 57 % (w/w).

An article presented by Mamleev et al. [124] in 2009, investigated a new kinetic model called a two-phase model. This
model incorporates the formation and growth of liquid cavities within the cellulose that contain high boiling point
compounds within the liquid phase (cellobiosan, cellotriosan). The condensed phase is an ionic medium, and promotes
the decomposition of the reduced cellulose chains (after depolymerization) by two competing reactions: β-elimination
and transglycolyzation. The first promotes char formation, while the second promotes the production of anhydrous
sugars and volatiles. They suggest that the quick removal of liquid compounds by evaporation at high heating rates
minimizes the effects of dehydration and crosslinking which are more prevelant at lower heating rates.

Dauenhauer et al. [149], presented convincing visual evidence (see Figure 6) on the formation of the condensed phase.
For the first time, a photographic record was made using a high-speed camera (1000 frames/second) to monitor the
evolution of cellulose during fast pyrolysis. For this study, 300 μm particles were pyrolyzed using a hot plate reactor at
700 °C. The melt phase begins to form at 50 ms (see figure 6), and it is completely formed after 120 ms. Inside the
melted phase, intense bubbling and evaporation promote aerosol ejection and shrining of the liquid phase .This phase
is completely consumed after 150 ms, without a trace of char formation.
Chapter 1 55

Figure 6. Photographic record of the evolution of the cellulose during fast pyrolysis. Taken from Dauenhauer et al.
[116].

Inside the intermediate liquid phase, gas bubbles are formed by chemical reactions and evaporation. The intensity of
bubbling depends of the liquid temperature, heating rate, pressure and nature of the liquid phase [235–237]. The
bubbles bursting promotes aerosol ejection by a reactive boiling ejection mechanism [133]. These aerosols can be
trapped inside the cell walls. Evidence of this was proved by Haas et al. [34] through in situ visualization inside the cell
walls which showed the presences of oligomeric liquids (see figure 7). In this study biomass pyrolysis was monitored in
real time using a microscope (Light microscopy) and TEM (Transmission Electron Microscopy). The pyrolysis
experiments were carried out using a 10x10 mm hot plate (INSTEC), at a heating rate of 150 °C/min and 700 °C.
Observations showed swelling and tissue collapse in wood fiber(Populus canadensis), and the appearance of droplets
of liquid encapsulated in the biomass pores, due to impingement of aerosols with the cell walls (see Figure 7). The
inclusion of the liquid phase within the cell walls has critical implications for attempts to define mass and energy transfer
during pyrolysis.

Figure 7. Images of the evolution of the biomass during the slow pyrolysis process. Figure taken from Thomas et al
[34]. : (A) 260 °C, (B) 299 °C, and (C) 501 °C. Red outlines of xylem cells track individual areas through each frame.
(D) Longitudinally sectioned control poplar showing cell wall faces of ray parenchyma, with white arrows pointing to
examples of cell wall pores (plasmodesma). (E) Longitudinally sectioned hot-stage poplar shows cell wall faces of ray
parenchyma with collected pyrolysis products (black arrows).

In 2011, Teixeira et al. [133] built upon the work of Dauenhauer et al. [149] using a high speed camera (1000 fps) to
monitor the pyrolysis of cellulose and sucrose. They visualized the mechanism of aerosol formation by bubbles bursting
in the intermediate liquid phase and devised a new mechanism called reactive boiling ejection to explain the results. It
56 Chapter 1
was not possible to track bubbles inside the cellulose particle due to high light absorption, so sucrose was used as
surrogate. Visualization of the ejection of droplets from the condensed phase of the sucrose during pyrolysis is
presented in Figure 8, along with cartoon represtations to more clearly ilustrate the process.

Figure 8. Ejection mechanism of liquid droplets dfrom the intermediate liquid phase originating from the fast pyrolysis
of sucrose (top) with cartoon representations of the observed behavior (bottom). Figure taken from Teixeira et al [133]

The ratio between bubble size before bursting and aerosol size ejected is close to 0.1. The ejection velocity of the liquid
droplets (0.1 m/s) and the mechanism of nucleation and evolution of these drops in the intermediate liquid phase were
experimentally estimated. The aerosol collected had an average size of 1 µm. The composition of the aerosol were
primarily carbohydrate dimers (Cellobiosan). With this information a mathematical model (VOF) was developed to
predict the mechanism for the formation of aerosols when a bubble bursts.

In 2012, three interesting experimental works were reported, which used in-situ viewing to monitor biomass tissues
during the pyrolysis process. The first, reported by Henrik et al. [238], examined the pyrolysis of two types of biomass,
American and European hardwood with a particle size of 300 µm. For this study a heating laser (Nd: YAG) of 2 kW was
used. Using a high-speed camera, they captured the behavior and evolution of the condensed phase (see Figure 9A)
as the sample was heated (fast pyrolysis conditions over a heating time of 0.1 s). They found that during biomass
pyrolysis, the tissues swelled, and formed a brown film that covered the biomass surface. This film was water-soluble
and had different levels of transparency as the pyrolysis process progressed.
Chapter 1 57
A. Hardwood pyrolysis. B. Soybean wastes

Figure 9. (A).Evolution of the intermediate liquid phase during the fast pyrolysis of hardwood. Figure taken from:
Carlsson et al [238]. (B). Evolution with temperature of a liquid phase during the slow pyrolysis of waste soybeans.
Figure taken from Zhu et al. [239].

The second paper was presented by Zhu et al. [239] who pyrolyzed waste soybeans (bean dregs) using a quartz
capillary (1.5 cm x 650 µm). Finely ground biomass was introduced into the capillary and was pyrolyzed using a hot-
plate reactor (INSTEC) at 30 °C/min at 400 °C. The process was monitored with a CCD (charge coupled device)
camera. This technique allows in-situ monitoring of biomass pyrolysis during heating (see Figure 9B).There was a
change in color as the reaction progressed, resulting in the formation, at 240 °C, of a liquid with an oily appearance.
This liquid became darker as the temperature increased (decomposition reactions and polymerization).

In 2013, Dufour et al. [171] pyrolyzed biomass (miscanthus) and its components (cellulose, lignin, hemicellulose) in a
parallel hot plate rheometer specifically adapted for in-situ tracking of the viscoelastic behavior of the intermediate liquid
phase formed during pyrolysis of biomass and its components. Experiments showed that melted lignin has an elastic
behavior at 200 °C, hemicellulose at 270 °C, cellulose at 300 °C, and biomass at 280 °C. The elastic behavior of the
biomass melt phase could not be estimated as a combination of each of its components. A valuable observation was
that viscous behavior was identified after the onset of elastic behavior, i.e. structural changes occur first (glass transition
point) and then bond breaking occurs to form a viscous fluid phase. The researchers also found that cellulose imposes
its viscoelastic behavior on biomass during thermal degradation. This viscoelastic behavior was found to be variable;
after 400 °C all materials increased their elastic and viscous moduli. This indicates that at higher temperatures a rigid
phase formed, which, in the case of pyrolysis, corresponds to the char formation. These results were validated and
compared with 1H-NMR.

In 2013 there were two studies, where the effect of cellulose crystallinity on the yield of the intermediate liquid phase
were evaluated experimentally, Wang et al. [240] and Liu et al.[102]. They found that amorphous cellulose produces a
higher content of the intermediate liquid phase, and that this formation begins at lower temperatures compared to
crystalline cellulose. Amorphous cellulose produces a variety of anhydrous sugars with high degrees of polymerization,
while the crystalline cellulose is rich in monomers and dimers of glucose. One explanation given by the authors was
that there is a protective effect of the crystal lattices (a collective effect that gives greater rigidity to the structure) that
prevents the breaking of glucose rings. Recently (2016), Westerhof, et al. [241], showed that depolymerization of
cellulose produces sugar oligomers in the liquid phase that are composed of anhydrosugars with a degree of
58 Chapter 1
polymerization > 2. At high heating rates and under vacuum (2 mbar) these compounds escape by evaporation and
aerosol ejection without producing char or permanent gases, indicating that for cellulose pyrolysis these fractions are
produced by secondary reactions. The yield of levoglucosan increases when operating at 1 bar, because the boiling
point of heavy sugars increases, reducing the evaporation rate of these compounds, allowing depolimerization reactions
to progress until forming levoglucosan, which evaporates.

Unfortunately, there are few papers regarding the intermediate liquid phase formed during lignin and hemicellulose
pyrolysis. Zhou et al. [132], showed the importance of heating rate on the intermediate liquid phase formation during
lignin pyrolysis. Using a fast speed camera (250 fps) the formation of intermediate liquid phase was visualized. In this
process 3 stages were recognized: Complete lignin melting, foaming and swelling of liquid phase by bubbles bursting
inside, and shrinking by evaporation and aerosol ejection.

In summary, it has been demonstrated that the intermediate liquid phase during biomass pyrolysis does exist, especially
under high heating rate conditions, however; this phase has not been adequately monitored experimentally. Particle-
level mathematical models for both the bubble formation and thermal aerosol ejection mechanisms have not yet been
completely explored.

1.11. Modeling intra-particle fast pyrolysis process


Since 1940, there have been significant developments in the modeling of biomass particles under pyrolysis conditions.
The complexity, and coupling of the phenomena involved (transport of mass, energy, momentum, reactivity), has led to
constant research and development of these models to achieve better representations of real processes. In general,
modeling can be as rigorous as desired, however; this does not imply that better representations of real phenomena
will result. Modeling can also be associated with different scales of study, for example, molecular, tissue, particle, or
reactor scale models can be designed [242].

The focus of this review are the particle-level models, which is one of the primary focuses of this doctoral thesis, however
some models that have developed for biomass fast pyrolysis at the reactor scale can be consulted [104,243–245].

Most articles published to date have considered one-dimensional non-steady state models, in one or two phases (solid-
gas) with thermal equilibrium between phases (solid/gas), and global reaction schemes as described in Section 3.2.2.
Some researchers, suggest that for small particles (< 1mm) the global kinetics scheme is sufficient to predict the product
distribution [22]. However for large particles, both physical and chemical changes occur simultaneously, and it is
essential to describe both phenomena to accurately obtain the global pyrolysis rate. A summary of models reported in
the literature is presented in table 4.

Table 4. Particle level models


Non dimensional Models
Author Characteristics Shortcomings
Explains the pyrolysis of cellulose. Developed to
This model does not include mass or energy
predict the behavior of cellulose under thermo
Diebold [156] transfer restrictions, evaporation of water or
gravimetric operating conditions such as
gas flow.
temperature, heating rate and pressure.
Evaluates the temperature and evolution of mass Employs a kinetic single step scheme, the
Moghtaderi et al. [246]
loss from biomass as a function of time. It models mass lost from the solid that generates
Chapter 1 59
only the solid phase, considering two areas of volatiles is modeled as a sublimation of the
interest: virgin biomass and char. virgin biomass,. The model only takes into
account the char-virgin biomass interface.
One-Dimensional Models
Cannot be used to predict product yields or
Predicts temperature and weight loss profiles. composition. No sensitivity analyses were
Bamford et al. [247]
Predicts the evolution rate of volatiles. carried out. No convection term. Assumes
constant physical properties.
Shows that the convection term is needed for the
Can only be used to investigate the
Kanury et al. [248] energy balance equation. Suggests that the
importance of convective heat transfer.
burning rate depends on particle size.
Predicts temperature and density profiles, as Cannot be used to predict product yields or
well as the volatile production rate. Suggests that composition of individual species. No solid
Kung [249] the thermal conductivity and the specific heat of shrinking is considered. Firsts order
char, heat of reaction and particle size are all reaction kinetics are used in 1 step.
important parameters in Wood pyrolysis. Developed for thick particles (1cm).
Predicts under what conditions heat transfer or
reaction kinetics are rate controlling. Predicts the
Cannot be used to predict product yields or
reaction rate and temperature profiles, as well as
composition. Does not predict volatile
reaction times. Utilizes simple derived
Maa and Baillie [250] evolution rate. Utilizes constant properties,
expressions to estimate reaction time by using
pressure, and single step kinetics. Balances
either effective thermal conductivity (large
are only performed for the solid phase.
particles) or effective activation energy (small
particles).
Introduces a new parameter, the Lewis Number,
defined as the ratio of thermal diffusivity and
Cannot be used to predict product yields or
mass diffusivity. Predicts the conversion and
chemical composition. Does not predict
temperature profiles. Suggests that the higher
Fan et al. [251] volatiles evolution rate, or consider solid
the Lewis Number the greater the conversion to
shrinking. Reaction rate is based on single
volatiles and the smaller the temperature
step reaction.
gradient within particle. Suggests that the heat of
reaction affects the rate of pyrolysis.
Predicts mass loss, pressure and temperature
profiles. Suggests that the heat of reaction, the Cannot be used to predict product yields or
thermal conductivity and specific heat of char chemical composition. Does not predict
Kansa et al. [252] and the permeability are constant and important volatiles evolution rate, or consider solid
parameters in wood pyrolysis. Includes porous shrinking. Utilizes constant physical
and permeable structural effects on the volatile properties.
flow.
Predicts the mass loss, as well as surface and
internal temperature profiles. Pyrolysis of white
Does not predict volatiles yields. Assumes
pine wood was studied using thin samples (1.6
constant physical properties. For large
Phillips et al. [253] mm thick slabs) exposed in an oven at 199-
particles, kinetics and heat transfer are not
365"C and thick samples (8-50 mm diam branch
coupled. Solid shrinkage is not considered.
segments) exposed in a wind tunnel at 365-
525"C with 3-18 m/s wind speed.
60 Chapter 1
Introduced two pyrolysis numbers, Py =
(K/kϱcpR2) or Py′= (h/kϱcpR). Evaluated the
Cannot be used to predict product yields.
importance of external and internal heat transfer.
Does not predicts volatile evolution rate. No
Derived fours models for Bi>1, Bi<1, Py>1, Py<1.
Pyle and Zaror [254]. convection term. Valid only for cylinders
Predicts the conversion and temperature
with diameters in the range of 0.03 to 2.2cm
profiles. Predicts conversion times. Particle size
and pyrolysis temperatures of 380 to 500°C.
affects conversion. A sensitivity study was also
performed.
Models heat transfer to the particle by radiation,
convection and conduction. Models diffusion of Cannot be used to predict product yields or
gaseous reactants through the unconverted chemical composition. Uses single step
Antal [255]
virgin solid. Evaluates chemical reaction of the kinetics., and does not consider particle
convertible solid, to the primary pyrolysis shrinkage.
products of.
Predicts product yields and composition of
permanent gases. Also predicts the volatiles
release rate and temperature profile. Predicts
Chan et al. [256] effects of moisture. Sensitivity studies were Limited to cylindrical pellets.
performed. Realistic values of variable physical
properties and reaction rate parameters are
utilized.
The model is able to predict temperature profiles
and weight loss in wet wood particles of variable Neglects water vapor diffusion, bound
geometry subjected to high temperatures in an water. The solid volume is assumed to be
inert atmosphere.Experimentally measured constant throughout the drying and
critical parameters, including thermal pyrolysis processes. Solid permeability is
Alves et al. [257]
conductivities and all kinetic parameters; other assumed to be sufficiently high for the
parameters were taken as average values from pressure gradients inside the particle to be
the literature. The drying model neglects bound- negligible. Valid only for thick particles.
water diffusion, air/vapor diffusion and pressure
gradients inside the solid.
A detailed single-particle model, which includes
a description of transport phenomena and a
global reaction mechanism. The model is
coupled with a plug-flow assumption for extra the
particle processes of tar cracking, in order to
predict the fast pyrolysis of wood in fluidized bed
reactors for liquid-fuel production. The average Does not take into account the effect of the
Di Blasi [56] heating rates are on the order of 45 - 455 K/s, formation of an intermediate liquid phase.
with reaction temperatures of between 640K and Thermal properties are constant.
770 K, particle sizes of 0.1- 6.0 mm and a reactor
temperature of 800 K. The pressure gradients
are estimated using Darcy's law (approximation
for porous media with low Reynolds numbers).
The model predicts the temperature profile, and
condensable volatile fraction in the gas phase,
Chapter 1 61
the escape velocity (dependant on particle size),
position (spatial gradients) and reaction time.
Transient, one-dimensional model for the fast
pyrolysis of a cellulosic cylinder. The model
accounts for transport phenomena through the
porous solid and the molten layer (adjacent to
the hot plate) whose properties and size change
Valid for ablative processes only. Devised
during thermal degradation. The effect of plate
for cylindrical pellets under pressure,
temperature and applied pressure on the
heated on a hot plate. No bubble dynamics
Di Blasi [258] process dynamics, ablation rate, thickness of the
within liquid layer are considered. Does not
reaction zone and product yields are simulated.
include secondary reactions, or moisture
The model considers thermal equilibrium
drying.
between the solid and gas phase; ideal gas
behavior; changes of properties with biomass
conversion; convective and diffusive transport of
gases; heat transfer by conduction, convection
and radiation; and particle shrinkage.
Wood flash pyrolysis has been performed and
model using single pellets. Mass transfer,
pressure gradients in the pellet, heat transfer via
Specific to flash heating of thick particles.
convection, conduction and radiation, and
Only solid conversion profiles are
structural changes occurring in the particle are
determined. Does not take into account the
Hastaoglu [259] included in the model. Wood particles are
formation of a liquid phase. The kinetic
assumed to consist of fibers with a hollow,
scheme used only represents lumped
cylindrical shape. Tar produced by primary
products.
reactions is assumed to further decompose by
secondary homogeneous cracking reactions to
yield gas.
A model of drying and devolatilization at
combustion temperatures. Employs one-
dimensional geometry with moving phase-
change boundaries, internal fluid flow and A simple single step scheme is used to
volatile formation. The model, includes steep describe the kinetics of pyrolysis.
Saastamoinen et al.
temperature and density profiles, frequently Secondary reactions, evolution of species
[260]
encountered in combustion. A transient model and a condensed phase are not
for the pyrolysis of solid fuels large (wood, coal, considered.
etc.) is used. Model presents a pore structure for
calculating the dependence of local thermal
conductivity on local porosity and temperature.
This model accounts for the combined effect of
Only available for spherical particles > 1mm.
various parameters such as particle shrinkage
Constant thermal properties are used. No
and drying, and was validated using available
Sharma et al. [261] liquid phase formation is considered. The
experimental data from the literature. A 5 step
model does not predict the evolution of
lumped kinetic model is employed.
individual chemical compounds.

Two Dimensional Models


62 Chapter 1
Energy transfer is valued by convection,
conduction and radiation. Mass transfer is
estimated by convection and momentum
diffusion using Darcy's law. In this model
physicochemical changes were estimated in two
dimensions (anisotropy), and finally the results
were compared with those of a non-steady state
This model proposes multiple steps in the
one-dimensional model. The comparison
overall reaction scheme, but does not
between the two models shows that the
include the formation of an intermediate
Di Blasi [262]. reactions are faster, and that more volatiles are
liquid phase, and does not include side
predicted by models with two dimensional
reactions, or the evolution of individual
configurations. An interesting feature of this work
chemical species.
is the evaluation of moisture evaporation and
study of different schemes for particle shrinkage,
including: (uniform shrinkage, shrinkage of
layers (cell shrinkage), and radial-cylindrical
shrinkage (Shrinking cylinders). The results
show that the shrinkage model used affects the
calculated heating rate.
This model examines the effect of particle size
and aspect ratio on intra-particle tar
Does not predict the rate of moisture
decomposition. The model also predicts
evaporation. Does not include mass transfer
Okenkunle et al.[95] temperature profiles and yields for the char,
by diffusion. Particle shrining is also not
volatiles and gases. The temperature profiles
included.
and char yields are compared using both
isotropic and anisotropic structures.
This model predicts the pyrolysis behavior of
large cylindrical (order of cm) biomass particles.
Does not consider the evolution of gaseous
Material properties are allowed to change with
species such as: CO,CO2,H2.. The formation
both conversion and the direction (anisotropy)
of intermediate liquid phase is also not
Bellais [263] Heat and mass transfer mechanisms are
considered. The model is limited to slow
evaluated by convection and diffusion. This
pyrolysis conditions ( slow heating rates).
model uses overall multi-step kinetics schemes,
neither include secondary reactions, either
evaluates the formation of the condensed phase.
Predict temperature profiles, drying rate, and
Isotropic particles are employed and mass
particle shrinkage. The carbohydrate profiles are
transfer by diffusion is neglected. Ideal
predicted inside the particle. The model
gases behavior is assumed with secondary
Paulsen et al. [264] classifies the solid particle as: unreacted
reactions omitted. Intermediate liquid phase
carbohydrate, virgin biomass and char. The
dynamics are neglected. The model was
simulated results are compared with an in-line IR
developed for thick square particles.
visualization.
Three Dimensional Models
This model examines pyrolysis of spherical The reaction scheme uses a single step
Yuen et al. [265] particles using a transient model for the pyrolysis global model without side reactions in the
of wet biomass. The0 model includes water gas phase, or gas-solid reactions. The
Chapter 1 63
evaporation, and spatial changes in development of condensable and non-
temperature, concentration, velocity, and condensable species, as well as the
physicochemical properties (anisotropy). formation of an intermediated liquid phase
are neglected. Mass and energy transport
resistance is also omitted.
An numerical study based on the Lattice
Boltzmann Method (LBM) is proposed here to
solve one, two and three dimensional heat and
mass transfer problems for the isothermal Does not include secondary reactions. The
carbonization of thick wood particles. The model physical properties are not depend of the
is used to study the effect of reactor temperature degree of biomass conversion. The model
Mahmoudi et al. [266]
and thermal boundary conditions, on the does not consider evolution of chemical
evolution of the local temperature and the mass species, or the effect of the heating rate on
distributions of products from the wood particle the reaction rate.
during carbonization. The model also considers
diffusive mechanisms for volatile through the
solid phase.
A biomass particle model with realistic
morphology and resolved microstructure for
simulations of intraparticle transport phenomena
is proposed. The morphology, and
microstructure are based on parameters
The model does not consider the enthalpy
obtained from quantitative image analysis.
of reaction, solid shrinking, secondary
Cieselinski et al. [267] Microstructural parameters, including cell wall
reactions, chemical compound evolution, or
thickness and cell lumen dimensions, are
intermediate liquid phase formation.
measured directly from micrographs of sectioned
biomass. A general, constructive solid geometry
algorithm is presented that produces models of
biomass particles based on these
measurements.
Models With Intermediate Liquid Phase
A one-dimensional model for the pyrolysis of This model does not consider the evolution
cylindrical cellulose cylindrical in an ablative of chemical species, or the effect of the
reactor (spinning-disk) as described by Ledé et heating rate on the reaction rate. It also
al [226]. A transient, one-dimensional, model disregards side reactions and the
Di Blasi [258]
was develoved using the Broido-Shafizadeh effects of diffusion and convection
kinetic scheme. This model also evaluates the of the volatiles. The bubbles dynamic
formation and evaporation of a condensed phase within the liquid phase are neglected. Model
intermediate. developed only for thick particles.
A model for the fast pyrolysis of cellulose pellets The model does not consider secondary
subjected to radiation heating. The model was reactions or loss of moisture from the
considers the formation and consumption of a biomass. Pressure gradients, convection
Boutin et al. [62] lumped intermediate liquid phase. This is a one- and mass transfer through each reaction
dimensional model, split into 3 areas (virgin zone are also not considered. Bubbles
biomass, liquid phase and char). Each zone dynamics and aerosol ejection are
utilizes separate mass and energy balances, neglected.
64 Chapter 1
and each interface boundary condition provides
continuity between the variables. The model only
predicts the temperature profile, and the
evolution of global species over time.
A dimensionless model for biomass fast
pyrolysis under non-steady state conditions. The
model includes the formation and consumption
of the intermediate liquid phase. Evaporation of
The model does not consider spatial
the liquid phase is modeled with by a Clausius-
changes to the global model parameters,
Clapeyron type expression and Darcy's law is
regarding the mechanisms of conduction
Dufour et al. [232] used to estimate volatile transport. The model
heat transfer, mass transfer by diffusion and
predicts the overall lumped yield and is
convection Does not include bubbles
dependent on the heating rate, and reaction
dynamics or aerosol ejection.
time. The time and temperature that yields the
maximum conversion to the intermediate liquid
phase is estimated.

A model of coal pyrolysis with high swelling


(softening coals), including the heat and mass
transfer through the condensed phase (meta There are no spatial gradients for
plastic). Bubble dynamics are considered temperature and concentration. The overall
Oh et al. [268] according to a discrete population balance. The kinetic scheme is single step, and does not
behavior is model under transient state evaluate the evolution of species. Aerosol
conditions. The yield of tar, light gases, char and ejection intensity is not included.
the evolution of condensed phase are predicted
as a function of time and temperature.
Presents a model for studying transient one
dimensional fast pyrolysis of spherical particles
of coal. This model is the most rigorous explored
in this review, and includes the effects of heat
and mass transfer in the condensed phase. The
The evolution of species, secondary
model also considers diffusive and convective
reactions and aerosol ejection intensity are
Sezen [269] transport of volatiles, heat transfer by conduction
not examined.
and convection, uses a multi-stage kinetic
mechanism. Bubble dynamics, including
coalescence, growth, and nucleation by
population balances are examined. The Swelling
and shrinking of the coal particles is also
considered.
2D-VOF modelling is performed to explain the
The model only applies to cellulose
boiling ejection of aerosols by bubble bursting.
pyrolysis, and was developed and is valid
Navier–Stokes equations are solved using the
only for one bubble burst. Does not take into
Teixeira et al. [133] finite volume method. Models were developed
account populations balances for other
for transport through the liquid/gas interface. The
bubbles, and does not describe other
simulation specifically considers ejection from
physical phenomena, including mass
the curved interface of a droplet, rather than an
Chapter 1 65
infinite pool with a flat surface. The model predict transfer by diffusion and convection, particle
the aerosol ejection velocity, aerosol droplet size shrinkage, or secondary reactions.
and distribution, and physical properties of the
molten phase.

1.12. Final Remarks

In this chapter, a general review of the effects of the main operational parameters on the global product distribution
(bio-oil, bio-char and syngas) was presented. This was done to provide a good understanding of the processes that
enable the maximization of pyrolysis products. From this review, the following conclusions and recommendations can
be drawn:

 Bio-oil production by pyrolysis is still an immature technology and it is not currently commercially viable. Bio-
oil production technologies still have to overcome many technical, economic and social barriers to compete
with traditional fossil fuels.

 Fast pyrolysis is a complex process, with many unexplored phenomena including chemical reactions; heating
rate effects on mass and energy transport inside particles; the effect of the real particle temperature during
pyrolysis; evolution of particle structural changes over time and as the process progresses; the evolution of
the intermediate phase (liquid phase) and its impact on product yields; and new experimental methodologies
to measure different parameters such as temperature, gas flow and chemical composition.

 The existence of the intermediate liquid phase has been demonstrated by many experimental techniques and
theoretical models for both cellulose and whole biomass. Although there is only very limited experimental
information regarding this phase and its incorporation into the particle kinetic models and how this phase
affects heat and mass transfer, and including the evolution of species condensable and non-condensable
volatile.

 Recent studies have shown that aerosols rich levoglucosan dimers and trimmers, and lignin oligomers are
formed directly from the ejection of fluid from the intermediate liquid phase. This phenomenon explains, at
fundamental level, the presence of heavy compounds in of bio-oil recovered from the fast pyrolysis of biomass,
as well as the agglomeration and crusting of biomass particles and inert material in the reactor. Experiments
at the laboratory level and rigorous models are still needed to understand many of the fundamentals associated
with bubble formation and aerosol ejection.

 The integration of reaction and transport phenomena in a single particle model is not a trivial undertaking, with
the validation the model an even more challenging task. Future pyrolysis models for particles must be able to
predict volatilization as a function of time for different biomass feedstock (necessary to design the reactor),
and predict the distribution of pyrolysis products. These models are likely to be based on existing transport
models for particles, but must include more detailed estimation methods for critical transport parameters (eg,
thermal conductivity, heat capacity, liquid and gas viscosity etc.), the condensed phase thermodynamics, and
the reaction kinetics.
66 Chapter 1

1.13. References
[1] IEA, Key World Energy Statistics 2010, OECD Publishing, 2010. doi:10.1787/9789264039537-en.
[2] G. Akhmat, K. Zaman, T. Shukui, F. Sajjad, Does energy consumption contribute to climate change? Evidence
from major regions of the world, Renew. Sustain. Energy Rev. 36 (2014) 123–134.
[3] International Energy Agency, I.E. Agency, Key World Energy, OECD Publishing, 2011.
[4] E.C. Marín, H.S. Mahecha, Biocombustibles y autosuficiencia energética, Dyna. 76 (2009) 101–110.
[5] WEC - World Energy Council, 2010 Survey of Energy Resources, World Energy Council, 2010.
[6] O.M. Hernandez, E. Humberto,Orduz,P. janneth, ZApata,L. Henry,Cardona,R. Maria Cecilia. Duarte, Atlas del
Potencial Energético de la Biomasa Residual en Colombia, 2013.
[7] J.I. Montoya, C. Valdés, F. Chejne, C.A. Gómez, A. Blanco, G. Marrugo, J. Osorio, E. Castillo, J. Aristóbulo, J.
Acero, Bio-oil production from Colombian bagasse by fast pyrolysis in a fluidized bed: An experimental study,
J. Anal. Appl. Pyrolysis. 112 (2015) 379–387.
[8] A. Dutta, R.M. Swanson, A. Aden, J.A. Satrio, M.M. Wright, A. Platon, D.E. Daugaard, R.C. Brown, D.D. Hsu,
R.P. Anex, F.K. Kazi, G. Kothandaraman, J. Fortman, Techno-economic comparison of biomass-to-
transportation fuels via pyrolysis, gasification, and biochemical pathways, Fuel. 89 (2010) S29–S35.
[9] E. Henrich, F. Schultmann, F. Trippe, R. Stahl, M. Fröhling, Techno-Economic Analysis of Fast Pyrolysis as a
Process Step Within Biomass-to-Liquid Fuel Production, Waste and Biomass Valorization. 1 (2010) 415–430.
[10] A. Dufour, C. Rogaume, Y. Rogaume, P. Girods, A. Zoulalian, Comparison of gasification and pyrolysis of
thermal pre-treated wood board waste, J. Anal. Appl. Pyrolysis. 85 (2009) 171–183.
[11] M.C. Tirado, M.J. Cohen, N. Aberman, J. Meerman, B. Thompson, Addressing the challenges of climate change
and biofuel production for food and nutrition security, Food Res. Int. 43 (2010) 1729–1744.
[12] P. De Wild, Biomass pyrolysis for chemicals, University Library Groningen, Groningen, 2011.
[13] M.-K. Bahng, C. Mukarakate, D.J. Robichaud, M.R. Nimlos, Current technologies for analysis of biomass
thermochemical processing: a review., Anal. Chim. Acta. 651 (2009) 117–138.
[14] W.R. Kang, K.B. Lee, Effect of operating parameters on methanation reaction for the production of synthetic
natural gas, Korean J. Chem. Eng. 30 (2013) 1386–1394.
[15] E. Ranzi, A. Cuoci, T. Faravelli, A. Frassoldati, G. Migliavacca, S. Pierucci, S. Sommariva, Chemical Kinetics
of Biomass Pyrolysis, Energy & Fuels. 22 (2008) 4292–4300.
[16] W.F. Degroot, M.D. Rahman, W.-P. Pan, G.N. Richards, First chemical events in pyrolysis of wood, J. Anal.
Appl. Pyrolysis. 13 (1988) 221–231.
[17] S. Salvador, C. Couhert, J.-M. Commandre, Is it possible to predict gas yields of any biomass after rapid
pyrolysis at high temperature from its composition in cellulose, hemicellulose and lignin?, Fuel. 88 (2009) 408–
417.
[18] C. Branca, C. Di Blasi, Multistep mechanism for the devolatilization of biomass fast pyrolysis oils, Ind. Eng.
Chem. Res. 45 (2006) 5891–5899.
[19] A. Sharma, V. Pareek, D. Zhang, Biomass pyrolysis—A review of modelling, process parameters and catalytic
studies, Renew. Sustain. Energy Rev. 50 (2015) 1081–1096.
[20] F. Tinwala, P. Mohanty, S. Parmar, A. Patel, K.K. Pant, Intermediate pyrolysis of agro-industrial biomasses in
bench-scale pyrolyser: Product yields and its characterization, Bioresour. Technol. 188 (2015) 258–264.
[21] A. Bridgwater, Fast Pyrolysis of Biomass: A Handbook, 3rd ed., PyNe, 2008.
[22] D. Meier, B. van de Beld, A. V. Bridgwater, D.C. Elliott, A. Oasmaa, F. Preto, State-of-the-art of fast pyrolysis
in IEA bioenergy member countries, Renew. Sustain. Energy Rev. 20 (2013) 619–641.
[23] R. Cuevas A, Lamas J, Advances and development at the Union Fenosa pyrolysis plant, in 8th
Conference on Biomass for Environment, in: O. Pergamom Press (Ed.), Agric. Ind., Chartier, Beenackers
and Grassi. Pergamom , 1995.
[24] H. Reith, P. de Wild, E. Heeres, Biomass pyrolysis for chemicals, Biofuels. 2 (2011) 185–208.
[25] Guo Xiu-juan, Wang Shu-Rong, Wang Kai-Ge, Liu Qian, Luo Zhong-Yang, Influence of extractives on
mechanism of biomass pyrolysis., J. Fuel Chem. Technol. 38 (2010).
[26] D.M. Alonso, S.G. Wettstein, J. a Dumesic, Bimetallic catalysts for upgrading of biomass to fuels and
chemicals., Chem. Soc. Rev. 41 (2012) 8075–98.
[27] M.W. Van de Weerdhof, Modeling the pyrolysis process of biomass particles, Eindhoven Universityof
Technology, 2009.
Chapter 1 67
[28] M.T. Klein, P.S. Virk, Modeling of Lignin Thermolysis, Energy & Fuels. 22 (2008) 2175–2182.
[29] M.W. Smith, Effect of Pyrolysis Temperature on Cellulose Char Aromatic Cluster Size by Quantitative Multi
Cross-Polarization 13C NMR with Long Range Dipolar.Characterization and modification of biomass pyrolysis
chars for environmental applications Dephasing, Washington State University, 2016.
[30] L. Li, P. P, X. Frank, K. Mazeau, A coarse-grain force-field for xylan and its interaction with cellulose, Carbohydr.
Polym. 127 (2015) 438–450.
[31] P.R. Patwardhan, J.A. Satrio, R.C. Brown, B.H. Shanks, Influence of inorganic salts on the primary pyrolysis
products of cellulose., Bioresour. Technol. 101 (2010) 4646–4655.
[32] J. Akhtar, N. Saidina Amin, A review on operating parameters for optimum liquid oil yield in biomass pyrolysis,
Renew. Sustain. Energy Rev. 16 (2012) 5101–5109.
[33] C. Zhengng, L. D, H. Yang, R. Yan, H. Chen, D. Lee, Influence of mineral matter on pyrolysis of palm oil wastes,
Combust. Flame. 146 (2006) 605–611.
[34] T.J. Haas, M.R. Nimlos, B.S. Donohoe, Real-Time and Post-reaction Microscopic Structural Analysis of
Biomass Undergoing Pyrolysis, Energy & Fuels. 23 (2009) 3810–3817.
[35] A. Sluiter, B. Hames, R. Ruiz, C. Scarlata, J. Sluiter, D. Templeton, D. Crocker, Determination of Structural
Carbohydrates and Lignin in Biomass, National Renewable Energy Laboratory, U. S., 2011.
[36] A. Trubetskaya, P.A. Jensen, A.D. Jensen, M. Steibel, H. Spliethoff, P. Glarborg, Influence of fast pyrolysis
conditions on yield and structural transformation of biomass chars, Fuel Process. Technol. 140 (2015) 205–
214.
[37] G.A. Zickler, W. Wagermaier, S.S. Funari, M. Burghammer, O. Paris, In situ X-ray diffraction investigation of
thermal decomposition of wood cellulose, J. Anal. Appl. Pyrolysis. 80 (2007) 134–140.
[38] Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, C.-Z. Li, A.G. McDonald, M. Garcia-Perez, Effect of
Cellulose Crystallinity on Solid/Liquid Phase Reactions Responsible for the Formation of Carbonaceous
Residues during Pyrolysis, Ind. Eng. Chem. Res. 53 (2014) 2940–2955.
[39] G. Cheng, P. Varanasi, C. Li, H. Liu, Y.B. Melnichenko, B.A. Simmons, M.S. Kent, S. Singh, Transition of
cellulose crystalline structure and surface morphology of biomass as a function of ionic liquid pretreatment and
its relation to enzymatic hydrolysis., Biomacromolecules. 12 (2011) 933–941.
[40] S. Andersson, R. Serimaa, T. Paakkari, P. Saranpaa, E. Pesonen, Crystallinity of wood and the size of cellulose
crystallites in Norway spruce (Picea abies), J. Wood Sci. 49 (2003) 531–537.
[41] G.R. Ponder, T.T. Stevenson, G.N. Richards, Influence of linkage position and orientation in pyrolysis of
polysaccharides: A study of several glucans, J. Anal. Appl. Pyrolysis. 22 (1992) 217–229.
[42] Matthew S.Mettler, Alex D Paulsen, Dionisios G Vlachos, Paul J. Dauenhauer, The chain length effect in
pyrolysis: bridging the gap between glucose and cellulose, Green Chem. 14 (2012) 1284–1294.
[43] S. Saka, M. Asmadi, H. Kawamoto, Gas- and solid/liquid-phase reactions during pyrolysis of softwood and
hardwood lignins, J. Anal. Appl. Pyrolysis. 92 (2011) 417–425.
[44] J.-Y. Kim, S. Oh, H. Hwang, U.-J. Kim, J.W. Choi, Structural features and thermal degradation properties of
various lignin macromolecules obtained from poplar wood (Populus albaglandulosa), Polym. Degrad. Stab. 98
(2013) 1671–1678.
[45] T. Hosoya, S. Saka, H. Kawamoto, Cellulose–hemicellulose and cellulose–lignin interactions in wood pyrolysis
at gasification temperature, J. Anal. Appl. Pyrolysis. 80 (2007) 118–125.
[46] T.J. Hilbers, Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, M.R. Pelaez-Samaniego, M. Garcia-Perez,
Cellulose-Lignin interactions during slow and fast pyrolysis, J. Anal. Appl. Pyrolysis. 114 (2015) 197–207.
[47] A. Dufour, M. Castro-Diaz, N. Brosse, M. Bouroukba, C. Snape, The origin of molecular mobility during biomass
pyrolysis as revealed by in situ 1H NMR spectroscopy, ChemSusChem. 5 (2012) 1258–1265.
[48] Broido A., Molecular weight decrease in the early pyrolysis of crystalline and amorphous cellulose, J. Appl.
Polym. Sci. 17 (1973) 3627–3635.
[49] D. Liu, Y. Yu, H. Wu, Differences in Water-Soluble Intermediates from Slow Pyrolysis of Amorphous and
Crystalline Cellulose, Energy & Fuels. 27 (2013) 1371–1380.
[50] J.R. Obst, T.K. Kirk, Isolation of lignin, Methods Enzymol. 161 (1988) 3–12.
[51] L. Yang, D. Wang, D. Zhou, Y. Zhang, Effect of different isolation methods on structure and properties of lignin
from valonea of Quercus variabilis., Int. J. Biol. Macromol. 85 (2016) 417–24.
[52] Custodis, C. Bährle, F. Vogel, J.A. van Bokhoven, Phenols and aromatics from fast pyrolysis of variously
prepared lignins from hard- and softwoods, J. Anal. Appl. Pyrolysis. 115 (2015) 214–223.
68 Chapter 1
[53] Montoya .J, B. Pecha, F. Chejne Janna, Micro-Explosion of Liquid Intermediates During the Fast Pyrolysis of
Sucrose and Organosolv Lignin, Submitt. to J. Anal. Appl. Pyrolysis.
[54] A. Bharadwaj, L.L. Baxter, A.L. Robinson, Effects of intraparticle heat and mass transfer on biomass
devolatilization: Experimental results and model predictions, Energy and Fuels. 18 (2004) 1021–1031.
[55] Alfred Bliek, W. M. Van Poelje, Effects of intraparticle heat and mass transfer during devolatilization of a single
coal particle, AIChE J. 31 (1985) 1666–1681.
[56] C. Di Blasi, Modeling Intra- and Extra-Particle Processes of Wood Fast Pyrolysis, AIChE J. 48 (2002) 2386–
2397.
[57] A. Melgar, D. Borge, J.F. Pérez, Estudio cinético del proceso de devolatilización de biomasa lignocelulósica
mediante análisis termogravimétrico para tamaños de partícula de 2 a 19 mm, Dyna. 75 (2008) 123–131.
[58] P. Basu, Biomass Gasification and Pyrolysis, © 2010 Elsevier Inc., 2010. doi:10.1016/B978-0-12-374988-
8.00011-8.
[59] R.J. Westerhof, Refining fast pyrolysis of biomass, Thesis Degree of Doctor of Science, University of Twente,
2011.
[60] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials., Curr. Opin.
Biotechnol. 24 (2013) 414–20.
[61] S. Zhou, M. Garcia-Perez, B. Pecha, A.G. McDonald, R.J.M. Westerhof, Effect of particle size on the
composition of lignin derived oligomers obtained by fast pyrolysis of beech wood, Fuel. 125 (2014) 15–19.
[62] O. Boutin, M. Ferrer, J. Lédé, Flash pyrolysis of cellulose pellets submitted to a concentrated radiation:
experiments and modelling, Chem. Eng. Sci. 57 (2002) 15–25.
[63] Gómez Alexander, Klose Wolframg, Rincón Sonia, Pirólisis de biomasa. Cuesco de palma de aceite, Bogotá,
2008.
[64] A. Bridgwater, Fast pyrolysis processes for biomass, Renew. Sustain. Energy Rev. 4 (2000) 1–73.
[65] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Effects of particle size on the fast
pyrolysis of oil mallee woody biomass., Fuel. 88 (2009) 1810–1817.
[66] Ö.M. Koçkar, Ö. Onay, S.H. Beis, Fixed-bed pyrolysis of safflower seed: influence of pyrolysis parameters on
product yields and compositions, Renew. Energy. 26 (2002) 21–32.
[67] M. Rafiqul Alam Beg, M. Nurul Islam, M. Rofiqul Islam, Pyrolytic oil from fixed bed pyrolysis of municipal solid
waste and its characterization, Renew. Energy. 30 (2005) 413–420.
[68] E. Hoekstra, W.P.M. Van Swaaij, S.R.A. Kersten, K.J.A. Hogendoorn, Fast pyrolysis in a novel wire-mesh
reactor: Decomposition of pine wood and model compounds, Chem. Eng. J. 187 (2012) 172–184.
[69] Hogendoorn, E. Hoekstra, S.R.A. Kersten, W.P.M. van Swaaij, Fast pyrolysis in a novel wire-mesh reactor:
Design and initial results, Chem. Eng. J. 191 (2012) 45–58.
[70] Alex D. Paulsen, Matthew S. Mettler, Paul J. Dauenhauer, The role of sample dimension and temperature
in cellulose pyrolysis, Energy Fuels. 27 (2013) 2126–2134.
[71] M. Sohrabi, M.R. Hajaligol, M. Nik-Azar, B. Dabir, Mineral matter effects in rapid pyrolysis of beech wood, Fuel
Process. Technol. 51 (1997) 7–17.
[72] K. Raveendran, A. Ganesh, K.C. Khilar, Influence of mineral matter on biomass pyrolysis characteristics, Fuel.
74 (1995) 1812–1822.
[73] Y. Yürüm, N.. Öztaş, Pyrolysis of Turkish Zonguldak bituminous coal. Part 1. Effect of mineral matter, Fuel. 79
(2000) 1221–1227.
[74] Alexander W. Williams, An Investigation of the kinetics for the fast pyrolysis of loblolly pine woody biomass, .
Georgia Institute of Technology , 2011.
[75] P.R. Patwardhan, Understanding the product distribution from biomass fast pyrolysis, Iowa State University,
2010. Doctoral Thesis.
[76] G. Gellerstedt, M. Kleen, Influence of inorganic species on the formation of polysaccharide and lignin
degradation products in the analytical pyrolysis of pulps, J. Anal. Appl. Pyrolysis. 35 (1995) 15–41.
[77] G.N. Richards, G. Zheng, Influence of metal ions and of salts on products from pyrolysis of wood: Applications
to thermochemical processing of newsprint and biomass, J. Anal. Appl. Pyrolysis. 21 (1991) 133–146.
[78] Z.A. Mayer, A. Apfelbacher, A. Hornung, A comparative study on the pyrolysis of metal- and ash-enriched wood
and the combustion properties of the gained char, J. Anal. Appl. Pyrolysis. 96 (2012) 196–202.
[79] F. Sulaiman, N. Abdullah, Optimum conditions for maximising pyrolysis liquids of oil palm empty fruit bunches,
Energy. 36 (2011) 2352–2359.
Chapter 1 69
[80] J.E. White, W.J. Catallo, B.L. Legendre, Biomass pyrolysis kinetics: A comparative critical review with relevant
agricultural residue case studies, J. Anal. Appl. Pyrolysis. 91 (2011) 1–33.
[81] J. Valette, J. Blin, F.-X. Collard, A. Bensakhria, Influence of impregnated metal on the pyrolysis conversion of
biomass constituents, J. Anal. Appl. Pyrolysis. 95 (2012) 213–226.
[82] K. Shimomura, H. Watanabe, K. Okazaki, Effect of high CO2 concentration on char formation through mineral
reaction during biomass pyrolysis, Proc. Combust. Inst. (2012).
[83] Z. Mayer, A. Apfelbacher, A. Hornung, Effect of sample preparation on the thermal degradation of metal-added
biomass, J. Anal. Appl. Pyrolysis. 94 (2012) 170–176.
[84] A. Jensen, K. Dam-Johansen, M.A. Wójtowicz, M.A. Serio, TG-FTIR Study of the Influence of Potassium
Chloride on Wheat Straw Pyrolysis, Energy & Fuels. 12 (1998) 929–938.
[85] Oudenhoven, R.J.M. Westerhof, N. Aldenkamp, D.W.F. Brilman, S.R.A. Kersten, Demineralization of wood
using wood-derived acid: Towards a selective pyrolysis process for fuel and chemicals production, in: J. Anal.
Appl. Pyrolysis, 2013: pp. 112–118.
[86] Oudenhoven, R.J.M. Westerhof, S.R.A. Kersten, Fast pyrolysis of organic acid leached wood, straw, hay and
bagasse: Improved oil and sugar yields, J. Anal. Appl. Pyrolysis. 116 (2015) 253–262.
[87] T. Iwasaki, S. Suzuki, T. Kojima, Influence of Biomass Pyrolysis Temperature, Heating Rate and Type of
Biomass on Produced Char in a Fluidized Bed Reactor, Energy Environ. Res. 4 (2014) 64–72.
[88] S.R.G. Oudenhoven, C. Lievens, R.J.M. Westerhof, S.R.A. Kersten, Effect of temperature on the fast pyrolysis
of organic-acid leached pinewood; the potential of low temperature pyrolysis, Biomass and Bioenergy. 89
(2016) 78–90.
[89] S.R.A. Kersten, X. Wang, W. Prins, W.P.M. van Swaaij, Biomass Pyrolysis in a Fluidized Bed Reactor. Part 1:
Literature Review and Model Simulations, Ind. Eng. Chem. Res. 44 (2005) 8773–8785.
[90] Bridgwater, A.V. and Peacocke, Fast pyrolysis processes for biomass, Renew. Sustain. Energy Rev. 4 (2000)
1–73.
[91] P. Hasler, T. Nussbaumer, P. Morf, Mechanisms and kinetics of homogeneous secondary reactions of tar from
continuous pyrolysis of wood chips, Fuel. 81 (2002) 843–853.
[92] Scott, P. Majerski, J. Piskorz, D. Radlein, A second look at fast pyrolysis of biomass—the RTI process, J. Anal.
Appl. Pyrolysis. 51 (1999) 23–37.
[93] Shurong Wang, Mengxiang Fang, Chunjiang Yu, Zhongyang Luo, Kefa Cen, Flash pyrolysis of biomass
particles in fluidized bed for bio-oil production, China Particuology. 3 (2005) 136–140.
[94] A.V. Bridgwater, Review of fast pyrolysis of biomass and product upgrading, Biomass and Bioenergy. 38 (2012)
68–94.
[95] Pious O. Okekunle, Hirotatsu Watanabe, and K. Okazak, Effect of Biomass Size and Aspect Ratio on Intra-
Particle Tar Decomposition during Wood Cylinder Pyrolysis, J. Therm. Sci. Technol. 7 (2012) 1–15.
[96] Boroson, J.B. Howard, J.P. Longwell, W. a. Peters, Product yields and kinetics from the vapor phase cracking
of wood pyrolysis tars, AIChE J. 35 (1989) 120–128.
[97] Hajaligol, Rapid pyrolysis of cellulose, MIT, 1980.
[98] M. Garcia-Perez, P. Lappas, P. Hughes, L. Dell, a Chaala, D. Kretschmer, C. Roy, Evaporation and
Combustion Characteristics of Biomass Vacuum Pyrolysis Oils, IFRF Combust. J. 200601 (2006) 1–27.
[99] J. Kizito, Vapor-Gas Bubble Evolution and Growth in Extremely, Methods. (2009) 1–11.
[100] M.M. Avulapati, L.C. Ganippa, J. Xia, A. Megaritis, Puffing and micro-explosion of diesel–biodiesel–ethanol
blends, Fuel. 166 (2016) 59–66.
[101] R.M. Baldwin, B.S. Donohoe, M.R. Nimlos, K.A. Magrini-Bair, J.E. Hensley, P. Pepiot, S.D. Phillips, Current
research on thermochemical conversion of biomass at the National Renewable Energy Laboratory, Appl. Catal.
B Environ. 115–116 (2012) 320–329.
[102] Y.Y., and H.W. Dawei Liu, Differences in Water-Soluble Intermediates from Slow Pyrolysis of Amorphous and
Crystalline Cellulose, Energy Fuels. 27 (2013) 1371–1380.
[103] J. Lédé, Biomass Pyrolysis: Comments on Some Sources of Confusions in the Definitions of Temperatures and
Heating Rates, Energies. 3 (2010) 886–898.
[104] C. Di Blasi, Modelling the fast pyrolysis of cellulosic particles in fluid-bed reactors, Chem. Eng. Sci. 55 (2000)
5999–6013.
[105] Y. Lin, J. Cho, G.A. Tompsett, P.R. Westmoreland, G.W. Huber, Kinetics and Mechanism of Cellulose Pyrolysis,
J. Phys. Chem. C. 113 (2009) 20097–20107.
70 Chapter 1
[106] J. Lédé, J.P. Diebold, G:V.C Peacocke, The nature and properties of intermediate and unvaporized biomass
pyrolisis materials, in: Springer Science+ Bussines Media (Ed.), Dev. Thermochem. Biomass Conversion.,
Dordrecht, 1997: pp. 27–42.
[107] J. Lédé, Cellulose pyrolysis kinetics: An historical review on the existence and role of intermediate active
cellulose, J. Anal. Appl. Pyrolysis. 94 (2012) 17–32.
[108] P. Pineda-Gomez, C.M. Bedoya-Hincapié, A. Rosales-Rivera, Estomación de los parámetros cinéticos y tiempo
de vida de la cascarilla de arroz y arcilla mediante la tecnica de análisis termogravimétrico (TGA), Dyna. 76
(2011).
[109] Jorge Montoya, Brennan Pecha, David Roman, Farid Chejne Janna, Effect of Temperature and Heating Rate
on Product Distribution from Pyrolysis of Sugarcane Bagasse in a Hot Plate Reactor, J. Anal. Appl. Pyrolysis.
Submited (2016).
[110] R.K. Singh, B. Ruj, Time and temperature depended fuel gas generation from pyrolysis of real world municipal
plastic waste, Fuel. 174 (2016) 164–171.
[111] P. Basu, Biomass gasification and pyrolysis practical design and theory, Academic Press, Burlington, MA, 2010.
[112] R. Narayan, M.J.J. Antal, Thermal Lag, Fusion, and the Compensation Effect during Biomass Pyrolysis, Ind.
Eng. Chem. Res. 35 (1996) 1711–1721.
[113] György Pokol, Gáor Várhegyi, David Dollimore, Kinetic aspects of thermal analysis, C R C Crit. Rev. Anal.
Chem. 19 (2008) 65–93.
[114] A. V Bridgwater, Fast pyrolysis of biomass: a handbook, CPL Press, Newbury, 1999.
[115] A.S. Kalgo, The Development and Optimisation of a Fast Pyrolysis Process for Bio-oil Production, Thesis
Degree of Doctor of Phylosophy, Aston University, 2011.
[116] M. Hajaligol, T. Fisher, B. Waymack, D. Kellogg, Pyrolysis behavior and kinetics of biomass derived materials,
J. Anal. Appl. Pyrolysis. 62 (2002) 331–349.
[117] P.J. Dauenhauer, J.L. Colby, C.M. Balonek, W.J. Suszynski, L.D. Schmidt, Reactive boiling of cellulose for
integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[118] Scott, J. Piskorz, M.A. Bergougnou, R.P. Overend, R. Graham, The role of temperature in the fast pyrolysis of
cellulose and wood, Ind. Eng. Chem. Res. 27 (1988) 8–15.
[119] Peng Fua, Song Hua, Jun Xianga, Peisheng Lib, Dan Huanga, Long Jianga, Anchao Zhanga, Junying Zhanga,
FTIR study of pyrolysis products evolving from typical agricultural residues, J. Anal. Appl. Pyrolysis. 88 (2010)
117–123.
[120] Manuel Garcia-Perez, The Formation of polyaromatic hydrocarbons anddioxins during pyrolysis: A review of
the literature with Descriptions of biomass composition, fast pyrolysis technologies and thermochemical
reactions, Washington, 2008.
[121] M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W.J. Lee, C.Z. Li, Effects of temperature on the formation of
lignin-derived oligomers during the fast pyrolysis of Mallee woody biomass, Energy and Fuels. 22 (2008) 2022–
2032.
[122] J. Diebold, J. Scahill, Ablative pyrolysis of biomass in solid-convective heat transfer environments, in: R. P.
Overend, T. A. Milne, L. K. Mudge (Eds.), Fundam. Thermochem. Biomass Convers., Springer Netherlands,
1985: pp. 539–555.
[123] A.S. Kalgo, The development and optimisation of a fast pyrolysis process for bio-oil production,
Aston University, 2011.
[124] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[125] Weerawut Chaiwat, Isao Hasegawa, Takaaki Tani, Kenshi Sunagawa, Kazuhiro Mae, Analysis of cross-
linking behavior during pyrolysis of cellulose for elucidating reaction pathway, Energy Fuels. 23 (2009) 5765–
5772.
[126] V. Agarwal, P.J. Dauenhauer, G.W. Huber, S.M. Auerbach, Ab Initio Dynamics of Cellulose Pyrolysis: Nascent
Decomposition Pathways at 327 and 600 °C, J. Am. Chem. Soc. 134 (2012) 14958–14972.
[127] Ozlem Onay, Influence of pyrolysis temperature and heating rate on the production of bio-oil and char from
safflower seed by pyrolysis, using a well-swept fixed-bed reactor, Fuel Process. Technol. 88 (2007) 523–531.
[128] S. Thangalazhy-gopakumar, S. Adhikari, R.B. Gupta, S.D. Fernando, Influence of Pyrolysis Operating
Conditions on Bio-Oil Components : A Microscale Study in a Pyroprobe, Energy and Fuels. 25 (2011) 1191–
1199.
Chapter 1 71
[129] G. Kumar, R. Singh, A.K. Panda, Optimization of process for the production of bio-oil from eucalyptus wood, J.
Fuel Chem. Technol. 38 (2010) 162–167.
[130] E. Bolat, K. Açıkalın, F. Karaca, Pyrolysis of pistachio shell: Effects of pyrolysis conditions and analysis of
products, Fuel. 95 (2012) 169–177.
[131] N. Ozbay, A.E. Putun, E. Putun, Bio-oil production from rapid pyrolysis of cottonseed cake: product yields and
compositions, Int. J. Energy Res. 30 (2006) 501–510.
[132] S. Zhou, B. Pecha, M. van Kuppevelt, A.G. McDonald, M. Garcia-Perez, Slow and fast pyrolysis of Douglas-fir
lignin: Importance of liquid-intermediate formation on the distribution of products, Biomass and Bioenergy. 66
(2014) 398–409.
[133] P.Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P. and
Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011)
4306.
[134] E. Mura, P. Massoli, C. Josset, K. Loubar, J. Bellettre, Study of the micro-explosion temperature of water in oil
emulsion droplets during the Leidenfrost effect, Exp. Therm. Fluid Sci. 43 (2012) 63–70.
[135] S. Das, D.S. Kumar, S. Bhaumik, Experimental study of nucleate pool boiling heat transfer of water on silicon
oxide nanoparticle coated copper heating surface, Appl. Therm. Eng. 96 (2016) 555–567.
[136] Peter R. Solomon, Michael A. Serio, Eric M. Suuberg, Coal pyrolysis: Experiments, kinetic rates and
mechanisms, Prog. Energy Combust. Sci. 18 (1992) 133–220.
[137] J. Lédé, O. Authier, Temperature and heating rate of solid particles undergoing a thermal decomposition. Which
criteria for characterizing fast pyrolysis?, J. Anal. Appl. Pyrolysis. 113 (2015) 1–14.
[138] X. Gong, Y. Yu, X. Gao, Y. Qiao, M. Xu, H. Wu, Formation of anhydro-sugars in the primary volatiles and solid
residues from cellulose fast pyrolysis in a wire-mesh reactor, Energy and Fuels. 28 (2014) 5204–5211.
[139] A.V. Sepman, L.P.H. de Goey, Plate reactor as an analysis tool for rapid pyrolysis of biomass, Biomass and
Bioenergy. 35 (2011) 2903–2909.
[140] S.S. Liaw, V. Haber Perez, S. Zhou, O. Rodriguez-Justo, M. Garcia-Perez, Py-GC/MS studies and principal
component analysis to evaluate the impact of feedstock and temperature on the distribution of products during
fast pyrolysis, J. Anal. Appl. Pyrolysis. 109 (2014) 140–151.
[141] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[142] Orfão, F.J.A. Antunes, J.L. Figueiredo, Pyrolysis kinetics of lignocellulosic materials—three independent
reactions model, Fuel. 78 (1999) 349–358.
[143] Morten Grunnar Gronli, A theorical and experimental study of the thermal degradation of biomass, Norwegian
University of Science and Technology, 1996.
[144] Gábor Várhegyi, Tamás Székely, Reaction kinetics in thermal analysis. Part 1. The sensitivity of kinetic
equations to experimental errors. A mathematical analysis, Thermochim. Acta. 57 (1982) 13–28.
[145] Gábor Várhegyi, Reaction kinetics in thermal analysis: a brief survey of fundamental research problems,
Thermochim. Acta. 110 (1987) 95–99.
[146] M.J. Antal, Cellulose Pyrolysis Kinetics : The Current State of Knowledge, Ind. Eng. Chem. Res. 34 (1995) 703–
717.
[147] I. Milosavljevic, E. Suuberg, Cellulose thermal-decomposition kinetics - global mass-loss kinetics, Ind. Eng.
Chem. Res. 34 (1995) 1081–1091.
[148] Scott, A. Grinshpun, D.S.T.A.G. Radlein, J. Piskorz, On the presence of anhydro-oligosaccharides in the sirups
from the fast pyrolysis of cellulose, J. Anal. Appl. Pyrolysis. 12 (1987) 39–49.
[149] Paul J Dauenhauer, Joshua L Colby, Christine M Balonek, Wieslaw J Suszynski, Reactive boiling of cellulose
for integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[150] W.J. Catallo, J.E. White, B.L. Legendre, Biomass pyrolysis kinetics: A comparative critical review with relevant
agricultural residue case studies, J. Anal. Appl. Pyrolysis. 91 (2011) 1–33.
[151] R. Capart, A.K. Burnham, L. Khezami, Assessment of various kinetic models for the pyrolysis of a microgranular
cellulose, Thermochim. Acta. 417 (2004) 79–89. doi:10.1016/j.tca.2004.01.029.
[152] O. Boutin, J. Lede, G. Olalde, A. Ferriere, Solar flash pyrolysis of biomass direct measurement of the optical
properties of biomass components, J. Phys. Iv. 9 (1999) 367–372.
[153] Broido, Speculations on the nature of cellulose pyrolysis, Pyrodinamica. (1965).
[154] Kilzer F.J, Broido A., The Modified Kilzer-Bioido Model, Pyrodynamics. 2 (1965) 151–163.
72 Chapter 1
[155] G. Varhegyi, E. Jakab, M.J. Antal Jr., Is the Broido-Shafizadeh Model for Cellulose Pyrolysis True, Energy &
Fuels. 8 (1994) 1345–1352.
[156] Diebold, An unified, global-model for the pyrolysis of cellulose , Biomass Bioenergy. 7 (1994) 75–85.
[157] F. Shafizadeh, M.L. Wolfrom, R.S. Tipson, Pyrolysis and combustion of cellulosic materials, Adv. Carbohydr.
Chem. 23 (1968) 419–474.
[158] C.A. Koufopanos, G. Maschio, A. Lucchesi, Kinetic modeling of the pyrolysis of biomass and biomass
components., Can. J. Chem. Eng. 67 (1989) 75–84.
[159] J.L. Banyasz, S. Li, J.L. Lyons-Hart, K.H. Shafer, Cellulose pyrolysis: the kinetics of hydroxyacetaldehyde
evolution, J. Anal. Appl. Pyrolysis. 57 (2001) 223–248.
[160] R.W. Agrawal, Kinetics of the reactions involved in pyrolysis of cellulose.The three reaction model., Can. J.
Chem. Eng. 66 (1988) 403–412.
[161] Bradbury, Y. Sakai, F. Shafizadeh, A kinetic model for pyrolysis of cellulose, J. Appl. Polym. Sci. 23 (1979)
3271–3280.
[162] G. Várhegyi, M.J. Antal, E. Jakab, P. Szabó, Kinetic modeling of biomass pyrolysis, J. Anal. Appl. Pyrolysis. 42
(1997) 73–87.
[163] A. Broido, M.W. Weinstein, Kinetics of solid-phase cellulose, in: Birkhauser Verleg (Ed.), Proc. 3 Rd Int. Conf.
Therm. Anal., 1972: pp. 285–296.
[164] C. Di Blasi, Comparison of semi-global mechanisms for primary pyrolysis of lignocellulosic fuels, J. Anal. Appl.
Pyrolysis. 47 (1998) 43–64.
[165] J. Lédé, J., Diebold, J.P., Peacocke, G.V.C. and Piskorz, The nature and properties of intermediate and
unvaporized biomass pyrolysis materials., Dev. Thermochem. Biomass Convers. (1997) 27–42.
[166] C. S¸erb˘anescu, Kinetic analysis of cellulose pyrolysis: a short review, Chem. Pap. 67 (2014) 847–860.
[167] H. Yang, R. Yan, H. Chen, C. Zheng, D.H. Lee, D.T. Liang, In-depth investigation of biomass pyrolysis based
on three major components: hemicellulose, cellulose and lignin, Energy and Fuels. 20 (2006) 388–393.
[168] R. Font, A. Marcilla, E. Verdú, Kinetics of the pyrolysis of almond shells impregnated with CoCl2in a fluidized
bed reactor and in a pyrotrobe 100., J. Ind. Eng. Chem. Res. 29 (1990) 1846–1855.
[169] T.R. Rao, A. Sharma, Pyrolysis rates of biomass materials, Energy. 23 (1998) 973–978.
[170] X. Guo, Z. Luo, K. Wang, S. Wang, Influence of the interaction of components on the pyrolysis behavior of
biomass, J. Anal. Appl. Pyrolysis. 91 (2011) 183–189.
[171] A. Dufour, M. Castro-Díaz, P. Marchal, N. Brosse, R. Olcese, M. Bouroukba, C. Snape, In Situ Analysis of
Biomass Pyrolysis by High Temperature Rheology in Relations with 1H NMR, Energy & Fuels. 26 (2012) 6432–
6441.
[172] T. Gebreegziabher, C.W. Hui, A.O. Oyedun, Mechanism and modelling of bamboo pyrolysis, Fuel Process.
Technol. (2012).
[173] R. Vinu, Linda J. Broadbelt., A mechanistic model of fast pyrolysis of glucose-based carbohydrates to predict
bio-oil composition., Energy Environ. Sci. 5 (2012) 9808–9826.
[174] M. Calonaci, R. Grana, E.B. Hemings, G. Bozzano, M. Dente, E. Ranzi, Comprehensive kinetic modeling study
of bio-oil formation from fast pyrolysis of biomass, Energy & Fuels. 24 (2010) 5727–5734.
[175] Liao Yan-fen, Wang Shu-rong, Ma Xiao-qian, Study of reaction mechanisms in cellulose pyrolysis.
[176] R.H. Venderbosch, W. Prins, Fast pyrolysis technology development, Biofuels, Bioprod. Biorefining. 4 (2010)
178–208.
[177] Scott, D. Radlein, J. Piskorz, L. Paterson, Pretreatment of poplar wood for fast pyrolysis: rate of cation removal,
J. Anal. Appl. Pyrolysis. 57 (2001) 169–176.
[179] S. Niksa, Predicting the rapid devolatilization of diverse forms of biomass with bio-flashchain, Proc. Combust.
Inst. 28 (2000) 2727–2733.
[180] http://www.afrinc.com/products/fgdvc/.
[181] C.P. Please, M.J. McGuinness, McElwain, Approximations to the distributed activation energy model for the
pyrolysis of coal, Combust. Flame. 133 (2003) 107–117.
[182] Q. Xiong, J. Zhang, F. Xu, G. Wiggins, C. Stuart Daw, Coupling DAEM and CFD for simulating biomass fast
pyrolysis in fluidized beds, J. Anal. Appl. Pyrolysis. 117 (2016) 176–181.
[183] Shen, S. Gu, B. Jin, M.X. Fang, Thermal degradation mechanisms of wood under inert and oxidative
environments using DAEM methods, Bioresour. Technol. 102 (2011) 2047–2052.
[184] Q.-V. Bach, K.-Q. Tran, Ø. Skreiberg, Combustion kinetics of wet-torrefied forest residues using the distributed
Chapter 1 73
activation energy model (DAEM), Appl. Energy. (2016).
[185] L. Fiori, M. Valbusa, D. Lorenzi, L. Fambri, Modeling of the devolatilization kinetics during pyrolysis of grape
residues, Bioresour. Technol. 103 (2012) 389–397.
[186] J. Cai, R. Liu, New distributed activation energy model: Numerical solution and application to pyrolysis kinetics
of some types of biomass, Bioresour. Technol. 99 (2008) 2795–2799.
[187] A.A. Jain, A. Mehra, V. V. Ranade, Processing of TGA data: Analysis of isoconversional and model fitting
methods, Fuel. 165 (2016) 490–498.
[188] S. Wang, H. Lin, B. Ru, W. Sun, Y. Wang, Z. Luo, Comparison of the pyrolysis behavior of pyrolytic lignin and
milled wood lignin by using TG–FTIR analysis, J. Anal. Appl. Pyrolysis. 108 (2014) 78–85.
[189] M. V Navarro, R. Murillo, A.M. Mastral, N. Puy, J. Bartroli, Application of the Distributed Activation Energy Model
to Biomass and Biomass Constituents Devolatilization, Aiche J. 55 (2009) 2700–2715.
[190] J. Cai, S. Yang, T. Li, Logistic distributed activation energy model--part 2: application to cellulose pyrolysis.,
Bioresour. Technol. 102 (2011) 3642–3644.
[191] A. Meng, H. Zhou, L. Qin, Y. Zhang, Q. Li, Quantitative and kinetic TG-FTIR investigation on three kinds of
biomass pyrolysis, J. Anal. Appl. Pyrolysis. 104 (2013) 28–37.
[192] A.N. Hayhurst, J.F. Davidson, S.A. Scott, J.S. Dennis, An algorithm for determining the kinetics of
devolatilisation of complex solid fuels from thermogravimetric experiments, Chem. Eng. Sci. 61 (2006) 2339–
2348.
[193] C. Ulloa, A.L. Gordon, X. Garcia, Distribution of activation energy model applied to the rapid pyrolysis of coal
blends, J. Anal. Appl. Pyrolysis. 71 (2004) 465–483.
[194] Y. Chen, R. Bassilakis, W.W. Smith, R.M. Carangelo, M.A. Wójtowicz, Modeling the evolution of volatile species
during tobacco pyrolysis, J. Anal. Appl. Pyrolysis. 66 (2003) 235–261.
[195] T. Sonobe, N. Worasuwannarak, Kinetic analyses of biomass pyrolysis using the distributed activation energy
model, Fuel. 87 (2008) 414–421.
[196] J. Cai, T. Li, R. Liu, A critical study of the Miura-Maki integral method for the estimation of the kinetic parameters
of the distributed activation energy model, Bioresour. Technol. 102 (2011) 3894–3899.
[197] K. Miura, T. Maki, A simple method for estimating f(E) and k0 (E) in the distributed activation energy model,
Energy & Fuels. 12 (1998) 864–869.
[198] Sione Paea, Coal pyrolysis distribution, Wellington University, 2008.
[199] M.S. Mettler, S.H. Mushrif, A.D. Paulsen, A.D. Javadekar, D.G. Vlachos, P.J. Dauenhauer, Revealing pyrolysis
chemistry for biofuels production: Conversion of cellulose to furans and small oxygenates, Energy Environ. Sci.
5 (2012) 5414–5424
[200] A. Dieguez-Alonso, A. Anca-Couce, N. Zobel, F. Behrendt, Understanding the primary and secondary slow
pyrolysis mechanisms of holocellulose, lignin and wood with laser-induced fluorescence, Fuel. 153 (2015) 102–
109.
[201] P.R. Patwardhan, D.L. Dalluge, B.H. Shanks, R.C. Brown, Distinguishing primary and secondary reactions of
cellulose pyrolysis, Bioresour. Technol. 102 (2011) 5265–5269.
[202] B.J. Vreugdenhil, R.W.R. Zwart., Tar formation in pyrolysis and gasification, 2009.
[203] J. Hayashi, T. Shoji, S. Kudo, K. Norinaga, Detailed chemical kinetic modelling of vapour-phase cracking of
multi-component molecular mixtures derived from the fast pyrolysis of cellulose, Fuel. (2011).
[204] Michael L. Borosont, Jack B. Howard, John P. Longwell, William A. Peters, Heterogeneous cracking of wood
pyrolysis tars over fresh wood char surfaces, Energy & Fuels. 3 (1989) 735–740.
[205] C. Font Palma, Modelling of tar formation and evolution for biomass gasification: A review, Appl. Energy. 111
(2013) 129–141.
[206] Pradeep Ahuja, Surendra Kumar, Prem Chand Singh, A model for primary and heterogeneous secondary
reactions of wood pyrolysis, Chem. Eng. Technol. 19 (1996) 272–282.
[207] Elly Hoekstra, Roel J. M Westerhof, Wim Brilman, Wim P.M. Van Swaaij, Sascha R. A. Kersten, Kees J. A.
Hogendoorn, Heterogeneous and homogeneous reactions of pyrolysis vapors from pine wood, AIChE J. 58
(2012) 2830–2842.
[208] Hugh N. Stiles, Rafael Kandiyoti, Secondary reactions of flash pyrolysis tars measured in a fluidized
bed pyrolysis reactor with some novel design features, Fuel. 68 (1989) 275–282.
[209] Jun-ichiro Hayashi, Shinobu Amamoto, Katsuki Kusakabe, Shigeharu Morooka, Evaluation of vapor-phase
reactivity of primary tar produced by flash pyrolysis of coal, Energy & Fuels. 9 (1996) 290–294.
74 Chapter 1
[210] Eun-Jae Shin, Mark R.Nimlos, RobertJ. Evans, Kinetic analysis of the gas-phase pyrolysis of carbohydrates,
Fuel. 80 (2001) 1697–1709.
[211] R.G. Graham, M.A. Bergougnou, B.A. Freel, the kinetics of vapor-phase cellulose fast pyrolysis reactions,
Biomass Bioenergy. 7 (1994).
[212] Y. Oki, M. Ashizawa, S. Kajitani, Y. Zhang, Tar destruction and coke formation during rapid pyrolysis and
gasification of biomass in a drop-tube furnace, Fuel. 89 (2010) 302–309.
[213] Heidi Böhm, Helga Jander, PAH formation in acetylene–benzene pyrolysis, Phys. Chem. Chem. Phys. 1 (1999)
3775–3781.
[214] J. Moon, J. Lee, U. Lee, J. Hwang, Transient behavior of devolatilization and char reaction during steam
gasification of biomass., Bioresour. Technol. 133 (2013) 429–36.
[215] S. Ma, Y. Tan, Y. Han, Methanation of syngas over coral reef-like Ni/Al2O3 catalysts, J. Nat. Gas Chem. 20
(2011) 440–435.
[216] Paulsen, Primary and Secondary Reactions of Cellulose Melt Pyrolysis, University of Massachusetts, 2014.
[217] X. Bai, P. Johnston, R.C. Brown, An experimental study of the competing processes of evaporation and
polymerization of levoglucosan in cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 99 (2013) 130–136.
[218] H. Kawamoto, M. Murayama, S. Saka, Pyrolysis behavior of levoglucosan as an intermediate in cellulose
pyrolysis: Polymerization into polysaccharide as a key reaction to carbonized product formation, J. Wood Sci.
49 (2003) 469–473.
[219] T. Hosoya, H. Kawamoto, S. Saka, Different pyrolytic pathways of levoglucosan in vapor- and liquid/solid-
phases, J. Anal. Appl. Pyrolysis. 83 (2008) 64–70.
[220] Westerhof, H.S. Nygård, W.P.M. van Swaaij, S.R.A. Kersten, D.W.F. Brilman, Effect of Particle Geometry and
Microstructure on Fast Pyrolysis of Beech Wood, Energy & Fuels. 26 (2012) 2274–2280.
[221] Goring, Thermal Softening of Lignin, Hemicellulose and Cellulose, Pulp Pap. Mag. Canada. 64 (1963) T517--
T527.
[222] S. Nordin, J. Nyren, E. Back, An Indication of Molten Cellulose Produced in a Laser Beam, Text. Res. J. 44
(1974) 915–917.
[223] Y. Halpern, S. Patai.Israel, Pyrolytic reactions of carbohydrates.Part V: isothermal decomposition of
cellulose in vacuo, J. Chem. 7 (1996) 673–683.
[224] O.P. Golova, Chemical Effects of Heat on Cellulose, Russ. Chem. Rev. 44 (1975) 687–697.
[225] J. Diebold, J. Scahill, Ablative pyrolysis of macroparticles of biomass, in: Proceedings of Specialists’ Workshop
on fast Pyrolysis of Biomass, Copper Mountain, in: R.P. Overend, T.A. Milne, L.K. Mudge (Eds.), Fundam.
Thermochem. Biomass Convers., Springer Netherlands, Copper Mountain, 1985: pp. 539–555.
[226] J. Lédé, J. Panagopoulos, H.Z. Li, J. Villermaux, Fast pyrolysis of wood: direct measurement and study of
ablation rate, Fuel. 64 (1985) 1514–1520.
[227] G.B. Eijkel, J.J. Boon, P.W. Arisz, A.D. Pouwels, Evidence for oligomers in pyrolysates of microcrystalline
cellulose, J. Anal. Appl. Pyrolysis. 15 (1989) 71–84.
[228] J. Lédé, M. Ferrer, O. Boutin, Radiant flash pyrolysis of cellulose—Evidence for the formation of short life time
intermediate liquid species, J. Anal. Appl. Pyrolysis. 47 (1998) 13–31.
[229] J. Piskorz, D.S. Scott, D. Radlein, Mechanisms of the fast pyrolysis of biomass: comments on some
sources of confusion, Meet. Front. Pyrolysis Biomass Convers. Polym. Recycl. (1995).
[230] J. Piskorz, A. Vladars-Usas, D. Radlein, P. Majerski, D.. Scott, Flash pyrolysis of cellulose for production of
anhydro-oligomers, J. Anal. Appl. Pyrolysis. 56 (2000) 145–166.
[231] T. McGrath, V. Baliga, D. Miser, M. Hajaligol, R. Sharma, Physical characterization of pyrolyzed tobacco and
tobacco components, J. Anal. Appl. Pyrolysis. 66 (2003) 191–215.
[232] R. Bounaceur, B. Ouartassi, A. Dufour, A. Zoulalian, Modelling intra-particle phenomena of biomass pyrolysis,
Chem. Eng. Res. Des. 89 (2011) 2136–2146.
[233] S.W. Qian Liu, Kaige Wang, Xiujuan Guo, Zhongyang Luo, Kefa Cen, Mechanism of Formation and
Consequent Evolution of Active Cellulose during Cellulose Pyrolysis, Acta Physico-Chimica Sin. 24 (2008)
1957–1963.
[234] H.Z. Li, J. Villermaux, H. Martin, J. Lédé, Fusion-like behaviour of wood pyrolysis, J. Anal. Appl. Pyrolysis. 10
(1987) 291–308.
[235] M.K. Koch, A. Voßnacke, J. Starflinger, W. Schütz, H. Unger, Radionuclide re-entrainment at bubbling water
pool surfaces, J. Aerosol Sci. 31 (2000) 1015–1028.
Chapter 1 75
[236] K. Butler, A numerical model for combustion of bubbling theroplastic materials in microgravity, (2002) 70.
http://fire.nist.gov/bfrlpubs/fire02/PDF/f02002.pdf.
[237] M. Shojaeian, A. Koşar, Pool boiling and flow boiling on micro- and nanostructured surfaces, Exp. Therm. Fluid
Sci. 63 (2015) 45–73.
[238] K. Voorhees, D.M. Blazsó, D.C. Schwarzinger, P. Carlsson, H. Lycksam, P. Gren, R. Gebart, H. Wiinikka, K.
Iisa, High-speed imaging of biomass particles heated with a laser, J. Anal. Appl. Pyrolysis. 103 (2013) 278–
286.
[239] F. Yi, G. Zhu, X. Zhu, R. Zhou, Z. Xiao, Pyrolysis characteristics of bean dregs and in situ visualization of
pyrolysis transformation, Waste Manag. (2012).
[240] Zhouhong Wanga, Armando G. McDonald, Roel J.M. Westerhofc, Sascha R.A. Kersten, Christian M.
Cuba-Torres, Su Had, Brennan Pechad, Manuel Garcia-Perez, Effect of cellulose crystallinity on the
formation of a liquid intermediate and on product distribution during pyrolysis, J. Anal. Appl. Pyrolysis.
100 (2013) 56–66.
[241] Westerhof, S.R.G. Oudenshoven, P.S. Marathe, M. Englen, M. Garcia-Pérez, Z. Wang, S.R.A. Kersten, The
Interplay between Chemistry and Heat/Mass Transfer during Fast Pyrolysis of Cellulose, Submitt. to Energy
Environ. Sci. (2016).
[242] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials, Curr. Opin.
Biotechnol. 24 (2013) 414–420.
[243] J. Ratte, F. Marias, J. Vaxelaire, P. Bernada, Mathematical modelling of slow pyrolysis of a particle of treated
wood waste., J. Hazard. Mater. 170 (2009) 1023–1040. doi:10.1016/j.jhazmat.2009.05.077.
[244] N. Bech, P.A. Jensen, K. Dam-Johansen, M.B. Larsen, Modelling solid-convective flash pyrolysis of straw and
wood in the Pyrolysis Centrifuge Reactor, Biomass and Bioenergy. 33 (2009) 999–1011.
[245] P.K. Agarwal, Transport phenomena in multi-particle systems—IV. Heat transfer to a large freely moving
particle in gas fluidized bed of smaller particles, Chem. Eng. Sci. 46 (1991) 1115–1127.
[246] B. Moghtaderi, V.N.D. Fletcher, J.H. Kent, An Integral Model for the Transient Pyrolysis of Solid Materials, FIRE
Mater. 21 (1997) 7–16.
[247] C.H. Bamford, J. Crank, D.H. Malan, A.H. Wilson, The combustion of wood. Part I, Math. Proc. Cambridge
Philos. Soc. 42 (1946) 166.
[248] A.M. Kanury, P.L. Blackshear, On the Combustion of Wood II: The Influence of Internal Convection on the
Transient Pyrolysis of Cellulose, Combust. Sci. Technol. 2 (1970) 5–9.
[249] H.-C. Kung, A mathematical model of wood pyrolysis, Combust. Flame. 18 (1972) 185–195.
[250] P.S. Maa, R.C. Bailie, Influence of Particle Sizes and Environmental Conditions on High Temperature Pyrolysis
of Cellulosic Material—I (Theoretical), Combust. Sci. Technol. 7 (1973) 257–269.
[251] T. Fan, Liang-Shih Fan, Kei Miyanami, A mathematical model for pyrolysis of a solid particle: Effects of the
lewis number, Can. J. Chem. Eng. 55 (1977) 47–53.
[252] E.J. Kansa, H.E. Perlee, R.F. Chaiken, Mathematical model of wood pyrolysis including internal forced
convection, Combust. Flame. 29 (1977) 311–324.
[253] H.A. Becker, A.M. Phillips∗, Pyrolysis of white pine, Combust. Flame. 58 (1984) 163–189.
[254] D.L. Pyle, C.A. Zaror, Heat transfer and kinetics in the low temperature pyrolysis of solids, Chem. Eng. Sci. 39
(1984) 147–158.
[255] M.J.J. Antal, Mathematical modelling of biomass pyrolysis phenomena: Introduction, Fuel. 64 (1985) 1483–
1486.
[256] B. ChanW-CR,Kelbon Mand Krieger, Product Formation in the Pyrolysis of Large Wood Particles, in: Fundam.
Thermochem. Biomass Convers., Elsevier Applied Science Publishers, New York, 1985: pp. 219–236.
[257] S.S. Alves, J.L. Figueiredo, A model for pyrolysis of wet wood, Chem. Eng. Sci. 44 (1989) 2861–2869.
[258] C. Di Blasi, Heat transfer mechanisms and multi-step kinetics in the ablative pyrolysis of cellulose, Chem. Eng.
Sci. 51 (1996) 2211–2220.
[259] M.A. Hastaoglu, A gas-solid reaction model for flash wood pyrolysis, Fuel. 68 (1989) 1408–1415.
[260] Richard, J. Saastamoinen, Simultaneous drying and pyrolysis of solid fuel particles, Combust. Flame. 106
(1996) 288–300.
[261] A. Sharma, V. Pareek, S. Wang, Z. Zhang, H. Yang, D. Zhang, A phenomenological model of the mechanism
of lignocellulosic biomass pyrolysis processes, Comput. Chem. Eng. 60 (2014) 231–241.
[262] C. Di Blasi, Physico-chemical processes occurring inside a degrading two-dimensional anisotropic porous
76 Chapter 1
medium, Int. J. Heat Mass Transf. 41 (1998) 4139–4150.
[263] Michel Bellais, Modelling of the pyrolysis of large wood particles, KTH - Royal Institute of Technology
Stockholm, 2007.
[264] J.P. Alex D. Paulsen, Blake R. Hough, C. Luke Williams, Andrew R. Teixeira, Daniel T. Schwartz, Fast Pyrolysis
of Wood for Biofuels: Spatiotemporally Resolved Diffuse Reflectance In situ Spectroscopy of Particles,
ChemSusChem. 7 (2014) 765–777.
[265] E. Leonardi, G. de Vahl Davis, R.K.K. Yuen, G.H. Yeoh, Modelling the pyrolysis of wet wood – I. Three-
dimensional formulation and analysis, Int. J. Heat Mass Transf. 50 (2007) 4371–4386.
[266] A.O. Ahmed Mahmoudi, Imen Mejri, Mohamed A. Abbassi, Lattice Boltzmann Simulation of the Carbonization
of Wood Particle, Int. J. Math. Comput. Phys. Quantum Eng. 8 (2014) 668–677.
[267] P.N. Ciesielski, M.F. Crowley, M.R. Nimlos, A.W. Sanders, G.M. Wiggins, D. Robichaud, B.S. Donohoe, T.D.
Foust, Biomass particle models with realistic morphology and resolved microstructure for simulations of
intraparticle transport phenomena, Energy and Fuels. 29 (2015) 242–254.
[268] Myongsook S., Oh William A., Peters Jack B. Howard, An experimental and modeling study of softening coal
pyrolysis, AIChE J. 35 (1989).
[269] Yusuf Sezen, Internal mass transfer considerations during the pyroly sis of an isolated spherical softening
coal particle, Lnt. J. Heat Mass Transf. 32 (1989) 1992–1987.
Chapter 2 77

2. Chapter 2. Identification of the Fractions Responsible for


Morphology Conservation in Lignocellulosic Pyrolysis:
Visualization Studies of Sugarcane Bagasse and its
Pseudo-Components
(Paper published in Journal of Analytical and Applied Pyrolysis)

Abstract: Pyrolysis of lignocellulosic materials typically form carbonaceous residues (chars) with structures similar to
the starting material. However, previous studies reported in the literature have shown that fast pyrolysis of cellulose
and lignin form a molten intermediate liquid resulting in chars with smooth surfaces and evidences of intense bubbling.
The fraction responsible for morphology conservation during pyrolysis is not known. In this article, pyrolysis visualization
studies were carried out with a fast camera on cellulose, xylan, Organosolv lignin (OL), milled wood lignin (MWL), and
lignocellulosic sugarcane bagasse to identify the component responsible for morphological conservation. The materials
were hearted using a modified Pyroprobe with measured sample heating rates close to 100 oC/s. This studies confirm
that cellulose, lignin, and deashed xylan are completely transformed into an intermediate liquid; the intensity of bubbling
for lignin was far superior that of cellulose. Xylan as received, however, only forms a small amount of intermediate
liquid; it does not fully melt and maintains its general shape during pyrolysis. These results suggest that the presence
of mineral matter is critical for lignocellulosic materials to retain their original morphology. It was repeated the studies
on raw bagasse and acid washed bagasse. The raw bagasse retains its morphology; the acid washed bagasse shrank
dramatically, but was not completely converted into a liquid. This results suggest that in addition to hemicellulose with
the presence of ash, other biomass fractions or the different temperatures at which the different fractions melt and
solidify could also be contributing to maintain the structure of the lignocellulosic material during pyrolysis.

Keywords: Fast pyrolysis, intermediate liquid phase, hemicellulose, Pyroprobe, visualization


78 Chapter 2

2.1. Introduction
Many physical and chemical phenomena occur simultaneously during thermochemical processes like pyrolysis,
gasification, and combustion, making it difficult to accurately predict the product distribution [1]. It has been recently
shown that during the pyrolysis of lignocellulosic components, a liquid intermediate phase is formed as cellulose and
lignin depolymerize. As such, this liquid phase is mostly composed of heavy oligomeric compounds with boiling points
higher than the temperatures of the depolymerization reactions [2–6]. The formation of excessive liquid intermediate is
problematic for the operation of fluidized bed reactors as it can lead to bed defluidization due to agglomeration of
particles [7]. In pyrolysis, the heavy components in the liquid intermediate can be thermally ejected in the form aerosols
through micro-explosions [5, 8, 9]. This phenomenon explains the presence of oligomeric products in bio-oil, which are
typically not well-characterized. Within the liquid phase many physiochemical phenomena occur simultaneously: solid
/ liquid reactions, mass and heat transfer in the liquid-solid-gas, bubbling (nucleation, coalescence, death, and growth),
and others which have not yet been identified. These reactions and phenomena are controlled by the type of material
studied, the final temperature to which the particle is heated, the particle size, and pressure [10]. Although there are
many parametric studies on the effects of these variables on the distribution of products [11–15], very little literature
exists regarding the effect of biomass composition on the formation and behavior of resulting liquid intermediate.

It is difficult to understand these phenomena mainly due to short reaction times in fast pyrolysis (< 1s) and limitations
with real-time sampling systems [16]. Some authors have used SEM, TEM [17–20] and other optical techniques [17,
20] to visualize the changes in the internal structure of particles undergoing pyrolysis. This information has been
correlated with surface area analysis [21, 22] to provide quantitative information on changes of the porous structure
before and after pyrolysis.

Two interesting studies on the morphological changes of biomass (Miscanthus) and its pseudo-components during
pyrolysis have been published by Dufour, et al. [23, 24]. In the first, the authors used 1H-MNR, to monitor changes to
each pseudo component during devolatilization with slow heating rates. The authors concluded that the lignin is
completely fluid between 200 and 225 °C, before the start of the mass loss. Hemicellulose has medium fluidity of 61%
between 200 and 275 °C, and cellulose only achieved 40% fluidity between 300 - 350 °C. In the second study, the
research extended to visco-elastic behavior pseudo components in a parallel plate reactor (viscometer). The authors
show that lignin has an elastic behavior between 150 - 350 °C; after this, re-solidification of the material (char
production) begins. Likewise, for cellulose between 250-340 °C a high viscosity liquid forms with the most fluidity at 340
°C. For xylan, the fluid phase forms between 200 - 250 °C, then re-solidifies at 270 °C. These studies were conducted
at low heating rates. Wang et al. [26,27] tested SEM images at low heating rates on cellulose. Under low heating rate
amorphous cellulose forms a liquid intermediate while crystalline cellulose forms a solid. Under high heating rates both
amorphous and crystalline cellulose forms a liquid intermediate [28–30].

The final morphology of the biomasses during fast pyrolysis can be affected by the intensity of two competitive
processes after the depolymerization reaction: oligomer evaporation and crosslinking reactions in the liquid phase.
Typically, it has been reported that at high heating rates (> 103 °C/s), a small biomass particle rapidly reaches high
temperatures where the evaporation of oligomers (sugars, anhydrosugars wit DP < 7, and pyrolytic lignin) are promoted,
leaving the solid matrix and avoiding carbonization reactions. However at low heating rates, the intermediate liquid
compounds, have more time inside the particle to react by crosslinking reaction in liquid phase. This liquid could act as
an electrolyte, catalyzing solid-liquid reactions that produce char [31,32]. This kind of behavior has been specially
proposed for pure cellulose [8,33–35], however it could be extrapolated to other biomass constituents. Dufour et al.
[24,25] have shown that during thermal degradation each biomass component passes through a thermoplastic state,
where visco-elastic properties depend strongly on the structure. The thermoplastic behavior of biomasses cannot be
Chapter 2 79
correlated like a linear combination of the behaviors of its components. At slow heating rates, melting on the char
surface competes to secondary cross-linking reactions. At high heating rates, biomass behaves similarly to intumescent
coals, with significant chemical bond breaking before the cross-linking starts to be important; its behavior is more like a
fluid phase [32,36].

The interactions of different components with each other and with mineral matter, specially alkali and alkali earth metals,
produce rigid structures due to cross-linking and dehydration reactions [37, 38]. Shafizadeh, et al. [39], suggest that
hemicellulose has a strong effect on maintaining of solid morphology, specially by association of minerals through ion
exchanges with acid functional groups belongs to pyran rings. In general the morphological changes in biomass during
fast pyrolysis are due to multiple effects such as heating rate, mineral matter composition, particle size, and
lignocellulosic composition.

Real-time visualization studies [5, 20, 24, 30, 31], show tissue level changes including the expansion of cell walls and
the appearance of a yellow resinous substance that fills the porous cavities as the reaction proceeds. It has been shown
[9, 42–44] that this substance is mainly constituted by oligomers. Dauenhauer, et al, [5, 31, 35, 36] using a torch
reactor reported the appearance of an ephemeral liquid intermediate (order of ms) during pyrolysis of cellulose, where
the appearance of microbubbles is evident. Bubbles are formed as products of decomposition reactions in the liquid.
These bubbles burst at the gas-liquid interface, promoting the ejection of aerosols (reactive boiling ejection). Their
visualization studies confirm that under high heating rates cellulose is fully converted into an intermediate fluid between
300 - 350 °C [5]. Studies with lignin are scarce, but literature confirms that this pseudo-component also forms a
intermediate liquid and the microexplosion intensity is much more violent than for cellulose [47].

Nevertheless, studies of 1H-NMR, rheological, and visualization have not fully explained why lignocellulosic materials
result in the formation of chars with a similar structure to the starting material [21, 24, 25]. For cellulose, 13C NMR
experiments coupled with computational chemistry tell us that until 300 °C, dehydration reactions create ether bonds
between cellulose chains. Between 300 and 400 °C, rings start to form, and above 400 °C aromatic ring formation
dominates [48]. It is critical to note that the depolymerization reactions occur in the same temperature range (400-500
°C), so there is a definitive competition between depolymerization and cross-linking reactions in cellulose pyrolysis;
future work may prove the same competition in hemicellulose and lignin as well.

In this article, it was examined through high speed visualization techniques the formation of the intermediate liquid
phase in cane bagasse and model pseudo-components during fast pyrolysis (heating rates> 100 ° C/s). The goal is to
identify the fraction responsible for the conservation of the morphology when biomass is pyrolyzed. This results also
provide insight into the bubbling phenomena and structural morphology of lignocellulose undergoing fast pyrolysis.

2.2. Materials and Methods


Morphological changes during pyrolysis of avicel cellulose, organosolv lignin, ball mil lignin, xylan, deashed xylan,
sugarcane bagasse and acid washed sugarcane bagasse were explored by fastcam visualization.

2.2.1. Materials and preparation

Sugarcane bagasse was obtained from the sugarcane mill Risaralda (Km 2 via La Virginia-Balboa, Risaralda-Colombia,
4°54'22.42"N, 75°53'45.46"O). The material was analyzed as received (proximate and elemental analysis, ash
composition and content of pseudo-components) and was then acid washed to reduce the content of alkalines and a
fraction of hemicellulose. For acid washing, sugarcane bagasse was soaked in 1.76 wt.% sulfuric acid solution (15 mL
80 Chapter 2
solution /g solid) at 152.6 oC for 21 minutes [49]. After acid wash, the solids were rinsed and centrifuged with nano-pure
water four times to remove residual H2SO4.

The model components used were avicel cellulose (Fluka 11365, Avicel PH-101,50μm, CAS 9004-34-6, with a degree
of crystallinity of 60%), organosolv lignin (371,017 Aldrich, CAS 8068-03-9) and beech wood xylan (Sigma Aldrich CAS
9014-63-5)). All samples were reduced to particle sizes less than 100 µm. The physicochemical characterization of
these materials is presented in Table 2. Xylan was subjected to acid washing using 1 M HCl, sonicated for 1 hr at 25
oC, then washed and centrifuged with water four times until the acid was removed, as verified using a silver nitrate

indicator.

The milled wood lignin herein studied was obtained from hybrid poplar heart wood using a method described elsewhere
[50–52]. Briefly, the wood was extracted with 9:1 acetone: water followed by 2:1 ethanol/benzene for 8 h each in a
Soxhlet extractor. The solids were oven dried at 45 oC, then ball milled (Planetary ball mill, Across International) for 50
h, 15 min on 15 min rest, at 600 rpm with ceramic balls and cup. Lignin dissolution was performed in dioxane: water
(DW) (96:4, 10 mL DW/g wood) for 24 h AT 50 oC. This was allowed to settle, then centrifuged and liquid was transferred
to a round bottom flask to evaporate dioxane. The residue in the evaporation flask was dissolved in acetic acid/water
(9:1) (20 mL/g) and added gradually into a 10 times larger volume of 4 oC distilled water directly into 6 250 mL centrifuge
tubes. Let sit in refrigerator. The acetic acid/water is removed by centrifugation (10 min 10,000 rpm, 4 oC). Decant and
discard liquids. Remaining acetic acid was removed by adding ethanol and rotovaping 3 times. This is crude MWL
(~10% carbohydrates). Dissolve crude lignin in dichloroethane: ethanol (2:1, v/v, minimal amount required for
dissolution ~20mL/g lignin). Drop into anhydrous diethyl ether (10 times volume of dichloroethane/ethanol) to precipitate
lignin. Centrifuge (10 min 10,000 rpm, 4 oC), decant ether. Wash 3 times with ether. Remove residual ether under
vacuum. This is MWL typically contain ~4% carbohydrate. The elemental and proximate analysis of all the materials
studied are shown in Table 2.

2.2.2. Characterization of materials

The lignocellulosic composition for bagasse was determined following the protocol NREL TP-510-42618, with results
shown in Table 1. The elemental analysis followed the manufacturer’s recommendation for biomass samples [53].
Proximate analysis followed ASTM D7582 – 15 [54], coupled with modifications for smaller sample size in a Mettler
Toledo TGA/SDTA851e. [55]. The results obtained are shown in Tables 2 and 3.

Metal analysis was conducted for the Douglas fir in an ICP-MS (Agilent 7500cx) following the same procedure described
by Pecha et al [56]. As it is seen in this table the content of catalytic metals is quite low; potassium, for example, only
has reported significant catalytic activity above 1% w / w [57,58]. An important fraction of the ash is silica.

Table 1. Composition of sugarcane bagasse as received and after acid wash on a dry ash-free basis. The sums slightly
exceed 100% due to experimental error. The sugarcane bagasse before acid washing has a much higher extractives
content than after acid washing, which contributes to relative changes in the glucan, xylan, arabinose, and lignin content.

Sugar cane bagasse as Acid washed sugarcane


received (wt.% DAF) bagasse (wt.% DAF)
Glucan 31.0 51.3
Xylan 19.2 18.8
Arabinose 2.5 0.00
Chapter 2 81
Lignin* 19.1 25.5
Extractives 29.3 5.8
* Lignin content was not nitrogen corrected.

Table 2. Physico-chemical characterization of the raw material. Moisture content is on as-received basis. Volatiles,
fixed carbon, ash, glucan, arabinose, lignin, and extractives are reported on a dry basis (w/w). CHN-O data is reported
on a dry, ash-free basis (DAF) [wt. %].

Sugarcane Acid- Avicel Xylan De-Ashed Organosolv Milled


bagasse Washed Cellulose Xylan Lignin Wood
sugarcane Lignin
bagasse
Moisture 5.25 2.71 3.65 7.95 6.48 1.19 2.70
Volatiles (dry) 82.6 86.8 87.7 71.9 82.6 70.6 75.8
Fixed Carbon 8.30 5.99 8.35 14.3 17.4 26.2 24.05
(dry)
Ash (dry) 3.90 4.53 0.314 5.85 0.0112 2.02 0.205
C (DAF) 46.6 49.3 44.4 43.0 50.4 69.7 60.2
H (DAF) 5.92 5.90 6.49 6.45 8.29 5.86 5.90
N (DAF) 0.14 0.0497 0.064 0.0074 0.0 0.11 0.050
O (DAF) 43.5 31.4 45.5 37.5 35.4 21.3 31.4

Table 3. Metal species analysis by ICP-MS for sugarcane bagasse and xylan before and after their respective acid
washings. The acid washing reduces the metal content for all samples. Notably, the Na content in xylan is reduced form
14 mg/g to sub-detection limit concentrations and K to near detection limits (mg/g).

Sugar cane Acid washed


Acid washed
Species (unit) bagasse as sugarcane Xylan
xylan
received bagasse*
Na (mg/g) - 0.32 14.2 -
Mg (mg/g) 0.30 - - -
K (mg/g) 2.32 0.03 20.3 0.02
Ca (mg/g) 2.18 0.54 2.17 0.37
Fe (mg/g) 0.40 0.00 0.15 0.06
Al (µg/g) 197. 216. 5.60 -
Zn (µg/g) 54.6 - 25.0 -
Cu (µg/g) 14.7 1.83 4.76 -
Ba (µg/g) 18.1 15.8 8.33 -
Pb (µg/g) 5.89 0.41 2.33 -
Mo (µg/g) 30.99 3.29 0.94 -
Mn (µg/g) 63.1 0.30 66.4 0.26
Ni (µg/g) 1.98 8.23 1.09 -
Sum of ICP analyzed
4.48 1.14 37.0 0.452
metals (mg/g)
82 Chapter 2
Silicates (calculated by
difference from ash 34.5 44.2
content) (mg/g)
*This had a small sample size, which increases error (such as sodium content).

2.2.3. Modified Pyroprobe pyrolysis visualization reactor

The formation of the intermediate liquid phase for both bagasse and the pseudo components studied, was visualized
with fast speed photography. The materials were placed on quartz discs which rested on the platinum wire resistor in
a modified Pyroprobe 5000 (CDS Analytical) described [59]. Briefly, the reactor chamber is a glass tube (Pyrex 9825)
open at both ends, screwed into a solid block of aluminum with a valve to regulate the continuous flow of nitrogen. The
display of the samples was performed by a fast speed camera (Troubleshooter LE Imaging Fastec) at 125 fps, using
two 25 mm extension tubes each connected in series with a macro lens (Tokina, 100 mm f 2.8). The camera was
positioned maintaining an approximate 45 ° angle to the main axis of the reactor and perpendicular thereto (see Figure
1).

Figure 1. Visualization setup for studying fast pyrolysis.

In the above picture, the quartz disc with sample is inserted within a glass pipe housing (D), and hold on a flat platinum
wire as heating element (E). In order to enhance the resolution of each video an external light source (B) is used,
lighting the housing surface. The videos were recorded with a slow motion camera (A) coupled with an array of macro
lens (Tokina 100 f 2.8) and 2 extensions tubes in order to reach magnifications until 100 times the original size. The
temperature and heating rate of the platinum wire were fixed by a Pyroprobe controller box (C). The sample temperature
was calibrated using a thin thermocouple inserted within the sample pile (see fig 3 in chapter 3). More details about the
reactor can be readed in chapter 3. For visualization, a Fastec-Sportscam 1000 ME (Fastec Imaging Corporation, San
Diego, CA) was used. An external source of white light (Solarc LB50, Ushio) was used for supplemental lighting. ImajeJ
and the video editing toolbox in MatLab R2014 were used to process the videos. The amount of sample used in each
experiment is approximately 0.5 mg with particle sizes <100 µm, positioned in the center of the quartz disc. Nitrogen
continuously flows at 5 ml / min to prevent fumes from blocking the field of view. In all experiments, pixel size calibration
was made considering the platinum filament diameter was 0.3556 mm.
Chapter 2 83

2.2.4. Temperature Calibration

The temperature was measured for both touching the sample-free quartz disc surface as well as inside the 0.5 mg
sample piles, but not actually touching the disc surface. Each measurement was performed in triplicate. The Pyroprobe
nominal heating rate set point was 800 °C/s with final temperature set point of 337 °C to reach a measured disc
temperature of 500 °C. The calibration was done with an Omega thermocouple (40 AWG), which has a response time
of 2.36 ms. A National Instruments USB-TC01 data acquisition device sends a record of the temperature every 0.25 s
in the MatLab for processing. The original shape of the platinum wire is a coil, and for this experimental setup was
necessary to modify the geometry by a flat coil, changing the electrical properties (resistance). For this reason when
the set point in the control box (C) is 337 °C the temperature on quatz disc is 500 °C. All temperatures in each sample
were calibrated fixing in 337 °C the control box, in order to keep the same initial conditions in each experiment.

Figure 2 shows the measurements of the temperatures of the blank disc and the samples. The maximum heating rate
for the blank disc was 181 °C / s, and average biomass and pseudo components were the order of 120 °C / s. In Figure
8, it can be seen that the final temperatures of most materials are similar (~ 380 ° C for lignin and cellulose, ~ 340 ° C
for biomass and Xylan). For other materials, differences in the final temperature correspond different physicochemical
properties of carbonaceous residues generated. At the end of each test the correlation between the temperature of the
sample and the image corresponding to each time point is to make the phenomenological description of each event
pyrolysis.

Figure 2. Calibration curves of the various components and the blank disc.

The behavior of milled wood lignin is notably different from the others. Unlike the organosolv lignin, the primary structure
of the MWL is not degraded and is quite similar to how it is in the native biomass [60,61]. MWL has been shown to have
more β-O-4 linkages than organosolv lignin [60]. These bonds are weak and are more likely to be broken at lower
temperatures. Alternatively, the Ball mill lignin possibly formed a char with a lower thermal conductivity, which results
in lower temperatures at the top surface of the char. Both explanations are simply hypotheses, but neither discredit the
experimental results that show that the MWL had a lower temperature.

The differences in final temperatures between organosolv lignin and MWL are likely due to Organosolv lignin forming a
final product with lower thermal conductivity; thus the heat transfer to surroundings is lower than for MWL char.
Visualization experiments (Figures 3 and 4) and SEM micrographs (Figure 10) bolster this hypothesis: the Ball mill lignin
produces bigger bubbles more intensely than MWL, and the final char has a porous structure with more voids that will
reduce the final thermal conductivity and heat transfer with surroundings.
84 Chapter 2

2.2.5. SEM analyzes of solid residues

SEM imaging was performed to analyze the char residue from all of the samples after pyrolysis on the quartz disc
surface as described above. Discs with solid residues were first sputter coated with 3 nm thick platinum/palladium
(Cressington Sputter Coater) followed by SEM imaging using an FEI Quanta 200F SEM at the Franceschi Microscopy
and Imaging Center at Washington State University.
Chapter 2 85

2.3. Results and discussion

2.3.1. Behavior of organosolv lignin during pyrolysis

The first lignocellulose pseudo-component to be discussed is lignin. Figure 3 shows the sequence of video snapshots
taken when Organosolv lignin was heated in the pyrolysis visualization reactor. The appearance of molten material can
be seen 1.88 s after the experiment begins. The temperature at which melting of the sample initiates is 176.9 °C (see
Frame A). This observation is close to results reported in the literature (150-154 °C) [25,62]. At 2.34 s the lignin is
partially melted. The portion of the sample that is in direct contact with the disc is completely melted, leaving the top
unchanged. The height of the lignin pile, estimated by image processing, is 0.31 mm, indicating that the process is
controlled by the heat transfer (“thermally thick”) [63].

The sample is 100% melted at 2.74 s, when the average temperature of lignin is 241.75 °C. A slight swelling of the pile
can be seen, as well as the appearance of small gas bubbles in the liquid (bubble diameter ~33 µm). The appearance
of these micro bubbles is indicative of vapor formation within the liquid phase via fragmentation reactions of β-O bonds
to form phenols, alkoxy-ketones, and light gases [64].

This behavior has been reported in previous studies by Zhou et al. [48, 49]. According to Dufour et al. [24,25], at this
temperature the mass loss is less than 12%; this is consistent with the video between 1.88 and 2.74 s, where no
significant outgassing is observed. When the sample exceeds 253 °C (2.86 s), the now continuous liquid slowly swells
and bubbles grow at the surface of the quartz disk, though the release of vapors is not clearly observable in the
videos. This indicates that decomposition reactions primarily are occurring in the liquid phase, but the bubbles do not
achieve sufficient energy to explode yet.

At 3.74 s (295 °C), growth of gas bubbles is continuous and becomes more intense, especially in areas close to the
disk surface. Here the gas formation rate has increased and they also overcome the barrier of the surface tension and
release the vapors via explosion (popping). The pile appears to swell because of coalescence of all the molten
material. In the video a white trail of gas can be seen, which corresponds to the release of vapors by the bubbles. Dufour
et al. [25] show that in this temperature range, both the viscous modulus and elasticity of lignin increase. This indicates
that the liquid phase begins to gain stiffness. Chemically, this likely reflects dehydration and crosslinking reactions which
produce char.

For temperatures near 300 °C, bubbling in the liquid phase is intense; there is continuous nucleation, bubble
coalescence, and breakup. Fluid expansion fills the entire surface of the quartz disc with film as result of viscous bubble
explosion and retraction. This dynamic continues until the sample reaches 342 °C (5.12s). This is a transition region
where the bubbling pattern has changed due to increases in viscosity, surface tension, and other properties in the liquid
phase.
A. Frame 236 Time=1.88s Tdisc=221°C B. Frame 293 Time=2.34s Tdisc=288°C C. Frame 343 Time=2.74s Tdisc=313°C
Tsample=176.9°C Tsample=223.39°C Tsample=241.75°C

Lignin partially melted Lignin completly melted

Raw lignin

D. Frame 358 Time=2.86 Tdisc=318°C


E. Frame 467 Time=3.736 Tdisc=355°C Tsample=295°C F. Frame 547 Time=4.37s Tdisc=381°C Tsample=314.7°C
Tsample=253.69°C

Lignin swelling Bubbles formation Bubbing intensification

G. Frame 640 Time=5.12s Tdisc=417°C H. Frame 980 Time=7.84s Tdisc=474°C


I. Frame 1245 Time=9.96s Tdisc=499°C Tsample=384°C
Tsample=342.8°C Tsample=377.5°C
Chapter 2 87

Bubbling diminution Liquid shrink Viscous liquid Thin char layer

Figure 3. Organosolv lignin melting and bubbling behavior during pyrolysis: still frames from video recordings.
88 Chapter 2

2.3.2. Behavior of milled wood lignin during pyrolysis

Although Organosolv lignin is commonly used by researchers because it is easy to purchase on the market, it is
generally accepted that milled wood lignin (MWL) is less degraded during the extraction process. The image sequence
for videos of pyrolysis of milled wood lignin (MWL) can be seen in Figure 4. Evidence of a liquid intermediate can be
seen as soon as 129 oC (4A) and the sample appears fully melted by 166 oC (4B). Structurally, MWL contains more β-
O-4 linkages than Organosolv lignin; these bonds are known to break at a lower temperature [60,66]. Other researchers
have reported similar results in slow-heating thermogravimetric analysis. Yang et al. [61] and Jakab et al. [67] found
that MWL decomposes in two regimes. The first, between 111 and 200 oC, corresponds to water and low molecular
weight gas release. The second regime, between 200 and 350 oC, corresponds to decarboxylation reactions which
release monomers. Mamdouh et al. [68] found that below 200 oC, endothermic phenomena are decomposition of weak
ether bonds, solid softening, melting, and evaporation of water. In Figures 4A and 4B, small bubbles (~24 µm) can be
seen, which indicate decomposition and evaporation inside the liquid phase. Between 166 and 230 oC (4C), a darkening
of the liquid intermediate phase can be seen. Other researchers have reported exothermic reactions at these
temperatures [69]. Here, the liquid intermediate swells due to expansion of gases and liquids. The bubbles are very
small and bubble bursting cannot be visualized in these recordings. The MWL has a broad molecular weight distribution
(400-5000 MW) [61, 70, 71] which means that the liquid phase should be quite viscous, limiting bubble coalescence
and transport [72, 73].

A. Frame 437 Time=3.49s B. Frame 488 Time=3.90s C. Frame 651 Time=5.20s


Tsample 129.5°C Tsample=166°C Tsample=230°C

Raw lignin Micro-bubbles within Liquid swelling and


Intermediate liquid phase darkening

Lignin melted
D. Frame 800 Time=6.4s E. Frame 954 Time=7.63s F. Frame 1141 Time=9.12s
Tsample=260°C Tsample=278°C Tsample=298°C

Liquid shrinking Liquid shrinking Liquid resolidification


and darkening

Figure 4. Milled wood lignin melting and bubbling behavior during pyrolysis: still frames from video recordings.

Because the liquid phase is very dark, it impedes observation of bubbles in the interior. In frames 837-838 and 872-878
of video B in the supplementary material, the collapse of a large bubble that covers the entire volume of the liquid phase
can be seen. A bubble of this size indicates a high rate of bubble coalescence which forms one large bubble that grows
Chapter 2 89
until the internal energy is sufficient to overcome the viscous and surface energy barrier to explode [74, 75]. After 260
oC (4E), the liquid phase swelling slows down, indicating a decrease in gas generation and an increase in viscosity by

polymerization reactions [24, 25]. Finally, after 298 o C (4F), volatile and gas release ceases and swelling and shrinking
cannot be observed because the carbon residue has formed. These observations match those obtained in previous
works [62].

2.3.3. Behavior of xylan (as received) during pyrolysis

Xylan, one of the many hemicelluloses in biomass, can be easily purchased from chemical vendors. Figure 5 shows
the behavior of xylan during pyrolysis. In the first 3.6 seconds of the reaction, when the sample has reached 231 °C,
no morphological change is apparent and there is no evidence of vapor release (see Figure 5 B). The same behavior
continues until 4.23 s when the surface of the xylan begins to darken. Darkening without formation of intermediate fluid
is a clear indication of solid phase reactions. At this time, the sample temperature reaches 280 °C. It has been reported
that the hemicellulose begins to decompose at 273 °C [76–78]. The xylan surface has completely darkened when the
sample reaches 323 °C. At this temperature cracking and material mass loss is observed as the sample shrinks a bit,
although there was no visualization of vapor release. Critically, the sample does not completely melt as with lignin and
cellulose. However, there is a liquid film just in contact with the quartz disk; due to optical properties and the low rate of
formation, it is difficult to distinguish through the images. With very careful analysis between frames 600-1000 (Video
C), the liquid at the heated surface is more evident.

Until 323 °C, not much change is seen from the original shape of the sample, except for some cracks in the central
area. The bubbles formed within the liquid phase are very small and are not possible to track with the camera due to
the low resolution (Figure 5, Frames D-F).

When the xylan has reached 366 °C, some swelling of the liquid phase can be seen throughout almost all the volume
of the sample (see videos in supplementary section). However, bubbling is minimal, and it is not possible to see any
coalescence or rupturing of them. Vapor formation is difficult to see in the videos. One explanation for this behavior is
that dehydration and crosslinking reactions between the different branches of xylan favor the production of char with
very low outgassing.
90 Chapter 2
A. Frame 341 Time 2.78s Tdisc=323°C B. Frame 450 Time=3.6s Tdisc=360°C. C. Frame 671 Time=5.36s Tdisc=430°C
Tsample= 188°C Tsample=231°C. Tsample=323°C

Crakcs formation

D. Frame 785 Time=6.28s Tdisc=460°C E. Frame 1051 Time=8.4s F. Frame 1500. Time=12s
Tsample=348°C Tdisc=493°C Tsample=366°C Tdisc 495°C Tsample=368°C

Cracks intensification and


Solid darkening Solid darkening
solid darkening

Figure 5. Xylan melting and bubbling behavior during pyrolysis: still frames from video recordings.

It is known that in the pyrolysis of cellulose, breakage of glycosidic bonds produces levoglucosan, by stabilizing the
cation at carbon 1, with the hydroxyl group on carbon 6. However for xylan, the nonexistence sixth carbon precludes
this chemical route; the carbon cation is rather stabilized by a dehydration reaction, generated water, char, and
furans. After 366 °C, swelling or shrinking of the liquid phase is not observed. Instead, the material resolidifies and no
changes can be seen after 1 minute. In the end the char retains most of the initial structure of xylan, though there was
a 36.3% shrinking of the surface area of the material.

2.3.4. Behavior of de-ashed xylan during pyrolysis

It is generally accepted that the mineral content has a strong influence on the behavior of biomass during pyrolysis [57,
70, 79]. In particular, sodium can catalyze dehydration and crosslinking reactions in sugars that help form char more
rapidly [38, 80, 81]. Normally, xylan is extracted from biomass using NaOH, so a significant quantity of Na remains in
the xylan purchased commercially. The xylan was thus subjected to an acid wash (1 M HCl) as described in the
materials and methods section to remove minerals and compare the behavior of xylan as received with ash-free xylan.
HCl was used to minimize hydrolysis of the hemicellulose. Table 1 shows that the ash content was reduced from 5.85%
Chapter 2 91
to 0.0112% by this acid washing procedure. Figure 8 presents the sequence of images obtained during pyrolysis of de-
ashed xylan.

A. Frame 575 Time=4.6s B. Frame 731 Time=5.84s C. Frame 805 Time=6.44s


Tsample 276°C Tsample=303°C Tsample=306°C

Cracks formation by solid


Deashed xylan melted entrainment Liquid swelling
D. Frame 864 Time =6.91s E. Frame 880 Time=7.04s F. Frame 1383 Time=11.06s
Tsample=308°C Tsample=309°C Tsample=335°C
Char layer

Liquid drops Liquid swelling


Liquid swelling Char layer
Figure 6. De-ashed xylan melting and bubbling behavior during pyrolysis: still frames from video recordings.

The behavior of de-ashed xylan is very different from the behavior of xylan as-received. In Figure 6A, the formation of
the liquid phase can be seen as the sample reaches 276 oC (4.6 s). In the corresponding video (video D, supplemental
material), physical ejection of partially reacted xylan from the pile can be seen. It is known that CO2 and formic acid are
rapidly produced from hemicellulose [70]. The violent release of these gases can expel aerosol particles. In Figure 6C,
the entire lower layer of xylan at the quartz surface is melted, and the top melts slowly because that bottom layer swells
and restricts heat transfer. The xylan accumulates internal energy until most of the sample is melted, forming a liquid
similar what is seen in melting of simple sugars. In this liquid phase, it is difficult to monitor the dynamics of bubble
nucleation, growth, coalescence, and other properties. The behavior is very chaotic. After 9.5 s, the liquid phase
formation has ceased and a carbonaceous residue remains.

The impact of removing the ash from xylan is striking. When the ash is present, the char formation appears to compete
much more heavily. Further, the ash physically takes up space in the material and will contribute to the solid structure
of the char. Without ash, the xylan looks much more like the amorphous, bubbling lignin.

2.3.5. Behavior of Cellulose

Cellulose is the most abundant component in most lignocellulosic biomass. The still video frames for cellulose can be
seen in Figure 7. During the first two seconds of heating, the sample reaches 229 °C, but no color changes, shrinkage,
swelling, or release of gases can be seen (see Figure 7 A-B). Only when the sample has exceeded 255 °C does it
92 Chapter 2
begin to develop small drops of liquid material on the surface of the hot disc. The top of the cellulose pile is unchanged,
due to poor heat transfer. The disk surface at this time (2.5 s) is 297 °C, at which point melting is evident.

Although the images do not fully tell the story, the video (see supplementary video E) shows that as soon as the liquid
droplets are formed, a release of white smoke can be seen. This could be attributed to ejection of aerosols [5, 41].

An interesting phenomenon is observed at 3.36 s, when the top of the sample is on average 290 °C (T disc 338
°C). Much of the sample has already been consumed by evaporation of the liquid phase, but a solid material remains
on the surface. This is likely due to the fact that that the material is in poor contact with the disk surface due to Leidenfrost
effects [82], in which the vapors rise solid sample, drastically reducing the rate of energy transfer by conduction (see
Frames C-D).

A. Frame 200 Time=1.6s B. Frame 250 Time=2s C. Frame 313 Time 2.5s
Td=196°C Ts=190°C Td=233.5°C Ts=229°C Td=297°C Ts=255°C

Melted cellulose Melted cellulose


D. Frame 420 Time=3.36s E. Frame 442 Time=3.53s F. Frame 464 Time=3.71s
Td=338°C Ts=290°C Td=342°C Ts=301°C Td=355°C Ts=308°C

Liquid swelling Solid shrink

Solid shrinking Liquid swelling Liquid swelling


G. Frame 495 Time=3.96s H. Frame 673 Time=5.38s I. Frame 700 Time=5.6s
Td=365°C Ts=314°C Td=425°C Ts=341°C Td=430°C Ts=344°C

Liquid shrinking and Char volume Final char volume


resolidification

Figure 7. Cellulose melting and bubbling behavior during pyrolysis: still frames from video recordings.
Chapter 2 93
The cellulose sample continues to evaporate and when the disc surface has reached 430 °C, the cellulose is completely
melted (sample temperature 344 °C). No obvious Leidenfrost phenomenon can be seen. Because of the optical
properties of the liquid phase, it is also difficult to track the dynamics of bubbles.

2.3.6. Behavior of sugarcane bagasse

Figure 8 shows the video frames of pyrolysis of sugarcane bagasse. During the first 1.64 s, (T = 140.12 °C) no change
is observed in the biomass. When the average pile temperature has reached 282 °C, darkening of the edges can be
seen where the temperature is higher than the average pile temperature (Figure 8B). At this temperature, the
hemicellulose has begun decomposing and lignin has changed phase. There is no dramatic change in the volume of
the sample, as with lignin.

A. Frame 206 Time=1.65s Td=220°C B. Frame 535 Time=4.28s Td=386.14°C C. Frame 848 Time=6.78s Td=467.73°C
Ts=140.12°C Ts=282.64°C Ts=332.08°C

Solid darkening Liquid phase

Solid shrinking
D. Frame 1066 Time=8.53s Td=489.39°C E. Frame 1351 Time=10.81s F. Frame 2050 Time=16.4s Td=496.39°C
Ts=344°C Td=491.56°C Ts=349.48°C Ts=357.13°C

Softened surface

Solid shrinking

Figure 8. Sugar cane bagasse melting and bubbling behavior during pyrolysis: still frames from video recordings.

More interesting behavior is seen when the sample has reached 332 °C. At this temperature there is a complete change
in the color of biomass, and the surface becomes smooth and molten. At this temperature, cellulose has now melted. No
volume changes (swelling or shrinking) or Leindefrost effects are observed. The behavior here is quite similar to the
xylan containing metals, with the difference that here the formation of fractures in the structure due to formation of small
bubbles trapped inside the micro molten phase. Although the behaviors of xylan containing metals and biomass are not
equal, they are more similar than they are to lignin or cellulose.

Figure 9 shows the thermal behavior of the acid-washed bagasse, which starts to degrade when the sample has
reached 260 °C. This is very close to the degradation temperature of the pure cellulose (255 °C). The biomass shrinks
quickly, and it is easy to see vapor release in the video (supplementary video G). The devolatilization looks similar to a
94 Chapter 2
solid that is sublimating. Although some swelling can be seen, especially after the sample exceed 300 °C, it is not as
vigorous as lignin. Instead, there is dramatic shrinking of the material, leaving only a small layer of char on the surface
of the disc at the end.

In conclusion, the behavior of acid washed biomass is completely different from its pseudo-components. It is possible
the acid wash has affected the crystallinity of cellulose and the lignin structure beyond removing alkali and alkali earth
metals, and ought to be studied in future work.

It is important to note that while xylan began melting at 348 °C, lignin was completely melted at this temperature. The
lignin acts as a base to catalyze dehydration reactions and crosslinking between chains xylan and cellulose; these
dehydration reactions lead to polyromantic ring formation upon further heating, which makes up the structure of char
[48]. These acid catalyzed reaction may explain why the biomass retains its structure throughout the pyrolysis process.
Furthermore, the presence of alkali and alkali earth metals dramatically effects the porosity and swelling of xylan and
biomass, as can be seen in the SEM micrographs in the upcoming section.

A. Frame 255 Time=2.04s B. Frame 485 Time=3.88s C. Frame 699 Time=5.59s


Td=280°C Ts=175°C Td=370°C Ts=260°C Td=433°C Ts=312°C

Solid shrinking
Solid darkening
D. Frame 715 Time=5.72s E. Frame 917 Time=7.34s F. Frame 1450 Time=11.6s
Td=441°C Ts=316°C Td=470°C Ts=335°C Td=492°C Ts=349°C

Solid shrinking
Solid Thin layer
Figure 9. Acid washed bagasse melting and bubbling behavior during pyrolysis: still frames from video recordings.

It is accepted that in the microstructure of lignocellulosic biomasses, hemicellulose surrounds the cellulose chains. The
lignin is present in the secondary cell walls as a binder between the stiff pore structures [83,84]. Research has shown
that hemicellulose tightens cellulose chains together and changes the dimensions of the micro-fibrils, as seen in the
crystal structure [84]. Of the various hemicelluloses, xylan in particular binds cellulose fibrils tighter in the paper-making
process [85,86]. Removal of hemicellulose in the cell walls would increase internal porosity and surface area of the pre-
treated biomass [85,86], thereby facilitating the escape of aerosols and volatile material that may contribute to the
formation of char. This would promote greater chain mobility for cellulose and lignin during pyrolysis. Apart from the
Chapter 2 95
above reasoning, it is important to note that the Organosolv lignin used in this study has a lower molecular weight
relative to the native lignin, so its fluidity starts at a lower temperature.

2.3.7. SEM imaging of solid residue from all materials studied

In order to compare the char structures after pyrolysis, SEM images were taken for all the materials, as shown in Figure
12. In the images, major and minor differences can be seen between the materials.

The organosolv lignin and milled wood lignin both melt completely and form chars that look like softened surfaces,
commonly attributed to liquid phase formation. The organosolv lignin (Figure 10 A) is very smooth, and does not appear
to bubble as much as the milled wood lignin (Figure 10 B-C), which has small holes on the order of 5-40 µm in diameter.
Organosolv lignin is produced at much higher temperatures than milled wood lignin, and its structure is likely more
degraded so there will be less fragmentation reactions, which produce gases and bubbles.

A. Organosolv lignin B. Milled wood lignin C. Milled wood lignin


(zoomed out)

50 µm 50 µm 500 µm

D. Xylan E. De-ashed xylan F. De-ashed xylan (zoomed


out)

50 µm 50 µm 500 µm

G. Avicel cellulose H. Sugarcane bagasse I. Acid washed bagasse

50 µm 50 µm 50 µm

Figure 10. SEM micrographs of (A) Organosolv lignin, (B-C) milled wood lignin, (D) xylan, (E-F) de-ashed xylan, (G)
cellulose, (H) sugarcane bagasse, and (I) acid washed sugarcane bagasse chars. Scale bars are indicated on the
figures.

The difference between xylan and de-ashed xylan is very dramatic. The xylan as-received (Figure 10 D) has a 3-
dimensional structure that looks like coral, with extrusions and holes of all sizes. This structure is evidence of substantial
96 Chapter 2
gas production, catalyzed by metals. The de-ashed xylan, however, looks completely different (Figure 10 E-F). There
is some evidence of bubbling, but it forms a very flat film that looks much more close to the Organosolv lignin than the
xylan with ash.

The Avicel cellulose (Figure 10 F) forms strings of interwoven char which is evidence of a liquid intermediate that melts
and moves around due to the Leidenfrost effect. Cross-linking reactions compete with depolymerization while the liquid
intermediate boils. This char looks very different from the chars from lignin and xylan.

The sugarcane bagasse (Figure 10 H) appears to retain its original structure during pyrolysis, and there is no strong
evidence of a non-intense liquid intermediate forming just on the surface. The acid washed bagasse, however, appears
to have small layers of amorphous char on the surface, which is evidence of liquid intermediate formation. As a result,
the char from acid washed bagasse appears to be less structured than the bagasse with ash. More work needs to be
done to assess the impact of removing ash on the formation of liquid intermediates.

2.4. Conclusions
It is well known that char produced from lignocellulosic biomass retains its general structure during pyrolysis. In this
work, it was performed visualization studies on cellulose, Organosolv lignin, milled wood lignin, xylan, de-ashed xylan,
sugarcane bagasse, and de-ashed sugarcane bagasse. The lignin melted, bubbled dramatically, but shrank again. The
cellulose melted and shrank. Ash-free xylan melted completely and shrank into a pool, then carbonized. However, when
ash content was high in the xylan, the xylan foamed dramatically and grew before carbonizing due to the catalytic
effects of ash on gas formation. In sugarcane bagasse, removing the ash appears to have enhanced the formation of
a liquid intermediate; SEM of the char indicates that more of the biomass melted (likely the hemicellulose and lignin.
Future work needs to be done to better understand the effect of metal content on the formation of products responsible
for char retaining the structure of the original biomass.
Chapter 2 97

2.5. References
[1] Bridgwater, A.V. and Peacocke, Fast pyrolysis processes for biomass, Renew. Sustain. Energy Rev. 4 (2000)
1–73.
[2] J. Lédé, Cellulose pyrolysis kinetics: An historical review on the existence and role of intermediate active
cellulose, J. Anal. Appl. Pyrolysis. 94 (2012) 17–32.
[3] C. Di Blasi, Heat transfer mechanisms and multi-step kinetics in the ablative pyrolysis of cellulose, Chem. Eng.
Sci. 51 (1996) 2211–2220.
[4] S. Nordin, J. Nyren, E. Back, An Indication of Molten Cellulose Produced in a Laser Beam, Text. Res. J. 44
(1974) 915–917.
[5] P.Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P. and
Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011)
4306.
[6] Goring, Thermal Softening of Lignin, Hemicellulose and Cellulose, Pulp Pap. Mag. Canada. 64 (1963) T517–
T527.
[7] D. Howe, D. Taasevigen, M. Gerber, M. Gray, C. Fernandez, L. Saraf, M. Garcia-Perez, M. Wolcott, Bed
Agglomeration during the Steam Gasification of a High-Lignin Corn Stover Simultaneous Saccharification and
Fermentation (SSF) Digester Residue, Energy and Fuels. 29 (2015) 8035–8046.
[8] Paul J Dauenhauer, Joshua L Colby, Christine M Balonek, Wieslaw J Suszynski, Reactive boiling of cellulose
for integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[9] S. Saka, M. Asmadi, H. Kawamoto, Gas- and solid/liquid-phase reactions during pyrolysis of softwood and
hardwood lignins, J. Anal. Appl. Pyrolysis. 92 (2011) 417–425.
[10] Zaror C, Pyle D. Sadhana, The pyrolysis of biomass: a general review, Proc. Eng. Sci. 5 (1982) 269–285.
[11] A. Demirbas, Effect of initial moisture content on the yields of oily products from pyrolysis of biomass, J. Anal.
Appl. Pyrolysis. 71 (2004) 803–815.
[12] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Effects of particle size on the fast
pyrolysis of oil mallee woody biomass. Fuel. 88 (2009) 1810–1817.
[13] A.V. Bridgwater, Effect of experimental conditions on the comopsition of gases and liquids from biomass.
Advances in thermochemical biomass conversion pyrolysis, Glasgow, 1994.
[14] A. Mittal, R. Katahira, M.E. Himmel, D.K. Johnson, Effects of alkaline or liquid-ammonia treatment on crystalline
cellulose: changes in crystalline structure and effects on enzymatic digestibility, Biotechnol. Biofuels. 4 (2011)
41.
[15] A. Debdoubi, A. El Amarti, E. Colacio, M.J. Blesa, L.H. Hajjaj, The effect of heating rate on yields and
compositions of oil products from esparto pyrolysis, Int. J. Energy Res. 30 (2006) 1243–1250.
[16] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[17] E. Bolat, K. Açıkalın, F. Karaca, Pyrolysis of pistachio shell: Effects of pyrolysis conditions and analysis of
products, Fuel. 95 (2012) 169–177.
[18] Haas, M.R. Nimlos, B.S. Donohoe, Real-Time and Post-reaction Microscopic Structural Analysis of Biomass
Undergoing Pyrolysis, Energy & Fuels. 23 (2009) 3810–3817.
[19] R. Bassilakis, R.M. Carangelo, M.A. Wójtowicz, TG-FTIR analysis of biomass pyrolysis, Fuel. 80 (2001) 1765–
1786.
[20] G.A. Zickler, W. Wagermaier, S.S. Funari, M. Burghammer, O. Paris, In situ X-ray diffraction investigation of
thermal decomposition of wood cellulose, J. Anal. Appl. Pyrolysis. 80 (2007) 134–140.
[21] G. Zhu, X. Zhu, Z. Xiao, F. Yi, Study of cellulose pyrolysis using an in situ visualization technique and
thermogravimetric analyzer, J. Anal. Appl. Pyrolysis. 94 (2012) 126–130.
[22] W. Li, K. Yang, J. Peng, L. Zhang, S. Guo, H. Xia, Effects of carbonization temperatures on characteristics of
porosity in coconut shell chars and activated carbons derived from carbonized coconut shell chars, Ind. Crops
Prod. 28 (2008) 190–198.
[23] L. Sun, P. Fu, S. Hu, S. Su, J. Wang, J. Xiang, Evaluation of the porous structure development of chars from
pyrolysis of rice straw: Effects of pyrolysis temperature and heating rate, J. Anal. Appl. Pyrolysis. 98 (2012)
177–183.
[24] A. Dufour, M. Castro-Diaz, N. Brosse, M. Bouroukba, C. Snape, The origin of molecular mobility during biomass
98 Chapter 2
pyrolysis as revealed by in situ 1H NMR spectroscopy, ChemSusChem. 5 (2012) 1258–1265.
[25] A. Dufour, M. Castro-Díaz, P. Marchal, N. Brosse, R. Olcese, M. Bouroukba, C. Snape, In Situ Analysis of
Biomass Pyrolysis by High Temperature Rheology in Relations with 1H NMR, Energy & Fuels. 26 (2012) 6432–
6441.
[26] Z. Wang, Armando G. McDonald, Roel J.M. Westerhofc, Sascha R.A. Kersten, Christian M. Cuba-Torres,
Su Had, Brennan Pechad, Manuel Garcia-Perez, Effect of cellulose crystallinity on the formation of a liquid
intermediate and on product distribution during pyrolysis, J. Anal. Appl. Pyrolysis. 100 (2013) 56–66.
[27] Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, C.-Z. Li, A.G. McDonald, M. Garcia-Perez, Effect of
Cellulose Crystallinity on Solid/Liquid Phase Reactions Responsible for the Formation of Carbonaceous
Residues during Pyrolysis, Ind. Eng. Chem. Res. 53 (2014) 2940–2955.
[28] H.W. Dawei Liu, Differences in Water-Soluble Intermediates from Slow Pyrolysis of Amorphous and Crystalline
Cellulose, Energy Fuels. 27 (2013) 1371–1380.
[29] J. Piskorz, A. Vladars-Usas, D. Radlein, P. Majerski, D.. Scott, Flash pyrolysis of cellulose for production of
anhydro-oligomers, J. Anal. Appl. Pyrolysis. 56 (2000) 145–166.
[30] J. Zhang, Fast pyrolysis behavior of different celluloses and lignocellulosic biopolymer interaction during fast
pyrolysis, MSc dissertation 2012. Iowa State University
[31] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[32] A. Trubetskaya, P.A. Jensen, A.D. Jensen, M. Steibel, H. Spliethoff, P. Glarborg, Influence of fast pyrolysis
conditions on yield and structural transformation of biomass chars, Fuel Process. Technol. 140 (2015) 205–
214.
[33] Teixeira, R.J. Hermann, J.S. Kruger, W.J. Suszynski, L.D. Schmidt, D.P. Schmidt, P.J. Dauenhauer,
Microexplosions in the upgrading of biomass-derived pyrolysis oils and the effects of simple fuel processing,
ACS Sustain. Chem. Eng. 1 (2013) 341–348.
[34] A. Paulsen, M.S. Mettler, P.J. Dauenhauer, The role of sample dimension and temperature in cellulose
pyrolysis, Energy Fuels. 27 (2013) 2126–2134.
[35] O. Boutin, M. Ferrer, J. Lédé, Flash pyrolysis of cellulose pellets submitted to a concentrated radiation:
experiments and modelling, Chem. Eng. Sci. 57 (2002) 15–25.
[36] E. Cetin, B. Moghtaderi, R. Gupta, T.. Wall, Influence of pyrolysis conditions on the structure and gasification
reactivity of biomass chars, Fuel. 83 (2004) 2139–2150.
[37] K. Chaiwat, W. Hasegawa, I. Kori, J. Mae, Examination of Degree of Cross-Linking for Cellulose Precursors
Pretreated with Acid/Hot Water at Low Temperature, Ind. Eng. Chem. Res. 47 (2008) 5948–5956.
[38] M.M. Ramirez-Corredores, Pathways and Mechanisms of Fast Pyrolysis: Impact on Catalyst Research, in: Role
Catal. Sustain. Prod. Bio-Fuels Bio-Chemicals, Elsevier, Pasadena-Texas, 2013: pp. 161–216.
[39] W.F. DeGroot, F. Shafizadeh, The influence of exchangeable cations on the carbonization of biomass, J. Anal.
Appl. Pyrolysis. 6 (1984) 217–232.
[40] F. Yi, G. Zhu, X. Zhu, R. Zhou, Z. Xiao, Pyrolysis characteristics of bean dregs and in situ visualization of
pyrolysis transformation, Waste Manag. (2012).
[41] Alex D. Paulsen, Blake R. Hough, C. Luke Williams, Andrew R. Teixeira, Daniel T. Schwartz, Fast Pyrolysis of
Wood for Biofuels: Spatiotemporally Resolved Diffuse Reflectance In situ Spectroscopy of Particles,
ChemSusChem. 7 (2014) 765–777.
[42] J. Lédé, J., Diebold, J.P., Peacocke, G.V.C. and Piskorz, The nature and properties of intermediate and
unvaporized biomass pyrolysis materials., Dev. Thermochem. Biomass Convers. (1997) 27–42.
[43] J. Diebold, J. Scahill, Ablative pyrolysis of macroparticles of biomass, in: Proceedings of Specialists’ Workshop
on fast Pyrolysis of Biomass, Copper Mountain, in: R.P. Overend, T.A. Milne, L.K. Mudge (Eds.), Fundam.
Thermochem. Biomass Convers., Springer Netherlands, Copper Mountain, 1985: pp. 539–555.
[44] J. Lédé, M. Ferrer, O. Boutin, Radiant flash pyrolysis of cellulose—Evidence for the formation of short life time
intermediate liquid species, J. Anal. Appl. Pyrolysis. 47 (1998) 13–31.
[45] M.S.Mettler, Alex D Paulsen, Dionisios G Vlachos, Paul J. Dauenhauer, The chain length effect in pyrolysis:
bridging the gap between glucose and cellulose, Green Chem. 14 (2012) 1284–1294.
[46] S. Agarwal, V., Dauenhauer, P.J., Huber, G.W. and Auerbach, Ab initio dynamics of cellulose pyrolysis: nascent
decomposition pathways at 327 and 600 °C., J. Am. Chem. Soc. 134 (2012) 14958–72.
[47] T.J. Hilbers, Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, M.R. Pelaez-Samaniego, M. Garcia-Perez,
Chapter 2 99
Cellulose-Lignin interactions during slow and fast pyrolysis, J. Anal. Appl. Pyrolysis. 114 (2015) 197–207.
[48] M.W. Smith, Effect of Pyrolysis Temperature on Cellulose Char Aromatic Cluster Size by Quantitative Multi
Cross-Polarization 13C NMR with Long Range Dipolar.Characterization and modification of biomass pyrolysis
chars for environmental applications Dephasing, Washington State University, 2016.
[49] T.S. Jeong, B.H. Um, J.S. Kim, K.K. Oh, Optimizing dilute-acid pretreatment of rapeseed straw for extraction
of hemicellulose, Appl. Biochem. Biotechnol. 161 (2010) 22–33.
[50] M.J. Goundalkar, D.B. Corbett, B.M. Bujanovic, Comparative analysis of milled wood lignins (MWLs) isolated
from sugar maple (SM) and hot-water extracted sugar maple (ESM), Energies. 7 (2014) 1363–1375.
[51] J.R. Obst, T.K. Kirk, Isolation of lignin, Methods Enzymol. 161 (1988) 3–12.
[52] A. Bjorkman, Isolation of Lignin from Finely Divided Wood with Neutral Solvents, Nature. 174 (1954) 1057–
1058.
[53] Carbon, Hydrogen, and Nitrogen in Biomass and Biofuel.Corporation, LECO, (2009).
[54] ASTM D7582-15, Standard Test Methods for Proximate Analysis of Coal and Coke by Macro
Thermogravimetric Analysis, ASTM International, West Conshohocken, PA, 2015, www.astm.org, (n.d.).
[55] R. García, C. Pizarro, A.G. Lavin, J.L. Bueno, Biomass proximate analysis using thermogravimetry, Bioresour.
Technol. 139 (2013) 1–4.
[56] B. Pecha, P. Arauzo, M. Garcia-Perez, Impact of combined acid washing and acid impregnation on the pyrolysis
of Douglas fir wood, J. Anal. Appl. Pyrolysis. 114 (2015)
[57] M. Sohrabi, M.R. Hajaligol, M. Nik-Azar, B. Dabir, Mineral matter effects in rapid pyrolysis of beech wood, Fuel
Process. Technol. 51 (1997) 7–17.
[58] K. Raveendran, A. Ganesh, K.C. Khilar, Influence of mineral matter on biomass pyrolysis characteristics, Fuel.
74 (1995) 1812–1822.
[59] J. Montoya, B. Pecha, F. Chejne Janna, M. Garcia-Perez. Micro-Explosion of Liquid Intermediates During the
Fast Pyrolysis of Sucrose and Organosolv Lignin, Submitt. to J. Anal. Appl. Pyrolysis. 2016.
[60] J.-Y. Kim, S. Oh, H. Hwang, U.-J. Kim, J.W. Choi, Structural features and thermal degradation properties of
various lignin macromolecules obtained from poplar wood (Populus albaglandulosa), Polym. Degrad. Stab. 98
(2013) 1671–1678.
[61] L. Yang, D. Wang, D. Zhou, Y. Zhang, Effect of different isolation methods on structure and properties of lignin
from valonea of Quercus variabilis., Int. J. Biol. Macromol. 85 (2016) 417–24.
[62] S. Zhou, B. Pecha, M. van Kuppevelt, A.G. McDonald, M. Garcia-Perez, Slow and fast pyrolysis of Douglas-fir
lignin: Importance of liquid-intermediate formation on the distribution of products, Biomass and Bioenergy. 66
(2014) 398–409.
[63] A.D. Paulsen, M.S. Mettler, P.J. Dauenhauer, The role of sample dimension and temperature in cellulose
pyrolysis, in: Energy and Fuels, 2013: pp. 2126–2134.
[64] V.B.F. Custodis, C. Bährle, F. Vogel, J.A. van Bokhoven, Phenols and aromatics from fast pyrolysis of variously
prepared lignins from hard- and softwoods, J. Anal. Appl. Pyrolysis. 115 (2015) 214–223.
[65] M. Zhou, S., Pecha, B., van Kuppevelt, M., McDonald, A.G. and Garcia-Perez, Slow and fast pyrolysis of
Douglas- Fir Lignin:Importance of liquid intermediate formation on distribution of products, Biomass and
Bioenergy. (2014) 398–409.
[66] Huijgen, G. Telysheva, A. Arshanitsa, R.J.A. Gosselink, P.J. de Wild, Characteristics of wheat straw lignins
from ethanol-based organosolv treatment, Ind. Crops Prod. 59 (2014) 85–95.
[67] E. Jakab, O. Faix, F. Till, Thermal decomposition of milled wood lignins studied by thermogravimetry/mass
spectrometry, J. Anal. Appl. Pyrolysis. 40-41 (1997) 171–186.
[68] Mamdouh, G.D.M. MacKay, Mechanism of Thermal Decomposition of Lignin, Wood Fiber Sci. 16 (1984) 441–
453.
[69] R.K. Sharma, J.B. Wooten, V.L. Baliga, X. Lin, W. Geoffrey Chan, M.R. Hajaligol, Characterization of chars
from pyrolysis of lignin, Fuel. 83 (2004) 1469–1482.
[70] Patwardhan, Understanding the product distribution from biomass fast pyrolysis, Iowa State University, 2010.
Doctoral Thesis.
[71] Y. Qu, H. Luo, H. Li, J. Xu, Comparison on structural modification of industrial lignin by wet ball milling and ionic
liquid pretreatment, Biotechnol. Reports. 6 (2015) 1–7.
[72] P.C. Duineveld, Bouncing and coalescence of bubble pairs rising at high Reynolds number in pure water or
aqueous surfactant solutions, Appl. Sci. Res. 58 (1998) 409–439.
100 Chapter 2
[73] S. Orvalho, M.C. Ruzicka, G. Olivieri, A. Marzocchella, Bubble coalescence: Effect of bubble approach velocity
and liquid viscosity, Chem. Eng. Sci. 134 (2015) 205–216.
[74] K. Butler, A numerical model for combustion of bubbling theroplastic materials in microgravity, (2002) 70.
[75] J.R. Grace, T. Wairegi, J. Brophy, Break-up of drops and bubbles in stagnant media, Can. J. Chem. Eng. 56
(1978) 3–8.
[76] H. Yang, D.H. Lee, H. Chen, C. Zheng, R. Yan, Characteristics of hemicellulose, cellulose and lignin pyrolysis,
Fuel. 86 (2007) 1781–1788.
[77] G. Varhegyi, M.J. Antal, T. Szekely, P. Szabo, Kinetics of the thermal decomposition of cellulose,
hemicellulose, and sugarcane bagasse. , Energy Fuels. 3 (1989) 329–335.
[78] H. Yang, R. Yan, H. Chen, C. Zheng, D.H. Lee, D.T. Liang, In-depth investigation of biomass pyrolysis based
on three major components: hemicellulose, cellulose and lignin, Energy and Fuels. 20 (2006) 388–393.
[79] K. Werner, L. Pommer, M. Broström, Thermal decomposition of hemicelluloses, J. Anal. Appl. Pyrolysis. 110
(2014) 130–137.
[80] J. Akhtar, N. Saidina Amin, A review on operating parameters for optimum liquid oil yield in biomass pyrolysis,
Renew. Sustain. Energy Rev. 16 (2012) 5101–5109.
[81] H. Kawamoto, S. Saito, W. Hatanaka, S. Saka, Catalytic pyrolysis of cellulose in sulfolane with some acidic
catalysts, J. Wood Sci. 53 (2007) 127–133.
[82] Teixeira, C. Krumm, K.P. Vinter, A.D. Paulsen, C. Zhu, S. Maduskar, K.E. Joseph, K. Greco, M. Stelatto, E.
Davis, B. Vincent, R. Hermann, W. Suszynski, L.D. Schmidt, W. Fan, J.P. Rothstein, P.J. Dauenhauer, Reactive
Liftoff of Crystalline Cellulose Particles, Sci. Rep. 5 (2015) 11238.
[83] L. Li, P. P, X. Frank, K. Mazeau, A coarse-grain force-field for xylan and its interaction with cellulose, Carbohydr.
Polym. 127 (2015) 438–450.
[84] T.M. Tenhunen, M.S. Peresin, P.A. Penttilä, J. Pere, R. Serimaa, T. Tammelin, Significance of xylan on the
stability and water interactions of cellulosic nanofibrils, React. Funct. Polym. 85 (2014) 157–166.
[85] E. Norström, L. Fogelström, P. Nordqvist, F. Khabbaz, E. Malmström, Xylan – A green binder for wood
adhesives, 67 (2015) 483–493.
[86] A. Ferraz, T.H.F. Costa, G. Siqueira, A.M.F. Milagres, Mapping of Cell Wall Components in Lignified Biomass
as a Tool to Understand Recalcitrance, Biofuels in Brazil. (2014) 173–202.
Chapter 3 101

3. Chapter 3. Micro-Explosion of Liquid Intermediates During


the Fast Pyrolysis of Sucrose and Lignin
(Paper published in Journal of Analytical and Applied Pyrolysis)

Abstract: A new methodology has been proposed to describe the dynamics of bubble formation in Organosolv lignin
and sucrose that combines fast speed visualization (125fps) with mathematical modeling. The model uses a population
balance to theoretically predict overall rates of bubble birth and death, bubble bursting, and aerosol ejection. The
experimental studies were performed on a uniquely modified CDS Analytical Pyroprobe 5000 to visualize the formation
of bubbles within the liquid intermediate phase at heating rates close 100 oC/s. Experimentally, it was observed that
gas bubbles follow a log-normal distribution versus bubble size within the liquid intermediate phase for both
materials. This distribution function changes over time due to increased viscosity from solidification reactions that
generate char and the changes in the rate of bubble formation. Micro-explosion intensity was used to estimate aerosol
ejection intensity. The model predicts aerosol ejection yields of 21.18% w / w from Organosolv lignin and 17.40% w /
w from sucrose during pyrolysis with an average droplet size of 10 µm.

Keywords: Fast Pyrolysis, liquid Intermediate Phase, Bubbling, micro-explosion, aerosol formation.
102 Chapter 3

3.1. Introduction
Biomass fast pyrolysis is a relatively simple process, but is very challenging to control. It is a fast reaction in which heat
transfer, mass transfer, and chemical reactions all contribute dramatically to the final distribution of products [1, 2].
Lignocellulosic biomass is made up of cellulose, hemicelluloses, lignin, extractives, and ash. Experimental and
theoretical evidence show that during neat cellulose and lignin depolymerization, a short-lived liquid phase is formed
before compounds are released from the biomass particle. This liquid phase is composed of heavy depolymerization
products [3 – 6] and explains why researchers commonly collect high molecular weight compounds (anhydro-sugars
DP > 2, pyrolytic lignin) in bio-oil [3, 7, 8].
In this liquid intermediate phase, volatile products of pyrolysis reactions (vapors and gases) form bubbles. These
bubbles grow, coalesce and collapse to release aerosols that are made up of whatever is in the liquid at that time [3,
8, 9]. When bubbles rise to the surface; the external liquid-gas interface swells, increasing the internal pressure of the
bubble (see Figure 1A). When the internal energy exceeds the surface’s free energy, the bubbles collapse (bursting)
(see Figure 1B) [8]. As the bubbles burst and release gases, the liquid surface geometry deforms to form “jets” in the
central region (see Figure 1C). The jets collapse and release large aerosols (0.1-1mm) [10–12].

Figure 1. Dynamic life of a bubble on a liquid surface [9].

The dynamics of bubble formation and collapse has been extensively studied in evaporative processes, droplet
combustion processes, and aeration in gas/liquid interfaces for water surfaces [10-16]. For example; the rapid boiling
processes (pool boiling, a mechanism used in the nuclear power plants for the fast transfer of energy); has received a
lot of attention [17–20]. Bubble dynamics are affected by temperature, fluid type, and the micro structure of the surface.

Micro-emulsions droplet combustion is a good example of a technology where microexplosions are important. During
combustion, volatiles may be released violently, totally or partially destroying the mother drop [21, 22]. Microexplosion
of droplets suspended on a thermocouple is commonly studied using fast speed cameras (> 1000 fps) to visualize
morphological changes as a function of temperature and time [23–27]. Pyrolysis oils (also known as bio-oil) are also
known to undergo microexplosions during combustion [25]. Bio-oil microexplosion intensity depends strongly on droplet
size, temperature, and heating rate [25]. Shuhn, et al. [28] described the swelling, bubbling, aerosol ejection phenomena
during micro-explosions of fast pyrolysis bio-oil blends with diesel.

Teixeira et al. [21] found that the presence of high molecular weight compounds (1 000 Da) in the bio-oil is important
for the formation of microbursts. Filtration of these compounds (probably pyrolytic lignins) reduces the frequency of
micro-explosions by 50 % [21]. The presence of solids such as char does not appear to alter the dynamics of bubble
formation or the intensity of microbursts [21]. The dynamics of bubble formation during pyrolysis has also been studied
for polymers [8] and coals [12, 29]. However, the authors were not able to find any studies on bubble dynamics for the
liquid intermediate produced during the pyrolysis of lignocellulosic materials.
Chapter 3 103

One metric to study aerosol formation is the intensity of ejection. The aerosol ejection intensity is defined as the average
total number of aerosol droplets ejected per unit time [3, 30–32]. Several studies report direct relationship between the
diameter of the gas bubbles within the liquid phase and the size of aerosol ejected. These relationships depend on the
physical and chemical nature of the fluid and conditions (temperature and pressure) of bubbling. Many researchers
have attempted to correlate the dynamic ejection aerosol, with the characteristics of the gas bubbles during the bursting
[30, 31, 33] . However, given the complexity of the phenomenon, it is difficult to make a universal description of aerosol
ejection. Recently, Zhang et al. [31] proposed correlations with dimensionless numbers to establish relationships
between the number of aerosols ejected as a function of the physicochemical properties of the medium (liquid phase),
and the diameter of the bubbles before bursting. Knowing the dynamics of bubble formation, the physico-chemical
properties of the liquid intermediate and existing correlations to calculate aerosol ejection intensity [31] could enhance
the predictive quality of models describing biomass pyrolysis.

Inherently connected to the outcome of the bubbling and thermal ejection phenomena is the biomass particle size.
Particle size affects the internal microstructure of biomass particles and plays a determinant role in heat and mass
transfer processes during biomass fast pyrolysis [34-36].

It is known that bio-oil yield decreases as the particle diameter increases, although it does not follow a linear trend. A
dramatic reduction in bio-oil and oligomeric yield occurs when the particle size increases from less than 0.1 mm to 1
mm. [37, 38]. The decrease in the lignin oligomer yield is mainly due to retention of these compounds inside the biomass
particles as they are released by thermal ejection within the pores [1]. Two pyrolysis regimes were suggested based on
particle size [5, 35]. The first regime is for very small particles (mostly fragments of cell walls) which release aerosols
directly off the cell walls into the reactor. In the second regime, the thermally ejected oligomers are trapped inside the
cell cavities leading to the formation of secondary char and bio-oils with a lower content of oligomers [5].

Based on a thorough literature review, there is lack of information regarding bubble formation in fast pyrolysis of
lignocellulosic materials and its contribution to aerosol ejection. In this study it is investigated this phenomena in
organosolv lignin and sucrose using experimental and modelling approaches. Organosolv lignin and sucrose act as
surrogates for biomass components lignin and cellulose, respectively [1]. The goal is to advance the understanding of
the processes of birth, growth, coalescence and microexplosion of bubbles in the liquid intermediate phase of pyrolysis.

3.2. Materials and Methods


Visualization of bubbling dynamics within sucrose and organosolv lignin as surrogates of lignocellulosic materials were
made by fast speed camera in a modified Pyroprobe reactor. Population balances was used to describe teoteically the
bubbles bursting coalescence and death rates.

3.2.1. Materials

In this work Organosolv lignin (371,017 Aldrich, CAS 8068-03-9) and sucrose (BioXtra, -Sigma ≥99.5%) were used. The
proximate analysis of the organosolv lignin was determined in a thermobalance Mettler Toledo TGA / SDTA 851
weighing 20 mg of sample in an open alumina capsule using the following heating method: N2 (25-107 oC at 100 oC/min,
hold 60 min, 107-950 oC at 45 oC/min, hold 7 min, 950-600 oC, -100 oC/min) O2 (600-750 oC at 2.5 oC/min, hold 60
min). Elemental analysis of both the organosolv lignin and the sucrose was performed on a Leco TruSpec following a
standard protocol for biomass. The results are shown in Table 1.
104 Chapter 3
Table 1. Physico-chemical characterization of the raw material. CHN-O data are reported on a dry basis.
Proximate Analysis [wt. %]
Compound Moisture Ash Fixed Carbon Volatile
Organosolv lignin 1.2 2.0 26.6 70.2
Ultimate Analysis [wt. %]
Compound C H N O*
Sucrose 42.1 6.4 - 51.5
Organosolv lignin 69.7 5.9 0.1 21.3
* by difference

3.2.2. Pyroprobe based fast pyrolysis visualization reactor

The formation of the intermediate liquid phase and bubble dynamics, were visualized with a fast camera. The studied
material is deposited on quartz disks supported on a modified CDS Pyroprobe 5000. A reactor consisting of a Pyrex
tube (No 9825), cut open at the end, screwed into a solid block of aluminum by connection ½ NPT and two lateral
connections ¼ NPT to install a pressure sensor and a valve constructed that regulates the continuous flow of
nitrogen. Shape platinum filament was modified to facilitate installation of a quartz disk of 1 cm in diameter and 0.5 mm
thick, on which rests the sample to be displayed. Originally the filament was 12.2 cm in length, 0.3556 mm thick and
forming a coil of 15 turns. The new filament has the same length and thickness, but arranged in a plane (See Figure 2-
A).

Figure 2. Pyroprobe-based fast pyrolysis visualization reactor. A: Modified heating element and quartz sample disc
with a cellulose sample. B: A view of the setup used for visualization, where the reactor on the right is lit up by a fiber-
optic light source. Nitrogen purges through the reactor, which is open on the end, and a thermocouple can be fed
through a tiny hole in the glass tube to measure temperature on the sample or quartz disc.

The visualization of the samples was performed with a fast camera (Troubleshooter Fastec LE Imagin Fastec 125
fps); using two 25 mm extension tubes each connected in series with a lens Tokina 100 mm f 2.8 framework. The
camera was positioned at an approximately 45 ° angle to the main axis of the quartz disc with the sample (see Figure
2-B). An external source of white light (Solac LB50) was coupled to improve the depth of field and focus on each
video. The amount of sample used in each experiment is approximately 0.5 mg, positioned in the center of the quartz
disc (see Figure 2-B). Analytical grade nitrogen flowed continuously at a rate of 5 ml/min to prevent the fumes
obstructing the field of view of the camera.
Chapter 3 105

3.2.3. Temperature Calibration

Calibration curves were obtained with the blank disc, and with 0.5 mg of each sample on it. Each assay was performed
in triplicate to ensure reproducibility. The Pyroprobe was programmed with heating rate of 800 °C/s, and a final
temperature of 337 °C such that the disc surface is 500 oC. In another experiment, the thermocouple was placed in the
sample pile but without touching the quartz surface to measure the sample heatup rate at the middle of the sample.
The measurements were performed with a 50 µm thermocouple (Type K, Omega), which has a response time of 2.36
ms (determined experimentally). A data acquisition system makes a record of the temperature every 0.25 s.

A
B

Figure 3. (A) Positioning the thermocouple and sample calibration curve. (B) Final temperatures on the lignin (~400
°C) and sucrose (~280 °C) are due to the fact that the thermocouple is not touching the quartz. Because char has a
low thermal conductivity, heat is not transferred well through the char, which has room-temperature nitrogen flowing
over the top.

Figure 3 shows the temperature evolution of the empty disc and the samples. The maximum heating rate achieved on
the surface of the blank disc was 181 °C/s, and an average for sucrose and lignin the order of 120 °C/s. Differences in
heating rate and final temperature are due to the physicochemical properties of carbonaceous residues formed. This
information was used to correlate the temperature of the sample and the image corresponding to each time point. This
information is critical to describe the pyrolysis events observed.

3.2.4. Processing of Visualization Data

This section describes the experimental methodology used for the characterization of bubble birth rate, growth rate,
size distribution, and bursting rate.

Control Volume: To study bubbling events within the intermediate liquid phase it was defined the control volume as the
initial volume occupied by the sample (see Figure 4). Since it is known the amount of sample used in the analysis
(typically 0.5 mg) and the actual density of both organosolv lignin (1300 kg/m3) and sucrose (1590 kg/m3), it is possible
to calculate the initial sample volume. This is the reference for the balance of populations. The cross-sectional area A is
determined using the software imageJ as the contour of the sample on the disc. The pixel area is estimated in the
region marked by the red line (Make-Binary / Analyze- Measure area). The sample volume is irregular, for purposes of
106 Chapter 3
developing balance populations an average height (ho) of the sample was estimated by dividing the volume by the
area.

Figure 4. Methodology to calculate the control volume using ImageJ using the platinum coil to set the scale.

Analysis of images: ImageJ and MatLab R 2014 were used to process the videos. The image sequences were modified
to improve the brightness and contrast, and in some cases the color map was modified to identify bubbles formed in
greater detail. In all experiments the conversion of pixels to units of length, is made taking as reference the platinum
filament diameter (0.3556 mm). Counting the bubbles in each image sequence is performed manually (see Figure 5),
using the image editor in MatLab.

Figure 5. Procedure for counting bubbles in Organosolv Lignin using Matlab’s image analysis software.

On average, each video is made up of 2 000 frames. However, in the first 200 and last 700 frames it was not possible
to observe bubbles. These frames correspond to the initial heating and char formation periods respectively. The
remaining 1100 intermediate frames is where most of the bubbling happens. It was selected the following frames for
the visualization and counting of bubbles (for sucrose: 362, 456, 556, 657, 856, 956 and 1156 for organosolv lignin:
631, 693, 725, 751, 768, 775, 812, 861, 1028, 1062). For each frame it was counted and measured each bubble using
the command ‘Measure distance’ in Matlab, this information is stored in a MatLab script for further processing. Only
bubbles that have clearly distinguishable and measurable edges were counted. As video capture speed is constant
(125 fps), it is known the exact time of each image, and the corresponding temperature value in both the quartz disk as
sample. Matrices with bubble diameter vs time were stored in MatLab.
Nucleation and growth of bubbles were monitored in frames where they could be clearly tracked. Organosolv lignin
bubble nucleation was distinguishable between frames 442 and 516 bubble growth between frames 436 - 447. In the
case of sucrose, it was studied bubble birth between the frames 514 and 559 and bubble growth between 703 and 715.
Chapter 3 107
Similarly, the number of new bubbles is counted, and the number of bubble vectors vs time are constructed to describe
the rate of nucleation of bubbles in each commodity.

Birth of new bubbles: For organosolv lignin the only period where it is possible to visualize the birth of new bubbles
without interference from the coalescence or death is between Frames 442 - 524 (3.5s - 4.13s); for sucrose this region
is between frames 514 - 559 (4.11s - 4.47s). In both cases, it was measured the number of bubbles formed per unit of
area. Birth rate of bubbles is reported as the number of bubbles per m2 s.

Bubble growth: It is difficult to track the size of each bubble because many of them interact, coalesce, and die. The
growth of the bubbles is best studied in the video regions between frames 436 - 447 for organosolv lignin and between
frames 703-715 for sucrose. It was assumed that the behavior of these bubbles is representative of the behavior of all
the bubbles. The diameter of each bubble versus time data was introduced into an Arrhenius type model (see equation
5). The derivative of the curve between bubble size vs. time represents the overall rate of bubble growth.

Bubble size distribution: In order to use this experimental information on population balances, histograms were built for
each frame using the bubble diameters, assigning bin width as the square root of the number of data. Using the
Experimental Fitting Toolbox in MatLab R2014, several distribution functions were tested, finding the best fit to the
histograms are lognormal type functions. The lognormal mean and standard deviation of the functions were calculated
in each of the selected frames. For the population balance, using continuous functions for the accurate calculation of
the partial derivatives is required. To achieve this, numerical interpolation (cubic) is made through time for the means
and standard deviations of the lognormal distribution functions to generate 1000 time steps (1000 new data between
the experimental values obtained of the videos). For each time step, there is a vector with the experimentally measured
bubble diameters, and the distribution function evaluated for each diameter. A description of bubble growth and
nucleation rate is made using the adjustable parameters. It is possible to obtain the partial derivatives of the population
balances (discretization methods) that correlate with each phenomenon.

3.2.5. Mathematical model based on global population balance to predict


microexplosion intensity and aerosol ejection intensity

Global balance of values for bubbles within the liquid phase: A mathematical model based on population balances, fed
with experimental data (rate of bubble birth, rate of bubble growth, rate of bubble coalescence and death) is proposed
to predict bubble bursting rate and aerosol ejection rate. Equation 1 is used to calculate bubble size distribution [39,40]
.

∂ (f(̅ t, D)) ∂
− (f(̅ t, D)v) = ̇ − Death
(Ḋf(̅ t, D)) + (Birth
̅̅̅̅̅̅̅ ̇ )
̅̅̅̅̅̅̅̅ (1)
∂t ∂D

Where the function f(̅ t, D) corresponds to the bubble distribution as a function of time per unit volume control and D
is diameter of bubbles (found experimentally as described in Section 2.4). v is the bubbles velocity within the molten
phase, described by Stockes law. The function f(̅ t, D) is shown in Equation 2. This represents the standard deviation
and average lognormal distribution function. The first and second term on the population balance refer to the rate of
change of the distribution function and convective transport of bubbles within the liquid phase. The control volume for
initial balance is the total volume of the sample was described in Section 2.4.
108 Chapter 3
2
(𝑙𝑜𝑔(𝐷)−𝛾(𝑡))
1 −
f(̅ t, D) = 𝑒 2𝜎(𝑡)2
(2)
Dσ(t)√2π

The third term of Equation 1 represents the growth rate of all the bubbles (𝐷̇ ). This parameter is measured
experimentally. Once the diameters of bubbles versus time are known, a parametric adjustment is made in an Arrhenius
expression for the rate of growth describer bubble versus time, as shown in Equation 3.

d(D)
= kDn (3)
dt

As a first approximation, it was supposed that this expression is valid for all bubbles. The growth rate of all bubbles
within the liquid phase was obtained by multiplying the above expression per the distribution function, and taking the
partial derivative respect to D at each time point. The terms of birth and death bubble, are quite complex to estimate
both experimentally and theoretically. Usually these terms are represented by differential expressions, which makes
the numerical solution rather complex for population balances [39,41]. Knowing the experimental information distribution
function, and growth rate of bubbles, it can be deconvoluted the main phenomena involved in the dynamics of bubble
formation, as shown in Figure 6. The partial derivatives are determined by discretization of the distribution function, as
appropriate (spatial or temporal).

Figure 6. Scheme of mathematical model diagram with the main phenomena used to describe bubble dynamics

Bubble bursting rate: Since there is no net flow of bubbles at the base of quartz disc, according to the balance sheet,
the output rate of bubbles is obtained knowing the distribution function at each instant of time (experimental), and the
rate of rise of the bubbles (stokes law). The expression representing the output rate of bubbles on the volume control
(the entire liquid sample) or equivalently the bubble bursting rate is shown in Equation 4:

Bubbles f(̅ t, D)v


Bursting rate ( )=− (4)
3
m s ho
Chapter 3 109
The above expression represents the output rate of bubbles of different diameter D at each time t, and ho is the height
of the control volume. The total number of bubbles per unit volume coming out (regardless of size) at each time “t” is
given by Equation 5:

Dmax
total bubbles f(̅ t, Di )vi
Net bursting rate ( 3
) = ∑ (− ) (5)
m s ho
i=0

Similarly, the total number of bubbles coming out at each instant of time, is obtained by integrating the above expression
through the control volume as shown in Equation 6:

Dmax
total bubbles f(̅ t, Di )vi
Total bursting ( ) = ∫( ∑ − ) dV (6)
s ho
i=0

Finally, the total number of bubbles ejected throughout the period of time is obtained by integrating the above expression
over the time domain as shown in Equation 7.

t Dmax
f(̅ t, Di )vi
Total bubbles bursted = ∫ ∫ ( ∑ − ) dV dt (7)
h
0 i=0

For bubbles with sizes between 1µm and 1 mm, velocity (v) can be estimated by Stokes law [42, 43], shown in Equation
8.

gR2 (ρl − ρb )
v= (8)
18μl

Where g is the gravity constant, R the radius of the bubble, densities liquid and bubbles, respectively, and the dynamic
viscosity of the liquid phase. This means that at each time it is possible to estimate the contribution to the rate of bubble
bursting by multiplying the rate of expression of a bubble by the distribution function.

Ejection rate: The intensity ejection aerosols, defined here as the volumetric flow of liquid droplets ejected from the
liquid phase, once collapsed gas bubbles (bubbles bursting). Paulsen et al. [44], Georgescu et al. [32], and Blanchard
et al. [33] have proposed mathematical models for the aerosol formation mechanism. However the proposed models,
although quite rigorous, only consider ejection due to the collapse of a single bubble in the gas-liquid interface, leaving
aside the dynamics of bubbles in the liquid phase that contribute to global aerosol ejection. In this proposal, it will be
used an expression to estimate the intensity of ejection of aerosols based on the dynamics of formation and transport
of gas bubbles within the liquid intermediate phase. This information is incorporated into bubble models to predict the
fraction of high molecular weight oligomers (liquid intermediate) ejected [31] as shown in Equation 9.

V Vamax
m3 aerosol 4π Naerosol 3
Iaerosol ( ) = ∫ ∑ ((Bursting rate)Di,t ( )R ) dV (9)
s 3 Nbubbles aerosol
0 i=0
110 Chapter 3
The total volume of droplets ejected as aerosols is determined by integrating vs. devolatilization time [1] as shown in
Equation 10.

V t Vamax
4π Naerosol 3
Vnet (m3 ) = ∫ ∫ ∑ ((Bursting rate)Di,t ( )R ) dV dt (10)
3 Nbubbles aerosol(i)
0 o i=0

Finally, the yield of aerosols, with respect to the initial mass of lignin or sucrose, is:

Vnetρl
Yieldaerosol = (11)
mo

3.3. Results
This section describes the experimental results first and then those obtained with the model. Experimental results
include visualization and description of the behavior of lignin and sucrose during pyrolysis at average heating rates of
120 oC/s. The model was used to calculate the dynamics of bubbling within the liquid phase and aerosol ejection
intensity.

3.3.1. Global behavior of Organosolv lignin and sucrose.

Figure 7 shows the sequence of photographs taken in for Organosolv lignin and sucrose. The first temperature value
within the parentheses corresponds to the surface of the disc, and the second is sample temperature.

Organosolv lignin Sucrose


A-Frame 236 (1.88s) (221°C-176.9°C) A-Frame 100 (0.8s) (131°C-86.9°C)

B-Frame293 (2.34s) (288°C-223.4°C) B-Frame 532 (4.25s) (386.1°C-225.9°C)


Chapter 3 111

Sucrose melted

C-Frame 343 (2.74s) (313°C-241.7°C) C-Frame 622 (4.97s) (421.9°C-253.7°C)

Gas bubble

D-Frame 547 (4.37s) (381°C-314.7°C) D-Frame 672 (5.38) (381°C-314.7°C)

Gas bubbles

Gas bubbles

E-Frame 1245 (9.96s) (499°C-384°C) E-Frame 1600 (12.8s) (484.7°C-292.9°C)


112 Chapter 3

Char layer
Char layer
Figure 7. Sequence of photographs with major changes during devolatilization of sucrose and Organosolv lignin (Disc
Temperature – Sample Temperature).

For Organosolv lignin, the appearance of the first film of molten material is seen at 1.88 s after the experiment
began. The temperature at which melting of the sample initiates is 176.9 °C (see Frame A). These data are reasonable
and close to those reported for Organosolv lignin by Dufour et al. (150 ° C) [45] and Zhou et al. (154 ° C) [46]. In frame
343 (2.74 s), the sample appears 100 % melted. At this time, the average temperature of lignin is 241.75 ° C; slight
swelling of this can be seen, and the appearance of small bubbles of gas into the liquid bubble diameter of about 33
microns (See Video 1-Supplementary material). The appearance of these microbubbles is indicative of vapor formation
within the liquid phase by fragmentation reactions of β-O bonds to form phenols, alkoxy-ketones and light gases
[47]. After this temperature, the whole dynamic of formation, growth, and death of bubble coalescence is evident. You
could observe bubbles clearly between 2.9-9.25s. (Frames 362-1156). The main transformations of sucrose during
pyrolysis are shown in Figure 7. Initially the specimen melts when it has reached 176 ° C (Frame 376-video
supplementary), and is completely melted at 225 °C (Box B). From this point on the dynamics of growth, coalescence
and death of bubbles, which can be described graphically as discussed in the next section is displayed. At the end, a
carbonaceous residue from re polymerization reactions and dehydration of fructose and glucose is produced.

3.3.2. Birth rate of bubbles in the liquid phase: Organosolv lignin

The birth of bubbles in Organosolv lignin can be seen in Figure 8. Once a group of molecules exceed a minimum
threshold energy (boiling point), the cluster is detached from the body of liquid and appears a gas-liquid interface. In
the case of pyrolysis the vapors can be produced by chemical reactions or by the vaporization of small molecules (e.g.
moisture evaporation).
A-Frame 442 Time=3.53s. B-Frame 471 Time=3.76s. C-Frame 493 Time=3.94s.
Td=312°C Ts=240°C Td=358°C Ts=295°C Td=365°C Ts=301°C

D-Frame 496 Time=3.96s. E-Frame 498 Time=3.98s. F-Frame 501 Time=4.0s


Chapter 3 113
Td=366°C Ts=301°C Td=366°C Ts=301°C Td=366°C Ts=302°C

G-Frame 505 Time=4.04 H-Frame 508 Time=4.06s I-Frame 516 Time=4.13s


Td=376°C Ts=302°C Td=376°C Ts=302°C Td=376°C Ts=306°C

Figure 8. Birth of bubbles in the intermediate stage of lignin pyrolysis.

Experimental analysis of the bubbling with number of bubbles and nucleation rates can be seen in Figure 9, where two
regimes of nucleation are evident. The first regime <430 ms (Frames A-D, Stage I) is where the bubble birth rate is
low. Only six bubbles have formed in the displayed area. In this first stage, the gas cluster formed by fragmentation of
volatile must overcome the barrier of the surface tension to form bubbles. This barrier begins to become less and less
important, as the temperature increases (Frames F-H), since the surface tension decreases, facilitating birth and
increasing growth of new bubbles (Stage II). The maximum bubble birth rate was observed once the sample has
reached 301 °C. From there, due to the intensity in the dynamics of bubbles, it was not possible to distinguish any
separate phenomenon. The birth of bubbles is homogeneous, i.e. bubbles broke off directly from within the liquid phase.
114 Chapter 3
Figure 9. Experimental and interpolated data for two stages in bubble growth and nucleation of Organosolv lignin.

The molecular theory of nucleation describes the behavior of birth of bubbles in terms of thermodynamic properties of
bulk [61] as shown in Equation 12.

∆Ecri
J = No e

k𝐵 T (12)

In the above equation J is the nucleation rate, No is a frequency factor of new bubble occurrence, ∆Ecri is the minimum
free energy required to form the first cluster gas, kB is the Boltzmann constant, and T is the temperature of the liquid
phase. By fitting the data, the frequency factor (No) is 62,166 bubbles ms -1 mm -2 and the minimum free energy is
43,317 kJ mol-1.

3.3.3. Birth rate of bubbles in the liquid phase: Sucrose

The nucleation process begins with the increase of internal energy of the bubble, which is proportional to the diameter
thereof. The thermodynamic driving force for nucleation and growth of bubbles is the vapor pressure of the volatiles
produced. During fast pyrolysis of cellulose and sugars, the major compounds released into the gas phase are
permanent gases (H2, CO2, CO), organic acids (acetic, formic), furfural, 5-HMF, and glyoxal. The anhydrosugars formed
are levoglucosan, cellobiosan, among others; they have very high boiling points and possibly are not part of the gas
inside the bubble. Instead, these sugars can be ejected as aerosols or evaporated directly from the liquid phase. Figure
10 shows the sequence of pictures describing the birth, growth and coalescence of bubbles in the liquid intermediate
phase of sucrose pyrolysis.

A-Frame 514 Time=4.11s. B-Frame 522. Time=4.17s. C-Frame 539. Time=4.31s.


Td=381°C. Ts=221°C. Td=383°C. Ts=223°C. Td=390°C. Ts=230°C.

D-Frame 554 Time=4.43s. E-Frame 556. Time=4.44s. F-Frame 559. Time=4.47s.


Td=391°C. Ts=231°C. Td=392°C. Ts=232°C. Td=395°C. Ts=235°C.

Figure 10. Sequence of frames to describe the birth of bubbles in the liquid phase intermediate formed during the fast
pyrolysis of sucrose.
Chapter 3 115
Similar to lignin, nucleation of sucrose bubbles is very fast and difficult to monitor. Bubbles merge very quickly with
other bubbles, making it difficult to count the birth of new bubbles. However, between frames 514-559 (see
supplementary video), it was possible to follow the birth of bubbles in isolation. Only seven frames in which the changes
shown in the number of bubbles (see Figure 10) were observed. Better visualization can be seen in the original video,
shown in Supplemental Material. The bubble birth rate is constant (see Figure 11). In the studied area and times used
to estimate the nucleation, the number of bubbles per mm 2 increases linearly. This behavior is that in the range studied
time, the sample temperature is practically constant (only increases 14 ° C).

Figure 11. Nucleation rate for Sucrose.

As for Organosolv lignin, the Arrhenius expression (equation 14) was fit to the experimental data to estimate the
nucleation rate at any time t. The results show that ∆Ecri is 69.45 kJ mol-1 and No is 2.86·10 5 bubbles ms-1 mm-2. These
results should be viewed as a starting point, since this is for lower temperature and bubble nucleation is certainly more
intense at higher temperatures, where it is difficult to track them with the equipment.

3.3.4. Bubble growth of Organosolv lignin and sucrose

It is difficult to track each bubble to estimate its growth rate. Once formed bubbles interact with others to coalesce and
die. For Organosolv lignin, a single bubble isolated from the others was identified between the frames 436-447 and was
tracked over time (see Figure 12). For sucrose, a solitary bubble was tracked between frames 703-713 (80 ms). In this
short period of time the temperature of the disc and the sample are on average 440 °C and 275 °C respectively.
116 Chapter 3

Figure 12. Monitoring the growth of a gas bubble in the molten lignin.

The Organosolv lignin bubble expanded for 88 ms before growth stopped. In this time interval, the temperature of the
disc, and sample have changed very little (< 4 °C), averaging 365 °C and 288 °C respectively (pseudo-isothermal
growth process). In the literature, the bubble growth data is normally reported for low temperature systems (mass
transfer or evaporation) [10, 13, 14, 54, 61]. It is difficult to find bubble growth studies in chemical reaction systems with
organic substances. Bubbles growth in thermoplastic materials, such as Organosolv lignin, is mostly due to the
decomposition of the liquid located in the internal interface of the bubble. The solubility of gases at high temperatures
and atmospheric pressure is very low, so growth is discarded due to mass transport from within the fluid. Due to lack
of data on the intermediate liquid properties (density, surface tension, chemical reaction rate, viscosity), it is difficult
to propose and solve a phenomenological level model that describes the growth of bubbles. However it was used
experimental data that was parametrically adjusted, assuming a linear behavior of order n (see Eq 5) to find n to be -2
and k to be 1575 um 3 ms -1.
Chapter 3 117

Figure 13. Experimental data for bubble size and bubble growth rate of Organosolv lignin versus time with model
fitting to describe the Arrhenius growth rate.

Previous studies on lignin pyrolysis have shown that after 250 ° C the liquid phase of the lignin begins to resolidify
(polycondensation reactions), substantially increasing its viscosity [45]. Based on this fact, and considering that the
solubility of gases is low these conditions [48,49], it was assumed that viscous stress predominates over gravitational,
surface tension, and expansion work to control the gas bubble growth dynamics. The overall performance is second
order, fulfilling the d 2 -law [8,50,51]. As the bubble growth was isothermal, the gas expansion effects should not be
important; instead, the growth is due to accumulation of gases from chemical decomposition reactions in liquid
phase. That is a direct relationship between the rate of bubble growth and the rate of generation of gas by decomposition
reaction. Adjust this kinetic coefficient Arrhenius type model, it can be estimated roughly the value of the activation
energy for the decomposition reaction in the liquid phase. Assuming values of 7.76 * 10 6 as reported by Dufour, et al.
[52], the activation energy is fairly low, but similar to that reported for organic liquids pyrolysis bio- oil [53]. For sucrose,
in the image sequence of Figure 14 (see supplementary video) in section I (Frame 712), the bubble has an average
diameter of 0.3894 mm. After this maximum size, the bubble ceases to grow.
118 Chapter 3

Figure 14. Monitoring the growth of a gas bubble into the molten sucrose.

For the experimental data of bubble sizes it was fitted an Arrhenius model to determine the growth rate. In Figure15, it
is observed that the bubble size increases almost linearly with time, with the highest growth rate in the first 5 ms,
decreasing very slowly after 20 ms. The experimental fitting as proposed in equation 5 gives parameters n = -0.15, k =
7.8193 (ms -1).

Figure 15. Experimental data for bubble size and bubble growth rate of sucrose versus time with model fitting.
Chapter 3 119

By comparison, the growth rates of organosolv lignin, and sucrose, are actually on the same order of magnitude.The
rate of bubble growth is dominated by many forces that interact simultaneously (viscous forces, drag forces, surface
tension, etc.). The differences in behavior between the growth rate of the Organosolv lignin and sucrose may be related
to changes in temperature depending on the physicochemical properties such as viscosity and surface tension and also
on the rate of volatile matter formation. It was not possible to find values of viscosity and surface tension of Organosolv
lignin and sucrose. Making a comparison between the intermediate phase of lignin to coal tars at 260 °C, the viscosity
and surface tension are on the order of 0.004 Pa·s and 0.015 N m-1 respectively [54]. By contrast, data on viscosity and
surface tension for the cellulose liquid intermediates are the order 10 -5 Pa·s and 10 -6-10 -5 N m-1 [4]. As such, it would
be expected that sucrose at these conditions has a lower surface tension and viscosity, which allow for more rapid
bubble growth.

3.3.5. Distribution of the bubbles within the liquid phase of Organosolv lignin

Histograms were created to visualize the distributions of bubble diameters in bins (see Figure 16). The number of
intervals (bins) was found by taking the square root of the number of total data. The distribution at each time step was
fit to a log-normal distribution curve. Initially, all the bubbles are of small diameter, with a mean of 183 μm.

In the range of temperatures studied, two regimes are observed. In the first 4 images (Figure 16 A-D), the number of
bubbles is approximately constant (50 bubbles), but the distribution is becoming wider (increased standard
deviation). This indicates that individual bubbles grow and increase in size. Bubbles escaping the system will likely be
replaced by new bubbles but the system is unlikely to achieve “steady state” since it is a small batch experiment. This
temperature range (250-380 °C) is known to have the maximum rate of volatiles release in lignin. The main products
are alkoxy phenols, alkoxylated phenol-ketone products as the β-O-4 lignin bonds break [47]. According Dufour et al.
[45,55], the viscosity of the lignin liquid intermediate increases in this temperature regime.
120 Chapter 3
Figure 16. Change in the distribution function of the bubbles to the molten phase of Organosolv lignin.

In the second regime (Figure 16 E-H), the total number of observable bubbles decreases continuously from 40 to 6
bubbles; the death and bursting rates are higher than the birth rate of bubbles. In this region dehydration reactions and
crosslinking should be very intense, promoting the production of char (i.e. re-condensation of the melt phase) which
dramatically increases the viscosity of the liquid intermediate phase as outgassing is reduced. After 9.25s, no new
bubbles are formed.

3.3.6. Distribution of the bubbles within the liquid phase to sucrose.

The same histograms were produced with the data for sucrose, which also exhibited Log-normal distribution (see Figure
17).

Figure 17. Change in the distribution function of the bubbles to the molten phase of sucrose.

At first, a few gas bubbles with an average diameter of 700 μm are formed (A-C) as a result of fragmentation or
dehydration reactions. In this time interval, both viscosity and surface tension of the liquid phase should be rather high,
slowing the birth and growth of bubbles. After 5 s (boxes D-F), the bubbles have a greater diameter, and the frequency
of formation also increases, partly because the surface tension and viscosity decreased by increase in temperature,
this favors the birth rate and growth bubble directly. The rate of bubble growth and the birth rate are important in this
period, in contrast to death rates and bursting (see Figure 17). There are abrupt changes in the distribution functions
(boxes F-G) caused by swelling and large bursting bubbles, after which new bubbles coalesce with smaller. Finally, as
with lignin, the liquid phase solidifies due to dehydration and polymerization reactions which increase viscosity, limiting
formation and growth of bubbles. The distribution is characterized by small diameter bubbles which give porous texture
end of carbonized. Importantly, after 7 s (Frame 916, supplemental video), the liquid phase begins to become dark
because of the polymerization reactions. The distributions of bubbles in the H-J boxes take into account only the bubbles
seen in the outer surface of the molten phase.
Chapter 3 121

3.3.7. Population balance to predict bubble bursting rate, bubbles coalescence, and
ejection intensity

In this section it is described the net rate of birth and death of bubbles using the population balance model. The
distribution functions for bubble birth and death and the rate of bubble growth obtained in the previous section are used.

Birth and death rates of bubbles: In Figure 18, the deconvolution of the distribution function is shown. The coalescence
and death events are very intense for very small bubbles (up to 60 µm). These results have been reported by other
investigators [56,57] for bubbles in viscous media. The curves shown in Figure 18 A and C, have positive and negative
peaks at each time t. When the peak is positive (birth rate > death rate), bubble nucleation or fragmentation
predominates. When curves are negative (birth rate < death rate), bubble coalescence dominates. It is quite interesting
to note that for each time t, the coalescence is favored for smaller bubbles in phase liquid for Organosolv lignin; however
the trend is opposite for sucrose (fig. 18-C) where the probability of bubble coalescence is inversely proportional to
the bubble diameter [39,40]. Similarly, the larger bubbles tend to fragment into multiple new bubbles. It is interesting
to note the behavior of the curve in Figure 18-A (9.5 s), where there is a huge positive peak for small diameter bubbles
but it is zero for diameters > 12 µm.

As mentioned previously, when the lignin melt phase exceeds 250 °C it begins to re-solidify, altering the dynamics of
bubbles inside. As the viscosity increases, the chances of meeting between bubbles decreases considerably [41]. Under
these conditions very small bubbles grow slowly and are virtually static. These bubbles are responsible for the final
formation of cenospheres (hollow spheres) after pyrolysis. Figure 18 B and D, represent the net rate of birth and death
averaged bubble. This is obtained by totaling all bubbles distribution at each time step. Analyzing the overall
performance of Figures 18 B and D, it is noted that on average and in the first 9.5s reaction the death rate of bubbles
is the predominant phenomenon (the rate of death by fragmentation and nucleation is more intense than bubble
coalescence). At the end, the liquid phase is very viscous as the liquid polycondenses to form a carbonaceous residue.
122 Chapter 3

Figure 18. Representation of the distribution functions of the net rate of birth and death of bubbles during
devolatilization of Organosolv lignin and Sucrose. A. Net rate of birth and death of bubbles within the liquid phase of
the Organosolv lignin. B. Total net rate of birth and death of bubbles within the liquid phase of the Organosolv
lignin. C. Net rate of birth and death of bubbles within the liquid phase of sucrose. D. Total net rate of birth and death
of bubbles within the liquid phase of sucrose.

Bubble Bursting: Experimentally, for lignin and sucrose it is difficult to track the aerosols as they leave the liquid
phase. This phenomenon is quite fast and difficult to capture with the fast speed camera. However, in Figure 19 a
sequence of images acquired every 8 ms shows gas bursting. For organosolv lignin, the frames 389-527, show the
morphological changes of the liquid surface once the bubbles collapse. Although there are many bubbles, it is only
possible to see substantial changes in the central bubble, which is larger. In the boxes B-E, the formation and growth
of a bubble in the lower right is observed. This bubble collapses (F), and deforms the entire surface of the liquid phase,
by shrinking (G-H). This process is continuous and regular, until re-condensation reactions.

There are two mechanisms for release of aerosols by breaking bubbles. The first is known as “film bursting”, in which
the film formed between the bubble and the liquid film at the liquid-gas interface atomizes when the bubble
collapses. Film bursting produces very small aerosols (< 1 µm) in very short times (< 300 μs) [32]. The second
mechanism (Frames I-O), is generated after the collapse of large bubbles, curving the outer surface of the liquid which
then contracts and converges axially, producing 'jets' that release large aerosols (order of mm in size) [30,32]. Frames J-
N show a region of high curvature generated by the collapse of the bubble in Frame I. At the beginning is a large
opening, which quickly closed (Frames K-I). Throughout the video sequence the formation of jets is not directly observed
(nor in any videos taken). Previous studies have shown that for a jet to be formed, the Ohnesorge (Oh) number should
not exceed a certain critical value [30,58]. This number is the ratio of viscous forces to inertial forces and tension
superficial.

𝜇𝑙
𝑂ℎ = (13)
√𝜌𝑙 Ω𝑅
Chapter 3 123

Where𝜇𝑙 , 𝜌𝑙 Ω , are the viscosity, density and the surface tension of the liquid phase, 𝑅 is the gas bubble radius within
the liquid phase. Systems with Ohnesorge number (Oh) over 0.037, do not produce jets because the capillary forces
that direct the movement of the liquid surface are damped by viscous forces [58]. It will be interesting to have an
estimate of what value it has for lignin oligomers. The Bond number is another parameter to consider when there is no
formation of jets because gravity is important, altering the balance, shape and movement [58].

Figure 19. Sequence of formation and collapse of bubbles in liquid-gas interface during the fast pyrolysis of
organosolv lignin.

Using data of viscosity (0.004 Pa·s), surface tension (0.015 N m -1) reported in Hwang et al. [54] for tar coal at high
temperatures and, the estimated density of the liquid phase [54] the bubble diameter was estimated as 1.025 mm (see
section 2.4). Figure 19 shows the bursting rate of lignin bubbles, which depends on diameter and time using the
methodology described in chapter 2 sections 2.4 and 2.5.
Figure 20 shows the sequence of frames that describe the collapse of bubbles within the liquid phase produced during
the pyrolysis of sucrose. Frames A shows the destruction of a large bubble and formation of a toroid shaped cavity,
which begins to contract gradually. As the shrinkage continues, gases are released due to a difference between the
vapor pressure and atmospheric pressure. In this sequence the formation of jets was not appreciated. However, in
Frames F a small jet is formed.
124 Chapter 3
A. Frame 687 Td=338ºC Ts=267ºC. B. Frame 688 Td=340ºC Ts=270ºC.

Bubble before explosion Hole left after bubble explosion

C. Frame 689 Td=338ºC Ts=267ºC. D. Frame 690 Td=340ºC Ts=270ºC.

Hole contraction Hole filled

E. Frame 766 Td=338ºC Ts=267ºC. F. Frame 768 Td=340ºC Ts=270ºC.

Before jet ejection Jet ejection

Figure 20. Sequence formation, breaking bubbles (Bursting) and release of jets, during the fast pyrolysis of sucrose.

A second event occurs after 8 ms wherein the central bubble collapses, releasing aerosols very small in diameter (12
µm) (see supplementary video). This has also been reported by other investigators [1,10,73]. During this stage, many
small aerosols are released in all directions at a very high speed, making it difficult to capture them, even with 1000 fps
[1]. The number of aerosols released by fragmentation of jets is directly linked to the diameter of the mother bubble. The
relationship between the size of aerosol and mother bubble is on average close to those reported elsewhere
[1,31,32]. Some researchers have proposed empirical correlations to relate these two variables [11, 21, 22, 48]. Using
equation 6, proposed by Zhang, et al [31]for a mother bubble 890 microns (Frame F), the number of theoretical aerosols
Chapter 3 125
formed is 2.7, very close to the two aerosols experimentally observed. Such correlations becomes important when
considering particle level models, as they are the bridge connecting the bubbling events with the release of
aerosols. Figure 21 shows the theoretical behavior of the predicted output bubbles with the methodology described in
section 2.5 for both Organosolv lignin and sucrose.

Figure 21. Bubble bursting rates for individual bubbles and the sum of bursting rates for Organosolv lignin and
sucrose. A. Bubble bursting rates for Organosolv lignin pyrolysis. B. Total Bubble bursting rate for Organosolv lignin
pyrolysis. C. Bubble bursting rates for sucrose pyrolysis. D. Total bubble bursting rate for sucrose pyrolysis .

Figure 21 clearly shows that the bubble bursting rate is fluctuating due to the balance between formation and explosion
of bubbles combined with liquid relaxation. Subplot A shows changes in bursting rate versus time and bubble diameter
for lignin. For example, the black curve (t = 4.5s) represents the distribution of bursting rates of bubbles in 4.5s. On the
same curve, the point of intersection of the two dotted lines represents the output rate of 0.15 mm diameter bubbles in
4.5 s. From the A and C curves, it is inferred that at low times when the generation of steam is very intense, gas bubbles
126 Chapter 3
tend to break while they are very small; when generation of vapors slows down, the bubbles tend to break at larger
diameters. In the region where there is intense release of volatiles, the viscosity of the liquid phase increases
dramatically so the bubbles require more energy to breach the surface. In these stages bubbles increase their internal
energy by coalescing with other bubbles or by expansion of stored gases. The rate of bubble bursting in lignin is highest
for bubbles with diameters smaller than 200 µm, 500 µm being the maximum. In sucrose however, the bubble bursting
rate is larger, contributing to larger bubbles (> 200 µm).

Subplots B and D shows the total intensity of bursting for both materials. Total bursting rate of bubbles was determined
by adding at each instant of time, exit rates of bubbles of different diameters present in the distribution function. Bursting
intensity is greater for the liquid phase of the Organosolv lignin, which produces more bubbles, but smaller bubbles
compared to sucrose. As a result the rate of ascent of the bubbles is increased, reducing the residence time within the
liquid phase. Bursting frequency factor bubble (1 / s), is linked to vb / h ratio, which determines the inverse of the
residence time of the bubbles within the liquid within. This is a first approximation which it is considered relevant and
valid for the study of the intensity of aerosol ejection due to bubble collapse in the liquid phase during pyrolysis of
biomass. This phenomenon is quite complex; other researchers have proposed other correlations depending on the
size of bubbles and fluid properties to estimate the bursting frequency of bubbles [8,9,12,17,59]

Aerosol ejection intensity: The formation and collapse of bubbles within the liquid phase generated during pyrolysis
intermediate biomass and / or pseudo-components is the driving force for ejecting aerosol [3]. Discussion related to
aerosol ejection is of vital importance in the field of fast pyrolysis of biomass. This phenomenon may explain the
presence of high molecular weight compounds in the bio-oil as oligomeric anhydro-sugars and lignin (lignin pyrolytic)
[4, 5, 60]. The actual content of oligomers in bio-oils is not well known. The oligomers derived from lignin can be easily
separated by cold water precipitation, however, there are no methods today to predict the oligomers derived from
cellulose. By understanding the formation of the liquid phase during devolatilization, one can predict theoretically, the
amount of these compounds which can be ejected. Here it is proposed a simple methodology to predict the rate of
thermal ejection based on the dynamics of bubble formation. Aerosol distribution is directly related to the bubbling
properties. The coalescence of bubbles, fragmentation, growth, evaporation, and other phenomena can alter the final
size distribution of aerosols [61]. Research related to the ejection of aerosols during pyrolysis of biomass is still in its
infancy and requires greater efforts to consolidate experimental and theoretical results. Figure 22 shows the aerosol
distribution function of time and diameter of the aerosol droplets ejected both Organosolv lignin and sucrose.
Chapter 3 127

Figure 22. Distribution aerosol function of time and droplet diameter.

Subplots A-D represent the distribution of aerosol ejection rate at different times depending on the diameter of aerosol.
For both materials, the maximum rate of ejection occurs with small diameter aerosols (<10 µm). Dauenhauer et al. [3]
found similar sizes for aerosols ejected during fast pyrolysis of cellulose.
128 Chapter 3
The B and E subplots represent the total aerosol ejection rate per unit volume of sample. The aerosol ejection intensity
of organosolv lignin higher after 6.3 s where bubbles are larger than 0.1 mm. As discussed, two mechanisms can
produce aerosols form sucrose. In each of these events, the distribution function is centered on bubble diameters 0.2
mm (see Figure 22). The C and F subplots represents the total volume of aerosols accumulated over time. This data is
useful to estimate the quantity of aerosols generated per mg of sample used. The amount of sample used for the
experiment was 0.5 mg, assuming sprays Sucrose lignin and have a density of 1000 kg/m3. The model predicts aerosol
ejection yields 21.18% w/w for devolatilization of organosolv lignin and 17.40 % w/w for sucrose.

Yields of lignin oligomers have been reported in the literature on the order of 8 and 30 wt. % from pyrolysis of biomass
and 14 - 24 wt. % for lignin [62, 63]. For cellulose, yields of oligomer anhydrosugars ( > 2 degree of polymerization) are
on the order of 8 % and 400 °C [64], and 20 % for Avicel cellulose (2 < Dp < 7) [60]. This reveals that the results are
in the correct range compared with experimental data.

3.4. Conclusions
A new methodology has been proposed to describe of the dynamics of bubble formation in Organosolv lignin and
sucrose that combines fast speed visualization (125 fps) with mathematical modeling. The model uses a population
balance to theoretically predict overall rates of bubble birth and death, bubble bursting, and aerosol
ejection. Experimentally, it was observed that gas bubbles follow a log-normal distribution versus bubble size within the
liquid intermediate phase for both materials. This distribution function changes over time due to increased viscosity
from resolidification reactions that generate char. The model predicts aerosol ejection yields of 21.18% w/w from
Organosolv lignin and 17.40 % w/w from sucrose during pyrolysis.
Chapter 3 129

3.5. References
[1] P. Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P. and
Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011)
4306.
[2] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[3] Paul J Dauenhauer, Joshua L Colby, Christine M Balonek, Wieslaw J Suszynski, Reactive boiling of cellulose
for integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[4] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[5] S. Zhou, M. Garcia-Perez, B. Pecha, A.G. McDonald, R.J.M. Westerhof, Effect of particle size on the
composition of lignin derived oligomers obtained by fast pyrolysis of beech wood, Fuel. 125 (2014) 15–19.
[6] J.R. Grace, T. Wairegi, J. Brophy, Break-up of drops and bubbles in stagnant media, Can. J. Chem. Eng. 56
(1978) 3–8.
[7] S. Saka, M. Asmadi, H. Kawamoto, Gas- and solid/liquid-phase reactions during pyrolysis of softwood and
hardwood lignins, J. Anal. Appl. Pyrolysis. 92 (2011) 417–425.
[8] K. Butler, A numerical model for combustion of bubbling theroplastic materials in microgravity, (2002) 70.
[9] Spiel, The number and size of jet drops produced by air bubbles bursting on a fresh water surface, J. Geophys.
Res. 99296 (1994) 289–10.
[10] M.K. Tripathi, K.C. Sahu, R. Govindarajan, Dynamics of an initially spherical bubble rising in quiescent liquid,
Nat. Commun. 6 (2015) 6268.
[11] B.R. Fu, C. Pan, Bubble growth with chemical reactions in microchannels, Int. J. Heat Mass Transf. 52 (2009)
767–776.
[12] Myongsook S., Oh William A., Peters Jack B. Howard, An experimental and modeling study of softening coal
pyrolysis, AIChE J. 35 (1989).
[13] S. Basu, A. Miglani, Combustion and heat transfer characteristics of nanofluid fuel droplets: A short review, Int.
J. Heat Mass Transf. 96 (2016) 482–503.
[14] A. Miglani, S. Basu, R. Kumar, Suppression of instabilities in burning droplets using preferential acoustic
perturbations, Combust. Flame. 161 (2014) 3181–3190.
[15] Williams, Ignition and burning of single liquid droplets, Acta Astronaut. 12 (1985) 547–553.
[16] Kim, D.I. Yu, D.W. Jerng, M.H. Kim, H.S. Ahn, Review of boiling heat transfer enhancement on
micro/nanostructured surfaces, Exp. Therm. Fluid Sci. 66 (2015) 173–196.
[17] S. Das, D.S. Kumar, S. Bhaumik, Experimental study of nucleate pool boiling heat transfer of water on silicon
oxide nanoparticle coated copper heating surface, Appl. Therm. Eng. 96 (2016) 555–567.
[18] M. Shojaeian, A. Koşar, Pool boiling and flow boiling on micro- and nanostructured surfaces, Exp. Therm. Fluid
Sci. 63 (2015) 45–73.
[19] C.K. Law, Recent advances in droplet vaporization and combustion, Prog. Energy Combust. Sci. 8 (1982) 171–
201.
[20] Avulapati, L.C. Ganippa, J. Xia, A. Megaritis, Puffing and micro-explosion of diesel–biodiesel–ethanol blends,
Fuel. 166 (2016) 59–66.
[21] Teixeira, R.J. Hermann, J.S. Kruger, W.J. Suszynski, L.D. Schmidt, D.P. Schmidt, P.J. Dauenhauer,
Microexplosions in the upgrading of biomass-derived pyrolysis oils and the effects of simple fuel processing,
ACS Sustain. Chem. Eng. 1 (2013) 341–348.
[22] S. Takeda, M. Fuchihata, T. Ida, Observation of microexplosions in spray flames of light oil-water emulsions
(3rd report, influence of the diameter of dispersed water droplets on the spray flame structure), Trans. Japan
Soc. Mech. Eng. Ser. B. 74 (2008) 1649–1654.
[23] H.Z. Sheng, L. Chen, Z.P. Zhang, C.K. Wu, C. An, C.Q. Cheng, The droplet group microexplosions in water-in-
oil emulsion sprays and their effects on diesel engine combustion, Symp. Combust. 25 (1994) 175–181.
[24] M.Y. Khan, Z.A.A. Karim, a. R.A. Aziz, I.M. Tan, Experimental investigation of microexplosion occurrence in
water in diesel emulsion droplets during the Leidenfrost effect, Energy & Fuels. 28 (2014) 7079–7084.
[25] M. Garcia-Perez, P. Lappas, P. Hughes, L. Dell, a Chaala, D. Kretschmer, C. Roy, Evaporation and
Combustion Characteristics of Biomass Vacuum Pyrolysis Oils, IFRF Combust. J. 200601 (2006) 1–27.
130 Chapter 3
[26] C. Di Blasi, Modelling the fast pyrolysis of cellulosic particles in fluid-bed reactors, Chem. Eng. Sci. 55 (2000)
5999–6013.
[27] C. Di Blasi, Physico-chemical processes occurring inside a degrading two-dimensional anisotropic porous
medium, Int. J. Heat Mass Transf. 41 (1998) 4139–4150.
[28] Hou, F.M. Rizal, T.-H. Lin, T.-Y. Yang, H.-P. Wan, Microexplosion and ignition of droplets of fuel oil/bio-oil
(derived from lauan wood) blends, Fuel. 113 (2013) 31–42.
[29] Yusuf Sezen, Internal mass transfer considerations during the pyroly sis of an isolated spherical softening
coal particle, Lnt. J. Heat Mass Transf. 32 (1989) 1992–1987.
[30] J.S. Lee, B.M. Weon, S.J. Park, J.H. Je, K. Fezzaa, W.-K. Lee, Size limits the formation of liquid jets during
bubble bursting., Nat. Commun. 2 (2011) 367.
[31] J. Zhang, J.J.J. Chen, N. Zhou, Characteristics of jet droplet produced by bubble bursting on the free liquid
surface, Chem. Eng. Sci. 68 (2012) 151–156.
[32] S.C. Georgescu, J.L. Achard, É. Canot, Jet drops ejection in bursting gas bubble processes, Eur. J. Mech.
B/Fluids. 21 (2002) 265–280.
[33] D.C. Blanchard, The size and height to which jet drops are ejected from bursting bubbles in seawater, J.
Geophys. Res. 94 (1989) 10999.
[34] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Effects of particle size on the fast
pyrolysis of oil mallee woody biomass., Fuel. 88 (2009) 1810–1817.
[35] Roel Johannes Maria Westerhof, Refining fast pyrolysis of biomass, Universtity of Twentee, 2011.
[36] Ciesielski, M.F. Crowley, M.R. Nimlos, A.W. Sanders, G.M. Wiggins, D. Robichaud, B.S. Donohoe, T.D. Foust,
Biomass particle models with realistic morphology and resolved microstructure for simulations of intraparticle
transport phenomena, Energy and Fuels. 29 (2015) 242–254.
[37] M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W.J. Lee, C.Z. Li, Effects of temperature on the formation of
lignin-derived oligomers during the fast pyrolysis of Mallee woody biomass, Energy and Fuels. 22 (2008) 2022–
2032.
[38] Westerhof, Refining fast pyrolysis of biomass, Thesis Degree of Doctor of Science, University of Twente, 2011.
[39] D. Ramkrishna, Population Balances, Academic Press, 2000.
[40] Rao, Simulations for modelling of particulate processes using discrete particle model (DPM), Otto-von-Guericke
Universität Magdeburg, 2009.
[41] S. Orvalho, M.C. Ruzicka, G. Olivieri, A. Marzocchella, Bubble coalescence: Effect of bubble approach velocity
and liquid viscosity, Chem. Eng. Sci. 134 (2015) 205–216.
[42] M.I. Urseanu, Guit, A. Stankiewicz, G. van Kranenburg, Lommen, Influence of operating pressure on the gas
hold-up in bubble columns for high viscous media, Chem. Eng. Sci. 58 (2003) 697–704.
[43] M.I. Urseanu, Scaling up bubble column reactors, University of Amsterdam (UvA), 2000.
[44] M.S. Mettler, S.H. Mushrif, A.D. Paulsen, A.D. Javadekar, D.G. Vlachos, P.J. Dauenhauer, Revealing pyrolysis
chemistry for biofuels production: Conversion of cellulose to furans and small oxygenates, Energy Environ. Sci.
5 (2012) 5414–5424.
[45] A. Dufour, M. Castro-Diaz, N. Brosse, M. Bouroukba, C. Snape, The origin of molecular mobility during biomass
pyrolysis as revealed by in situ 1H NMR spectroscopy, ChemSusChem. 5 (2012) 1258–1265.
[46] M. Zhou, S., Pecha, B., van Kuppevelt, M., McDonald, A.G. and Garcia-Perez, Slow and fast pyrolysis of
Douglas- Fir Lignin:Importance of liquid intermediate formation on distribution of products, Biomass and
Bioenergy. (2014) 398–409.
[47] Custodis, C. Bährle, F. Vogel, J.A. van Bokhoven, Phenols and aromatics from fast pyrolysis of variously
prepared lignins from hard- and softwoods, J. Anal. Appl. Pyrolysis. 115 (2015) 214–223.
[48] Aydin Akgerman, Prediction of diffusivity of the gas-liquid, Ind. Eng. Chem. Fundam. 17 (1972) 372–377.
[49] M. Zirrahi, H. Hassanzadeh, J. Abedi, Prediction of CO2 solubility in bitumen using the cubic-plus-association
equation of state (CPA-EoS), J. Supercrit. Fluids. 98 (2015) 44–49.
[50] J. Lambert, I. Cantat, R. Delannay, R. Mokso, P. Cloetens, J.A. Glazier, F. Graner, Experimental growth law for
bubbles in a moderately “wet” 3D liquid foam, Phys. Rev. Lett. 99 (2007).
[51] R.S. Subramanian, R. Balasubramaniam, G. Wozniak, Fluid mechanics of bubbles and drops, Phys. Fluids
Microgravity. (2002) 29.
[52] R. Bounaceur, B. Ouartassi, A. Dufour, A. Zoulalian, Modelling intra-particle phenomena of biomass pyrolysis,
Chem. Eng. Res. Des. 89 (2011) 2136–2146.
Chapter 3 131
[53] S. Liu, M. Chen, Q. Hu, J. Wang, L. Kong, The kinetics model and pyrolysis behavior of the aqueous fraction
of bio-oil., Bioresour. Technol. 129 (2013) 381–6.
[54] S. Hwang, C. Tsonopoulos, Density , Viscosity , and Surface Tension of Coal Liquids at High Temperatures
and Pressures, (1982) 127–134.
[55] A. Dufour, M. Castro-Díaz, P. Marchal, N. Brosse, R. Olcese, M. Bouroukba, C. Snape, In Situ Analysis of
Biomass Pyrolysis by High Temperature Rheology in Relations with 1H NMR, Energy & Fuels. 26 (2012) 6432–
6441.
[56] P.C. Duineveld, Bouncing and coalescence of bubble pairs rising at high Reynolds number in pure water or
aqueous surfactant solutions, Appl. Sci. Res. 58 (1998) 409–439.
[57] L. Doubliez, The drainage and rupture of a non-foaming liquid film formed upon bubble impact with a free
surface, Int. J. Multiph. Flow. 17 (1991) 783–803.
[58] Walls, L. Henaux, J.C. Bird, Jet drops from bursting bubbles: How gravity and viscosity couple to inhibit droplet
production, Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 92 (2015) 4–7.
[59] E. Ghabache, A. Antkowiak, C. Josserand, T. Séon, On the physics of fizziness: How bubble bursting controls
droplets ejection, Phys. Fluids. 26 (2014).
[60] J. Piskorz, A. Vladars-Usas, D. Radlein, P. Majerski, D.. Scott, Flash pyrolysis of cellulose for production of
anhydro-oligomers, J. Anal. Appl. Pyrolysis. 56 (2000) 145–166.
[61] L. Ghazaryan, Aerosol dynamics in porous media, University of Twentee, 2000.
[62] Y. Wang, S. Wang, F. Leng, J. Chen, L. Zhu, Z. Luo, Separation and characterization of pyrolytic lignins from
the heavy fraction of bio-oil by molecular distillation, Sep. Purif. Technol. 152 (2015) 123–132.
[63] C.A. Mullen, A.A. Boateng, Characterization of water insoluble solids isolated from various biomass fast
pyrolysis oils, J. Anal. Appl. Pyrolysis. 90 (2011) 197–203.
[64] X. Gong, Y. Yu, X. Gao, Y. Qiao, M. Xu, H. Wu, Formation of anhydro-sugars in the primary volatiles and solid
residues from cellulose fast pyrolysis in a wire-mesh reactor, Energy and Fuels. 28 (2014) 5204–5211.
132 Chapter 4

4. Chapter 4. Methodology for estimating the distribution and


intensity of ejection of aerosol ejected during pyrolysis of
bagasse and models of their components
(Paper submitted to Journal of Analytical and Applied Pyrolysis)

Abstract: In this chapter, it is proposed a new methodology to capture and visualize aerosols that are thermally ejected
from the liquid intermediate of pyrolysis for sugarcane bagasse and its pseudo-components (Avicel cellulose, xylan and
organosolv lignin). Tests were conducted on a hot plate reactor at 500 °C with two different pressures (150 mbar and
900 mbar) and two heating rates (10 °C/s and 1200 °C/s). The aerosols were collected on a heated slide located across
from the sample. The slides were visualized with a scanning electron microscope (SEM) to determine the number and
diameters of the droplets. Results showed that vacuum and the high heating rate (150 mbar and 1200 °C/s) favored
thermal ejection of aerosols. The aerosol yield for sugarcane bagasse was lower to those obtained for cellulose and
Organosolv lignin, and heating rate enhanced aerosol release more dramatically than pressure.

Keywords: Aerosol, Distribution of aerosols, Cellulose, Organosolv lignin, Xylan, Sugarcane bagasse, Hot plate
reactor, SEM.
Chapter 4 133

4.1. Introduction
Biomass pyrolysis is a thermochemical process with great potential for the production of liquid fuels [1,2]. A brownish-
black, viscous, highly oxygenated liquid called bio-oil is the main product of fast pyrolysis [1,3–5]. The yield and
composition of this bio-oil depend on the operating conditions of the reactor (temperature, heating rate, carrier gas), as
well as on the particle size and nature of the feedstock. High bio-oil yields (more than 60 wt. %) are typically obtained
when the particles are heated very fast (> 100 °C/s) and with small biomass particles (< 2 mm) [3,6].

The chemical compositions of these oils are very complex, covering a wide range of molecular weights and
functionalities [7–9]. The most commonly characterized fraction of the oil is made up of water and volatile compounds
detectable by GC/MS. These molecules are released from the reacting particle by evaporation. Another important
fraction of the bio-oil is made up of oligomeric molecules with boiling points higher than the pyrolysis temperature; these
molecules cannot evaporate and are understood to be released by thermal ejection [10,11]. Although the oligomeric
products are not well characterized, researchers have reported yields between 12 and 20 wt. % (on biomass basis) for
the lignin derived oligomers [12,13], and between 10 and 15 wt. % for the oligomeric products derived from the
carbohydrate fraction [13,14]. One recent study [11] has divided the oligomeric products into four major groups: (1)
Water insoluble lignin derived oligomers (better known as pyrolytic lignin), (2) Water soluble heavy phenols, (3) Slightly
dehydrated carbohydrates, and (4) Highly dehydrated carbohydrates rich in carbonyl groups.

The primary and secondary reactions that produce or consume oligomers during fast pyrolysis are still very poorly
known [15,16]. The authors were not able to find any parametric studies on the effect of pyrolysis conditions on the
yield of the water soluble oligomers. All the published studies focus on the yield of water insoluble lignin oligomers
[12,17–20]. The effect of particle size on the yield and composition of lignin oligomers has been studied in the literature
[18–20]. Two reaction regimes have been identified [18–21]. In the first one, (for particles smaller than 2 mm) the
quantity of oligomers decreases dramatically as particle size increases, mostly due to the retention of oligomers inside
the biomass particles or Intraparticle heat transfer. In the second regime (particles bigger than 3 mm), the yield of lignin
oligomers did not change much as a function of particle size. The yield of pyrolytic lignin is also drastically affected by
the pyrolysis temperature [12,17]. At temperatures over 500 oC degradation reactions dominate dramatically reduce the
yield of this fraction [17]. New modelling and experimental tools are needed to describe the complex phenomena
happening during pyrolysis that control the yield and composition of pyrolysis oils [21,22].

One mechanism that explains the release of heavy oligomeric compounds is thermal ejection [10]. Thermal ejection
occurs in pyrolysis when bubbles inside the liquid intermediate burst (Figure 1). The main method used to study reactive
boiling ejection is visualization of morphological changes in the liquid intermediate via a fast speed camera. The collapse
of the bubble in the liquid intermediate interface results in the ejection of aerosols. Teixeira et al. [10] modeled the
dynamic of bubble collapse and its effect on ejection intensity and found that, for cellulose, the diameter of these
aerosols is close to 1 mm. However, our current understanding of micro-explosion and aerosol formation during biomass
pyrolysis is still in its infancy compared with the deep understanding of these phenomena in polymer and coal
devolatilization [23,24], raining [25,26], during cigarette smoking [27,28], champagne aerosol formation [29], formation
of aerosols by bursting of air bubbles in marine water [30–32], aerosol formation in cooling towers [33], and micro-
explosion of emulsified fuels during combustion [34–36]. A better understand of the mechanism of liquid intermediate
formation, micro explosions, and aerosol thermal ejection could contribute to improve the better control the operation
of fast pyrolysis reactors [37].
134 Chapter 4

Figure 1. Multi-scale scheme for the fast pyrolysis of biomass.

Thus, the main goal of this chapter is to develop a methodology for collection and analysis of aerosols thermally ejected
during the pyrolysis of sugarcane bagasse and its model fractions (Cellulose, Organosolv Lignin, and Xylan).

4.2. Materials and Methods

Characterization by SEM visualization of size distribution of aerosol ejected during pyrolysis of avicel cellulose, xylan,
organosolv lignin and sugarcane bagasse were made, also correlate this distribution with ejection intensity at different
heating rates and pressures.

4.2.1. Feedstock characterization

The sugarcane bagasse sample was obtained from the sugarcane mill Risaralda (Colombia, 4°54'22.42"N,
75°53'45.46"O). The proximate and elemental analysis of the material studied can be found elsewhere [38]. The model
compounds used were Avicel cellulose (Fluka 11365, Avicel PH-101,50μm, CAS 9004-34-6, with a degree of
crystallinity of 60%), Organosolv lignin (371,017 Aldrich, CAS 8068-03-9) and beech wood xylan (Sigma Aldrich CAS
9014-63-5)). All samples were reduced to particle sizes less than 100 µm.

The samples were suspended (~10 wt.%) in distilled water and applied with a brush to the central region of a 6 x 4 x
0.05 cm stainless steel plate. The central region covered by the suspension had a 4 x 1 cm dimension (Figure 2).
Chapter 4 135

Figure 2. Impregnation of the samples on the cold-rod stainless steel plate and average film thickness. (A) Finely
milled sample and distilled water. (B) Brush used for impregnation (C). Empty steel plate. (D). Steel plate impregnated
with a water / biomass ≈10% w / w suspension (E). Steel plate impregnated with a water / biomass suspension after
drying in halogen moisture balance at 80 °C.

The plate was dried in a Halogen Mettler Toledo HB34S scale at 80 oC to form a film an average thickness of 60 µm
and mass of 10 mg.

4.2.2. Hot plate reactor

Films of different materials were pyrolyzed in a hot plate reactor as shown in Figure 3.

Figure 3. Layout of Hot Plate Reactor.

The experimental set-up used for the pyrolysis studies is shown in Figure 1. The hot plate was built of stainless steel
304. Two copper electrodes (1/4 NPT) press the stainless steel plate with the biomass samples and transfer the 10 kV
current from the transformer. The sample is located on the hot stage face looking downwards, so that the aerosols can
136 Chapter 4
be collected on a glass slide (of plastic for lignin) (2 x 2 x 0.2 mm) heated to 200 oC. The distance between the glass
slide and the hot plate was 1 cm. The glass slide is heated to avoid the collection of light volatile compounds (water,
acetic acid, formic acid, hydroxyacetaldehyde, and monophenols) that can act as aerosol solvent. This is critical to
ensure that the compounds collected on the heated glass slide are high molecular oligomeric compounds ejected in the
form of aerosols. The reactor has two ¼” NPT inlets for purging with N2. A calibrated pyrometer (µ-Epsilon CTZ SF 25),
installed in the center of the reactor cap, is part of a control loop that measures and controls the temperature and
heating rate in each experiment. A µ PID-controller was designed to read and control the temperature of the plate. The
experiments were conducted at heating rates between 10 °C/s and 1200 °C/s, at pressures from 160 mbar to 1 bar,
and a final temperature of 500 °C with a holding time at the final temperature of 2 s. The pressure of the reactor was
measured with a Druck-DPI 100 sensor with an average error of 0.07 mbar. At the beginning and end of each
experiment was recorded the weight of the glass slide and the stainless steel plate and the data collected is used to
estimate the yield of char and aerosols collected. More details of reactor configuration are presented in chapter 5.

4.2.3. Slides visualization by SEM.

Each of the slides with pyrolytic aerosols was visualized by Scanning Electron Microscopy (EVO MA10 with a detector
from Oxford Instruments). The samples were covered with gold to facilitate the transport of electrons. The pictures were
processed with ImageJ software [39]. Each picture was formatted in binary mode adjusting the threshold of whites and
blacks to better quantify the size of the aerosol droplets as shown in Figure 4.

Figure 4. Tone setting in the image to be processed. (A) Original image cellulose 1200 ° C / s vacuum. (B) Image A
converted into binary scale (black-white) using ImajeJ. (C) Manual Processing highlighting aerosols outline (D)
Processed image with outline highlighted.

The number of the droplets and the Feret diameters of the droplets were calculated with the ImageJ Analyze/Analyze
Particle command. Some pictures could not be processed in ImageJ due to low contrast. In these cases, we did a
pretreatment of the pictures to highlight the borders (Figure 4D) followed by the conversion of the picture to binary mode
with ImageJ. Finally, with the information was visualized via histogram for the droplet size distribution. In order to
confirm that most of the droplets collected were from aerosols of oligomeric products, we conducted some preliminary
tests using Eugenol, phenol, Furfuryl alcohol, 5-HMF, sucrose, Pyrolytic lignin and xylose. All these molecules should
be evaporated under the pyrolysis conditions studied. The tests were conducted at 500 oC, 900 mbar, heating rates of
1,200 oC and the glass slide at 200 oC, except for pyrolytic lignin (plastic surface).
Chapter 4 137

4.2.4. Aerosol Thermal Ejection Intensity

Unfortunately, the hydrodynamic characteristics of our set-up do not allow to collect all the aerosols produced. However,
by choosing a region for our analyses and supposing that the visualized picture is representative of the distribution of
aerosols in the slides it is possible to compare the relative thermal ejection intensity at different experimental conditions.
Equation 1 is used to estimate the relative aerosols thermal ejection intensity.

mo − mf Aslide
AI = ( )( ) F(D) (1)
mslide−after − mslide−before Aimage

𝐴𝐼 , represents the distribution of the number of aerosols per unit of initial mass of material used; F(D) is the distribution
function of the number of aerosols per unit of the initial biomass (mo) less solid residue after pyrolysis (mf ). Aslide is the
total area of the glass slide and Aimage is the area visualized by SEM. The summation of all the droplet size distributions
results on the number of total ejected aerosols collected on the slide per unit of biomass used (see Eq. 2).

Atotal = ∫ AI dD (2)
0

Atotal is the total number of aerosols collected per mass unit of initial biomass pyrolyzed, D is the aerosol diameter
(µm).

4.3. Results
Some preliminary visualization tests were conducted to confirm that only aerosols form droplets on our heated glass
slides. The molecules studied are known to evaporate during pyrolysis. Figure 5 shows SEM pictures of the hot glass
slides of each of the compounds studied.

20µm 200µm 200µm

A. Empty slide B. Water C. 5-HMF

200µm 200µm 200µm

D. Furfuryl alcohol E. Phenol F. Iso-Eugenol


138 Chapter 4

1mm 20µm 20µm

G. Sucrose H. Xylose I. Monomers on plastic


Figure 5. SEM visualization of some pure compounds in bio-oils.

The pictures for experiments with water, 5-HMF, Furfuryl alcohol, Phenol, Eugenol are similar to those with the slides
before the tests. This means that at 200 °C (the temperature at which the aerosols were collected), most of the material
collected is from the aerosols. Most of the small molecules in vapor phase that touch the slide will not condense. In the
case of sucrose, the melting point is 186 °C, but it appears to undergo secondary reactions at temperatures over 200
°C. The box I shows the micrograph obtained by pyrolysis of pyrolytic lignin on plastic surface. In pyrolytic lignin
pyrolysis, the main products are monomers. The micrograph does not show evidence of any monomer stuck on. The
extreme vacuum of SEM chamber (<10-6 torr) is able to evaporate monomers at atmospheric pressure. Thus, the
ejected sucrose formed a thin film on the glass slide.

4.3.1. Aerosols ejection during the pyrolysis of Avicel cellulose

Figure 6 shows that the yield of char obtained during the pyrolysis of avicel cellulose was in all cases lower than 5 wt.
%.

Figure 6. Char yield obtained during avicel cellulose pyrolysis.

The yield under vacuum at very high heating rates was lower than 1 wt.%. The rest of mass is volatile material,
composed by anhydrosugars oligomers and a small fraction of permanent gases. Some researchers have suggested,
that during cellulose fast pyrolysis, cellulose conversion to oligomers is almost complete [40,41]; possible minimum
char is generated due to re-polymerization reactions and liquid phase dehydration [42–45]. The extent of these reactions
depends largely on the competition between evaporation and ejection rate of these oligomers and crosslinking reactions
in liquid phase. Evaporation and bubble bursting (micro explosion) are enhanced under vacuum, where intermediate
liquid phase viscosity is reduced, therefore viscous stresses acting on bubbles are weaker, intensifying the escape of
bubbles through the gas-liquid interface [26] and generating higher aerosol ejection rate. Figure 7 shows the aerosols
images visualized by SEM for the avicel cellulose pyrolysis.
Chapter 4 139

20 µm 20 µm

A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C

20 µm 20 µm

C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C
Figure 7. SEM images obtained by ejected aerosol during avicel cellulose pyrolysis.

In all cases, recovered aerosols have rounded shapes with soft edges, suggesting that these droplets, composed of
intermediate liquid fractions, remained melted before cooling and later slide visualization, as has been suggested by
Teixeira et al [46]. Under vacuum condition, droplets appear to have a bigger relief, indicating higher hydrophobicity
(high contact angle), this is a possible indicator of the presence of higher molecular weight oligomers. At low heating
rates (10 °C/s) are observed some large droplets (60 - 80 μm). Analyzing the droplets shape, apparently they were
formed by coalescence with neighbor droplets. The exposure time of aerosols caught in the slide is high, increasing the
probability of collision with other aerosols released in before. Although this is not part of the core of this article, in
Appendix is presented the visualization of organosolv at vacuum and atmospheric pressure (Video A,H,J,K), high
heating rate (> 100 °C/s) and final temperature of 400 °C, this in order to show, in a illustrative way, bubbles collapse
dynamic and aerosol ejection. In videos it can be seen very clearly that under vacuum conditions bubbles are smaller
than at atmospheric pressure, also in vacuum escape quickly, releasing great amount of aerosol in the process. This
information allows to infer, that somehow in vacuum conditions and high heating rates the intensity of aerosol ejection
is important. Figure 8 shows the distributions as a function of aerosols diameter recovered during Avicel cellulose
pyrolysis in different reaction conditions.

A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C
140 Chapter 4

C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C

Figure 8. Aerosols distribution during Avicel cellulose pyrolysis.

In Figure 8 it can be seen that at high heating rates (1200 °C/s), the particle density is higher between 0 - 3μm, being
the distribution less dispersed under vacuum condition. As we showed before, under vacuum condition, gas bubbles
formed do not grow as much as under atmospheric condition, thus bubbles that collapse have smaller diameter and
therefore tends to generate aerosols of smaller diameter. The relationship between bubble diameter and ejected aerosol
diameter has been well reported in literature [10,26,32,47]. Teixeira, et al [10], suggests a bubble/aerosol diameter ratio
equals to 0.1 for sucrose pyrolysis.

The intermediate liquid formation dynamic, therefore bubble formation and aerosol ejection are not unique in fast
pyrolysis processes. Zhang et al, and Zhou et al and F. Yi et al [48–50] have reported the existence of the liquid phase
at low heating rates, which can be seen small aerosols trapped inside the cell wall, although there are not many details
about bubble formation and aerosol ejection dynamics. In slow pyrolysis (10 °C/s) and vacuum conditions, it can
Chapter 4 141
appreciate a double aerosols distribution, the first one between 0 – 6 μm and the second one between 6 – 20 μm, with
a maximum at 10 μm. As is shown in Figure 7C, there are small and large aerosols, some of them are a product of
multiple droplet coalescence. Because the heating is slow and aerosols that can be collected in the slide are a
cumulative fraction of all ejected aerosols, it is possible that aerosols released at different reaction times can impact at
the same point in the slide, forming new larger diameter particles. It is also possible that, because the process is too
slow, bubbles inside the intermediate liquid phase could have greater volume, therefore ejected aerosol droplets are
large and can collide more likely with other aerosols caught in the slide. This behavior is also observed at atmospheric
pressure, however at this condition it could be observed larger aerosol droplets (40 µm). According to char yield seen
in Figure 5 and aerosols distributions shown in Figure 6, it may suggest that most aerosol ejection occurs under vacuum
conditions and high heating rates. Figure 9 shows the relative aerosol ejection intensity for Avicel cellulose pyrolysis.

Figure 9. Relative aerosol ejection intensity during Avicel cellulose pyrolysis.

As shown in Figure 9, relative aerosol ejection intensity is more intense in vacuum and high heating rates. Under these
conditions both, depolymerization and anhydrosugar oligomers evaporation are intense minimizing crosslinking and
dehydration reactions that generate char [40,42,45]. At 1200 °C/s and 900 mbar, cellulose depolymerization is quite
fast, but oligomers evaporation is not as intense as in vacuum conditions, therefore the number of aerosols visualized
per mass unit is smaller. An interesting result is that at low heating rates (10 °C/s), regardless of pressure,
depolymerization is so slow that oligomers remain enough time inside the intermediate liquid phase to react by
crosslinking, dehydration and fragmentation, generating char. Although at 10 °C/s and at vacuum conditions less char
is recovered (≈1.8% Wt.) compared to the recovery at atmospheric pressure (≈2.7% Wt.), the difference is not too
representative. Small difference obtained in intensities at low heating rates may be because vapors diffusivity is much
lower at atmospheric pressure than at vacuum, therefore the probability of collisions of these vapors is higher,
generating somehow larger droplets. There are no reports in literature to contrast some of the results presented here.
Teixeira, et al [10] showed avicel cellulose aerosols distribution at atmospheric pressure and 700 °C, finding the
maximum sizes between 0.5 – 1 μm. Comparing with this results under similar conditions (1200 °C, 500 °C, 900 mbar),
the distribution is between 0 – 3 μm. Strictly speaking, in both cases, aerosols are in the same size range. Differences
may be due to the visualization method used (Optical microscope vs SEM), operating temperature (700 °C vs. 500 °C),
and sample preparation (powder 300 μm particle vs 60 μm thin film).
142 Chapter 4

4.3.2. Aerosols ejected during the pyrolysis of organosolv lignin

Char yield in organosolv lignin pyrolysis are presented in Figure 10. The error bars in the experiments are quite small,
so they cannot be seen in figure (sd ± 0.3).

Figure 10. Char yield obtained in Organosolv lignin pyrolysis.

The lowest char yield (26 wt. %) is obtained under vacuum and at the highest heating rates. Evaporation of oligomers
and some monomer is more intense at vacuum conditions than at atmospheric pressure, therefore probability of
secondary fragmentation, re-polymerization and crosslinking reactions in liquid phase are reduced. At low heating rates
regardless of the pressure, char yield are virtually equal. Heating is so slow and intermediate liquid phase so heavy,
that when the last one is already melted, temperature is low and it is not hot enough to vaporize monomers or lignin
oligomers (molecular weight > 500 Da), therefore re-solidification reactions to form char are important. At high heating
rates, liquid phase is at high temperatures (500 °C) in a very short time, so evaporation and bubble bursting is important,
and even more under vacuum. Char yield obtained at low heating rates reported by Patwardhan [43] using PyGC-MS
are equal to those reported by Zhou, et al [51] using TGA and similar to those reported by Hoekstra, et al [38] in a Wire
Mesh Reactor at 60 °C/s and 0.3 mbar. Figure 11 shows ejected aerosol distribution during organosolv lignin pyrolysis
at different reaction conditions on glass slide.

A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C
Chapter 4 143
C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C

Figure 11. Images obtained by SEM of ejected aerosols during Organosolv lignin pyrolysis.

Figure 11 shows ejected and captured aerosol distribution for organosolv lignin. Compared with cellulose, aerosol shape
is no circular; in all pyrolysis conditions is amorphous. This condensation pattern may be linked with multiple factors
including surface tension and viscosity of ejected fluid. If ejected aerosol viscosity is low, the impact with the glass
deform the aerosol once the collision ends. Also, if solid-liquid interfacial tension is low, aerosols tend to settle with high
contact angles, thus it would not be visible a clear droplet relief [52,53]. Several researchers [54–58] have proposed a
dimensionless number ξ, which is a function of contact angle θ between fluid and surface, the number of Webber (We)
and fluid viscosity µF. This parameter relates aerosol Feret diameter before and after a collision. If ξ = 1.0 aerosol does
not change its shape due to the collision, and the bigger this parameter more dispersed is shape relation. This
phenomenon is commonly called splashing. If We is big and We*µF >= 100, usually contact angles are small and
splashing is intense. Another possibility to explain aerosols shape is coalescence of multiple droplets in the slide. To
rule out this possibility, an experiment was performed reducing 10 times the initial mass of lignin, operating the reactor
at 500 °C, 1200 °C/s and 900 mbar, and then visualizing the captured aerosols geometry. An additional experiment
was performed under the same conditions but using a plastic slide. Figure 12 shows SEM images obtained by each of
these new experiments.

A-1200°C/s-0.7mg-glass B-Plastic blank C-1200°C/s-900mbar


144 Chapter 4
Figure 12. Effect of the amount of sample and surface type over aerosol captured geometry during organosolv lignin
pyrolysis at 500 °C, 1200 °C/s and 900 mbar.

Figure 12 shows that, regardless of the amount of mass, oligomeric lignin droplets over glass are amorphous, and tend
to be flat (small contact angle), which are difficult to visualize. Changing the surface material, droplet geometry
drastically changes (Figure 12C). The behavior of aerosol droplets between this material is hydrophobic, forming high
relief droplets; they are easy to identify by microscope. Analysis of aerosols distribution for lignin oligomers was carried
out from that moment over over plastic, in order to have better characterization. Figure 13 shows the image sequence
of organosolv lignin aerosols trapped on the new plastic surface.

20 µm 20 µm

A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C

20 µm 20 µm

C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C
Figure 13. Aerosols distribution during organosolv lignin pyrolysis.

Droplets collected have an uniform, high relief, almost spherical geometry, indicating a hydrophobic interaction between
the plastic surfaces. At the SEM chamber operating pressure (10-4 Pa), the monomers (Eugenol, Vanillin, Phenol, and
others) are evaporated; therefore visible droplets correspond to high molecular weight compounds. At high heating
rates are visualized bigger aerosols; perhaps under these conditions gas bubbles size are bigger, so bigger aerosols
are released. At high heating rates the gas bubbles expansion inside the liquid phase is intense, generating violent and
destructive liquid micro-explosions [36,59,60]. As Teixeira suggests [36], the presence of filterable polymers (oligomers)
promote the frequency and intensity of micro-explosions. This may explain big aerosol droplets size found at high
heating rates. In the other hand, at low heating rates, the intensity of micro-explosions is not intense, and ejection
mechanisms are mainly film droplets and jet droplets, generating smaller aerosols. Captured aerosols size distribution
for organosolv lignin is shown in Figure 14.
Chapter 4 145
A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C

C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C

Figure 14. Aerosols distribution during organosolv lignin pyrolysis.

At high heating rates, average aerosol size is between 12 -14 μm; at low heating rates, average size is between 4 -6
μm. Unlike avicel cellulose and biomass aerosols, aerosol distribution is uniform, with a dispersion of ±10 μm, being
the maximum aerosol size ≈30 μm (150 mbar-1200 °C/s). The relative aerosol ejection intensity is presented in Figure
15.
146 Chapter 4

Figure 15. Relative aerosols ejection intensity during organosolv lignin pyrolysis.

As for avicel cellulose, under vacuum conditions and high heating rates, aerosol ejection intensity is high compared
with the intensity at low heating rates. Under these conditions, micro-explosions intensity are intense for two reasons:
first, vacuum promotes oligomers evaporation, second is the effect of high heating rate which promotes violent gas
bubbles explosion, destroying the liquid phase core in droplets liquid form. At low heating rates and atmospheric
pressure, relative aerosols ejection intensity is the lowest of all. Being the liquid heating so slow, it is expected there
were no noticeable micro-explosions, instead the main ejection mechanism is bubbles bursting at the gas/liquid
interface. Intermediate liquid phase has enough time to re-polymerization reactions, in which viscosity increases enough
to reduce gas bubbles collapse, increase its molecular weight, and therefore decrease its evaporation rate. There are
no appreciable differences in aerosol ejection intensity between 900 mbar, 1200 °C/s conditions and 150mbar, 10 °C/s
conditions. Apparently evaporation contributes more significantly to aerosol ejection than bubble micro-explosion at
high heating rates. Future studies are required to evaluate with more detail this type of behavior, also relating them to
changes in physicochemical properties (molecular weight, density, viscosity and surface tension).

4.3.3. Aerosols ejected during the pyrolysis of Xylan

There are not many studies for hemicellulose pyrolysis; normally these studies are made with reference to a model
compound, generally xylan. But still, there are few studies with Xylan as a model compound. Some studies have shown
that in forest fires, during pyrolysis and combustion stages, aerosols rich in levoglucosan and its stereoisomers
(manosan, galactosan) are released from cellulose and hemicellulose decomposition, respectively [61,62]. As a first
indicator of xylan devolatilization intensity, Figure 16 shows char yield for xylan pyrolysis at different pressures and
heating rates.
Chapter 4 147

Figure 16. Char yield obtained for xylan pyrolysis.

Unlike cellulose, char yield it is quite high and slightly lower than the lignin. Other studies have shown yield in the same
order of magnitude like yields reported by Hoekstra [40] in a Wire mesh reactor (30 %w/w) at 0.3mbar, 6000 °C/s, and
those reported by [63,64], but higher than those reported by Patwardhan [43] for pyrolysis of hemicellulose extracted
from demineralized switchgrass (≈10 %w/w). The mineral material effect over char yield is fairly reported [65–69]. More
detailed and future studies would involve the mineral material effect over yields and aerosol ejection intensity. For this
study, using the same raw material in all experiments, it was evaluated the effect of heating rate and pressure over
aerosol ejection intensity. Figure 17 shows the sequence of images of recovered aerosols obtained by SEM over glass
slides.

A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C

C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C
148 Chapter 4

Figure 17. Images obtained by SEM of ejected aerosols during xylan pyrolysis.

Unlike cellulose, organosolv lignin and bagasse, it was not possible to visualize xylan aerosols. Only under vacuum
and high heating rates, small aerosols, similar to those recovered cellulose were observed, however the number of
intercepted droplets is quite few. Given the high content of mineral material in xylan, it is expected than cross-linking
and dehydration reactions are favored as opposed to depolymerization and pentoses evaporation. In Figures 17 A-C,
dark gray spots are observed, which at first sight could be aerosols. However these spots form no relief and are 100%
flat, indicating deposition and immediate evaporation of a compound. As it was seen in Figure 5, for each sugar aerosols
droplets can be seen, even for xylose, which is the constituent unit of xylan. This means that there was no sugar
deposition (Xylopyranose). Instead other compounds like methyl-furan, 2-furaldehyde, acetol, and others could deposit,
which are xylose fragmentation products [43]. A detailed analysis by EDS to one of these spots shows that there is
organic matter (small film of evaporated compound). However this is only a very thin film, since the C/Si ratio is very
low. This means that at this point, the composition is almost entirely glass Si.

A-150mbar-1200°C/s-500°C B. Cumulative aerosol intesnsity

C-SEM image for for xylan gay stin D-EDS Spectrum for xylan gray stin
Chapter 4 149

%C=2.49; %O=43.80; %Na=5.76; %Mg=1.68; %Si=34.43.


Figure 18. Distribution of aerosols from the pyrolysis of xylan.

It was only possible to determine xylan aerosol distribution at 1200 °C/s and 150 mbar. In Figure 18A, the highest
aerosols density is between 0 – 8 μm, with no aerosols > 30 μm. As it was shown for cellulose, under vacuum and high
heating rate conditions, xylan depolymerization and sugars evaporation are more intense than at atmospheric pressure
and low heating rates, therefore more aerosols are expected. Other studies are needed, using hemicellulose native
from biomasses or demineralizing xylan to minimize cross-liking and deshydratation that produce char.

4.3.4. Aerosols ejected during the pyrolysis of sugarcane bagasse

Finally, it was analyzed aerosols distribution for sugarcane bagasse as lignocellulosic material. For these biomass type,
aerosols can come from anhydrosugar oligomers, lignin oligomers and some pentoses from lignin hemicellulose. Figure
19 shows char yield obtained for sugarcane bagasse pyrolysis under different reaction conditions.

Figure 19. Char yield obtained for sugarcane bagasse pyrolysis.

As with previously studied materials, less char was obtained under vacuum and high heating rates. These same
behaviors have been reported by Hoekstra, et al [40,70] for Pine Wood. Although at low heating rates, char yield is
virtually the same, regardless of the reactor pressure. Interactions of all biomass components modified aerosol ejection
behavior. For example, at low temperatures (T < 250 °C), lignin and hemicellulose may be melted, but not cellulose,
150 Chapter 4
forming a liquid-solid system, in which solid can act as a basic catalyst, or vice versa when intermediate liquid can act
as an acid electrolyte. Char yield obtained with these experiments are quite similar to those reported by Drummond et
al [71] at 1000 °C/s, 1 °C/s, and atmospheric pressure (7 %w/w at 1000 °C/s, 15 % w/w at 1 °C/s). The sequence of
images with captured aerosols for sugarcane bagasse are shown in Figure 20.

20 µm 20 µm

A. 150mbar-1200°C/s-500°C B. 900mbar-1200°C/s-500°C

20 µm 20 µm
C. 150mbar-10°C/s-500°C D. 900mbar-10°C/s-500°C
Figure 20. Images obtained by SEM of ejected aerosols during sugarcane bagasse pyrolysis.

Under vacuum and high heating rates conditions, aerosol shape is very well defined, showing reliefs in the droplets,
very similar to those obtained for cellulose and xylan (sugars). In this extreme condition, despite the chemical
interactions of each pseudo-component, the oligomers ejection for each component is possible. More studies are
required to determine which of them predominate in bio-oil. Although it is not part of this chapter core, it was analyzed
an image using a confocal microscope coupled to a Raman HORIBA LabRam HR detector, to see the chemical
composition of a representative droplet of a captured aerosol during bagasse pyrolysis at 500 °C, 1200 °C/s and
atmospheric pressure, as shown in Figure 21.
Chapter 4 151

B
A

Figure 21. Raman spectra of a representative bagasse aerosol.(A). Confocal visualization of an aerosol of sugarcane
bagasse. (B). Raman spectrum of sugarcane aerosol and avicel cellulose.

Sugarcane bagasse aerosols consist mostly of phenols and aromatic compounds that fluoresce (see Figure 21), but
also appear in a lower intensity groups that belong to sugars, perhaps derived from cellulose oligomers. This indicates
that dominant sugarcane bagasse aerosols come from lignin oligomers. Nevertheless, this is only a starting point, and
further studies are needed to better characterize the chemical composition of these aerosols. Captured and visualized
aerosols at 1200 °C/s, atmospheric pressure and low heating rates are very similar to each other, however, unlike
organosolv lignin aerosols, these have very definite circular shapes. There may be differences in aerosols depending
on the type of lignin. In the first case, organosolv lignin was studied as a model compound, which structurally is more
degraded (lower molecular weight) than lignin native from biomass. In that order of ideas, it is possible that lignin
oligomers released by sugarcane bagasse pyrolysis have a higher viscosity and thus lower splashing when droplets
collide with the glass surface. Another possibility is that because of interactions between lignin, cellulose and
hemicellulose, it is possible that oligomers have different molecular weight and physicochemical properties [72–74],
which make them behave differently when hitting the surface. Figure 22 shows aerosols distribution functions obtained
by sugarcane bagasse pyrolysis.

A-150mbar-1200°C/s-500°C B-900mbar-1200°C/s-500°C
152 Chapter 4
C-150mbar-10°C/s-500°C D-900mbar-10°C/s-500°C

Figure 22. Aerosol distribution during sugarcane bagasse pyrolysis.

In contrast of distributions found for model compounds described above, for sugarcane bagasse aerosol distribution,
aerosols size are smaller and lesser disperse. At high heating rates, highest aerosol density corresponds sizes between
0 – 4 μm, and at low heating rates, aerosols are distributed mostly between 0 – 2 µm. These sizes are similar to those
reported for forest fire aerosols [75–77]. Differences in aerosols distribution are very small. Under vacuum conditions,
a small aerosol fraction is slightly bigger than at atmospheric pressure, but for practical purposes, these differences are
not important. Although aerosol distribution does not change, relative intensity is different, because the number of
aerosols per mass unit change depending on pyrolysis conditions, as shown in Figure 23.
Chapter 4 153
Figure 23. Relative aerosols ejection intensity during sugarcane bagasse pyrolysis.

At high heating rates, high aerosol ejection intensity is achieved. In other heating conditions, aerosol ejection intensity
drops drastically. Under slow pyrolysis conditions, cellulose and lignin depolymerization rate is slow, being the time of
crosslinking, deshydratación and fragmentation reactions in liquid phase extensive enough to compete with oligomers
evaporation and to promote interactions between components of biomass, giving way to hydrogen bonds and inter-
chain covalent bonds [78]. Hilbers, et al [79], have shown the effect of interactions between the lignin and cellulose over
levoglucosan production, suggesting an increase in yield for avicel cellulose/organosolv lignin mixtures 20 % w / w. This
increase is due to the action of the generated bubbles in the lignin intermediate liquid phase that helps formed oligomers
to escape from the cellulose liquid phase. Although for lignocellulosic material, lignin and cellulose are bonded by
covalent and hydrogen bonds, that are a stronger interaction type than a simple mixing of its components, it is a very
interesting result, since it can propose new pathways in reaction engineering to promote a product selectivity with high
added value. For example, it can prepare mixtures of organic foaming agents with biomasses to promote aerosol
ejection rich in anhydrosugars that have a high added value in pharmaceutical industry.

4.4. Aerosol yield theoretical estimation


To achieve a first estimation in ejected aerosol yield, it is important to know the representative volume of a aerosol
droplet over the impact surface (glass, plastic) and the compound density of the aerosol droplet. This information is
difficult to find, and also is hard to obtain experimentally because is difficult to recover enough oligomers
(Anhydrosugars and pyrolytic lignin) to be analyzed. As a first approximation, it was measured the contact angle of a
honey droplet as a representative compound of sugars oligomers from cellulose depolymerization, the contact angle of
an insoluble bio-oil heavy fraction from bagasse (soluble in dichloromethane, insoluble in water) as representative
compound of lignin oligomers and the contact angle of sugarcane bagasse bio-oil as representative compound of
oligomers released during sugarcane bagasse pyrolysis. The contact angle measurements were performed taking
pictures with a Sony Alpha 3000 camera with a Tokina AT-X f2.8 100mm macro lens, focal length of 5 cm. Softly, it was
put a small droplet (with the tip of a 1 ml syringe) of each compound on the edge of a glass slide (for cellulose and
biomass) or plastic slide (for lignin). A picture was taken with a f 22 aperture, and then it was processed in ImageJ
software with the Angle Tool option to measure the contact angle. Figure 24 shows the geometric relations the volume
of a spherical sector as a function of contact angle and diameter, measured by SEM visualization.

𝜋 𝑎3
𝑉≈ (2 − 3𝑐𝑜𝑠(𝜃) + 𝑐𝑜𝑠 3 (𝜃)) (3)
3 𝑠𝑖𝑛3 (𝜃)

Figure 24. Geometric representation of volume of a spherical sector as a function of contact angle and diameter.

Contact angles measured experimentally for each of compound are shown in Figure 25.

A-Honey B-Heavy fraction of Bio-oil C-Bio-oil


Glass Plastic Glass
154 Chapter 4

θ≈44.42° θ≈65° θ≈7.58°


Figure 25. Contact angles measured experimentally for each model compound of ejected oligomers during pyrolysis.

For the mathematical approximation for aerosol yield, it was assumed that each droplet of the compound of interest has
the same contact angle that its model compound (Figure 25) and each droplet density is approximately 1200 kg/m3.

V
mo − mf Aslide
Aero/yield = ( )( ) 𝜌 ∫ F(V)dV (4)
mslide−after − mslide−before Aimage
0

Where F(V), represents the aerosol distribution function as a function of droplet volume, obtained by combining Eq (3)
and measured diameters by SEM visualization. Yield of each aerosols is shown in Table 1.
Chapter 4 155
Table 1. Estimated aerosol yield for each material.

These aerosol yields are only an approximation. For future studies, more accurate results would be obtained quantifying
oligomers yield captured on the slide. However, these yields are higher than those reported by Teixeira, et al [46] for
cellulose pyrolysis (3 %Wt) at atmospheric pressure and high heating rates. At very low heating rates (10 °C/s),
oligomers yield drastically decreases, and it is consistent with char increase shown in Figure 6. For xylan, the high
mineral material content catalyzes carbonization reactions. Patwardhan [43] showed that for hemicellulose fast
pyrolysis, adding 5% of mineral material reduces the sugar content down to 1.6 %Wt. At low heating rates, these
reactions are even more important than sugars vaporization, therefore both char and permanent gases yield are very
high, so it is expected that aerosol yield is quite low, consistent with the observations. For organosolv lignin aerosols,
lignin oligomers aerosol yield are lower than those obtained for avicel cellulose. For this type of lignin, mineral content
is relatively high (see Table 1), this greatly contributes to char production (Figure 16) and the reduction of condensable
volatiles generation. It has been reported bio-oil oligomers yield between 10 – 30 % w/w [12,43]. At atmospheric
pressure and 1200 °C/s, aerosols estimated yield drops drastically and is much lower than yield reported for biomass
fast pyrolysis in fluidized bed reactors [12]. In fluidized bed reactors, the carrier gas can remove rapidly aerosols,
compared to ejection in a static atmosphere used in experiments. Westerhof [37], shows the decrease in lignin
oligomers yield as gases residence time increases in a fluidized bed reactor at 500 °C.

4.5. Final comments


Understanding aerosol ejection dynamics during lignocellulosic material pyrolysis open a new opportunity for new
recovery systems product (condenser) design based on physicochemical properties of recovered oligomers. The design
of these new recovery systems could mitigate some problems related to bio-oil upgrading (reduction of heavy oligomers
that increase the bio-oil viscosity by aging), recovery of high added value in pharmaceutical industry (levoglucosan and
anhydrosugars), or improve bio-oil burners performance. (See Figure 26).
156 Chapter 4
Figure 26. Importance of aerosol characterization in condenser design.

In Figure 26 is clear that for oligomers selective condenser design, it is important to know aerosol distribution function
and aerosol ejection intensity (bulk property) in order to know the required droplet deposition area. Besides knowing
part of aerosol chemical structure, it is possible to select a type of material in which collection efficiency would be high
for a particular oligomer. As it was discussed in this chapter, sugar and pyrolytic lignin oligomers are well deposited
over two very different materials (glass vs. plastic). The type of material can be a key parameter when selective
oligomers separation is performed.

4.6. Conclusions
For the first time a simple methodology based on SEM visualization for morphological characterization of ejected
aerosols during biomass and model compounds pyrolysis is proposed. Under vacuum (150 mbar) and high heating
rates (1200 °C/s) conditions aerosol ejection intensity for all studied materials is favored, by favoring oligomers
evaporation, bubbles bursting and micro-explosions at the intermediate liquid phase. Higher relative yield was obtained
for Avicel cellulose at 150 mbar and 1200 °C/s (71.78 % Wt.). This yield corresponds to high molecular weight
oligomers; conditions in which aerosols ware captured and visualized minimize levoglucosan contribution. For xylan,
aerosols were obtained only at vacuum and high heating rates. The presence of mineral material drastically reduces
aerosol yield by catalyzing cross-linking and dehydration reactions. Sugarcane bagasse aerosol yield cannot be
expressed as a linear combination of model compounds aerosol yield; interactions between its components and mineral
material contribute to the final aerosol ejection.
Chapter 4 157

4.7. References
[1] A. Bridgwater, Fast Pyrolysis of Biomass: A Handbook, 3rd ed., PyNe, 2008.
[2] M-K. Bahng, C. Mukarakate, D.J. Robichaud, M.R. Nimlos, Current technologies for analysis of biomass
thermochemical processing: a review., Anal. Chim. Acta. 651 (2009) 117–138.
[3] A. Bridgwater, European Commission, R. and D. Directorate-General XII - Science, Agro-Industrial Research
Division, Thermal biomass conversion and utilization - Biomass information system, EUR-OP, Luxembourg,
1996.
[4] S. Thangalazhy-gopakumar, S. Adhikari, R.B. Gupta, S.D. Fernando, Influence of Pyrolysis Operating
Conditions on Bio-Oil Components : A Microscale Study in a Pyroprobe, Energy and Fuels. 25 (2011) 1191–
1199.
[5] P. Basu, Biomass Gasification and Pyrolysis, © 2010 Elsevier Inc., 2010.
[6] A.V. Bridgwater, Review of fast pyrolysis of biomass and product upgrading, Biomass and Bioenergy. 38 (2012)
68–94.
[7] T. Cheng, Y. Han, Y. Zhang, C. Xu, Molecular composition of oxygenated compounds in fast pyrolysis bio-oil
and its supercritical fluid extracts, Fuel. 172 (2016) 49–57.
[8] M.R. Rover, Analysis of sugars and phenolic compounds in bio-oil, Grad. Theses Diss. (2013).
[9] T. Chen, C. Deng, R. Liu, Effect of selective condensation on the characterization of bio-oil from pine sawdust
fast pyrolysis using a fluidized-bed reactor, Energy and Fuels. 24 (2010) 6616–6623.
[10] Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P. and
Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011)
4306.
[11] and M.G.-P. Filip Stankovikj, Armando G. McDonald, Gregory L. Helms, Quantification of Bio-oil Functional
Groups and Evidences of the Presence of Pyrolytic Humins, (2016).
[12] M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W.J. Lee, C.Z. Li, Effects of temperature on the formation of
lignin-derived oligomers during the fast pyrolysis of Mallee woody biomass, Energy and Fuels. 22 (2008) 2022–
2032.
[13] M.R. Rover, Analysis of sugars and phenolic compounds in bio-oil, Iowa State University, 2013.
[14] A. Oasmaa, D. Meier, Norms and standards for fast pyrolysis liquids, J. Anal. Appl. Pyrolysis. 73 (2005) 323–
334.
[15] X. Bai, K.H. Kim, R.C. Brown, E. Dalluge, C. Hutchinson, Y.J. Lee, D. Dalluge, Formation of phenolic oligomers
during fast pyrolysis of lignin, Fuel. 128 (2014) 170–179.
[16] T. Kotake, H. Kawamoto, S. Saka, Mechanisms for the formation of monomers and oligomers during the
pyrolysis of a softwood lignin, J. Anal. Appl. Pyrolysis. 105 (2014) 309–316.
[17] Westerhof, D.W.F. (Wim) Brilman, W.P.M. van Swaaij, S.R.A. Kersten, Effect of Temperature in Fluidized Bed
Fast Pyrolysis of Biomass: Oil Quality Assessment in Test Units, Ind. Eng. Chem. Res. 49 (2010) 1160–1168.
[18] S. Zhou, M. Garcia-Perez, B. Pecha, A.G. McDonald, R.J.M. Westerhof, Effect of particle size on the
composition of lignin derived oligomers obtained by fast pyrolysis of beech wood, Fuel. 125 (2014) 15–19.
[19] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Effects of particle size on the fast
pyrolysis of oil mallee woody biomass., Fuel. 88 (2009) 1810–1817.
[20] Westerhof, H.S. Nygård, W.P.M. van Swaaij, S.R.A. Kersten, D.W.F. Brilman, Effect of Particle Geometry and
Microstructure on Fast Pyrolysis of Beech Wood, Energy & Fuels. 26 (2012) 2274–2280.
[21] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials., Curr. Opin.
Biotechnol. 24 (2013) 414–20.
[22] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[23] Myongsook S., Oh William A., Peters Jack B. Howard, An experimental and modeling study of softening coal
pyrolysis, AIChE J. 35 (1989).
[24] K. Butler, A numerical model for combustion of bubbling theroplastic materials in microgravity, (2002) 70.
[25] Y.S. Joung, C.R. Buie, Aerosol generation by raindrop impact on soil., Nat. Commun. 6 (2015) 6083.
[26] J.S. Lee, B.M. Weon, S.J. Park, J.H. Je, K. Fezzaa, W.-K. Lee, Size limits the formation of liquid jets during
bubble bursting., Nat. Commun. 2 (2011) 367.
[27] M. Li, J.-G. Dong, Z.-X. Huang, L. Li, W. Gao, H.-Q. Nian, Z. Fu, P. Cheng, Z. Zhou, Analysis of Cigarette
158 Chapter 4
Smoke Aerosol by Single Particle Aerosol Mass Spectrometer, Chinese J. Anal. Chem. 40 (2012) 936–939.
[28] R. Yadav, K. Saoud, F. Rasouli, M. Hajaligol, R. Fenner, Study of cigarette smoke aerosol using time of flight
mass spectrometry, J. Anal. Appl. Pyrolysis. 72 (2004) 17–25.
[29] G. Liger-Belair, T. Seon, a Antkowiak, Collection of collapsing bubble driven phenomena found in champagne
glasses, Bubble Sci. Eng. Technol. 4 (2012) 21–34.
[30] F.J. Resch, P. De Toulon, D. Sciences, J.S. Darrozes, Marine Liquid Aerosol Production From Bursting of Air
Bubbles, 91 (1986) 1019–1029.
[31] Spiel, The number and size of jet drops produced by air bubbles bursting on a fresh water surface, J. Geophys.
Res. 99296 (1994) 289–10.
[32] S.C. Georgescu, J.L. Achard, É. Canot, Jet drops ejection in bursting gas bubble processes, Eur. J. Mech.
B/Fluids. 21 (2002) 265–280.
[33] I. Kataoka, M. Ishii, Mechanistic modeling of pool entrainment phenomenon, 27 (2014).
[34] M. Fuchihata, T. Ida, Y. Mizutani, Observation of microexplosions in spray flames of light oil-water emulsions
(2nd report, influence of temporal and spatial resolution in high speed videography), Trans. Japan Soc. Mech.
Eng. Ser. B. 69 (2003) 1503–1508.
[35] M.Y. Khan, Z.A.A. Karim, a. R.A. Aziz, I.M. Tan, Experimental investigation of microexplosion occurrence in
water in diesel emulsion droplets during the Leidenfrost effect, Energy & Fuels. 28 (2014) 7079–7084.
[36] Teixeira, R.J. Hermann, J.S. Kruger, W.J. Suszynski, L.D. Schmidt, D.P. Schmidt, P.J. Dauenhauer,
Microexplosions in the upgrading of biomass-derived pyrolysis oils and the effects of simple fuel processing,
ACS Sustain. Chem. Eng. 1 (2013) 341–348.
[37] Westerhof, Refining fast pyrolysis of biomass, Thesis Degree of Doctor of Science, University of Twente, 2011.
[38] Montoya .J, B. Pecha, F. Chejne Janna, Micro-Explosion of Liquid Intermediates During the Fast Pyrolysis of
Sucrose and Organosolv Lignin, Submitt. to J. Anal. Appl. Pyrolysis. (n.d.).
[39] S.J. Abramoff, M.D.; Magalhães, Paulo J.; Ram, Image processing with ImageJ, Biophotonics Int. 11 (2004)
36–42.
[40] E. Hoekstra, W.P.M. Van Swaaij, S.R.A. Kersten, K.J.A. Hogendoorn, Fast pyrolysis in a novel wire-mesh
reactor: Decomposition of pine wood and model compounds, Chem. Eng. J. 187 (2012) 172–184.
[41] E. Hoekstra, W.P.M. van Swaaij, S.R.A. Kersten, K.J.A. Hogendoorn, Fast pyrolysis in a novel wire-mesh
reactor: Design and initial results, Chem. Eng. J. 191 (2012) 45–58.
[42] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[43] Patwardhan, Understanding the product distribution from biomass fast pyrolysis, Iowa State University, 2010.
[44] G.B. Eijkel, J.J. Boon, P.W. Arisz, A.D. Pouwels, Evidence for oligomers in pyrolysates of microcrystalline
cellulose, J. Anal. Appl. Pyrolysis. 15 (1989) 71–84.
[45] Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, C.-Z. Li, A.G. McDonald, M. Garcia-Perez, Effect of
Cellulose Crystallinity on Solid/Liquid Phase Reactions Responsible for the Formation of Carbonaceous
Residues during Pyrolysis, Ind. Eng. Chem. Res. 53 (2014) 2940–2955.
[46] Teixeira, R. Gantt, K.E. Joseph, S. Maduskar, A.D. Paulsen, C. Krumm, C. Zhu, P.J. Dauenhauer, Spontaneous
Aerosol Ejection: Origin of Inorganic Particles in Biomass Pyrolysis, ChemSusChem. 0132 (2016) 1322–1328.
[47] J. Zhang, J.J.J. Chen, N. Zhou, Characteristics of jet droplet produced by bubble bursting on the free liquid
surface, Chem. Eng. Sci. 68 (2012) 151–156.
[48] S. Zhou, B. Pecha, M. van Kuppevelt, A.G. McDonald, M. Garcia-Perez, Slow and fast pyrolysis of Douglas-fir
lignin: Importance of liquid-intermediate formation on the distribution of products, Biomass and Bioenergy. 66
(2014) 398–409.
[49] F. Yi, G. Zhu, X. Zhu, R. Zhou, Z. Xiao, Pyrolysis characteristics of bean dregs and in situ visualization of
pyrolysis transformation, Waste Manag. (2012).
[50] G. Zhu, X. Zhu, Z. Xiao, F. Yi, Study of cellulose pyrolysis using an in situ visualization technique and
thermogravimetric analyzer, J. Anal. Appl. Pyrolysis. 94 (2012) 126–130.
[51] M. Zhou, S., Pecha, B., van Kuppevelt, M., McDonald, A.G. and Garcia-Perez, Slow and fast pyrolysis of
Douglas- Fir Lignin:Importance of liquid intermediate formation on distribution of products, Biomass and
Bioenergy. (2014) 398–409.
[52] J.Q. Feng, A computational study of high-speed microdroplet impact onto a smooth solid surface, (2016).
[53] C. Motzkus, F. Gensdarmes, E. Géhin, Study of the coalescence/splash threshold of droplet impact on liquid
Chapter 4 159
films and its relevance in assessing airborne particle release, J. Colloid Interface Sci. 362 (2011) 540–552.
[54] A.L. Yarin, Drop Impact Dynamics: Splashing, Spreading, Receding, Bouncing..., Annu. Rev. Fluid Mech. 38
(2006) 159–192.
[55] C. Mundo, M. Sommerfeld, C. Tropea, Droplet-wall collisions: Experimental studies of the deformation and
breakup process, Int. J. Multiph. Flow. 21 (1995) 151–173.
[56] Moreira, A.S. Moita, M.R. Panão, Advances and challenges in explaining fuel spray impingement: How much
of single droplet impact research is useful?, Prog. Energy Combust. Sci. 36 (2010) 554–580.
[57] G.E. Cossali, a. Coghe, M. Marengo, The impact of a single drop on a wetted solid surface, Exp. Fluids. 22
(1997) 463–472.
[58] S. Mandre, M.P. Brenner, The mechanism of a splash on a dry solid surface, J. Fluid Mech. 690 (2012) 148–
172.
[59] F.A. Williams, Ignition and burning of single liquid droplets, Acta Astronaut. 12 (1985) 547–553.
[60] M.L. Botero, Y. Huang, D.L. Zhu, A. Molina, C.K. Law, Synergistic combustion of droplets of ethanol, diesel and
biodiesel mixtures, Fuel. 94 (2012) 342–347.
[61] Simoneit, J.J. Schauer, C.G. Nolte, D.R. Oros, V.O. Elias, M.P. Fraser, W.F. Rogge, G.R. Cass, Levoglucosan,
a tracer for cellulose in biomass burning and atmospheric particles, Atmos. Environ. 33 (1999) 173–182.
[62] C.-L. Hsu, C.-Y. Cheng, C.-T. Lee, W.-H. Ding, Derivatization procedures and determination of levoglucosan
and related monosaccharide anhydrides in atmospheric aerosols by gas chromatography–mass spectrometry,
Talanta. 72 (2007) 199–205.
[63] P. Giudicianni, G. Cardone, R. Ragucci, Cellulose, hemicellulose and lignin slow steam pyrolysis: Thermal
decomposition of biomass components mixtures, J. Anal. Appl. Pyrolysis. 100 (2013) 213–222.
[64] J. Gosh, Rapid pyrolysis of sweet gum xylan with applications in modelling wood, MIT, 1983.
[65] M. Sohrabi, M.R. Hajaligol, M. Nik-Azar, B. Dabir, Mineral matter effects in rapid pyrolysis of beech wood, Fuel
Process. Technol. 51 (1997) 7–17.
[66] Y. Yürüm, N.. Öztaş, Pyrolysis of Turkish Zonguldak bituminous coal. Part 1. Effect of mineral matter, Fuel. 79
(2000) 1221–1227.
[67] B. Pecha, P. Arauzo, M. Garcia-Perez, Impact of combined acid washing and acid impregnation on the pyrolysis
of Douglas fir wood, J. Anal. Appl. Pyrolysis. 114 (2015) 127–137.
[68] Patwardhan, J.A. Satrio, R.C. Brown, B.H. Shanks, Influence of inorganic salts on the primary pyrolysis
products of cellulose., Bioresour. Technol. 101 (2010) 4646–4655.
[69] S.R.G. Oudenhoven, R.J.M. Westerhof, N. Aldenkamp, D.W.F. Brilman, S.R.A. Kersten, Demineralization of
wood using wood-derived acid: Towards a selective pyrolysis process for fuel and chemicals production, in: J.
Anal. Appl. Pyrolysis, 2013: pp. 112–118.
[70] Hogendoorn, E. Hoekstra, S.R.A. Kersten, W.P.M. van Swaaij, Fast pyrolysis in a novel wire-mesh reactor:
Design and initial results, Chem. Eng. J. 191 (2012) 45–58.
[71] Drummond, I.W. Drummond, Pyrolysis of Sugar Cane Bagasse in a Wire-Mesh Reactor, Ind. Eng. Chem. Res.
35 (1996) 1263–1268.
[72] Custodis, C. Bährle, F. Vogel, J.A. van Bokhoven, Phenols and aromatics from fast pyrolysis of variously
prepared lignins from hard- and softwoods, J. Anal. Appl. Pyrolysis. 115 (2015) 214–223.
[73] S. Wang, H. Lin, B. Ru, W. Sun, Y. Wang, Z. Luo, Comparison of the pyrolysis behavior of pyrolytic lignin and
milled wood lignin by using TG–FTIR analysis, J. Anal. Appl. Pyrolysis. 108 (2014) 78–85.
[74] M. Brebu, T. Tamminen, I. Spiridon, Thermal degradation of various lignins by TG-MS/FTIR and Py-GC-MS, J.
Anal. Appl. Pyrolysis. 104 (2013) 531–539.
[75] X. Yu, C. Shi, J. Ma, B. Zhu, M. Li, J. Wang, S. Yang, N. Kang, Aerosol optical properties during firework,
biomass burning and dust episodes in Beijing, Atmos. Environ. 81 (2013) 475–484.
[76] Q. He, X. Zhao, J. Lu, G. Zhou, H. Yang, W. Gao, W. Yu, T. Cheng, Impacts of biomass-burning on aerosol
properties of a severe haze event over Shanghai, Particuology. 20 (2015) 52–60.
[77] S. Samaras, D. Nicolae, C. Böckmann, J. Vasilescu, I. Binietoglou, L. Labzovskii, F. Toanca, A. Papayannis,
Using Raman-lidar-based regularized microphysical retrievals and Aerosol Mass Spectrometer measurements
for the characterization of biomass burning aerosols, J. Comput. Phys. 299 (2015) 156–174.
[78] V. Agarwal, G.W. Huber, W.C. Conner, S.M. Auerbach, Simulating infrared spectra and hydrogen bonding in
cellulose Iβ at elevated temperatures., J. Chem. Phys. 135 (2011) 134506.
[79] T.J. Hilbers, Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, M.R. Pelaez-Samaniego, M. Garcia-Perez,
160 Chapter 4
Cellulose-Lignin interactions during slow and fast pyrolysis, J. Anal. Appl. Pyrolysis. 114 (2015) 197–207.
Chapter 5 161

5. Chapter 5. Effect of Temperature and Heating Rate on


Product Distribution from Pyrolysis of Sugarcane Bagasse
in a New Hot Plate Reactor
(Paper published to Journal of Analytical and Applied Pyrolysis)

Abstact: This chapter reports studies on the film pyrolysis of sugarcane bagasse in an electrically heated hot plate
reactor. A new experimental set up used is described. The bagasse studied was ground to particle sizes less than 37
µm (mesh 400) and was mixed with water to form the paste to be used to form the biomass film (10 mg of sugarcane
bagasse distributed evenly on a steel plate, 60 µm thick). All the tests were conducted under vacuum (150 mbar). The
effect of heating rate (between 5 and 1200 °C/s) and final temperatures (between 250 and 650 °C) on product yields
(permanent gases, water, light volatile, pyrolytic lignin, sugars and char) were studied. The yield of fragmentation and
dehydration products increased as heating rates augment. The effect of heating rate is only important till 100 °C/s, no
significant changes in product yields are observed at higher heating rates. Bio-oils enriched in lignin and sugars
oligomers were obtained at high temperatures and heating rates. These results suggest that good yields of fermentable
sugars may be obtained in pyrolysis reactors working at intermediary pyrolysis conditions.

Keywords: Fast Pyrolysis, Intermediate Phase Liquid, Bubbles, distribution function, Pyroprobe.
162 Chapter 5

5.1. Introduction
The sugarcane industry generates 100 tons of bagasse per 1000 tons of sugarcane processed [1]. This material is
typically combusted at the mills for steam or power production. Sugarcane bagasse contains high amounts of volatile
material (about 80 wt.% DAF) [2]. This can be an excellent feedstock to produce value-added compounds via fast
pyrolysis [3]. The final product distribution of fast pyrolysis depends on reactor operating conditions (temperature,
pressure, gas flow entrainment, and reactor type) and on feedstock characteristics (particle size, nature of biomass
material, and mineral content) [4–7]. The main targeted product of fast pyrolysis is a viscous polar liquid with high acidity
known as Bio-oil [8–12]. Pyrolysis reactors are typically characterized as “fast” or “slow” depending on the heating rate.
At high heating fluxes (0.1 – 1.0 MW m-2 [13]) it is possible to obtain bio-oil yields between 40 and 70 wt. %. Much lower
bio-oil yields (between 20 and 50 wt. %) have been reported at lower heating rates [14–16]. However, the exact heating
rate at which the transition between slow and fast pyrolysis happens is unknown.

On an industrial scale, high heating rates are typically achieved in fluidized bed reactors with particle sizes between 0.5
and 2 mm [4,6]. These tests are typically conducted between 500 and 600 °C [14,17]. There are many parametric
studies on the effect of reaction conditions in fluidized bed reactors on the yield [8,9,13,16,18–29] and composition of
pyrolysis oils [6,30–35]. For fundamental studies, wire mesh [2,36–38] reactors have been used to achieve very high
heating rates (> 1000 °C/s) at the lab scale [2,36–38] with particle sizes typically between 100 and 150
microns. Although some researchers [36] have reported that between 42 µm to 2 mm the overall product yields are not
affected, recent experimental evidences [37, 39] suggest that even in small particles (<780 microns),
there are appreciable internal temperature gradients. The authors [37] found differences in the vapor products
(levoglucosan, HMF, light oxygenates) caused by the formation of inter-particle temperature gradients.

Thus, the main goal of this chapter is to describe a new analytical hot plate pyrolysis reactor and to use this experimental
set up to study the effect of final pyrolysis temperature and heating rates on the yield of sugarcane bagasse pyrolytic
products.

5.2. Materials and Methods

The study of primary products released during sugarcane bagasse pyrolysis was studied in a new hot plate reactor,
which allow controlled temperatures and heating rates. Thin film of biomass was created by impregnation on steel plate.

5.2.1. Hot Plate Reactor configuration

A schematic representation of the reactor used is shown in Figure 2. The body of the reactor is made of stainless steel
304, schedule 40. Two copper electrodes were connected to the steel plate (6 x 4 x 0.45 mm). In the top reactor cover
a μ-Epsilon CT SF 25pyrometer, spectral range 8 µm -14 µm, 25 was installed to measure the temperature on the
surface of the steel sheet. A laterally connected ¼ NPT fitting connects a DRUCK-DPI 104 pressure sensor to record
pressure changes (error of 0.07 mbar). A thermocouple type K (0.6 mm thickness) was installed 3 cm above the plate
to record the temperature of the gas phase. The connection indicated by the number 5 in Figure 1 is coupled to a
vacuum pump (Pfeiffer MVP 006-4) to reduce pressure to 150 mbar.
Chapter 5 163

Figure 1. Schematics of the hot plate reactor (1) External housing for cooling (2) Glass window (3) Port for gas
phase thermocouple (4) Pyrometer (5) Valve for sampling gases and pulling vacuum (6) Top cover reactor (7)
Reaction chamber (8) Pressure sensor port (9) Copper electrode (10) Gas tight electrical feed (11) Teflon insulation
(12) Screws for electrodes (13) Steel plate heated by electrodes (14) Fittings for electrode screws (15) Cold Plate (16)
Fittings to adjust the plate and feed cooling or hot fluids (CO2) (17) Copper sinks. (18) Port for carrier gas (Purge).

The outer shell of the reactor allows the addition of ice to maintain 0 °C on reactor walls, minimizing possible secondary
reactions. Two glass windows were installed on top of the cover for the pyrometer and for visualization. A heat
exchanger was formed as a flat plate, in order to quickly collect reaction products generated. Inside the exchanger
CO2 (g) is expanded from a tank to keep the surface at -10 °C. The reaction chamber has 5 ¼ NPT connections. In
one, a pressure sensor is installed. The remaining four are for input and output flows of nitrogen (passing through the
reaction chamber) and CO2 flowing through the cold plate heat exchanger. Copper electrodes have a diameter of 1/8”,
and a length of 10 cm. These pass through two cylindrical Teflon blocks for electrically isolating the housing and
electrodes. A set of 4 copper plates (2 at each end, 5 mm thick, 5 cm long x 2 cm wide) are used as terminals to transmit
power to the steel plate. The power system and control was designed in the laboratory facilities to have two independent
circuits. The first supplies power to the reactor and the second records data and feeds information to the PID
temperature control loop. Heating is accomplished by feeding AC current from a transformer (10 kW 220 V) to a 1/20
power increaser. Figure 2 shows an overview of the control system.

Figure 2. Power and control scheme for temperature control.


164 Chapter 5

The heart of the control system is a 16-bit micro-processor responsible for controlling the temperature measured by the
pyrometer. For reliable reading, the microprocessor record 50 times the temperature during the pyrometer response
time (10 ms) and average it before any control action. This signal is sent every 10 ms to a computer using USB
protocol. The temperature reading from the pyrometer is filtered by a digital filter (fourth order) to eliminate disturbances
and noise caused by the electrical field generated at the electrodes. Finally, the software calculates the power to be
delivered by the transformer using a PID control algorithm and sends the data to the micro-processor. The entire routine
is performed every 10 ms. At the maximum heating rate of this study (1200 °C/s) the system has a maximum
temperature lag of 12 °C and an overall response time close to 10 ms.

5.2.2. Plate Temperature Calibration and Temperature Distribution

Because of changes in emissivity of the surface of the steel sheet the measurements in the pyrometer have to be
corrected. Calibration was done by adjusting the pyrometer emissivity to match the reading of a 0.022” type K
thermocouple (Omega). The plate was heated at 50 °C/s and held at final temperature for 2 min between 100 and 800
°C. Readings with the thermocouple and pyrometer were made in the central point (1 in Figure 3).

Figure 3. Verification of the temperature distribution in the central area of the steel sheet cold rod. Figure A shows
what the plate looks like after heating, which illustrates that there are distinct heating regions. The pyrometer looks at
point 1 (B) on the plate. The other numbered points indicate placement of the thermocouple for comparison to IR
camera readings. At the end the calibration was verified by comparing melting points of 5 pure metals (indium, tin,
zinc, lead, and aluminum) to the corrected temperature from the pyrometer.

The temperature in the biomass film was also measured by inserting a standard thermocouple type K chal 0.022 " on
the film. These additional experimients, are made to validate the hypothesis of Biot <0.1, and confirm that the
temperature of the film biomass is the same as that registered on the surfaces of the steel sheet. These hypotheses
were confirmed. The temperature distribution was checked recording 2 videos with thermal camera FLIR E60, one
video at 50 °C/s and Tf=600 °C, the second at 200 °C/s and Tf=400 °C. (see figure 8).

5.2.3. Materials

Sugarcane bagasse was obtained from the sugarcane mill Risaralda , Colombia (4 °53'48 "N 75 °53'02 "O), ground in
a micro-hammer mill (IKA MF-10 3, 500 rpm), and sieved with 250 mesh (63 microns). In Tables 1-3 of chapter 2, the
physicochemical characterization of this raw material and the methods used for compositional analyzes are shown.

5.2.4. Sample preparation


Chapter 5 165
Biomass films were prepared by suspending finely ground sugarcane bagasse (Figure 4A) in distilled water. The
suspension is prepared with a solids concentration of 10 wt.%. A thin brush is dipped into the suspension of the biomass,
and painted over the central strip (63 mm long) of a steel plate (6 cm x 4 cm x 0.45 mm) (Figure 4B). The water from
the suspension was removed by drying the sheet in a Mettler Toledo moisture balance (HB 34) under halogen slowly
at 80 °C. The resulting dry film of biomass adhered to the plate surface is shown in Figure 1C. To confirm the thickness
of the biomass on the surface of the sheet, the edge was imaged with an optical microscope at 100 X [37]. On average,
in each experiment 10 mg of bagasse was applied. The weight of the impregnated biomass is obtained by difference
between the weight of the steel sheet before and after being impregnated with dry biomass.

Figure 4. Preparation of thin film of bagasse on the steel plate (A) Raw ground biomass (B) Steel plate (C) Biomass
impregnated on the steel plate (D) Char after pyrolysis.

The thickness of the biomass on the plate was measured using a Nikon Eclipse TE-200 microscope with a 20X objective
lens using the film methodology described in Section 2.2. With the film, the sheet was cut in half and placed vertically
under the microscope. This procedure was performed with three different impregnations. In each image, five different
height measurements were taken and averaged (see Figure 10). The distribution of biomass in the plate was examined
with a Wild M3 stereoscope. Direct views of biomass objectives between 10X/21B with 0.6-1-1.6-2.5-4 magnification
and image were made last seen 20 X under a Nikon Eclipse TE-200 microscope, in order to see the distribution of
biomass. Tests were performed to determine the effect of the amount of mass and hold time on the yield of char obtained
and thus verify kinetic control conditions and reaction time. Figure 5 shows the distribution of the impregnated biomass
on the steel sheet at various magnifications. From the chart (A) - (F) a fairly smooth surface with very few irregularities
is observed.

Figure 5. Viewing the distribution of biomass particles on the steel sheet. (A) 0.64X (B) 1.0X (C) 1.6X (D) 2.5X (E)
4.0X (F) Increase 20X
166 Chapter 5

5.2.5. Run Sequence

Each experiment was performed in triplicate to ensure reproducibility. The weight of the steel plates was recorded
before and after impregnation and drying with a Mettler Toledo analytical balance (XS205DU, d = 0.01 mg). On average
10 mg bagasse was deposited on the central part of the plate where there is homogenous temperature (reaction area
of 1 cm x 4 cm) (see Figure 8).

mass sample = (Plate + Biomass)after drying − (Empty Plate)Before impregnation (1)

The plate is then clamped between the two electrodes. The reactor is sealed and ice is added around the outer casing
to cool the outer walls of the reactor. Nitrogen (> 99.996 % purity) is fed for 2 minutes at a rate of 5 l/min to remove any
air contained within the reactor. The vacuum pump evacuates nitrogen until the absolute pressure was 150 mbar. The
software is set for heating rates between 5 and 1200 °C /s and final temperature between 250 and 650 °C. The heat
exchanger is chilled using CO2 until the outlet temperature is -10 °C. At the beginning and end of each experiment the
gas temperature and pressure are recorded. At the end of each experiment, the ice casing is removed, the CO2 cooling
valve closed, and the reactor is brought to room temperature. After reaching room temperature (measured by the
thermocouple mounted inside the reactor), the reactor is pressurized to just above 1 bar using nitrogen.

5.2.6. Permanent gases

The gas sample is taken into a 200 ml Tedlar bag through valve 5 (see Figure 1) and is analyzed using a μ-GC (Agilent
3000) with two independent channels. The first contains a Molsieve 5 oA 10 m capillary micro column to quantify CO,
H2 and CH4. The second module contains a micro-capillary column 10 m QP100 for the analysis of CO2. The carrier gas
was helium for both channels with a 5.0 grade pressure in both columns at 20 psi. The suction time of the sample was
500 ms and the temperature in both ovens was maintained at 50 °C and the total analysis time is 3 min. The chemical
composition of the generated species were normalized in a nitrogen free basis. Gas yields were estimated by knowing
the mass of the pyrolyzed biomass temperature and initial and final pressure. The final yield of each chemical species
in the gas phase is determined with the total estimated yield and their respective normalized mole fractions (without
nitrogen). The micro-GC was calibrated using certified mixtures of gases, in concentration ranges 0, 100, 200, 400, 800
ppm. The reactor volume was 897 ml. It was measured filling the reactor with water.

Gas yield was determined by ideal gas equation of state:

w
G𝑎s Yield (% )
w
n=4 Vreactor %moli (2)
(Pafter experiment − Pbefore experiment ) RT Mwi (
100
)
room
= 100 ∑
mbiomass
i=CO,CH4,CO2 ,H2

Yield of each quantified species in analysis was obtained by multiplying the total gas yield (Ec 2) and the volume fraction
of each quantified species in the μ-GC and corrected by the ratio of molecular weights of the species to analyze and
molecular weight of the gas mixture:

%moli Mwi
Yieldi (%w/w) = Gas Yield ( ) (3)
Mwmixture
Chapter 5 167

5.2.7. Char

Char in these experiments is defined as the carbonaceous material remaining after each experimental test. The biochar
yield was determined by weighing the steel sheet with a final carbonaceous material, and knowing the initial weight of
the biomass.
Char yield was determined by the weights of steel plate without sample, with sample after impregnation and
char at the end of each experiment.

(Plate + char)after pyrolysis − Platebefore impregnation


Char yield (%w/w) = 100 (4)
(Plate + biomass)after drying − Platebefore impregnation

The morphology of some char samples was studied using SEM (ZEISS model EVO MA 10 with a tungsten filament,
Oxford Instruments X-act detector coupled to an EDX spectrometer). Before each analysis the samples were gold
sputter coated (40 s, 5 mA). EDX analysis was used to estimate the presence of mineral material through the sample.

5.2.8. Bio-oil

Because the diffusion of gases in vacuum is high, part of the bio-oil is condensed through all internal walls of the reactor,
making it quite difficult to make a full recovery. Thus the bio-oil yield was determined by difference from the overall
mass balance.

Bio oil Yield (% w/w) = 100 − Gas Yield − Char Yield (5)

A bio-oil sample condensed on the surface of the condenser was collected with quartz gauze, the weight of the gauze
before and after collecting the sampling was recorder to estimate the mass for the sample. The bio-oil collected was
dissolved in methanol and the resulting solution analyzed by GC/MS. The yield of the individual condensable species
was calculated by multiplying their mass fraction in the bio-oil by the bio-oil yield.

Analysis of the condensable species, was made with a spectrometer Shimadzu QP 2010 using a capillary column Shrxi-
MS of 0.25 mm ID, 30 m long and 0.25 µm stationary phase. The chromatograph oven was programmed with two
temperature ramps beginning at 50 °C and heated at 10 °C/min to 250 °C, then at 20 °C/min to 320 °C with a holding
time of 20 min. The flow of carrier gas is He at 54 l/min, power of 0.1 kV filament, split ratio 1, injection volume 1 µl. The
sample was injected into an injection port at 250 °C. Before and after each injection the syringe washes five times with
methanol. Calibration solutions were prepared with 5 different mass fractions of selected pyrolysis products
(levoglucosan, 5-HMF, furan, formic acid, acetic acid, acetoin, glycoaldehyde, furfuryl alcohol, phenol, eugenol, vanillin,
naphthalene, pyrocatecol, benzofuran and water). The compounds used for calibration were all analytical grade purities
greater than 98 wt.%. The moisture content was determined by quantifying the ion m/z 18 by mass spectrometry using
a method described elsewhere [39]. The water mass fraction results were validated by comparing the results obtained
by GC/MS for samples with those obtained by Karl Fisher titration. Identification of each species was achieved knowing
the retention time of each compound and comparing the spectra with those reported by the NIST database library.

Anhydro-sugar analysis was performed with a Varian ProStar HPLC (refractive index detector, Aminex HPX-87H
column). The furnace was maintained at 80 °C throughout the analysis with nano-pure water (18 mS) (flow rate: 2
ml/min). The injection volume was 40 µl. Calibration curves were made for celotriosan, cellobiosan, 1,6 AGF, 1.4-3.6
DGP, levoglucosenone with 4 mass fraction levels.
168 Chapter 5

An important fraction of bio-oil is oligomeric lignin (also known as pyrolytic lignin), which can be up to 10 wt.% of bio-oil
[40]. The yield of oligomeric lignin was determined by a modified UV fluorescence methods [40] calibrated with a
pyrolytic lignin obtained by cold water precipitation. Briefly, 1 liter of bio-oil obtained from the fast pyrolysis of sugarcane
bagasse (particle size 1 mm, fluidized bed reactor at 500 °C, with flow biomass feed 10 kg/h, entraining gas flow 50
l/min) was precipitated in cold water [2]. The lignin oligomeric fraction collected by filtration was used to prepare
standard solutions in methanol (concentrations between 0.04 to 0.5 ppm for quantification of pyrolytic lignin yield (these
low concentrations prevent self-absorption effects and redshift). The methanol solutions were analyzed on a Perkin
Elmer LS 55 spectrofluorometer with a quartz cell of 1 cm nominal length. Data were obtained in synchronous mode
with an energy gap of 2 800 cm-1 and a scan speed of 50 nm/min. Each spectrum corresponds to the average of 3
scans. The content of lignin was calculated using the calibration curve obtained with the pyrolytic lignin separated from
the bio-oil produced in the fluidized bed [2].

5.3. Results and Discussion

5.3.1. Temperature Calibration and Distribution

To perform thermochemical reaction studies, the temperature of the sample and the experimental set up need to be
known. The fit between the thermocouple temperature measured at the midpoint of the steel sheet and the uncalibrated
pyrometer (emissivity: 1) is shown in Figure 5-A. The average temperature difference between the uncalibrated
pyrometer and the thermocouple is 90 °C. The curve shown in Figure 5 was used to correct the measurements obtained
with the pyrometer. The corrected values of pyrometer temperature were compared with the melting point of known
metals with good agreements (see Figure 5B).

Figure 6. (A) Uncalibrated pyrometer vs centerpoint thermocouple temperature readings used to develop the
calibration curve. (B) Calibrated pyrometer temperature vs melting points of pure metals at atmospheric pressure.
Differences between the melting points and thermocouple were less than 10 °C.

The temperature distribution on the steel plate is shown in Figure 7. The central strip of 1 x 4 cm (central strip of Figure
7-A) is the hottest; the average temperature throughout this range is 383 °C ± 7 °C with a difference of 23 °C between
the center and the edge. This variation is quite small, and it is important to note that this is still in the heating up period.
Figure 6-B shows the thermal image at the same time point as data in Figure 6A (400 °C). This image also shows
relatively even temperature distribution along the plate.
Chapter 5 169
A. K-type measurements B. Thermography-FLIR

Figure 7. Comparison of temperature distribution in the steel sheet. (A) Temperature measurements with type K
Thermocouple at 9 points around the central region. Tf = 400 °C, holding time 60s. (B) Frame 240 of sheet
temperature Tf = 400 °C.

The temperature distribution during heating is uniform in the central strip. Unfortunately, for higher rates the camera is
not fast enough to follow the temperature.

A. Temperature distribution when plate is heated at 50°C/s and Tf=600°C.

B. Temperature distribution when plate is heated at 200°C/s and Tf=400°C.


170 Chapter 5
Figure 8. Temperature gradients at different times taken with infrared camera when the plate is heated at 50 °C/s
and 600 ºC.B. Temperature distribution at different times when plate is heated at 200 °C/s and final temperature of
400 °C.

The temperature was uniform in the central area (see Figure 8). At lower heating rates (8A), the steel plate has more
time to achieve a uniform distribution. The temperature distribution at moderate heating rates (200 °C/s) is too uniform.
In both cases the area with uniform temperature is 2 cm*4 cm, in this way we assure that sample is heated at the same
conditions in all points in the central frame. Similar results were obtained by Prins et al [41]. The author shows uniform
temperature distribution in a Ni80/Cr20 foil of 1.6 cm*1 cm*0.1mm heated at 725 °C (between 300 °C/s-600 °C/s) with
differences in temperature around 50 °C. The authors recommend the use of foils instead of meshes to assure uniform
temperature distributions.

The next step was to evaluate the effect of biomass film loading on temperature measurements. Figure 9 shows that
variation between the loaded and unloaded plate is very small, indicating that biomass load has no effect on the
temperature as recorded by the pyrometer.

Figure 9. Comparison of records temperature on the surface of the steel sheet and on the biomass surface.

The temperature history, is practically the same between hot plate surface and biomass thin film surface. With amount
of biomass loaded and spread over the plate, a very thin film is formed with very low resistance to the heat transfer by
conduction. In other words the Biot number is small (Bi <0.01). So, for practical purposes it is considered that the
biomass sample and the plate surface temperatures are the same.

5.3.2. Impact of holding time and film thickness on char yield

The study of the effect of particle thickness and hold time was determined at 500 ° C/s and 500 °C, 150 mbar pressure,
varying thicknesses of biomass between 29 – 291 μm. Figure 10 shows an optical microscope visualization of
sugarcane bagasse films with thickness of 60 ± 13 microns.
Chapter 5 171

Figure 10. Film thickness impregnated biomass on the surface of cold steel rod. (A) Visible Light. (B) Fluorescent
Light.

Char yields obtained with different thicknesses (different mass) are shown in Figure 11A. Particles <100 microns
produce the same char yields (≈ 10 wt. %). These results indicate that secondary reactions solid-gas within the film
(thickness <100 microns) do not alter the product distribution. So it was decided to conduct the rest of experiments with
films having an average thickness of 60 ± 13 microns. Figure 11B shows that the yield of char is the same for different
hold times.

A. Effect of film thickness B. Effect of Holding time

Figure 11. A. Char yield versus film thickness for pyrolysis at 500 °C (500 °C/s, 150 mbar pressure, hold time 0s). B.
Char yield versus hold time for film thickness of 60 ± 13 µm at 500 ° C (500 °C/s, 150 mbar pressure).

5.3.3. Species studied in these experiments

In Table 1, the major compounds identified by GC-MS, GC-μ, and HPLC are listed. The names of molecules, retention
time, molecular ions, were identified by NIST library.

Table 1. Main Compounds identified by GC-MS, HPLC, UV-Fluorescence, µ-GC.


GC-MS
Peak Compound time (min) Formula m/z MW (g/mol)
1 Water 1.045 H2O 18 18.015
2 Furan 1.230 C4H4O 39 68.074
3 Formic Acid 1.250 CH2O2 29 46.025
4 Acetic Acid 1.500 C2H4O2 43 60.052
5 Glycoaldehyde 1.950 C2H4O2 31 60.052
6 Acetoin 2.020 C4H8O2 45 88.105
7 Furfuryl Alcohol 3.500 C5H6O2 98 98.099
172 Chapter 5
8 Phenol 5.290 C6H6O 94 94.111
9 Naphthalene 8.415 C8H10 128 128.170
10 Pyrocatechol 8.535 C6H6O2 110 110.110
11 Benzofuran 8.825 C8H6O 118 118.132
12 5-Hydroxymethyl-Furfural 9.025 C6H6O3 97 126.110
13 Eugenol 10.835 C10H12O2 164 164.201
14 Vanillin 11.395 C8H8O3 151 152.147
15 Levoglucosan 12.500 C6H10O5 60 162.140
µ-GC
1 Hydrogen 0.200 H2 2.016
2 Carbon Dioxide 0.440 CO2 44.010
3 Carbon Monoxide 1.125 CO 28.010
4 Methane 1.575 CH4 16.040
HPLC

1 Cellotriosan 10.948 C18H30O15 486.420


2 Cellobiosan 13.855 C12H20O10 324.280
3 1,6 Anhydro β-D-Glucofuranose 16.248 C6H10O5 162.140
4 1,4:3,6 Dianhydro α-D-Gluco Pyranose 20.005 C6H8O4 144.130
5 Levoglucosenone 21.865 C6H6O3 126.110
UV-Fluorescence
1 Lignin Oligomers 300 - 380 nm

5.3.4. Effect of Peak Temperature over volatile species yield identified by GC/MS

A parametric study was performed to determine the effects of temperature and heating rate on product yields (water,
acetoin, acetaldehyde, formic acid, acetic acid, furfuryl alcohol). Two heating rates were used (5 and 1200 °C/s) and
temperature between 177 to 677 °C. Results are shown in Figure 12. Most of these products are mainly formed by
fragmentation reactions, decarboxylation, and pyranose ring formation. Furanose comes from cellulose and
hemicellulose [42,43]. Yields of all these products increase with temperature in agreement with results in the literature
[44,45]. At a lower heating rate (5 °C/s), the yields of these products are higher than at high heating rates (1200 °C/s). At
low heating rates, the dehydration and crosslinking reactions in the solid phase or in the liquid intermediate are
enhanced [42].

In this study, glycoaldehyde has a maximum yield of 5 wt. % at 5 °C/s. Paulsen et al. [37] found for fast pyrolysis of
cellulose glycoaldehyde yield increases with temperature until being constant at 6.7 wt.% after 450 ° C. It has been
reported that the presence of mineral material, could reduce yields because it promotes fragmentation reactions and
ring opening of the pyranose and furanose [43]. Similarly, yields of other light oxygenates such as acetoin, formic acid,
acetic acid, furfuryl alcohol increase with temperature; this is in agreement with other investigations on fast pyrolysis of
biomass [42,44–46]. Acetic acid can be formed by various reaction routes, mostly through de-acetylation of
hemicellulose at 300 °C [43]; some also comes from rearrangement of the pyranose rings in cellulose [47,48]

A. Water B. Acetoin
Chapter 5 173

C. Formic Acid D. Acetic Acid

E. Furfuryl Alcohol F. Glycoaldehyde

Figure 12. Distribution of light oxygenates analyzed by GC-MS and obtained by sugarcane bagasse pyrolysis in a hot
plate reactor at 1200 °C/s and 5 °C/s and 0 holding time. (A) Water yield (wt.%). (B) Acetoin yield (wt.%). (C) Formic
acid yield (wt.%). (D) Acetic acid yield (wt. %). (E) Furfuryl alcohol yield (wt.%). (F) Glycoaldehyde yield (wt.%).

The water yield (Figure 12A) increases with temperature, reaching a peak of 19.0 wt. % at the low heating rate, mostly
due to enhanced crosslinking/dehydration reactions [49]. At high heating rates, the samples depolymerize mostly over
350 °C, at these temperatures depolymerization reactions dominate over crosslinking reactions. At 500 °C with 5 °C/s,
the largest amounts of water are produced. At this temperature (200 - 300 °C), hemicellulose is known to decompose
and form water [47,50–52]. An important amount of water is also produced from the dehydration of cellulose primary
products in the liquid intermediate [53]. Lignin is degraded in a wide temperature range (300 - 500 °C); its main products
are oligomers and phenols, but water can also be produced in small quantities [54]. Yields of Furan and 5-HMF (Figure
13) increase with temperature. These compounds are also released from fragmentation reactions (ring-openings of
cellulosic pyranose and hemicellulosic furanose rings).
174 Chapter 5
A. Furan Yield B. 5-HMF Yield

Figure 13. Yields of 5 member ring compounds analyzed by GC-MS obtained by sugarcane bagasse pyrolysis in a
hot plate reactor at 1200 °C/s and 5 °C/s and 0 holding time. (A) Furan yield (wt. %). (B) 5-HMF yield (wt.%).

The formation of 5-HMF is promoted by acid catalysis [55]. During pyrolysis the sugar rings open and reorganize to
form new sugars such as fructose. Dehydration of fructose produces 5-HMF directly [50]. Mettler et al. [47] found
however that both 5-HMF and glycoaldehyde are primary reaction products of cellulose which are formed by breaking
the glycosidic homolitic bonds and subsequently breaking pyranose rings. 5-HMF yields for fast pyrolysis of cellulose
are typically in the range of 3 - 8.7 wt. % (cellulose basis) [37],[56],[56]. Results at high heating rates are lower (1 wt.
% at 1200 °C/s). Both temperature and lignin content increase the yield of 5-HMF [60]. Zhang, et al. [45] found that
lignin interactions has an effect on the performance of furans, increasing it 1.4 wt. %, due to the presence of lignin-
cellulose covalent bonds at carbon 6 in pyranose ring.

Hemicellulose is another possible source of furanic molecules [51]. The interaction of hemicellulose and lignin furans
increases its yield [45]. In results, furan yield increases with temperature, in correspondence with other studies [40,57].

The distribution of the main lignin monomeric products is shown in Figure 14. As temperature increases, yields of
naphthalene, phenol, pyrocatechol, eugenol, vanillin, and benzofuran also increase. This is in agreement with the
literature [44,45,58]. It has been reported that the primary products of lignin decomposition are high molecular weight
oligomers and a smaller quantity of monomers [59–62]. At high heating rates, the residence time of these oligomers
within the particle is low (< 2 s) so the mono-phenolic yield is lower compared to low heating rates. The phenolic
monomers are produced by both primary reactions of lignin and from the further degradation of the oligomers.

A. Naphthalene Yield B. Phenol Yield

C. Pyrocatechol Yield D. Eugenol Yield


Chapter 5 175

E. Vanillin Yield F. Benzofuran Yield

Figure 14. Distribution of aromatics products analyzed by GC-MS, and obtained by devolatilization from sugar cane
bagasse (A) Naphtalene (B) Phenol (C) Pyrocatechol (D) Eugenol (E) Vanillin (F) Benzofuran.

5.3.5. Lignin oligomers formation

The UV fluorescence spectra of the bio-oil obtained at 300, 400 and 600 (500 °C/s) are shown in Figure 15A. Two
representative peaks are shown: Peak A is between 296 - 326 nm and peak B is between 326 - 380 nm. Tests with
solutions of eugenol, vanillin, 1, 2, 3, -trimethoxy benzene (single-ring phenols) in methanol at 4 ppm showed that these
peaks appear in between 280 - 290 nm. Levoglucosan barely shows any intensity at 280 nm. The results shown in
Figure 15 confirm that Peak B is mostly due to the presence of oligomeric compounds. Peak B, increases dramatically
with temperature. This indicates the formation of high molecular weight compounds with more than two aromatic rings.

A B
176 Chapter 5
Figure 15. A. UV-fluorescence spectra of bio-oil samples obtained by sugarcane bagasse pyrolysis in a hot plate
reactor at 300 °C, 400 °C, 600 °C heating rate of 500 °C/s, 0 holding time. B. Standards of eugenol, vanillin, 1, 2, 3-
trimethoxy benzene, and levoglucosan.

Oligomers lignin phenols are known to be mixture of two to four aromatic rings (average molecular weights 500 Da [63],
645 Da [60]). Using DFT, researchers have concluded that excitation wavelength increases as rings grow to clusters of
four, but beyond four rings there is no difference in the excitation wavelength (plateau) [61]. It is also important to
mention that experiments were conducted in vacuum, so it is possible that oligomers with high molecular weight quickly
escape through vaporization or thermal ejection [40].

Figure 16 shows the yields of total oligomers obtained depending on the temperature and heating rate. The yield of
lignin oligomers increases with heating rate and temperature. It is known that during pyrolysis lignin in the biomass is
in a melt phase [64,65], in which bubbles are formed responsible for ejection of aerosols [66]. At higher temperatures
and heating rates the intensity of bubble formation and microexplosion increased. This claims have been validated with
pure liquids [67–70].

Figure 16. Effect of temperature and heating rate on pyrolytic lignin yields measured by UV-Fluorescence
Spectroscopy and obtained by devolatilization from sugar cane bagasse in a hot plate reactor at 1200 °C/s and 5
°C/s and 0 holding time.

5.3.6. Temperature effect on anhydrosugar yields

Figure 17 shows the distribution of levoglucosan (GC-MS) and total sugars determined by HPLC. The profile of
identifiable sugars by HPLC shows only cellotriosan, cellobiosan in very small amounts (about 200 ppm) and 1,4:3,6
dianhydroglucopyranose (DGP). The levoglucosan yield averages 8.65 wt. % at high heating rates (1200 °C/s), and is
double that obtained at low heating rates (4.42 wt.% at 5 °C/s). Biomasses with low content of mineral material can
reach levoglucosan yields of between 8 and 18 wt.% [65], [73], [71].

A. Levoglucosan Yield B. Sugars Yield


Chapter 5 177

Figure 17. Effect of temperature and heating rate on sugars yields measured by HPLC and obtained by
devolatilization from sugarcane bagasse in a hot plate reactor at 1200°C/s and 5 °C/s and 0 holding time. The sugars
yield represents cellobiosan, cellotriosan, and 1,4:3,6 dianhydroglucopyranose (DGP).

At low heating rates, the temperature at which levoglucosan formation begins is approximately 277 °C (yield 1 wt. %).
Constant yield is achieved beyond 350 °C (623 K). Levoglucosan boils at 292 °C [68]. The high molecular
anhydrosugars content is low, indicating that they are decomposed into low molecular weight sugars such as
levoglucosan and 1,4:3,6 DGP. Final yield of levoglucosan is greater at high heating rates. Many researchers suggest
that levoglucosan yield reduces by effect of alkali cations, especially potassium. Alkaline metals can promote
fragmentation and disproportiation reactions, increasing he amount of acetaldehyde, glyoxal and other light oxygenated
compounds. Alkali metals might inhibit levoglucosan yield by capping the free ends groups of cellulose chain and thus
avoiding the unzipping reactions continues. Potassium can also inhibit levoglucosan yield by disrupting the Trans
glycosylation reaction, promoting the formation of carbonyl groups, double bonds. In the same way, Patwardah shows
the catalytic effect of alkali metal and anions over formation of formic acid, glycoaldehyde and acetol; there are a
competitive mechanism where alkali suppress the reaction leading to levoglucosan formation, but it is possible that
alkali accelerates the rate of dehydration in liquid phase in contrast with evaporation. Although some researches argue
that acid or alkali effect has secondary importance because the catalytic effect is likely through ionic medium [72].
However, in this particular study, there are two competing effect, the first one is the possible reduction in levoglucosan
yield by catalytic effect of potassium and the second one is the possible increase of levoglucosan yield by rapid remotion
of levoglucosan favored by vacuum. The final result depends of which of both parameters is more sensitive. For
instance, for cellulose pyrolysis at 300 ºC [73]. there ir 53 % more levoglucosan at 1.5 mmHg and 5% SbCl3 in
comparison with atmospheric pressure without catalyst. Likewise in Cellulose impregnated with acid [74] during vacuum
pyrolysis (130 Pa) at 2 °C/min levoglucosan yield increases, contrary to results at atmospheric pressure [75].
178 Chapter 5

5.3.7. Temperature effect over Permanent Gases

Figure 18 shows the yields of permanent gases detected during pyrolysis of bagasse. The yield of carbon monoxide
increases with temperature and remains constant after 773 K (500 °C). Carbon monoxide comes primarily lignin
decomposition at low temperatures (T < 500 K). In lignin, CO is generated by fragmentation reactions of carbonyl and
carboxyl groups attached to γ-carbons [76–78]. The α-O-4 and β-O4 bonds tend to fragment after 200 °C and 245 °C
releasing CO and CO2 respectively. At high temperatures (T > 773 K), the source of CO is ether bond breaking 4-O-5
[76,79]. Hemicellulose decomposition at low temperatures (T < 250 °C) produces CO2 [44,47] and char. At low heating
rates, pyranose rings in the intermediate liquid phase release CO2 [40,42]; this partly explains the decrease in yield with
increasing heating rate. Similar results were obtained by Hoekstra et al. [80] pyrolyzing pinewood at 6000 °C/s. At low
heating rates, the majority of the CO2 is released before 600 K (327 °C) with carbonyl and carboxyl decomposition of
hemicellulose [56]. Lignin is another precursor of CO2, especially by decarboxylation reactions at low temperatures (T
< 300 ° C). Lignin oligomers produce large amounts of CO2 [81], at high heating rates these oligomers rapidly leave the
solid matrix of the biomass, thereby reducing their contribution to the production of CO2.

A. Hydrogen Yield B Methane Yield

C Carbon Monoxide Yield D. Carbon Dioxide Yield

Figure 18. Distribution of permanent gases at different temperatures and heating rates.

The yields of hydrogen and methane are quite low (< 0.5 wt.%), similar to other researchers [80]. Methane and hydrogen
likely come from lignin [82]. Generally, it is observed that the effect of the heating rate is most dramatic for CO2.
Chapter 5 179

5.3.8. Temperature effect on char yield

Figure 19 shows yields of char as a function of temperature at two heating rates.

Figure 19. Distribution of Char depending on the temperature and heating rate.

As the temperature increases, char yield decreases until 740 K (467 °C) at 1 200 °C/s, and 673 K (400 °C) at 5 °C/s.
There is a difference of 25 % in char yields produced at 1200 °C/s and 5 °C/s, respectively. As the heating rate
increases, char decreases dramatically. As discussed in the previous section, the content of oligomers and lignin
anhydro-sugar increases with the heating rate due to rapid evaporation and aerosol ejection [49,59,71]. These
oligomers are precursors of dehydration and crosslinking reactions which lead to the generation of char [49,62,80]

Morphological changes of biomass pyrolyzed at high (1200 °C/s) and low rates of heating (5 °C/s) and different
temperatures (500 °C - 300 °C) are presented in Figure 20. In addition, it shows the distribution of major minerals in
the carbonaceous residue obtained at 1200 °C and 500 °C. At low temperatures (300 °C), irrespective of the heating
rate, no morphological changes were observed. From 500 °C, both high and low heating rates are softening the solid
matrix. Finally the spectrogram obtained by EDX 1200 °C/s and 500 °C, shows the distribution of minerals and carbon
species present in the char analyzed. Spectrograms show that in the analyzed area, the ratio between the Si and the
catalytically active cations (K, Ca, Mg, Na) is very high, with a ratio of 60/1 (by fractions obtained by EDX).

A 1 200 oC/s – 300 oC B 5 oC/s – 300 oC


180 Chapter 5

C 1200 oC/s – 500 oC D 5oC/s – 500 oC

E 1 200 oC/s – 500 oC

Figure 20. SEM visualization of morphological changes of sugarcane bagasse chars.

5.3.9. Heating rate effect on main products yield at 500 °C

The effect of the heating rate on the yield of the main products analyzed for pyrolysis of bagasse at 500 °C are shown
in Figure 21. It is clear that after 100 °C /s it can not be seen any significant effect of the yield of the products studied.

A. Light oxygenates B. Phenolic compounds

Heating rate (°C/s)


Heating rate (°C/s)

C. Permanent gases D. Oligomers-Char-Water


Chapter 5 181

Heating rate (°C/s) Heating rate (°C/s)

Figure 21. Effect of heating rate on product distribution during the pyrolysis of sugarcane bagasse in a hot plate
reactor at 500 °C (773K).

Above 100 °C /s, for most compounds, changes in yields are minimal. For heating rates below 100 °C /s, the oligomeric
primary products can undergo secondary reactions [82]. For heating rates at 100 °C /s, the secondary dehydration and
crosslinking reaction in condensed stages seems to be minimized [85-87]. In the range of heating rates used in this
study, no major changes were observed in yields of hydrogen and methane, as seen by others [80]. With respect to
CO2 and CO, results show a decrease in the yield of CO2 for heating rates above 100 °C/s. An important source of
CO2 is the hemicellulose and cellulose via decarboxylation reactions of the acetyl groups (hemicellulose) and
fragmentation reactions of the pyranose ring (sugars). As the heating rate increases, the fragmentation reactions of
sugars are minimized. The increase in CO is related to depolymerization of cellulose and lignin.

5.4. Conclusions
In this article a new reactor and methodology was presented to study sugarcane bagasse with direct impregnation of a
biomass sample on a steel sheet. The film used for pyrolysis studies had an average thickness of 60 µm, uniform over
the entire surface distribution. This allowed for the same reaction conditions each point on the film. Parametric studies
across temperature and heating rate were performed with sugarcane bagasse. The results showed that beyond 100
°C/s, no significant changes are observed in the yields of the products due to the low thermal conductivity of the biomass
[36]. At high heating rates (> 100 °C/s), the highest yields of oligomers of sugars and lignin and lowest production of
char were observed. The bio-oil obtained at low heating rates were high in water, organic acids, aldehydes, and ketones
with a lower concentration of oligomers compared recovered with high heating rates. Overall yields obtained at high
heating rates are on the same order of magnitude as those previously reported by [2] for bagasse in a fluidized bed
reactor.
182 Chapter 5

5.5. References
[1] Drummond, I.W. Drummond, Pyrolysis of Sugar Cane Bagasse in a Wire-Mesh Reactor, Ind. Eng. Chem.
Res. 35 (1996) 1263–1268.
[2] J.I. Montoya, C. Valdés, F. Chejne, C.A. Gómez, A. Blanco, G. Marrugo, J. Osorio, E. Castillo, J. Aristóbulo,
J. Acero, Bio-oil production from Colombian bagasse by fast pyrolysis in a fluidized bed: An experimental
study, J. Anal. Appl. Pyrolysis. 112 (2015) 379–387.
[3] J.I. Montoya Arbeláez, F. Chejne Janna, M. Garcia-Pérez, Fast pyrolysis of biomass: A review of relevant
aspects. Part I: Parametric study, DYNA. 82 (2015) 239–248.
[4] A. Murty Kanury, Thermal decomposition kinetics of wood pyrolysis, Combust. Flame. 18 (1972) 75–83.
[5] H.S. Heo, H.J. Park, Y.-K. Park, C. Ryu, D.J. Suh, Y.-W. Suh, J.-H. Yim, S.-S. Kim, Bio-oil production from
fast pyrolysis of waste furniture sawdust in a fluidized bed., Bioresour. Technol. 101 Suppl (2010) S91–S96.
[6] H.J. Park, J.-I. Dong, J.-K. Jeon, Y.-K. Park, K.-S. Yoo, S.-S. Kim, J. Kim, S. Kim, Effects of the operating
parameters on the production of bio-oil in the fast pyrolysis of Japanese larch, Chem. Eng. J. 143 (2008)
124–132.
[7] J. Lehto, A. Oasmaa, Y. Solantausta, M. Kytö, D. Chiaramonti, Fuel oil quality and combustion of fast
pyrolysis bio-oils, VTT Publ. (2013) 79.
[8] A. Oasmaa, D. Meier, Norms and standards for fast pyrolysis liquids: 1. Round robin test, J. Anal. Appl.
Pyrolysis. 73 (2005) 323–334.
[9] D. Meier, B. van de Beld, A. V. Bridgwater, D.C. Elliott, A. Oasmaa, F. Preto, State-of-the-art of fast pyrolysis
in IEA bioenergy member countries, Renew. Sustain. Energy Rev. 20 (2013) 619–641.
[10] A. V Bridgwater, Fast pyrolysis of biomass: a handbook, CPL Press, Newbury, 1999.
[11] J. Akhtar, N. Saidina Amin, A review on operating parameters for optimum liquid oil yield in biomass pyrolysis,
Renew. Sustain. Energy Rev. 16 (2012) 5101–5109.
[12] and M.G.-P. Filip Stankovikj, Armando G. McDonald, Gregory L. Helms, Quantification of Bio-oil Functional
Groups and Evidences of the Presence of Pyrolytic Humins, (2016).
[13] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for
biofuels, Energy Environ. Sci. 5 (2012) 7797.
[14] A.. Bridgwater, Fast Pyrolysis of Biomass: A Handbook, 3rd ed., PyNe, 2008.
[15] F. Sulaiman, N. Abdullah, Optimum conditions for maximising pyrolysis liquids of oil palm empty fruit
bunches, Energy. 36 (2011) 2352–2359.
[16] A.S. Kalgo, The Development and Optimisation of a Fast Pyrolysis Process for Bio-oil Production, Thesis
Degree of Doctor of Phylosophy, Aston University, 2011.
[17] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Effects of particle size on the fast
pyrolysis of oil mallee woody biomass., Fuel. 88 (2009) 1810–1817.
[18] A. Demirbas, Effect of initial moisture content on the yields of oily products from pyrolysis of biomass, J. Anal.
Appl. Pyrolysis. 71 (2004) 803–815.
[19] Kim, S. Oh, H. Hwang, Y.-H. Moon, J.W. Choi, Assessment of miscanthus biomass (Miscanthus
sacchariflorus) for conversion and utilization of bio-oil by fluidized bed type fast pyrolysis, Energy. (2014).
[20] M. Asadullah, R. Anisur M., A. Mohsin M., M. Abdul M., S. Borhanus M., A. Robiul M., R. Sahedur M.,
Jute fiber pyrolysis for bio-oil production in fluidized bed reactor., Bioresour. Technol. 99 (2007) 44–50.
[21] P.S. Maa, R.C. Bailie, Influence of Particle Sizes and Environmental Conditions on High Temperature
Pyrolysis of Cellulosic Material—I (Theoretical), Combust. Sci. Technol. 7 (1973) 257–269.
[22] G. Kumar, R. Singh, A.K. Panda, Optimization of process for the production of bio-oil from eucalyptus wood,
J. Fuel Chem. Technol. 38 (2010) 162–167.
[23] S. Thangalazhy-gopakumar, S. Adhikari, R.B. Gupta, S.D. Fernando, Influence of Pyrolysis Operating
Conditions on Bio-Oil Components : A Microscale Study in a Pyroprobe, Energy and Fuels. 25 (2011) 1191–
1199.
[24] C. Zhengng, L. D, H. Yang, R. Yan, H. Chen, D. Lee, Influence of mineral matter on pyrolysis of palm oil
wastes, Combust. Flame. 146 (2006) 605–611.
[25] M. Sohrabi, M.R. Hajaligol, M. Nik-Azar, B. Dabir, Mineral matter effects in rapid pyrolysis of beech wood,
Fuel Process. Technol. 51 (1997) 7–17.
[26] Patwardhan, J.A. Satrio, R.C. Brown, B.H. Shanks, Influence of inorganic salts on the primary pyrolysis
Chapter 5 183
products of cellulose., Bioresour. Technol. 101 (2010) 4646–4655.
[27] G. Gellerstedt, M. Kleen, Influence of inorganic species on the formation of polysaccharide and lignin
degradation products in the analytical pyrolysis of pulps, J. Anal. Appl. Pyrolysis. 35 (1995) 15–41.
[28] P. De Wild, Biomass pyrolysis for chemicals, University Library Groningen, Groningen, 2011.
[29] T. Chen, C. Deng, R. Liu, Effect of selective condensation on the characterization of bio-oil from pine sawdust
fast pyrolysis using a fluidized-bed reactor, Energy and Fuels. 24 (2010) 6616–6623.
[30] C.C. Chang, S.R. Wu, C.C. Lin, H.P. Wan, H.T. Lee, Fast pyrolysis of biomass in pyrolysis gas: Fractionation
of pyrolysis vapors using a spray of bio-oil, in: Energy and Fuels, 2012: pp. 2962–2967.
[31] a J. Sherwood-Pollard, Comparison of bio-oil produced in a fractionated bio-oil collection system, Mechnical
Eng. (2009) 182.
[32] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials, Curr. Opin.
Biotechnol. 24 (2013) 414–420.
[33] Y. Chen, R. Bassilakis, W.W. Smith, R.M. Carangelo, M.A. Wójtowicz, Modeling the evolution of volatile
species during tobacco pyrolysis, J. Anal. Appl. Pyrolysis. 66 (2003) 235–261.
[34] Z. Ma, D. Chen, J. Gu, B. Bao, Q. Zhang, Determination of pyrolysis characteristics and kinetics of palm
kernel shell using TGA–FTIR and model-free integral methods, Energy Convers. Manag. 89 (2015) 251–259.
[35] Z. Ma, Q. Sun, J. Ye, Q. Yao, C. Zhao, Study on the thermal degradation behaviors and kinetics of alkali
lignin for production of phenolic-rich bio-oil using TGA–FTIR and Py–GC/MS, J. Anal. Appl. Pyrolysis. 117
(2016) 116–124.
[36] Kersten, X. Wang, W. Prins, W.P.M. van Swaaij, Biomass Pyrolysis in a Fluidized Bed Reactor. Part 1:
Literature Review and Model Simulations, Ind. Eng. Chem. Res. 44 (2005) 8773–8785.
[37] Paulsen, M.S. Mettler, P.J. Dauenhauer, The role of sample dimension and temperature in cellulose
pyrolysis, in: Energy and Fuels, 2013: pp. 2126–2134.
[38] M. Reinmöller, M. Klinger, E. Thieme, B. Meyer, Analysis and prediction of slag-induced corrosion of
chromium oxide-free refractory materials during fusion of coal and biomass ash under simulated gasification
conditions, Fuel Process. Technol. 149 (2016) 218–230.
[39] M. Windt, a M. Azeez, D. Meier, Novel Gas Chromatography-Mass Spectrometry Methods for
Characterization of Volatile Organic Compounds and Water in Fast Pyrolysis Liquids, Energy & Fuels. 27
(2013) 7413–7423.
[40] M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W.J. Lee, C.Z. Li, Effects of temperature on the formation of
lignin-derived oligomers during the fast pyrolysis of Mallee woody biomass, Energy and Fuels. 22 (2008)
2022–2032.
[41] M.J. Prins, J. Lindén, Z.S. Li, R.J.M. Bastiaans, J.A. Van Oijen, M. Aldén, L.P.H. De Goey, Visualization of
biomass pyrolysis and temperature imaging in a heated-grid reactor, Energy and Fuels. 23 (2009) 993–1006.
[42] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[43] D. Shen, W. Jin, J. Hu, R. Xiao, K. Luo, An overview on fast pyrolysis of the main constituents in
lignocellulosic biomass to valued-added chemicals: Structures, pathways and interactions, Renew. Sustain.
Energy Rev. 51 (2015) 761–774.
[44] A. Bridgwater, Fast pyrolysis processes for biomass, Renew. Sustain. Energy Rev. 4 (2000) 1–73.
[45] J. Zhang, Fast pyrolysis behavior of different celluloses and lignocellulosic biopolymer interaction during fast
pyrolysis, (2012).
[46] K. Werner, L. Pommer, M. Broström, Thermal decomposition of hemicelluloses, J. Anal. Appl. Pyrolysis. 110
(2014) 130–137.
[47] M.S. Mettler, S.H. Mushrif, A.D. Paulsen, A.D. Javadekar, D.G. Vlachos, P.J. Dauenhauer, Revealing
pyrolysis chemistry for biofuels production: Conversion of cellulose to furans and small oxygenates, Energy
Environ. Sci. 5 (2012) 5414–5424.
[48] J. Kusters, D. Zurwerra, Supporting Information, 1 (2007) 2–7.
[49] Weerawut Chaiwat, Isao Hasegawa, Takaaki Tani, Kenshi Sunagawa, Kazuhiro Mae, Analysis of cross-
linking behavior during pyrolysis of cellulose for elucidating reaction pathway, Energy Fuels. 23 (2009) 5765–
5772.
[50] C. Li, Z. Zhang, Z.K. Zhao, Direct conversion of glucose and cellulose to 5-hydroxymethylfurfural in ionic
liquid under microwave irradiation, Tetrahedron Lett. 50 (2009) 5403–5405.
184 Chapter 5
[51] S. Neupane, S. Adhikari, Z. Wang, A.J. Ragauskas, Y. Pu, Effect of torrefaction on biomass structure and
hydrocarbon production from fast pyrolysis, Green Chem. 17 (2015) 2406–2417.
[52] Zhouhong Wanga, Armando G. McDonald, Roel J.M. Westerhofc, Sascha R.A. Kersten, Christian M.
Cuba-Torres, Su Had, Brennan Pechad, Manuel Garcia-Perez, Effect of cellulose crystallinity on the
formation of a liquid intermediate and on product distribution during pyrolysis, J. Anal. Appl. Pyrolysis.
100 (2013) 56–66.
[53] R. Vinu, Linda J. Broadbelt., A mechanistic model of fast pyrolysis of glucose-based carbohydrates to predict
bio-oil composition., Energy Environ. Sci. 5 (2012) 9808–9826.
[54] R.J.. Westerhof, Refining fast pyrolysis of biomass, Thesis Degree of Doctor of Science, University of Twente,
2011.
[55] H.B. Mayes, M.W. Nolte, G.T. Beckham, B.H. Shanks, L.J. Broadbelt, The Alpha–Bet(a) of Glucose Pyrolysis:
Computational and Experimental Investigations of 5-Hydroxymethylfurfural and Levoglucosan Formation
Reveal Implications for Cellulose Pyrolysis, ACS Sustain. Chem. Eng. (2014) 140515083343008.
[56] P.R. Patwardhan, Understanding the product distribution from biomass fast pyrolysis, Iowa State University,
2010.
[57] Patwardhan, D.L. Dalluge, B.H. Shanks, R.C. Brown, Distinguishing primary and secondary reactions of
cellulose pyrolysis, Bioresour. Technol. 102 (2011) 5265–5269.
[58] X.-L. Hou, Z. Yang, K.-S. Yeung, H.N.C. Wong, Five-Membered Ring Systems: Furans and Benzofurans, 1st
ed., Elsevier Ltd., 2009.
[59] Hogendoorn, E. Hoekstra, S.R.A. Kersten, W.P.M. van Swaaij, Fast pyrolysis in a novel wire-mesh reactor:
Design and initial results, Chem. Eng. J. 191 (2012) 45–58.
[60] S. Zhou, N.B. Osman, H. Li, A.G. McDonald, D. Mourant, C.-Z. Li, M. Garcia-Perez, Effect of sulfuric acid
addition on the yield and composition of lignin derived oligomers obtained by the auger and fast pyrolysis of
Douglas-fir wood, Fuel. 103 (2013) 512–523.
[61] F. Barsotti, G. Ghigo, D. Vione, Computational assessment of the fluorescence emission of phenol oligomers:
A possible insight into the fluorescence properties of humic-like substances (HULIS), J. Photochem.
Photobiol. A Chem. 315 (2016) 87–93.
[62] T. Kotake, H. Kawamoto, S. Saka, Mechanisms for the formation of monomers and oligomers during the
pyrolysis of a softwood lignin, J. Anal. Appl. Pyrolysis. 105 (2014) 309–316.
[63] X. Bai, K.H. Kim, R.C. Brown, E. Dalluge, C. Hutchinson, Y.J. Lee, D. Dalluge, Formation of phenolic
oligomers during fast pyrolysis of lignin, Fuel. 128 (2014) 170–179.
[64] H. Auracher, W. Marquardt, Heat transfer characteristics and mechanisms along entire boiling curves under
steady-state and transient conditions, Int. J. Heat Fluid Flow. 25 (2004) 223–242.
[65] F. Shafizadeh, T.T. Stevenson, Saccharification of douglas-fir wood by a combination of prehydrolysis and
pyrolysis, J. Appl. Polym. Sci. 27 (1982) 4577–4585.
[66] P.. Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P.
and Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4
(2011) 4306.
[67] Oudenhoven, R.J.M. Westerhof, N. Aldenkamp, D.W.F. Brilman, S.R.A. Kersten, Demineralization of wood
using wood-derived acid: Towards a selective pyrolysis process for fuel and chemicals production, in: J. Anal.
Appl. Pyrolysis, 2013: pp. 112–118.
[68] T. Shoji, H. Kawamoto, S. Saka, Boiling point of levoglucosan and devolatilization temperatures in cellulose
pyrolysis measured at different heating area temperatures, J. Anal. Appl. Pyrolysis. 109 (2014) 185–195.
[69] A. Aho, N. Kumar, K. Eranen, B. Holmbom, M. Hupa, T. Salmi, D.Y. Murzin, Pyrolysis of softwood
carbohydrates in a fluidized bed reactor, Int. J. Mol. Sci. 9 (2008) 1665–1675.
[70] F.X. Collard, J. Blin, A review on pyrolysis of biomass constituents: Mechanisms and composition of the
products obtained from the conversion of cellulose, hemicelluloses and lignin, Renew. Sustain. Energy Rev.
38 (2014) 594–608.
[71] IEA Bioenergy, Levoglucosan: Historical Developments and Current Status of Levoglucosan Production and
Uses, (2015).
[72] P.T. Williams, P.A. Horne, The role of metal salts in the pyrolysis of biomass, Renew. Energy. 4 (1994) 1–13.
[73] F. Shafizadeth, Alternative Pathways for Pyrolysis of Cellulose., ACS Div. Fuel Chem. Prepr. 28 (1983) 285–
290.
Chapter 5 185
[74] S. Julien, E. Chornet, R.P. Overend, Influence of acid pretreatment (H2SO4, HCl, HNO3) on reaction
selectivity in the vacuum pyrolysis of cellulose, J. Anal. Appl. Pyrolysis. 27 (1993) 25–43.
[75] Z. Wang, S. Zhou, B. Pecha, R.J.M. Westerhof, M. Garcia-Perez, Effect of pyrolysis temperature and sulfuric
acid during the fast pyrolysis of cellulose and douglas fir in an atmospheric pressure wire mesh reactor,
Energy and Fuels. 28 (2014) 5167–5177.
[76] W.F. DeGroot, F. Shafizadeh, The influence of exchangeable cations on the carbonization of biomass, J.
Anal. Appl. Pyrolysis. 6 (1984) 217–232.
[77] A. Debdoubi, A. El Amarti, E. Colacio, M.J. Blesa, L.H. Hajjaj, The effect of heating rate on yields and
compositions of oil products from esparto pyrolysis, Int. J. Energy Res. 30 (2006) 1243–1250.
[78] Ozlem Onay, Influence of pyrolysis temperature and heating rate on the production of bio-oil and char from
safflower seed by pyrolysis, using a well-swept fixed-bed reactor, Fuel Process. Technol. 88 (2007) 523–531.
[79] N. Ozbay, A.E. Putun, E. Putun, Bio-oil production from rapid pyrolysis of cottonseed cake: product yields and
compositions, Int. J. Energy Res. 30 (2006) 501–510.
[80] E. Hoekstra, W.P.M. Van Swaaij, S.R.A. Kersten, K.J.A. Hogendoorn, Fast pyrolysis in a novel wire-mesh
reactor: Decomposition of pine wood and model compounds, Chem. Eng. J. 187 (2012) 172–184.
[81] S. Wang, H. Lin, B. Ru, W. Sun, Y. Wang, Z. Luo, Comparison of the pyrolysis behavior of pyrolytic lignin and
milled wood lignin by using TG–FTIR analysis, J. Anal. Appl. Pyrolysis. 108 (2014) 78–85.
[82] R.J. Evans, T.A. Milne, M.N. Soltys, Direct mass-spectrometric studies of the pyrolysis of carbonaceous fuels,
J. Anal. Appl. Pyrolysis. 9 (1986) 207–236.
186
Chapter 6
6. Chapter 6. Distribution of Activation Energy Model to
Describe Sugarcane Bagasse Thermal Depolymerization
Kinetics
(Paper to be submitted to Journal of Analytical and Applied Pyrolysis)

Abstract: This paper reports kinetics studies using a distribution of activation energy model (DEAM) of the main
species released during the thermal degradation of a thin film of sugarcane bagasse in an electrically heated hot plate
reactor. The biomass film studied was prepared with milled bagasse with particle sizes less than 37 µm (mesh 400). It
was mixed with water to form a paste to form the biomass film (10 mg of sugarcane bagasse distributed evenly on a
steel plate, 60 µm thick). All the tests were conducted under vacuum (150 mbar) to minimize secondary reactions. The
kinetics parameters were fitted using all experimental data of each species at heating rates between 5 °C/s – 1200 °C/s
and temperatures between 250 °C – 650 °C by the Pattern Search algorithm, with the latin hypercube sampling using
the Simulink-Toolbox (MatLab R2014). All pre-exponential factors were between 106-107 s-1 and the Activation energies
between 150 and 220 kJ/mol with standard deviations 15 - 21 kJ/mol. The activation energy obtained was very close
to the values reported in the literature for cellulose pyrolysis.

Keywords: Kinetics, DAEM, Fast pyrolysis, hot plate reactor, Sugarcane bagasse
Chapter 6 187

6.1. Introduction

Biomass fast pyrolysis has gained interest in recent years as an alternative for liquid fuels and high added value
production [1]. Pyrolysis is also the first stage other important thermochemical processes such as gasification and
combustion [2,3]. Thus, understanding biomass pyrolysis is critical to control and optimize biomass thermochemical
conversion processes. A key element to study thermal degradation reactions is to be able to measure the evolution of
primary products. Thermogravimetric analyzers (TGA) monitors sample weight changes as a function of pyrolysis
temperature. The nature of volatile products generated by this technique can be studied by coupling it with other
analytical instruments (TGA-FTIR [4–11] or TG-MS [12–16]). TGA is a very popular tool to obtain global kinetic
parameters for these reactions [5,17–25]. Integral and differential methods have been widely used to obtain pseudo-
kinetic parameters from TGA data. However, this technique cannot be used to study reactions at heating rates over
100 °C/min, limiting its applicability for processes of high practical importance such as fats pyrolysis in fluidized bed
reactors. In the market there are micro-reactors coupled to a chromatograph and / or a mass spectrometer, such as
Pyroprobe CDS, and Frontier-Lab micro furnaces (Py-GC-MS), that can be used to study pyrolysis at high heating rates.
In these applications, small amounts of sample (mg) are normally pyrolyzed under controlled temperature (200 °C-1000
°C) and heating rates (1 °C/s - 20000 °C/s) conditions. These techniques are useful for quantify light volatile compounds
released during devolatilization, however they are not able to recover and analyze high molecular weight compounds
like lignin and cellulose oligomers [26–33]. These commercial systems can be modified to collect and analyze the
oligomeric products [34].

To analyze the reactions of primary products, the operating conditions should be set in order to ensure minimum contact
time of volatile with the solid matrix. It has been suggested that the best approach to this goal is operating at vacuum
to improve oligomeric products removal (by thermal ejection) [35]. Hot Plate Reactors [21,36], Wire Mesh Reactors [37–
40], Pulsed Reactors [41], Radiant Reactors [42] and Ablative Reactors [43,44] have gained prominence to study the
products of pyrolysis reactions under high heating rates. Hoekstra et al [39,45] studied overall distribution of total
product at vacuum and high heating rates (> 5000 °C/s) conditions in a Wire Mesh Reactor. More information on these
systems can be found elsewhere [40,46]. Despite these developments, the methods to mathematically describe
devolatilization kinetics at high heating rates are still in its infancy [8,35,36] . Drummond et al [37], Goey, et al [36] and
Alexander, et al [21] have proposed global kinetic studies for biomass fast pyrolysis in Wire Mesh Reactors and Hot
Plate Reactors. In these studies, authors describe overall char and total condensable volatile yield at different final
mesh temperatures (peak temperature) and heating rates with zero holding time; the results of the experiments are
adjusted to first order reaction models. However, there are no studies involving main species release under these
conditions, nor that link these product release with a mathematical model that represents reaction rate.

Biomass pyrolysis is a complex multiphase reacting system for which much uncertainty exist about the main controlling
mechanisms and phenomena [60–64]. There are also difficulties associated to finding the best configuration,
experimental conditions and reactor analytical techniques to study and to differentiate primary and secondary reaction
products [26, 65, 66]. Interactions between cellulose, hemicellulose and lignin during devolatilization are not clear, much
less the effect of mineral material over these interactions [30, 67, 68]. With this uncertainty, the models adopted for
studying the kinetic behavior are global models (lumped models) in one or multiple stages. These models are used to
describe the overall products composition (char, total gas and total tar) and in some cases to describe the behavior of
chemical species [54,55]. In these models there are many theoretical approximations from isoconversionals methods
[25,56–58] to distribution of activation energies methods (DAEM) [6,10,73–78]. More information on these models can
be found elsewhere [64–69].
188
Chapter 6
In this paper, we propose a methodology for studying the kinetics of main compounds released during sugarcane
bagasse devolatilization under high heating rates. The species present in permanent gases, light volatile, lignin
oligomers and cellulose oligomers will be analyzed. The results obtained for each species are numerically adjusted to
a reaction model based on first order activation energies distribution (Distributed Activation Energy Model DAEM).

6.2. Materials and Methods


Kinetics parameters of primary products released during thin film Sugarcane bagasse in a hot plate reactor were fitted
by DAEM Models. The materials and methods are descripted below.

6.2.1. Material

Sugarcane bagasse obtained at Ingenio Risaralda, Colombia (Vía Balboa, La Virginia, Risaralda, Colombia 4°53′48″N
75°53′02″O) was milled in a IKA 3500rmp MF-10 hammer micro-mill, and then sieved using a 250-mesh (63 μm).
Lignocellulosic material analysis was carried out at the University of Antioquia (Colombia), following the methodology
described elsewhere [70].

6.2.2. Sample preparation

Biomass film was prepared using the methodology described elsewhere [71]. Briefly, the film was prepared with
suspension made of 10 wt.% of a finely milled bagasse sugarcane (63 μm) in distilled water. A thin brush was immersed
into the biomass suspension, and subsequently it was spread over the central surface of the steel plate (like painting).
The water from of the suspension was removed by drying the plate in a Mettler Toledo HB 34S Halogen Moisture
Balance. At the end, a biomass film adhered to the steel plate was obtained. On average, 10 mg of bagasse were
impregnated in each experiment.

6.2.3. Hot Plate Reactor configuration

The experimental data used in this paper was obtained with a Hot Plate Reactor capable of controlling temperature and
heating rate of a steel plate on which rests biomass developed by the National University of Colombia. This experimental
set up has been described elsewhere [72]. Briefly, the reactor was operated under vacuum (150 mbar), and the pyrolysis
tests were conducted with a uniform thin films (≈ 60 μm) that ensures small Biot number (< 1) and uniform reaction
conditions. The special design of a flat plate heat exchanger was made to collect quickly the products generated.

6.2.4. Operation sequence and experimental plan

The operational sequence used to obtain our experimental results was described elsewhere [72]. Briefly, initial steel
plate weight was recorded before impregnation with the biomass / water suspension in a Mettler Toledo XS205DU (d
= 0.01mg) analytical balance. Then biomass suspension is impregnated to form a thin film, which is dried in a Mettler
Toledo HB 43-S halogen moisture balance at 80 ° C, until final mass loss is less than 0.01%. Once this is done, steel
plate is pressed with the copper clamps of the two reactor electrodes. The reactor is sealed, and ice is added in the
external shell to cool outer reactor walls. The reactor is purged with analytical grade nitrogen (> 99.996% purity) for 2
minutes at a rate of 5 l/min. Vacuum pump is turned on to evacuate nitrogen and to generate a vacuum pressure of 150
mbar. In the reactor control software reaction conditions (heating rate, final reaction temperature and holding time) are
Chapter 6 189
set (Final hot plate Temperature: 250 - 650 oC, Heating rate: 5 - 1200 oC/s, Pressure: 150 mbar, No holding time at the
final temperature). Due to technical limitations, it is quite difficult to recover all bio-oil produced in the pyrolysis. A portion
of bio-oil is recovered wiping with a quartz gauze the surface of the cold plate. The weight difference of the gauze before
and after wiping is the amount of bio-oil used in analysis. Yield of each released species is obtained by multiplying mass
fractions of measured compounds and the theoretical bio-oil yield calculated by overall mass balance. Each
experimental test was performed in triplicate to ensure reproducibility of the results. Char yield was determined by the
weights of steel plate without sample, with sample after impregnation and the amount of char at the end of each
experiment.

Gas sample recovered with the Tedlar bag was analyzed in a μ-GC Agilent 3000a with two independent channels as
described elsewhere [72]. Briefly, the chemical composition of generated species was normalized in a nitrogen-free
basis. Yield of each quantified species in analysis was obtained by multiplying the total gas yield and the volume fraction
of each quantified species in the μ-GC and corrected by the ratio of molecular weights of the species to analyze and
molecular weight of the gas mixture: Overall yield of the recovered bio-oil was determined by overall mass balance.
The analysis of volatile species was made in a Shimadzu QP 2010 spectrometer. The chromatograph oven was
programmed with two temperature ramps, starting at 50 °C and heating at 10 °C/min up to 250 °C, then at 20 °C/min
up to 320 °C with a holding time of 20 min. The carrier gas was helium at 54 l/min, filament power of 0.1 kV, split ratio
1, injection volume 1 µl. The calibration was made preparing 5 solutions with different dilutions of main compounds
identified in pyrolysis. Bio-oil moisture content was determined quantifying the ion m/z 18 by mass spectrometry, with
the same method used for volatile quantification, following the methodology of Windt et al [73]. Identification of each
species was achieved knowing the retention time of each compound and comparing the spectra with those reported in
NIST database library as described elsewhere [72]. The compounds used for calibration (levoglucosan 5-HMF, furan,
formic acid, acetic acid, Acetoin, glycoaldehyde, furfuryl alcohol, phenol, eugenol, vanillin, naphthalene, Pyrocatechol,
Benzofuran and water) were all analytical grade.

Anhydrosugar were analyzed in a Varian ProStar HPLC with detector RID, column Aminex HPX-87 H following the
method described elsewhere [72]. Briefly, the carrier fluid was distilled water (18μS) at 2 ml/min. The injection volume
was 40µl. The calibration curves were made for cellotriosan, cellobiosan, 1,6 AGF, 1,4-3,6 DGP and levoglucosenone,
preparing 4 dilution levels of each of these components. The total sugars correspond to the sum of the above
compounds. The lignin oligomers were quantified by UV-Fluorescence following the method described elsewhere [72].

6.2.5. Distribution of Activated Energy Model

In this paper we will use the distribution of activation energy model (DAEM) to fit our experimental results. This type of
model considers that the solid matrix or the species under analysis reacts by an infinite number of stages and each of
these stages may represent the transformation of the multiple bonds present in original biomass [74]. First order
irreversible reaction rate expression for a differential element of reaction (reaction i) is given by the expression:

∂[dm(E)] −E
= k o eRT(t) [dm(E)] (1)
∂t

In these type of models must agree the mathematical relationship:

dm(E) = m(E)f(E)dE (2)


190
Chapter 6
Where m represents the total mass of the reacting species. dm is the amount of differential reacted mass. 𝐟(𝐄)
represents the activation energy distribution function of the reaction. Combining equations (1) and (2) and integrating
with respect to E, it is obtained the final expression for reaction rate:


dm(t) dm(T) dT −E
= ( ) = k o m(T) ∫ e RT(t) f(E)dE (3)
dt dT dt
0

There is an activation energy for each reaction stage. In the literature there are many ways to estimate this distribution
function, by mathematical adjustments [19,61,62] or graphical methods [59,75]. In this study it is not possible to use
graphical methods, because there is not enough information (mass loss continuous curve) for accurately estimation of
mass loss rates. For practical purposes, in order to facilitate mathematical adjustment, a normal distribution of activation
energies is assumed [33,76,77].

1 (E−E0 )2

f(E) = e 2σ2 (4)
σ√2π

Where σ and E0 are the standard deviation and average activation energy of the distribution function, respectively.
Now it is defined the mass of volatile species V (T), written as follows:

m(T) = V ∗ − V(T) (5)

When the interest is to study the solid residue kinetics, istead of use m(T), is convenient to use the solid conversion
X(T), written as follows:

mo − m(T)
𝑋(𝑇) = (6)
mo − mf

Being mf carbonaceous residue final mass (when stady state is achieved), mo solid residue initial mass, and
m(T) carbonaceous residue mass in each peak temperature.The equation for reaction rate in terms of each gaseous
species yield is:


dV(t) dV(T) dT −E

= ( ) = k o (V − V(T)) ∫ e RT(t) f(E)dE = K DAEM (V ∗ − V(T)) (7)
dt dT dt
0

Where V ∗ is the maximum or ultimate mass of volatile for each species. The above equation can be reorganized to be
expressed in terms of products yield, dividing both sides of the equation by 𝑚0 (initial mass):

1 dV(t) 1 dY(t)
= K DAEM (V ∗ − V(T)) ↔ = K DAEM (Y ∗ − Y(T)) (8)
𝑚0 dt 𝑚0 dt

Where Y(T) is the yield for each species at temperature T, and Y ∗ is the final or ultimate yield for the specie
i.
Chapter 6 191
There are many theoretical approaches to solve the differential equation (6), many of them based on quadrature
methods [19,56,61–63,77]. These approaches are useful to solve the equation (7) integrated. However, there are other
approaches to reduce the above expression in a set of two coupled ordinary differential equations. Consider the
function:


−E
g(E, t) = ∫ eRT(t) f(E)dE (9)
0

The derivative of this function with respect to E is estimated by calculus first fundamental theorem:

−E
d(g(E, t)) = eRT(t) f(E)dE (10)

Dividing at both sides by dt, to obtain a differential ratio of the activation energy in each time t

d(g(E, t)) −E dE
≈e RT(t) f(E) (11)
𝑑𝑡 𝑑𝑡

There are many theoretical approximations for the derivative of E with respect to t, as reported in [78–84].

dE −E
≈ k o RT(t)eRT(t) (12)
dt

Finally it is resolved simultaneously the system of differential equations described by equations (6) - (11). In this system
of equations, experimental conversion data X(T) are optimized of each studied species to determine the kinetic
parameters 𝑘𝑜 , 𝐸𝑜 , 𝜎. The input data for numerical optimization were temperature vectors of each experimental test
(Peak temperatures) for each heating rate (7 input vectors), and output information corresponds to the yield V (T) of the
species under study in each Peak Temperature (7 output vectors). For simulation and optimization of the differential
equations, a program was developed in Simulink-Matlab R2014 using Solver ODE15s (Numerical Discretization
Formula) with implicit differentiation with respect to time, to solve the system of DE, and using the Pattern-Search
numerical method with Latin Hypercube algorithm for data sampling. Optimization was made minimizing the cost
function SSE (Sum of Square Errors). This analysis was performed for each species released and quantified in
sugarcane bagasse devolatilization.

6.2.6. Lignin oligomers decomposition

The decomposition of lignin oligomers (pyrolytic lignin) was studied to examine the secondary reactions associated with
these compounds that occur during biomass pyrolysis. The oligomers used for this study were extracted from a
sugarcane bagasse fast pyrolysis bio-oil produced in a fluidized bed [85], by precipitation in cold water, following the
procedure described in [86]. Secondary reactions were tested by TGA using a Linseis STA PT 1600 TG-DSC balance.
Each experiment utilized 10 mg of pyrolytic ligin (𝑚𝑜 ); a 50 ml/min flow of nitrogen carrier gas; heating ramps of 10
°C/min, 20 °C/min and 40 °C/min; a final temperature of 800 °C and a hold time of 20min. Each experiment was
replicated in triplicate at each heating rate. A Tedlar bag was coupled to the balance output for gas sampling during
each experiment.
192
Chapter 6
The collected gases were then analyzed using an Agilent 3000A the micro-GC, to determine the evolution of permanent
gases, specifically CO, CO2, CH4, and H2, during devolatilization. The results are based on an average of 10 gas
samples, taken by filling the Tedlar bag at a rate of 40 ml/min for 1 min. The yield of each species generated during
devolatilization is calculated based on equation (12):

(Ci)(mg)
Yield(i) = (13)
mo

Where Ci is the mass fraction (% w/w) obtained by GC, mg is the mass of the gas mass collected during sampling with
Tedlar bag, and mo is the initial mass of pyrolytic lignin. The operating parameters of the micro-GC are as follow: a
helium carrier gas is used at a flow rate of 40 ml/min the oven temperature is maintained at 80 °C, the sample suction
time is 1 min with a 10 s back flush and a column pressure of 30 kPa. Calibration was performed using four certified
mixtures of the permanent gases.

Elemental analysis was performed on the original pyrolytic lignin sample and the carbonized residues obtained after
decomposition in the TGA using an EXETER-Analytical CE-480 elemental analyzer. Each analysis used 1500 μg of
sample, at a combustion temperature of 850 °C and a reduction temperature of 650 °C with a purge time of 20 s.

6.2.7. Levoglucosan decomposition

Thermal decomposition of levoglucosan was studied as a model compound, to describe the decomposition reactions
of sugar and anhydrosugar oligomers released during biomass fast pyrolysis. Approximately 20 mg of levoglucosan
was used for each test, following the same procedures described for the analysis of lignin oligomers. Thermal
decomposition was studied at heating rates of 10 °C/min, 20 °C/min, and 30 °C/min. A 50 ml/min nitrogen carrier gas
flow was used with a final set temperature of 800 °C and hold time of 20 min. Here, the gas composition was analyzed
using the same Agilent micro-GC and methods used for the analysis of permanent gases from the pyrolysis of lignin
oligomers (see 6.2.11). Other researchers [87] have shown that for small amounts of material (< 10 mg), there no
carbonization by secondary reactions occurs, instead the levoglucosan simply evaporates. In order to determine this
rate of evaporation for use in future models, a second set of analyses was conducted using 5 mg of levoglucosan and
heating rates of t 10 °C/min, 30 °C/min and 50 °C/min.

6.2.8. Biomass Drying

Finely milled sugarcane bagasse particles (passing through Tyler 100-mesh) were used for drying studies to provide
additional information for the particle level models. 1 g of milled sugarcane bagasse was placed in a Mettler Toledo HB
34S halogen moisture balance, which was programmed to reach a final temperature of 110 °C at 3 °C/s (180 °C/min).
There are many types of mathematical approaches for biomass drying [88–90]. The recorded mass loss data were
fitted to a first order Arrhenius type model to find the parameters that describes the biomass drying rate. A model of this
nature assumes that evaporation is a thermally activated process that can be described by a first order rate equation.
The advantage of this assumption is that is allows drying to be easily implemented as an additional first order chemical
reaction in particle models and results in smooth intraparticle gradients [91]. The rate of levoglucosan evaporation was
determined by fitting the experimental data obtained at 10 °C/min, 30 °C/min and 50 °C/min to an Arrhenius first order
model, equation (14):
Chapter 6 193
dLevevap dLevevap dT
= ( ) = −K vap (e−E/RT )Levevap (14)
dt dT dt

Where Levevap represents the mass remaining of levoglucosan at each time and K vap evaporation kinetic
coefficient.

6.3. Results
Figure 1 shows comparative parity plot comparing the water content of various methanol water mixtures as measured
by Karl Fisher titration and by GC-MS. The water content measured by Karl Fisher titration was performed using
hydranal as titrant, by a coulometric method using 10 mg of sample.

Figure 1. Humidity comparison measured by GC-MS vs. Karl Fisher. The red dashed line represents the best fit line
for the parity plot.

Other researchers have reported similar methods for estimating the moisture content of bio-oil samples by GC-MS,
obtaining very good correlations with respect to the results achieved by Karl Fisher titration [73,92]. In Appendix B a
sample chromatogram obtained at 1200 °C/s and 500 °C is shown. A list of analyzed and quantified compounds is
given in Table 2. The retention time for the family of light oxygenated compounds is distributed below 10 minutes,
phenols and lignin monomers (Eugenol, Vanillin, Phenol, and Pyrocatechol) are distributed between 13 min - 25 min
and Levoglucosan has a retention time of 12 min.

A general inspection of chromatogram given in appendix B shows a high distribution of light oxygenated compounds in
relationship to other compounds. The kinetic parameters obtained by fitting this data to the main compounds analyzed
in this study are also presented in Table 1.

Table 1. Summary of kinetic parameters found by fitting experimental data in distribution of activation energies model.
Compound m/z Ko (s-1) Eo (kJ/mol) σ (kJ/mol) **R2

Char ---- 9.08E(6) 171.19 16.926 0.95


CO ---- 8.683E(6) 220 23.57 0.89
CO2 ---- 9.05E(6) 219.76 25.48 0.90
H2 ---- 1.00E(7) 179.04 17.96 0.88
CH4 --- 8.137E(6) 175.18 20.81 0.92
194
Chapter 6
H2O 18 1.00E(7) 214.55 20.97 0.93
Acetoin 45 1.07E(7) 181.37 16.59 0.83
Formic Acid 29 1.07E(7) 171.64 16.19 0.88
Acetic Acid 43 1.07E(7) 175.07 14.81 0.91
Furfuryl Alcohol 98 8.16E(6) 171.95 16.08 0.89
Glycoaldehyde 31 1.31E(7) 187.63 15.59 0.76
Furan 68 1.43E(7) 173.94 17.12 0.70
5-HMF 97 1.07E(7) 176.25 15.66 0.90
Benzofuran 118 1.07E(7) 180.37 16.77 0.88
Naphtalene 128 9.08E(6) 172.33 16.48 0.87
Phenol 94 1.07E(7) 182.11 17.18 0.90
Eugenol 164 1.07E(7) 219.00 21.01 0.76
Vanillin 151 1.07E(7) 194.13 17.30 0.81
Pyrocatechol 110 1.07E(7) 219.44 20.94 0.88
Levoglucosan 60 8.82E(7) 167.02 13.46 0.92
Sugars --- 1.07E(7) 148.56 15.57 0.63
Pyrolytic Lignin --- 1.07E(7) 219.00 23.53 0.90
Pyrolytic Lignin(S) --- 1.07E(7) 220 31.9 0.90
Levoglucosan(S) --- 1.42E(7) 103.01 8.99 0.92
Levoglucosan(E) --- 8.90E(10) 113.21 ----- 0.95
Water(E) --- 1.06E(6) 67 ----- 0.99
𝒏 𝒏
**R calculated as:
𝟐 𝟐
∑(𝒀𝒆𝒙𝒑 − ̅̅̅̅̅
𝒀𝒆𝒙𝒑 ) ⁄∑(𝒀𝒇𝒊𝒕 − ̅̅̅̅̅
𝒀𝒇𝒊𝒕 )
𝟏 𝟏

The kinetic coefficients shown in Table 2 are on the order of 106 - 107 (s-1) for all primary reactions. Because there are
few kinetics studies reported at high heating rates, and even fewer studies that monitor species during devolatilization,
it is difficult to identify an appropriate comparison for these results. However, Nunn et al. [93] found coefficients in the
order of 105 (s-1) and activation energies of 82.1 kJ/mol for pyrolysis of milled Wood in a Wire Mesh Reactor at 100
°C/s. Di Blasi, et al. [68] reported coefficients in the order of 106 - 1010 (s-1) and activation energies of between 112 -
153 kJ/mol for biomass pyrolyzed in a Fluidized Bed Reactor. For the specific case of sugarcane bagasse, Drummond,
et al. [37] reported kinetic coefficients of 106 (s-1), activation energy of 92 kJ/mol, for pyrolysis in a Wire Mesh Reactor
at 1000 °C/s, 1 atm and temperatures between 300 and 900 °C. Aiman, et al. [94] reported kinetic parameters on the
order of 104 (s-1) and activation energy of 59.5 kJ/mol for sugarcane bagasse pyrolysis in a Wire Mesh Reactor at
atmospheric pressure with heating rates between 200 and 1000 °C/s, particle sizes between 0.42 and 0.064 mm, and
final temperatures between 300 and 1100 °C. The proposed kinetic models are shown in relation to the related
experimental data in Figures 2-4.

A-Glycoaldehyde fitting B-Acetoin fitting


Chapter 6 195

C-Formic acid fitting D-Acetic acid fitting

E-5-HMF fitting F-Furan fitting

Figure 2. Mathematical fitting of specified light oxygenated compounds released during Sugarcane bagasse pyrolysis
in a hot plate reactor at different temperatures and Heating rates at 150 mbar. Experimental data at 1200 °C/s (o);
Experimental data at 500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o);
Experimental data at 50 °C/s (o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at 1200
°C/s (--); Fitting at 500 °C/s (--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting at 10
°C/s (--); Fitting at 5 °C/s (--).

Levoglucosan has been identified as the primary anhydro-monosaccharide released during cellulose devolatilization
[95]. Based on the observed behavior of levoglucosan, the appearance of light oxygenated compounds can be
explained based on ring-opening reactions [96,97] and char formation by polycondensation reactions in the liquid phase
[98,99]. Heating rate has been found to have a substantial effect on levoglucosan and sugar yields. Under high heating
rates and at pyrolysis temperatures > 550 K (levoglucosan boiling point 533 K), sugars oligomers and levoglucosan
196
Chapter 6
vaporization is favored over cross-linking reactions in liquid phase [100]. In figure 3 is presented the fitting for cellulose,
and lignin oligomers and water.

A-Levoglucosan fitting B-Sugars fitting

C-Water fitting D-Lignin Oligomers fitting

Figure 3. Mathematical fitting of specified oligomers compounds and water released during Sugarcane bagasse
pyrolysis in a hot plate reactor at different temperatures and Heating rates at 150 mbar. Experimental data at 1200
°C/s (o); Experimental data at 500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o);
Experimental data at 50 °C/s (o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at 1200
°C/s (--); Fitting at 500 °C/s (--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting at 10
°C/s (--); Fitting at 5 °C/s (--).

Heating rate not only alters the dynamics of oligomer evaporation, but also intensifies the bubbles bursting and micro-
explosions in the liquid phase, increasing aerosol ejection intensity and the yield of high molecular weight compounds
[101]. Contrary to this, at low heating rates, fragmentation and cross-linking reactions in the liquid phase are dominant
over evaporation and ejection, mainly due to the long lifetime of the intermediate liquid phase. Under these conditions,
the yield of water, a primary product of cross-linking and sugar dehydration reactions, increases dramatically [102,103],
as Figure 3-C shows. The kinetic coefficients for levoglucosan and sugars formation are similar to those reported by
Ranzi, et al. [54]. Unfortunately, there is limited information related to lignin oligomers formation, and this data is far
Chapter 6 197
less quantitative. The most abundant permanent gases produced were CO2 and CO, with only traces of hydrogen and
methane formed. Figure 4 presents experimental data fit to the proposed DAEM model for permanent gas.
198
Chapter 6
A-Hydrogen fitting B-Methane fitting

C-Carbon monoxide fitting D- Carbon dioxide fitting

Figure 4. Mathematical fitting of permanent gases released during Sugarcane bagasse pyrolysis in a hot plate reactor
at differen temperatures and heating rates at150 mbar. Experimental data at 1200 °C/s (o); Experimental data at 500
°C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o); Experimental data at 50 °C/s (o);
Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at 1200 °C/s (--); Fitting at 500 °C/s (--);
Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting at 10 °C/s (--); Fitting at 5 °C/s (--).

During the decomposition of biomass, each pseudo-component contributes to CO2 and CO production. CO2 yield
remains constant at temperatures above 800 K and at high heating rates (1200 °C/s), unlike CO, which goes greater
viability. At low heating rates, CO2 and CO production is favored because of fragmentation of sugar oligomers f and
dehydration reactions occurring in liquid phase. The kinetic coefficients for the formation of these permanent gases are
an order of magnitude lower than for other compounds (106 s-1), however the activation energies and dispersion
coefficients are higher than most compounds.
Chapter 6 199

Figure 5. Solid residue conversion at different Heating rates and temperatures. Experimental data at 1200 °C/s (o);
Experimental data at 500 °C/s (o); Experimental data at 200 °C/s (o); Experimental data at 100 °C/s (o);
Experimental data at 50 °C/s (o); Experimental data at 10 °C/s (o); Experimental data at 5 °C/s (o); Fitting at 1200
°C/s (--); Fitting at 500 °C/s (--); Fitting at 200 °C/s (--); Fitting at 100 °C/s (--); Fitting at 50 °C/s (--); Fitting at 10
°C/s (--); Fitting at 5 °C/s (--).

The solid residue conversion at different heating rates and peak temperatures is presented in Figure 5. A delay of ≈
100 K, in the onset of solid residue formation exists when comparing pyrolysis at 5 °C/s and 1200 °C/s. This delay is
associated with reaction times, because at low heating rates the solid material remains at decomposition temperature
for an extended time (≈ 120 s at 5 °C/s to reach 900 K) compared with shorter reaction times at high heating rates (0.5s
at 1200 °C/s to reach 900 K). As with most compounds, the biomass conversion is constant at temperatures greater
than 800 K. The kinetic coefficients found in this work for the formation of solid residues from sugarcane bagasse,
(k=9.08*106 and Ea=171.19 kJ/mol), are higher than those reported by Drummond, et al. [37] for the same feedstock
in a Wire Mesh Reactor at atmospheric pressure and 1000 °C/s (2.13 * 106 s-1, 92.6 kJ/mol). Other researchers have
reported higher [104] and lower [105,106] coefficients based on experiments performed using a thermobalance at very
low heating rates. Differences in all kinetic coefficients reported in literature are mainly due to pyrolysis conditions such
as particle size, sample quantity, mineral content, initial moisture, carrier gas flow, and furnace volume, among others.
Typically, in tests carried out in thermobalance, secondary reactions of volatiles with the solid residue (char) and the
capsule walls containing thermally ejected residues are favored. In this work, it is proposed that vacuum operation (150
mbar) rapidly removes oligomers and volatile from the particle and thus minimizes secondary reactions. The use of
biomass thin films also allows for uniform heat distribution over the sample and good energy transfer by conduction
with very low Biot numbers (Biot << 0.1 for a 60 μm film), eliminating heat transfer limitation that could create localized
areas of reduced heating rates.

6.3.1. Evaporation and secondary reactions

The intermediate liquid phase formed by cellulose decomposition is composed of high molecular weight anhydrosugars
and levoglucosan [107,108]. Unlike other sugars, levoglucosan has high vapor pressure, so evaporates easily at low
temperatures [18,87]. In the early stages of pyrolysis, when the liquid phase is first formed; the consumption of
levoglucosan can be controlled by two competing mechanisms depending on reaction conditions (pressure, heating
rate, temperature): evaporation and polymerization. The description of levoglucosan evaporation and polymerization
mechanisms are important to quantify secondary char formation and predict pyrolysis products at the particle level
under different reaction conditions. According to Bai et al [99] levoglucosan evaporation is strongly linked to the initial
200
Chapter 6
quantity used for experiments, carrier gas flow and type of capsule. Figure 6-B shows the evaporation of 5 mg
levoglucosan at three different heating rates. In these experiments, a nitrogen flow of 40 ml/min and use of an open
capsule, allowed total evaporation of the levoglucosan leaving no trace of carbonized material in the thermobalance
capsule at the end of the experiment.

A-Water evaporation B-Levoglucosan evaporation

C-Sec. reaction of Pyrolytic lignin D-Sec. reaction of Levoglucosan

Figure 6. Evaporation and secondary reactions of lignin oligomers and levoglucosan. (A).Sugarcane bagasse drying
by halogen balance. (B). Evaporation of 5 mg of levoglucosan at different heating rate in Thermobalance (C).
Thermogravimetric behavior of Pyrolytic lignin (lignin oligomers) at different heating rates. (D). Thermogravimetric
behavior of 20 mg of levoglucosan at different heating rates.

Two additional experiments were carried out to analyze the gases released during the decomposition of levoglucosan
with an initial load of 5 mg. No permanent gases were detected by micro-GC analysis. The kinetics parameters that
describe levoglucosan evaporation were 8.9*1010 s-1 and 113.21 kJ/mol; the same order of magnitude as results
reported by Bai et al. [99]. Figure 6A shows a sugarcane bagasse drying curve, with the related drying kinetics of 1.06
*106 s-1 and 67 kJ/mol and secondary reactions of lignin oligomers and levoglucosan (model compound for cellulose
intermediate liquid phase) are presented in Figures 6C and 6D. For mass balance purposes, and to propose a reaction
for pyrolytic lignin decomposition, the elemental analysis of pyrolytic lignin before and after pyrolysis are given in Table
2.
Chapter 6 201

Table 2. Ultimate Analysis of Pyrolytic lignin and its solid residue after pyrolysis by TGA.
Compound %C %H %N %O*

Pyrolytic lignin 43.58±0.45 6.22±0.18 0.78±0.10 49.41

Char residue* 78.51±3.08 1.29±0.19 1.22±0.33 18.99


*Char residue left after thermogravimetric analysis at 10°C/min.

With the above information and the yield of char, CO2 and H2, a single step reaction is proposed using Vanillin
(𝐶8 𝐻8 𝑂3 ) as a model compound for lignin monomers generated during the decomposition of pyrolytic lignin, see
equation (15).

C3.63 H6.22 O3.09 N0.055 → 0.11(CO2 ) + 1.97(H2 O) + 0.41(C6.54 H1.29 O1.19 N0.09 ) + 0.105(C8 H8 O3 ) + 0.46(H2 ) (15)

The above reaction can be used for both particle and reactor level models to describe decomposition reactions for lignin
oligomers and estimate their contributions to secondary char and permanent gases. There are, to date, no publications
describing the reaction rates of lignin oligomers. The activation energy and kinetic coefficient for this reaction is given
in Table 2.

Similarly, from the study of the decomposition of 20 mg of levoglucosan in a TGA, a reaction is proposed that can be
used as a model for secondary reactions of anhydrosugar oligomers formed by cellulose depolymerization in the liquid
phase. In this study, the yield of permanent gases, such as CO, H2, CH4, is practically zero, only CO2 (4.94% w/w) and
char (5.73% w/w) is pronounced. An important product in the decomposition of cellulose and carbohydrates is water
[55,102,103] which was not measured in the TGA experiments, but has been included in the balance equations by
theoretical estimations. For the 20 mg TGA experiments, the entire levoglucosan sample could not evaporate; part of
the sample reacted in the liquid phase, therefore, in Figure 6D, a fraction of the sample can be seen to form char, CO2
and water. To explain this behavior, the reaction model proposed by Milosavljevic, et al. [109] was adopted as a first
approximation. Elemental analysis of solid residue remaining after levoglucosan pyrolysis is presented in Table 3. As
with the analysis of pyrolytic lignin pyrolysis, the balance of the stoichiometric equation was performed using the CO 2
and char yield (experimentally measured) and the elemental analysis of generated char, see equation (16).

Table 3. Elemental Analysis of levoglucosan and its solid residue after pyrolysis by TGA.
Compound %C %H %N %O*

Char residue* 76.9±1.9 4.05±0.15 --- 19.05


*Char residue left after thermogravimetric analysis at 10°C/min.

C6 H10 O5 → 0.871(C6 H10 O5 )v + 0.18CO2 + 0.458H2 O + 0.11C5.38 H3.40 O(s) (16)

The kinetic parameters for this reaction were 1.42*107 s-1 and 103.01 kJ/mol; lower than those reported by Bai et al.
[87] for levoglucosan polymerization. According to the expression above, a water yield of 5%, relative to the initial
levoglucosan mass is obtained. However, considering that part of the mass is evaporated and only a fraction reacts,
when the water yield is recalculated with respect to only the levoglucosan that reacted, a yield of 39.45% w/w is
obtained. Theoretically, for each mole of levoglucosan, the three –OH groups can each participate in cross-linking
reactions to form up to 3 moles of water. From this, the maximum yield of water formed by cross-linking is 33.33% w/w.
202
Chapter 6
The higher than predicted water yield is a result of additional oxygen and hydroxyl groups present in molecules formed
by ring opening and fragmentation reactions.

6.4. Conclusions
A new methodology and reactor configuration for kinetic studies of the primary species released during the fast pyrolysis
of biomasses were presented. Good fittings of the experimental data was achieved for each of the primary products
formed during bagasse devolatilization with a first order distribution of activation energies model (DAEM). The kinetic
parameters were on thej order of 106-107 s-1, with activation energies between 170-220 kJ/mol and standard deviation
of between 15 - 21 kJ/mol. A single stage reaction was proposed and analyzed for the behavior of pyrolytic lignin and
levoglucosan during biomass pyrolysis based on results obtained from the secondary reactions of model oligomeric
compounds. These reactions produce water, carbon dioxide and char as the primary products.
Chapter 6 203

6.5. References
[1] H. Reith, P. de Wild, E. Heeres, Biomass pyrolysis for chemicals, Biofuels. 2 (2011) 185–208.
[2] M.-K. Bahng, C. Mukarakate, D.J. Robichaud, M.R. Nimlos, Current technologies for analysis of biomass
thermochemical processing: a review., Anal. Chim. Acta. 651 (2009) 117–138.
[3] P. Basu, Biomass gasification and pyrolysis practical design and theory, Academic Press, Burlington, MA, 2010.
[4] L. Tian, B. Shen, H. Xu, F. Li, Y. Wang, S. Singh, Thermal behavior of waste tea pyrolysis by TG-FTIR analysis,
Energy. 103 (2016) 533–542.
[5] A. Meng, H. Zhou, L. Qin, Y. Zhang, Q. Li, Quantitative and kinetic TG-FTIR investigation on three kinds of
biomass pyrolysis, J. Anal. Appl. Pyrolysis. 104 (2013) 28–37.
[6] Peng Fua, Song Hua, Jun Xianga, Peisheng Lib, Dan Huanga, Long Jianga, Anchao Zhanga, Junying Zhanga,
FTIR study of pyrolysis products evolving from typical agricultural residues, J. Anal. Appl. Pyrolysis. 88 (2010)
117–123.
[7] X. Peng, X. Ma, Y. Lin, Z. Guo, S. Hu, X. Ning, Y. Cao, Y. Zhang, Co-pyrolysis between microalgae and textile
dyeing sludge by TG–FTIR: Kinetics and products, Energy Convers. Manag. 100 (2015) 391–402.
[8] J. Giuntoli, W. de Jong, S. Arvelakis, H. Spliethoff, A.H.M. Verkooijen, Quantitative and kinetic TG-FTIR study
of biomass residue pyrolysis: Dry distiller’s grains with solubles (DDGS) and chicken manure, J. Anal. Appl.
Pyrolysis. 85 (2009) 301–312.
[9] H.J. Park, J.-I. Dong, J.-K. Jeon, Y.-K. Park, K.-S. Yoo, S.-S. Kim, J. Kim, S. Kim, Effects of the operating
parameters on the production of bio-oil in the fast pyrolysis of Japanese larch, Chem. Eng. J. 143 (2008) 124–
132.
[10] W. DeJong, G. Dinola, B. Venneker, H. Spliethoff, M. Wojtowicz, TG-FTIR pyrolysis of coal and secondary
biomass fuels: Determination of pyrolysis kinetic parameters for main species and NOx precursors, Fuel. 86
(2007) 2367–2376.
[11] R. Bassilakis, R.M. Carangelo, M.A. Wójtowicz, TG-FTIR analysis of biomass pyrolysis, Fuel. 80 (2001) 1765–
1786.
[12] Y. Qiao, S. Chen, Y. Liu, H. Sun, S. Jia, J. Shi, C.M. Pedersen, Y. Wang, X. Hou, Pyrolysis of chitin biomass:
TG-MS analysis and solid char residue characterization., Carbohydr. Polym. 133 (2015) 163–70.
[13] E. Părpăriţă, M.T. Nistor, M.-C. Popescu, C. Vasile, TG/FT–IR/MS study on thermal decomposition of
polypropylene/biomass composites, Polym. Degrad. Stab. 109 (2014) 13–20.
[14] J. Yang, H. Chen, W. Zhao, J. Zhou, TG–FTIR-MS study of pyrolysis products evolving from peat, J. Anal. Appl.
Pyrolysis. 117 (2016) 296–309.
[15] D. Yu, M. Chen, Y. Wei, S. Niu, F. Xue, An assessment on co-combustion characteristics of Chinese lignite and
eucalyptus bark with TG–MS technique, Powder Technol. 294 (2016) 463–471.
[16] T. Ahamad, S.M. Alshehri, TG-FTIR-MS (Evolved Gas Analysis) of bidi tobacco powder during combustion and
pyrolysis., J. Hazard. Mater. 199–200 (2012) 200–8.
[17] Z. Ma, D. Chen, J. Gu, B. Bao, Q. Zhang, Determination of pyrolysis characteristics and kinetics of palm kernel
shell using TGA–FTIR and model-free integral methods, Energy Convers. Manag. 89 (2015) 251–259.
[18] I. Milosavljevic, E. Suuberg, Cellulose thermal-decomposition kinetics - global mass-loss kinetics, Ind. Eng.
Chem. Res. 34 (1995) 1081–1091.
[19] A.N. Hayhurst, J.F. Davidson, S.A. Scott, J.S. Dennis, An algorithm for determining the kinetics of
devolatilisation of complex solid fuels from thermogravimetric experiments, Chem. Eng. Sci. 61 (2006) 2339–
2348.
[20] Orfão, F.J.A. Antunes, J.L. Figueiredo, Pyrolysis kinetics of lignocellulosic materials—three independent
reactions model, Fuel. 78 (1999) 349–358.
[21] Alexander W. Williams, An Investigation of the kinetics for the fast pyrolysis of loblolly pine woody biomass, .
Georgia Institute of Technology , 2011.
[22] L. Fiori, M. Valbusa, D. Lorenzi, L. Fambri, Modeling of the devolatilization kinetics during pyrolysis of grape
residues, Bioresour. Technol. 103 (2012) 389–397.
[23] C. Branca, C. Di Blasi, Multistep mechanism for the devolatilization of biomass fast pyrolysis oils, Ind. Eng.
Chem. Res. 45 (2006) 5891–5899.
[24] M.J. Antal, G. Varhegyi, E. Jakab, Cellulose Pyrolysis Kinetics : Revisited, Ind. Eng. Chem. Res. 37 (1998)
1267–1275.
204
Chapter 6
[25] J.E. White, W.J. Catallo, B.L. Legendre, Biomass pyrolysis kinetics: A comparative critical review with relevant
agricultural residue case studies, J. Anal. Appl. Pyrolysis. 91 (2011) 1–33.
[26] P.R. Patwardhan, Understanding the product distribution from biomass fast pyrolysis, Iowa State University,
2010. Doctoral Thesis.
[27] S.S. Liaw, V. Haber Perez, S. Zhou, O. Rodriguez-Justo, M. Garcia-Perez, Py-GC/MS studies and principal
component analysis to evaluate the impact of feedstock and temperature on the distribution of products during
fast pyrolysis, J. Anal. Appl. Pyrolysis. 109 (2014) 140–151.
[28] Q. Lu, X.C. Yang, C.Q. Dong, Z.F. Zhang, X.M. Zhang, X.F. Zhu, Influence of pyrolysis temperature and time
on the cellulose fast pyrolysis products: Analytical Py-GC/MS study, J. Anal. Appl. Pyrolysis. 92 (2011) 430–
438.
[29] B. Pecha, P. Arauzo, M. Garcia-Perez, Impact of combined acid washing and acid impregnation on the pyrolysis
of Douglas fir wood, J. Anal. Appl. Pyrolysis. 114 (2015) 127–137.
[30] T.J. Hilbers, Z. Wang, B. Pecha, R.J.M. Westerhof, S.R.A. Kersten, M.R. Pelaez-Samaniego, M. Garcia-Perez,
Cellulose-Lignin interactions during slow and fast pyrolysis, J. Anal. Appl. Pyrolysis. 114 (2015) 197–207.
[31] J.D. Murillo, E.A. Ware, J.J. Biernacki, Characterization of milling effects on the physical and chemical nature
of herbaceous biomass with comparison of fast pyrolysis product distributions using Py-GC/MS, J. Anal. Appl.
Pyrolysis. 108 (2014) 234–247.
[32] X. Gu, X. Ma, L. Li, C. Liu, K. Cheng, Z. Li, Pyrolysis of poplar wood sawdust by TG-FTIR and Py–GC/MS, J.
Anal. Appl. Pyrolysis. 102 (2013) 16–23.
[33] T. Chen, J. Zhang, J. Wu, Kinetic and energy production analysis of pyrolysis of lignocellulosic biomass using
a three-parallel Gaussian reaction model., Bioresour. Technol. 211 (2016) 502–8.
[34] F. Ronsse, D. Dalluge, W. Prins, R.C. Brown, Optimization of platinum filament micropyrolyzer for studying
primary decomposition in cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 95 (2012) 247–256.
[35] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials, Curr. Opin.
Biotechnol. 24 (2013) 414–420.
[36] A.V. Sepman, L.P.H. de Goey, Plate reactor as an analysis tool for rapid pyrolysis of biomass, Biomass and
Bioenergy. 35 (2011) 2903–2909.
[37] Drummond, I.W. Drummond, Pyrolysis of Sugar Cane Bagasse in a Wire-Mesh Reactor, Ind. Eng. Chem. Res.
35 (1996) 1263–1268.
[38] X. Gong, Y. Yu, X. Gao, Y. Qiao, M. Xu, H. Wu, Formation of anhydro-sugars in the primary volatiles and solid
residues from cellulose fast pyrolysis in a wire-mesh reactor, Energy and Fuels. 28 (2014) 5204–5211.
[39] E. Hoekstra, W.P.M. Van Swaaij, S.R.A. Kersten, K.J.A. Hogendoorn, Fast pyrolysis in a novel wire-mesh
reactor: Decomposition of pine wood and model compounds, Chem. Eng. J. 187 (2012) 172–184.
[40] Hogendoorn, E. Hoekstra, S.R.A. Kersten, W.P.M. van Swaaij, Fast pyrolysis in a novel wire-mesh reactor:
Design and initial results, Chem. Eng. J. 191 (2012) 45–58.
[41] C. Krumm, J. Pfaendtner, P.J. Dauenhauer, Millisecond Pulsed Films Unify the Mechanisms of Cellulose
Fragmentation, Chem. Mater. (2016)
[42] J. Lédé, M. Ferrer, O. Boutin, Radiant flash pyrolysis of cellulose—Evidence for the formation of short life time
intermediate liquid species, J. Anal. Appl. Pyrolysis. 47 (1998) 13–31.
[43] K. Voorhees, D.M. Blazsó, D.C. Schwarzinger, P. Carlsson, H. Lycksam, P. Gren, R. Gebart, H. Wiinikka, K.
Iisa, High-speed imaging of biomass particles heated with a laser, J. Anal. Appl. Pyrolysis. 103 (2013) 278–
286.
[44] J. Diebold, J. Scahill, Ablative pyrolysis of biomass in solid-convective heat transfer environments, in: R. P.
Overend, T. A. Milne, L. K. Mudge (Eds.), Fundam. Thermochem. Biomass Convers., Springer Netherlands,
1985: pp. 539–555.
[45] E. Hoekstra, W.P.M. van Swaaij, S.R.A. Kersten, K.J.A. Hogendoorn, Fast pyrolysis in a novel wire-mesh
reactor: Design and initial results, Chem. Eng. J. 191 (2012) 45–58.
[46] Peter R. Solomon, Michael A. Serio, Eric M. Suuberg, Coal pyrolysis: Experiments, kinetic rates and
mechanisms, Prog. Energy Combust. Sci. 18 (1992) 133–220.
[47] P. Ranganathan, S. Gu, Computational fluid dynamics modelling of biomass fast pyrolysis in fluidised bed
reactors, focusing different kinetic schemes, Bioresour. Technol. (2016).
[48] M. Zhang, Z. Geng, Y. Yu, Density functional theory (DFT) study on the dehydration of cellulose, Energy and
Fuels. 25 (2011) 2664–2670.
Chapter 6 205
[49] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[50] A. Dieguez-Alonso, A. Anca-Couce, N. Zobel, F. Behrendt, Understanding the primary and secondary slow
pyrolysis mechanisms of holocellulose, lignin and wood with laser-induced fluorescence, Fuel. 153 (2015) 102–
109.
[51] P. Hasler, T. Nussbaumer, P. Morf, Mechanisms and kinetics of homogeneous secondary reactions of tar from
continuous pyrolysis of wood chips, Fuel. 81 (2002) 843–853.
[52] J.S. Srndovic, Chalmers tekniska högskola, Institutionen för kemi- och bioteknik, Biopolymertechnologi,
Innventia AB, Interactions between wood polymers in wood cell walls and cellulose/hemicellulose
biocomposites, Chalmers University of Technology, Göteborg, 2011.
[53] T. Hosoya, S. Saka, H. Kawamoto, Cellulose–hemicellulose and cellulose–lignin interactions in wood pyrolysis
at gasification temperature, J. Anal. Appl. Pyrolysis. 80 (2007) 118–125.
[54] E. Ranzi, A. Cuoci, T. Faravelli, A. Frassoldati, G. Migliavacca, S. Pierucci, S. Sommariva, Chemical Kinetics
of Biomass Pyrolysis, Energy & Fuels. 22 (2008) 4292–4300.
[55] R. Vinu, Linda J. Broadbelt., A mechanistic model of fast pyrolysis of glucose-based carbohydrates to predict
bio-oil composition., Energy Environ. Sci. 5 (2012) 9808–9826.
[56] M. Hu, Z. Chen, S. Wang, D. Guo, C. Ma, Y. Zhou, J. Chen, M. Laghari, S. Fazal, B. Xiao, B. Zhang, S. Ma,
Thermogravimetric kinetics of lignocellulosic biomass slow pyrolysis using distributed activation energy model,
Fraser–Suzuki deconvolution, and iso-conversional method, Energy Convers. Manag. 118 (2016) 1–11.
[57] A.A. Jain, A. Mehra, V. V. Ranade, Processing of TGA data: Analysis of isoconversional and model fitting
methods, Fuel. 165 (2016) 490–498.
[58] R. Hilten, J.P. Vandenbrink, A.H. Paterson, F.A. Feltus, K.C. Das, Linking isoconversional pyrolysis kinetics to
compositional characteristics for multiple Sorghum bicolor genotypes, Thermochim. Acta. 577 (2014) 46–52.
[59] T. Sonobe, N. Worasuwannarak, Kinetic analyses of biomass pyrolysis using the distributed activation energy
model, Fuel. 87 (2008) 414–421.
[60] C. Ulloa, A.L. Gordon, X. Garc??a, Distribution of activation energy model applied to the rapid pyrolysis of coal
blends, J. Anal. Appl. Pyrolysis. 71 (2004) 465–483.
[61] J. Cai, R. Liu, New distributed activation energy model: numerical solution and application to pyrolysis kinetics
of some types of biomass., Bioresour. Technol. 99 (2008) 2795–2799.
[62] C.P. Please, M.J. McGuinness, D.L.S. McElwain, Approximations to the distributed activation energy model for
the pyrolysis of coal, Combust. Flame. 133 (2003) 107–117.
[63] J. Cai, R. Liu, New distributed activation energy model: Numerical solution and application to pyrolysis kinetics
of some types of biomass, Bioresour. Technol. 99 (2008) 2795–2799.
[64] F.X. Collard, J. Blin, A review on pyrolysis of biomass constituents: Mechanisms and composition of the
products obtained from the conversion of cellulose, hemicelluloses and lignin, Renew. Sustain. Energy Rev. 38
(2014) 594–608.
[65] M.J.J. Antal, Mathematical modelling of biomass pyrolysis phenomena: Introduction, Fuel. 64 (1985) 1483–
1486.
[66] A. Sharma, V. Pareek, D. Zhang, Biomass pyrolysis—A review of modelling, process parameters and catalytic
studies, Renew. Sustain. Energy Rev. 50 (2015) 1081–1096.
[67] C. Di Blasi, Comparison of semi-global mechanisms for primary pyrolysis of lignocellulosic fuels, J. Anal. Appl.
Pyrolysis. 47 (1998) 43–64.
[68] C. Di Blasi, C. Branca, Kinetics of Primary Product Formation from Wood Pyrolysis, Ind. Eng. Chem. Res. 40
(2001) 5547–5556.
[69] A. Anca-Couce, Reaction mechanisms and multi-scale modelling of lignocellulosic biomass pyrolysis, Prog.
Energy Combust. Sci. 53 (2016) 41–79.
[70] A. Sluiter, B. Hames, R. Ruiz, C. Scarlata, J. Sluiter, D. Templeton, D. Crocker, Determination of Structural
Carbohydrates and Lignin in Biomass, National Renewable Energy Laboratory, U. S., 2011.
[71] Alex D. Paulsen, Matthew S. Mettler, Paul J. Dauenhauer, The role of sample dimension and temperature
in cellulose pyrolysis, Energy Fuels. 27 (2013) 2126–2134.
[72] Jorge Montoya, Brennan Pecha, David Roman, Farid Chejne Janna, Effect of Temperature and Heating Rate
on Product Distribution from Pyrolysis of Sugarcane Bagasse in a Hot Plate Reactor, J. Anal. Appl. Pyrolysis.
Submited (2016).
206
Chapter 6
[73] M. Windt, a M. Azeez, D. Meier, Novel Gas Chromatography-Mass Spectrometry Methods for Characterization
of Volatile Organic Compounds and Water in Fast Pyrolysis Liquids, Energy & Fuels. 27 (2013) 7413–7423.
[74] M.L. Boroson, J.B. Howard, J.P. Longwell, W. a. Peters, Product yields and kinetics from the vapor phase
cracking of wood pyrolysis tars, AIChE J. 35 (1989) 120–128.
[75] J. Cai, T. Li, R. Liu, A critical study of the Miura-Maki integral method for the estimation of the kinetic parameters
of the distributed activation energy model, Bioresour. Technol. 102 (2011) 3894–3899.
[76] Q.-V. Bach, K.-Q. Tran, Ø. Skreiberg, Combustion kinetics of wet-torrefied forest residues using the distributed
activation energy model (DAEM), Appl. Energy. (2016).
[77] Q. Xiong, J. Zhang, F. Xu, G. Wiggins, C. Stuart Daw, Coupling DAEM and CFD for simulating biomass fast
pyrolysis in fluidized beds, J. Anal. Appl. Pyrolysis. 117 (2016) 176–181.
[78] D.K. Shen, S. Gu, B. Jin, M.X. Fang, Thermal degradation mechanisms of wood under inert and oxidative
environments using DAEM methods, Bioresour. Technol. 102 (2011) 2047–2052.
[79] N.A. Soto, W.R. Machado, D.L. López, Determinación de los parámetros cinéticos en la pirólisis del pino ciprés,
Quim. Nova. 33 (2010) 1500–1505.
[80] T. Maki, A. Takatsuno, K. Miura, Analysis of Pyrolysis Reactions of Various Coals Including Argonne Premium
Coals Using a New Distributed Activation Energy Model, Energy & Fuels. 11 (1997) 972–977.
[81] H. Teng, C.-T. Hsieh, Activation Energy for Oxygen Chemisorption on Carbon at Low Temperatures, Ind. Eng.
Chem. Res. 38 (1999) 292–297.
[82] Z. Du, A.F. Sarofim, J.P. Longwell, Activation Energy Distribution in Temperature-Programmed Desorption:
modeling and Application of the Soot-Oxygen System, Energy & Fuels. 4 (1990) 296–302.
[83] K. Miura, T. Maki, A simple method for estimating f(E) and k0 (E) in the distributed activation energy model,
Energy & Fuels. 12 (1998) 864–869.
[84] A.A. Rostami, M.R. Hajaligol, S.E. Wrenn, A biomass pyrolysis sub-model for CFD applications, Fuel. 83 (2004)
1519–1525.
[85] J.I. Montoya, C. Valdés, F. Chejne, C.A. Gómez, A. Blanco, G. Marrugo, J. Osorio, E. Castillo, J. Aristóbulo, J.
Acero, Bio-oil production from Colombian bagasse by fast pyrolysis in a fluidized bed: An experimental study,
J. Anal. Appl. Pyrolysis. 112 (2015) 379–387.
[86] M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W.J. Lee, C.Z. Li, Effects of temperature on the formation of
lignin-derived oligomers during the fast pyrolysis of Mallee woody biomass, Energy and Fuels. 22 (2008) 2022–
2032.
[87] X. Bai, R.C. Brown, Modeling the physiochemistry of levoglucosan during cellulose pyrolysis, J. Anal. Appl.
Pyrolysis. 105 (2014) 363–368.
[88] J.-R. Richard, J. Saastamoinen, Simultaneous drying and pyrolysis of solid fuel particles, Combust. Flame. 106
(1996) 288–300.
[89] A.H. Mahmoudi, F. Hoffmann, B. Peters, Semi-resolved modeling of heat-up, drying and pyrolysis of biomass
solid particles as a new feature in XDEM, Appl. Therm. Eng. 93 (2016) 1091–1104.
[90] J. Ratte, F. Marias, J. Vaxelaire, P. Bernada, Mathematical modelling of slow pyrolysis of a particle of treated
wood waste, J. Hazard. Mater. 170 (2009) 1023–1040.
[91] Michel Bellais, Modelling of the pyrolysis of large wood particles, KTH - Royal Institute of Technology
Stockholm, 2007.
[92] K. Smets, P. Adriaensens, J. Vandewijngaarden, M. Stals, T. Cornelissen, S. Schreurs, R. Carleer, J. Yperman,
Water content of pyrolysis oil: Comparison between Karl Fischer titration, GC/MS-corrected azeotropic
distillation and 1H NMR spectroscopy, J. Anal. Appl. Pyrolysis. 90 (2011) 100–105.
[93] T.R. Nunn, Rapid pyrolysis of sweet gum wood and milled wood lignin, MIT, 1981.
[94] J.F. Stubington, S. Aiman, Pyrolysis kinetics of bagasse at high heating rates, Energy & Fuels. 8 (1994) 194–
203.
[95] H. Kawamoto, M. Murayama, S. Saka, Pyrolysis behavior of levoglucosan as an intermediate in cellulose
pyrolysis: Polymerization into polysaccharide as a key reaction to carbonized product formation, J. Wood Sci.
49 (2003) 469–473.
[96] X. Zhang, W. Yang, W. Blasiak, Thermal decomposition mechanism of levoglucosan during cellulose pyrolysis,
J. Anal. Appl. Pyrolysis. 96 (2012) 110–119.
[97] J. Huang, C. Liu, G. Zeng, Y. Xie, H. Tong, W. Li, A density functional theory study on the mechanism of
levoglucosan pyrolysis, J. Fuel Chem. Technol. 40 (2012) 807–815.
Chapter 6 207
[98] X. Bai, K.H. Kim, R.C. Brown, E. Dalluge, C. Hutchinson, Y.J. Lee, D. Dalluge, Formation of phenolic oligomers
during fast pyrolysis of lignin, Fuel. 128 (2014) 170–179.
[99] X. Bai, P. Johnston, R.C. Brown, An experimental study of the competing processes of evaporation and
polymerization of levoglucosan in cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 99 (2013) 130–136.
[100] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[101] Montoya .J, B. Pecha, F. Chejne Janna, Micro-Explosion of Liquid Intermediates During the Fast Pyrolysis of
Sucrose and Organosolv Lignin, Submitt. to J. Anal. Appl. Pyrolysis.
[102] Weerawut Chaiwat, Isao Hasegawa, Takaaki Tani, Kenshi Sunagawa, Kazuhiro Mae, Analysis of cross-
linking behavior during pyrolysis of cellulose for elucidating reaction pathway, Energy Fuels. 23 (2009) 5765–
5772.
[103] K. Chaiwat, Weerawut; Hasegawa, Isao; Kori, Junichi; Mae, Examination of Degree of Cross-Linking for
Cellulose Precursors Pretreated with Acid/Hot Water at Low Temperature, Ind. Eng. Chem. Res. 47 (2008)
5948–5956.
[104] D. Surroop, I.R. Energy, A. Responses, Climate-Smart Technologies, in: 2013: pp. 415–424.
[105] E.R. Zanatta, T.O. Reinehr, J.A. Awadallak, S.J. Kleinübing, J.B.O. dos Santos, R.A. Bariccatti, P.A. Arroyo,
E.A. da Silva, Kinetic studies of thermal decomposition of sugarcane bagasse and cassava bagasse, J. Therm.
Anal. Calorim. 125 (2016) 437–445.
[106] A. Ounas, A. Aboulkas, K. El harfi, A. Bacaoui, A. Yaacoubi, Pyrolysis of olive residue and sugar cane bagasse:
Non-isothermal thermogravimetric kinetic analysis, Bioresour. Technol. 102 (2011) 11234–11238.
[107] Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P. and
Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011)
4306.
[108] J. Lédé, J.P. Diebold, Peacocke, The nature and properties of intermediate and unvaporized biomass pyrolisis
materials, in: Springer Science+ Bussines Media (Ed.), Dev. Thermochem. Biomass Conversion., Dordrecht,
1997: pp. 27–42.
[109] I. Milosavljevic, V. Oja, E.M. Suuberg, Thermal Effects in Cellulose Pyrolysis: Relationship to Char Formation
Processes, Ind. Eng. Chem. Res. 35 (1996) 653–662.
208
Chapter 7
7. Chapter 7. Single Particle Model for Biomass Pyrolysis:
Dynamics of Bubble Formation Inside the Liquid
Intermediate and Their Contribution to Aerosol Formation
by Thermal Ejection
Abstract: A mathematical model was developed to describe the product yields obtained from the fast pyrolysis of
biomass particles. The model proposes coupled mass and energy balances with multiple chemical reactions
responsible for the generation of key species as they evolve through different stages of the pyrolysis process. The
model also includes volumetric particle shrinking as well as generation and destruction of secondary products. For the
first time, a biomass particle model is proposed that includes the evolution of a liquid phase trapped inside the solid
matrix, with bubble dynamics and aerosol ejection due to bubble bursting also presented. This new model was used to
evaluate the effect of particle size, temperature, heating rate, and reaction time on the product distribution. The highest
aerosol yield was 20 wt. % and was obtained for pyrolysis in thermal conditions such as fluized bed reactor at high
heating rate (1000 °C/s), using 0.5 mm particles. Particle size was found to have the greatest impact on aerosol yield,
larger 1 mm particles produced substantially lower yields of aerosols at the same conditions (5.79 wt. %). Aerosol yield
was also found to be sensitive to heating rate. Aerosol yields for pyrolysis conducted with 1 mm particles at a lower
heating rate of 50 °C/s, was only1. 67 wt. %.The parameters that most strongly affected the thermal behavior during
pyrolysis were the biomass thermal conductivity and global heat transfer coefficients. In the intermediate liquid phase,
the bubble density was found to be affected by liquid properties, including viscosity and surface tension.

Keywords: Fast pyrolysis, single particle model, liquid intermediate, bubbles, aerosol ejection.
Chapter 7 209

7.1. Introduction
The interest in modeling biomass thermochemical processes has grown in recent years, with the goal of developing
and supplying alternative energy sources [1, 2]. In Colombia, 72 million tons of sugarcane bagasse are produced every
year. This amount of material represents a potential energy of 332 million GJ yr-1, which is equivalent to 7.4 % of the
total energy consumed in the country [3]. Fast pyrolysis is a promising way to produce high value products from this
material. The main product of this technology (between 60 and 70 wt. %) is a brown, cloudy liquid called bio-oil [4–6].
In fast pyrolysis biomass is rapidly heated to around 500 °C in an inert atmosphere to form a condensable vapor (rich
in oxygenated compounds), a carbonaceous residue (char), and light gases (CO, CO2, H2, CH4) [7,8]. Fast pyrolysis is
a complex process in which chemical reactions, heat transfer and mass transfer happen simultaneously in different
phases (solid, liquid and gases) [3,9–13].

There is substantial experimental and theoretical evidences showing that a short-lived intermediate liquid phase is
formed during the thermal degradation of cellulose and lignin. This intermediate liquid phase is made of high molecular
weight oligomer products derived from cellulose, hemicellulose and lignin [13–17]. The existence of this intermediate
liquid phase, and the associated thermal ejection of aerosols, is critical for explaining the presence of high molecular
weight (oligomer) compounds derived from cellulose and lignin in bio-oil [14,17,18]. This liquid phase has been
observed under a microscope inside biomass cellular matrix [19,20] Many, poorly known, physicochemical processes
(reactions, bubble formation, nucleation, coalescence, bursting) occur inside this liquid phase. Of these phenomena,
thermal ejection has been the most extensively studied via high-speed photography [14,18]. This phenomena is
controlled by formation and bursting of bubbles formed by pyrolysis volatile products [21–23].

Many models have been developed at several scales (reactor, particle, atomic) to predict the global pyrolysis products
yield as a function of temperature, particle size, carrier gas flow and heating rate [2,24–28]. However, none of these
models explicitly consider the presence of an intermediate liquid phase. For particles below 1 mm in average diameter,
most of literature considers that temperature of the whole particle is uniform and equal to the rector temperature. In this
case, the fast pyrolysis process is considered to be kinetically controlled, so typically only simple kinetic models are
used to describe the product distribution. When larger particles are used, important temperature gradients are formed,
requiring the integration of heat and mass transfer equations with reactions kinetics [24,25].

Most models are one dimensional [2,25,27,28], with constant transport properties and global reaction rates that do not
explicitly consider secondary reactions. Such one-dimensional models are valid for relatively long particles with
length/diameter ratio higher than 3 (l/d > 3) [24,29]. Some authors have used two-dimensional [29–31] and three-
dimensional models [32,33] to represent the anisotropy of real biomass. Most models are pseudo-homogeneous,
considering average physicochemical and transport properties for the solid. There are very few models that have
included the presence of an intermediate liquid phase [14,26,34]. Very few models also explicitly consider the effect
over biomass internal structure [32,35,36]. In some cases, the internal microstructure is used to estimate effective
transport properties (effective conductivity, effective diffusivity), which are used in pseudo-homogeneous models.
[30,36,37]. Ciesielinki et al. [35] developed a model that explicitly considers biomass microstructure obtained from SEM
visualization studies and demonstrated that temperature profiles and gas diffusion do not change significantly between
models that include the actual microstructure of biomass and those that consider effective properties in pseudo-
homogeneous models.

The role of thermal ejection as the primary mechanism for collecting oligomer products has been highlighted recently
[15, 19, 41, 42]. Although the intermediate liquid is formed on both the internal and external surfaces of the particle,
thermal ejection products formed on internal surface are typically retained inside the particle, where they are more
210
Chapter 7
susceptible to secondary reactions. Because of this, the biomass microstructure plays an important role controlling
aerosol yields [40–43]. None of the biomass particle models currently available in the literature include thermal ejection.
This is mostly due to the lack of knowledge regarding the reactive bubbling phenomena responsible for thermal ejection
[18]. This type of mechanism is similar to the one reported for thermoplastic materials such as polymers and coals
[21,34].Thus, the main goal of this paper is to develop a mathematical model of a single biomass particle that describes
the formation of aerosols by the bursting of bubbles formed in the intermediate liquid phase.

7.2. Development of the Mathematical Model

7.2.1. Physical Model

Biomass pyrolysis modeling is a multi-scale problem [9,44]. This model is developed at the millimeter scale, and utilizes
an abstraction of the biomass structure to highlight the key phenomena responsible for aerosol formation. This physical
model is based on the structure of randomly cut biomass (see Figure 1).

A-Real object. Biomass structure (views and cuts)

B-Abstract object. Representation of random cut

Figure 1. Scheme for the differential element of a particle of biomass in a slab.(A).Representation of biomass
structure (views and cuts). (B).Abstract Representation of biomass structure in order to represent the mathematical
model.

In this model, it is assumed that a biomass particle is a heterogeneous porous media which can be completely described
by a random mixture of four phases: biomass, char, liquid, and gas (see Figure 1) at every point. The liquid phase is
Chapter 7 211
formed by products of depolymerization reactions [13,18,45,46]. The gases formed by pyrolysis reactions result in
bubbles in the intermediate liquid phase that nucleate, coalesce, and burst. The bubbles grow at each point inside the
liquid phase, while the bursting phenomena happens locally. The final aerosol yield is the sum of all local ejections and
is the maximum yield that a single particle can generate. It is well known that actual aerosols yields are much lower
because aerosols produced in internal cell walls are retained inside the particle. To take this phenomenon into account,
it is recommended that the maximum aerosol yield be multiplied by the external wall/total wall (internal and external)
surface ratio.

Aside from bubbles, gases also are transported by diffusion and convection. In this model, biomass particles are
considered to lie on a metal plate that is heated by a known heat flux (Newman-condition). Heat is transported by
conduction to the biomass particle. This heat transfer mechanism is very relevant for ablative reactors [47–50], auger
reactors, rotary cone reactors [4,51], and fixed bed reactors [52]. The geometry is assumed to be rectangular, which is
a good approximation of the shape of natural fibers after milling. The length/thickness ratio of the fiber is greater than
three, allowing approximation by a one dimensional model [24]. Based on these considerations, the following
assumptions are made for this model: (1) one-dimension analysis, (2) non-steady state, (3) gas phase acts as an ideal
gas (low pressures, high temperatures), (4) particle geometry does not change during pyrolysis, though size is allowed
to reduce, (5) no solid fragmentation, (6) gas velocity through the solid phase can be described by Darcy’s law (low
Reynolds number) (7), liquid phase is considered static (high density and viscosity) (8), aerosol formation by ejection
is considered (9) bubble movement is described through a population balance in the liquid phase to determine the
number of bubbles and distribution, (10) CO, CO2, H2, CH4, C2H6, N2 are considered non-condensable gases, (11) H2O,
5-HMF, levoglucosan, furan, acetone, glycoaldehyde, furfuryl alcohol, acetic acid, formic acid, benzofurane, eugenol,
phenol, naphthalene, pyrocathechol, and vanillin are considered condensable vapors, (12) oligomer products are
produced from aerosol ejection (pyrolytic lignin and sugars oligomers), (13) secondary reactions are considered to be
cracking, gasification, and decomposition of sugar oligomers and pyrolytic lignin. (14) Aerosols movement inside the
particle is not described. In order to take into account the effect of particle size on aerosol retention, a correction factor
used in aerosol filtration in porous media is proposed. This factor relates retention efficiency to particle thickness and
the ratio of inner and outer total biomass area.

The thermal ejection intensity predicted by this model does not consider internal micro-explosions (bubble bursting).
The model assumes that only micro-explosions that happen on external walls lead to the formation of collectable
aerosols.

7.2.2. Particle geometry and differential element

A slab type geometry was assumed for this model because the fibrous nature of biomass tends to result in slab like
particles after milling process (see Figure 2). To verify the validity of this geometry for the proposed model, observations
of milled sugarcane bagasse fibers were made using a microscope, see figure 2. For fiber retained on a 200-mesh
screen, the fiber geometry is similar to a rectangular flat plate, which coincides with the flat form adopted in the model
212
Chapter 7

173 pixels

83 pixels
65 pixels
185 pixels

39 pixels
117 pixels
155 pixels

69 pixels

Figure 2. Dimensions of sugarcane bagasse fibbers.

An abstract representation of the model is shown in figure 3, at each point inside the particle, liquid, solid and gas
phasse coexist.

Figure 3. Differential element for global mass and energy balance.

7.2.3. Pyrolysis reactions

Mathematical expressions for the pyrolysis reaction rate were obtained experimentally using a hot plate reactor,
described in Chapter 6, following a distribution of activation energies model. Kinetic studies were carried out between
5 °C/s and 1200 °C/s to cover the complete heating profile at all points in the biomass sample. Reaction rate
expressions have also be considered for the solid residue, CO, CO 2, H2, CH4, levoglucosan, glycoaldehyde, acetoin,
furufuryl alcohol, formic acid, acetic acid, furan, 5-HMF, vanillin, eugenol, phenol , naphtalene, benzofuran, pyrolytic
lignin, water, and sugars released during devolatilization.

In this model, the reaction rate expression for each chemical species is given by equation 1:
Chapter 7 213
d(mi )
ri = = K DAEM−i (Yield∗i − mi ) (1)
dt

Where 𝐊 𝐃𝐀𝐄𝐌−𝐢 the mathematical expression for kinetic coefficient obtained by DAEM is, presented in Chapter 6, for
each specie i. The mass of species (i) per unit volume of the unreacted feed particle is given by 𝑚𝑖 is. 𝐘𝐢𝐞𝐥𝐝∗𝐢 Is the
final yield of specie i per volume unit. This value is easy obtained by multiplying the final mass yield by the biomass
density (real particle density), equation 2:

Yield∗i = ρbiomass Yi∗ (2)

The mass balance is related to volume (kg/m3) such as 𝐦𝐢 = 𝛆𝐣 𝛒𝐢 , where εj is the volumetric fraction of the phase j
(gas, solid or liquid), and ρi is the mass of species i per initial volume unit. Chapter 6 gives the kinetic parameters
obtained by the DAEM for each species.

7.2.4. Secondary reactions

This model considers homogenous cracking reactions of primary products and solid phase reactions, including;
gasification, methanation, and the Boudouard reaction, as well as liquid phase secondary cross-linking reactions where
oligomers react to produce char. Kinetic parameters for homogeneous reactions in the gas phase and solid/gas
heterogeneous reactions where take from the literature, and are given in Table 1.

Table 1. Secondary reactions and its kinetic parameters.


Reaction ko s-1 Ea kJ/mol Ref
Glycoaldehyde → 2CO + 2H2 4.48·106 106.00 [53]
5 − HMF → 3CO + 1.5C2 H4 4.48·106 106.00 [53]
Phenol → 0.5CO2 + 1.5C2 H4 + 2.5C 4.48·106 106.00 [53]
C + CO2 ↔ 2CO 36.1 77.3 [54]
C + H2 O ↔ CO + H2 15 170 121 [54]
C + 2H2 ↔ CH4 4.18·10-3 19.2 [54]
CO + H2 O → CO2 + H2 1389 125.00 [55]
CH4 + H2 O → CO + 3H2 1.65·1011 329.00 [55]
Acetic acid → CO2 + CH4 1.32·1011 244.44 [56]
Formic acid → CO + H2 O 7.49·1014 287.48 [56]
Formic acid → CO2 + H2 4.46·1013 285.51 [56]

Secondary reactions of lignin and sugar oligomers lead to the formation of additional char, water and CO 2. Yields and
kinetic parameters for these reactions are presented in chapter 6. Secondary reactions for volatile cracking, presented
in table 1, are described by the Arrhenius equation, see equation 3.

−Eai
ṙ sec−i = δij koi exp ( ) ρi ε k (3)
RT

Where 𝛿𝑖𝑗 represents the stoichiometric coefficient, on a mass basis, of species i in reaction j, 𝜌𝑖 represents the density
of spicies i, and 𝜀𝑘 is the volumetric fraction of the phase where the secondary reaction occurs.
214
Chapter 7
Many reports argue that these secondary reactions are generally most important at temperatures higher than 500°C
[8,24,57]. The most important secondary reactions that occur during biomass pyrolysis at moderate temperatures (300-
600°C) are cross-linking and depolymerization of oligomers in the liquid phase [58,59]. Secondary reactions of
oligomers in liquid phase were studied experimentally using a TGA, as are described in Chapter 6. In these experiments
levoglucosan was used as a surrogate to describe the behavior of sugar oligomers formed during cellulose pyrolysis.
Secondary reactions of lignin oligomers were studied using cold water precipitates obtained from the heavy fraction of
bio-oil obtained from sugarcane bagasse pyrolysis conducted in a fluidized bed reactor at 500 oC. Kinetic parameters
and stoichiometry for both sets of secondary reactions are given in Chapter 6, Table 2.

7.2.5. Drying

There are several models describing biomass drying for different scales, particle sizes, moisture contents and pyrolysis
conditions [27,28,60–62]. In this work, an Arrhenius type equation was used, as proposed by Sharma et al. [28] (see
equation 4). Because this is a global model, desorption, diffusion and evaporation effects are all considered.

dρwater −Edrying
= −ρwater k drying exp ( ) (4)
dt RT

Kinetic parameters were obtained experimentally using a halogen moisture balance and are presented in Table 2 in
Chapter 6. ρwater is the moisture mass per particle volume.

7.2.6. Particle shrinking

Particle shrinking during devolatilization reduces the temperature gradient and volatile residence time inside the particle
[63,64]. For this model, shrinking was considered based on the rate of solid consumption, assuming that the cross-
sectional area remained constant. Equation 5 shows how this is incorporated in the model.

ρs + ρl − ρ0
V = Vo + (Vf − V0 ) (5)
ρcf − ρ0

Where: Vf =Sh ∗ V0 and V0 , Vf , Sh, are the initial volume, final volume and shrinkage factor, respectively.
ρs , ρl , ρ0 are the solid, liquid and particle densities, respectively.

7.2.7. Mass balances

The mass balance considers that the particle is not deformed and that there is no movement of the solid phase through
the particle (static body). The mass balances for biomass and char are given in equations 6 and 7, respectively.

∂(εb ρb )
= −K DAEM−b εb ρb (6)
∂t
j
∂(εc ρc )
= K DAEM−b (Yield∗𝑐 − εc ρc ) + ∑ ṙ secj (7)
∂t
1

The left side of these equations represents the change in mass with time for each species in one particle volume.
Reaction rate constants (KDAEM-b) are expressed in units of s-1, εb and εc are biomass and char volumetric fractions,
Chapter 7 215
respectively, while ρb and ρc are biomass and char intrinsic densities. Secondary reactions are expressed using an
Arrhenius equation as described in section 7.2.2. For each species i in the gas phase (except for inert water vapor and
levoglucosan), the mass balance is expressed by equation (8).
j
∂(εg ρi ) ∂(𝑣𝑔 ρi ) ∂(𝑣𝑏 ρi ) ∂ ∂(ρi ) εgs ρi ∂V
= K DAEM−i (Yieldi∗ − ε𝑔 𝜌𝑖 ) + ∑ ṙ secj − εgs − εgl + (Deffi εgs )− (8)
∂t ∂x ∂x ∂x ∂x V ∂t
1

Where 𝜌𝑖 , 𝜀𝑔 , 𝜀𝑔𝑠 , 𝜀𝑔𝑙 , 𝑣𝑔 , 𝑣𝑏 are the density of specie i, volume fraction of gas phase, gas fraction contained within
the solid phase, the gas fraction in contact with the liquid phase, the gas velocity through the pores in the solid phase,
and the average bubble velocity in the liquid phase, respectively. This equation describes the evolution of each gas
species in a differential element in the particle. There are two convection terms: one represents gases transport in the
solid phase (third term in eq. 8), and the other represents transport in the liquid phase (fourth term in eq. 8). Each term
is related to the volume fraction of each phase, 𝜀𝑔𝑠 𝑎𝑛𝑑 𝜀𝑔𝑙 . The sum of these terms is equal to the total volumetric
fraction of gases, that is 𝜀𝑔𝑠 + 𝜀𝑔𝑙 = 𝜀𝑔 . As a first approximation, because of low the solubility of gases at high
temperatures, mass transfer between the liquid and gas phases was not considered. The effective change of particle
volume due to the diffusion of each species inside the particle was also considered. A complete mathematical
description of gas diffusivity, as used in this model, is given in Appendix A.2. For this model, diffusion of each species
through the liquid phase was not considered, because of the very low associated diffusivity coefficients (~10-6 cm2 s-1).
The mass balance for water is similar to the one proposed for each species in gas phase, but has an additional term to
describe biomass drying, where vap means water, see equation (9).

j
∂(εg ρv ) ∂(𝑣𝑔 ρv ) ∂(𝑣𝑏 ρv ) ∂ ∂(ρv ) εg ρv ∂V dρwater
=K DAEM−v (Yieldv∗ − εg ρv )+ ∑ ṙ secj - εgs - εgl + (Deffv εgs ) - + (9)
∂t ∂x ∂x ∂x ∂x V ∂t dt
1

Where 𝜌𝑣 and 𝐷𝑒𝑓𝑓𝑣 are the density and effective diffusivity of vapor, respectively. The mass balance for nitrogen (the
carrier gas in this system) was also considered. For this species there is no reaction term for transfer through the liquid
phase, equation (10).
∂(εg ρN2 ) ∂(𝑣𝑔 ρN2 ) ∂ ∂(ρN2 ) εg sρN2 ∂V
=− εgs + (DeffN2 εgs ) − (10)
∂t ∂x ∂x ∂x V ∂t

Where 𝐷eff𝑁2 and 𝜌𝑁2 represent the effective diffusivity and density of nitrogen, respectively. The nitrogen balance
was included to account for its displacement by new gases formed during heating. Furthermore, inclusion of this species
was necessary to accurately model the internal pressure of particle, which will be higher than the external pressure.

For the liquid phase, the balance is similar to the ones proposed previously for char and biomass and considers a static
liquid phase, equation (11).

j
∂(εl ρl )
= ∑ K DAEM−j (Yield∗j − εl ρl ) − K lvg εl ρl Ylvg + ∑ ṙ secj − İaerosol (11)
∂t
j=Oligomers,levo 1

Where Ylvg is the levoglucosan yield, 𝐊 𝐥𝐯𝐠 is levoglucosan evaporation rate and İaerosol, represents the intensity of
thermal aerosol ejection during bubble collapse per volume unit of particle. This balance is valid for particle nodes. The
third term in equation (9) correspond to levoglucosan evaporation given in Table 1 of chapter 6. The mass balance for
levoglucosan in the gas phase is shown in equation (12).
216
Chapter 7

j
∂(εg ρlvg−v ) ∂(𝑣𝑔 ρlvg−v ) ∂(𝑣𝑏 ρlvg−v ) ∂ ∂(ρlvg−v ) εg ρlvg−v ∂V
= K lvg εl ρl Ylvg − εgs − εgl + (Defflvg εgs ) + ∑ ṙ secj − (12)
∂t ∂x ∂x ∂x ∂x V ∂t
1

Where 𝜌𝑙𝑣𝑔−𝑣 is the density of levoglucosan in gas phase. The aerosols ejection rate is determined by the bubble
population balance using the mathematical expression given in equation 13:

Ra/max
Naerosol 4πρl 3
İaerosol = ∑ ((𝐵̇ 𝑏 )ri,t ( ) R ) (13)
Nbubbles 3 aerosol
i=0

Where 𝐵̇𝑏 is the rate of bubbles leaving the liquid per unit volume of the particle (bubble bursting), Naerosol, Nb and
Naerosol
R aerosol are the number of aersols, number of bubbles, and aerosols radius, respectively. The ratio ( ) is
Nb
obtained by an empirical correlation reported in [65]. During bubble collapse, the ratio between the parent bubble radius
𝑅𝑎𝑒𝑟𝑜𝑠𝑜𝑙
and the radius of the ejected aerosol is on the order of 0.1 ≤ ≤ 0.3. Experimentally, Paulsen et al. [14,18]
𝑅𝑏𝑢𝑏𝑏𝑙𝑒
found f a ratio of 0.1 or cellulose [14,22,65,66].

7.2.8. Population balances in the liquid phase

Bubble dynamics are described for a differential element of the particle, in the liquid phase, with a generation with a
population balance, as shown in equation 14.

∂(F(t, r, x)) ∂(F(t, r, x)𝑣𝑏 ) ∂


+ ̇ − Death
= (ḊF(t, r, x)) + (Birth ̇ ) + J̇nuc δ(rb − r ∗ ) (14)
∂t ∂x ∂r

Where the function F(t, r, x) represents the number of bubbles per unit volume in node x. The first term of the balance
represents the change in the distribution function. The second term represents the convective transport of bubbles in
liquid phase. The third and fourth term represent bubble birth and growth/death rates, respectively. The value of the
delta function is zero when the radius of the bubble 𝒓𝒃 is different than 𝒓∗. Where 𝒓∗ represents the minimum radius
of gas clusters needed to form a bubble, as estimated using the Young-Laplace equation. Ḋ is the bubble growth rate,
described in appendix C-13. There are many techniques to solve this type of balance, however, the moments method
has been adopted here because the balance equation is part of a group of differential equations that can be solved
simultaneously with other balance equations. In appendix C-11 each terms involved in population balance is given.
Equation (15) is obtained after multiplying each term equation14 by 𝑟 𝑛 and the definition of momentum:


∂Mn ∂(r n F(t, rb , x))
=∫ dr (15)
∂t ∂r
0

This expression can be transformed into a system of three differential equations (more detail are in Appendix C), see
equations (16)-(18):

∂(M0 ) 1 ∂
− J̇nuc − φαM02 + φM02 + (𝑣𝑏 M0 ) (16)
∂t 2 ∂x
Chapter 7 217
∂(M1 ) ∂
+ 〈Ḋ〉M0 −
̇ Jnuc rf − φα2 M0 M1 + φM0 M1 + (𝑣𝑏 M1 ) (17)
∂t ∂x
∂(M2 ) ∂
+ 2〈Ḋ〉M1 −
̇ Jnuc rf 2 − φα3 M0 M2 − α3 φM12 + φM0 M2 + (𝑣𝑏 M1 ) (18)
∂t ∂x

Where 𝑀𝑖 (with i=0, 1, 2) represents the momentum zero, one and two, respectively. Experimental studies [67] confirm
that The distribution function, F(t, rb , x), for bubbles inside the melted lignin and cellulose model compound (sucrose),
follow a Log-normal distribution, equation 19.

1 (log(𝑟𝑏 ) − 𝑟̅𝑏 )2
F(t, rb , x) = exp (− ) (19)
√2πσ 2σ2

The mathematical expression for 𝑟̅𝑏 and the standard deviation 𝜎 of the bubbles distribution can be found in Appendix-
11. The total output of bubbles from the last node (bubbles bursting), necessary for estimating the aerosol ejection rate,
is described by equation 20:

𝑟𝑏/𝑚𝑎𝑥 𝑟𝑏𝑚𝑎𝑥
1
Ḃ b = ∑ (Bursting rate)ri,t = ∑ F(t, rb(i) , N) (20)
tbi
𝑖=0 𝑖=0

Where t b represents the bubble burst time after it has reached the surface (see Appendix C). Normally this time is
related to the bubble size, bubble velocity and physicochemical properties of the liquid phase [21,68–70]. Li, et al. [68]
presented empirical correlations for small bubbles (<2 mm) in different fluids. Using these correlations [68] in
combination with previously reported surface tension and viscosity of the liquid phase [18], it was possible to estimate
the bubble bursting time at the gas-liquid interface. Nucleation molecular theory (equation 21), describes the birth rate
of bubbles in terms of the thermodynamic properties of the bulk [71].

∆Ecri
J̇nuc = No e− KbT (21)

Where Jnuc is the nucleation rate, (#bubbles⁄𝑠𝑚3 ), 𝐍𝐨 is a formation frequency factor of new bubbles, ∆𝑬𝒄𝒓𝒊 is the
minimum free energy required to form the first gas cluster, Kb is the Boltzmann constant and T is the temperature of
liquid phase. The frequency factor 𝐍𝐨 is theoretically estimated based on the number of molecules in the liquid phase.
This equation assumes that each molecule is a possible precursor of a bubble during chemical decomposition of the
liquid [21]. The minimum radius of the cluster of molecules to form a bubble is given by rf . This radius is theoretically
determined by correlations of thermophysical properties of the liquid at pyrolysis conditions [65]. The parameter
φ represents the odds of two adjacent bubbles growing sufficiently to coalesce, and 𝛼 relates the size of a new bubble
formed by coalescence. In Appendix C are presented correlations in order to calculate the kernels for bubble
coalescence and death (φ, 𝛼).

7.2.9. Global energy balance

At any point in the porous particle, all three phases are assumed to have the same temperature The contribution of gas
diffusion to energy transport and pressure and viscosity to energy accumulation have been neglected because the
impact of all these contributions are very small. The global energy balance as implemented is given by equation 22:
218
Chapter 7
∂T ∂ ∂T ∂ ∂
∑ ρi ̅̅̅̅
Cpi εi = (K ) − (ρg Cpg εg 𝑣𝑔 T) − (ρg Cpg εgl 𝑣𝑏 T) + ṙ water ∆Hvap + ∑(ṙ j ρj ∆Hrxnj ) + K lvg εl ρl Ylvg ∆Hvaplvg (22)
∂t ∂x eff ∂x ∂x ∂x
i=g,s,l j

The first term of this equation represents internal energy accumulation, where 𝜀𝑖 the volumetric fraction of each phase
inside the particle is ̅̅̅̅
Cpi , is the average heat capacity (gas, liquid, solid phases),ρi is the density phase i (gas, liquid,
solid). The second term represent energy transport by conduction, where K eff is the effective thermal conductivity (see
eq. 23) accounts for the contribution of all three phases present, as well as the contribution of radiation inside the pores.

solid,gas,liquid

K eff = ( ∑ εi K i ) + K rad (23)


i

Where Ki is the thermal conductivity of phase I (gas, solid, liquid) and Krad is the pore radiation contribution (Appendix
C). Third and fourth terms are the heat transport by gas convection within solid matrix, and due to bubbles motion within
liquid phase. Where, vg 𝑣𝑏 and εgl are the gas, bubble velocity respectively and volumetric fraction of gas within liquid
phase. Gas velocity within solid phase are calculated by momentum balance as descripted in sec 7.10. Bubbles velocity
within liquid phase is calculated by Stockes law approximation as is presented In Appendix C-8. εgl is estimated by
gases production within liquid phase due to reactions in liquid phase (levoglucosan and pyrolytic lignin presented in
chapter 6). The fifth term corresponds to the enthalpy for water vaporization. Sixth term represent the enthalpy of the
reaction, including primary and secondary reactions. Finally, the last term, are the enthalpy contribution due to
levoglucosan vaporization. Where K lvg, Ylvg , ∆Hvaplvg are the levoglucosan vaporization rate (chapter 6), fraction of
levoglucosan in liquid phase and enthalpy of vaporization respectively. Physicochemical properties in the solid phase
(thermal conductivity, heat capacity, density, emissivity, pore diameter, permeability) change as biomass degrades. In
this model these properties are estimated by a linear interpolation from the conversion function (see eq. (24)).

Psolid = XPbiomass + (1 − X)Pchar (24)

Where 𝑋 represents the extent of conversion, and P is the value for each independent property for the biomass, char
and average solid, respectively. The authors are aware of the existence of anisotropy in biomass due to the cellular
structure and the fact that pyrolysis reactions modify these structures [72,73], however, for simplicity, these effects were
not taking into account.

7.2.10. Momentum balance

Darcy’s law has been used to calculate the average gas speed through the solid structure, see equation 25.

Ug ∂P
𝑣𝑔 = − (25)
μg ∂x

This equation is valid for laminar flow regimes, when the pore diameter is small (on the order of Å) and Reynolds number
is very low (<<1). Ug and μg are the solid permeability and gas viscosity respectively. Viscosity is calculated according
Chapman-Enskog model, as presented in appendix C-3. The pressure can be calculated using the ideal gas law,
equation 26.
Chapter 7 219
ρg RT
P= (26)
Mw

At each point in the bioass particle, the pressure is linked to the temperature and chemical composition of the gases.
The density of the gases (𝜌𝑔𝑎𝑠𝑒𝑠 ) and the molecular weight of mixture is estimated by a weighted average based on
gas composition, as shown in equations 27-29.

∑ni ρi
ρg = (27)
εg
ρi /Mwi
yi = (28)
∑n1 ρi⁄Mw
i
n

Mw = ∑ yi Mwi (29)
1

Where: 𝜌𝑖 is the mass of each species per unit volume, 𝜀𝑔 is the volume fraction of each gas, and yi is the mole
fraction of each gas species in gas phase. Mwi and Mw are the molecular weigh of species i and gas mixture.

7.2.11.Aerosol retention efficiency

Since ejected aerosols movement inside the particle is not evaluated in this model, a correction factor is proposed to
take into account internal aerosol retention based on particle thickness. It is well known that once aerosols have been
released, they move in a chaotic way, and tend to collide with the inner walls of porous medias [42,43,74]. The
probability of these collisions depends strongly on the thickness of porous medium through which aerosols travel, pore
sizes and aerosols ejection velocities [42]. Gazaryan, Fallah et al., Serrano et al. and Yue et al. [74–77], have presented
a simple Arrhenius type model to describe aerosol retention describer by an impact mechanism with the inner walls of
porous media. For simplicity, an expression for aerosol filtration is adopted for this model as described in equation 30.

ϑr = exp−ω (30)

Where 𝝑𝒓 represents aerosols retention efficiency and 𝝎 is associated with the degree of compaction of the porous
medium; typically dependent on porosity, fiber diameter, tortuosity, fiber orientation, and other properties. Zhou et al.
[39], have suggested that for large particles, lignin oligomers (aerosols), formed during the pyrolysis of Douglas Fir
lignin, are retained on the inner walls of the biomass, blocking pores. The intensity of this effect increases with particle
diameter between 0.3 mm-3 mm. Similar results have also been reported by Westerhof, et al. [78]. Based on this, an
expression for 𝝎 is proposed, see equation 31:

d̂pore d̂pore
ω ≈ a( ) ≈ (ABET ρb) ( ) ho ≈ 1000ho (31)
d̂aerosol d̂aerosol

Where 𝑎 is the ratio between the total internal area and external area of the sugarcane bagasse, and ℎ𝑜 represents
the thickness of biomass particle. 𝐴𝐵𝐸𝑇 is the specific surface area of the sugarcane bagasse (m2/g) as determined by
a BET isotherm. A correction factor (d̂pore ⁄d̂aerosol ) is needed to fully address the relationship between pore
diameter and aerosol diameter because the travel of aerosols through biomass pores is heavily dependent on the
220
Chapter 7
relative diameters, i.e. if a biomass pore are sufficiently large, aerosols can move through the particle without restriction,
regardless of particle thickness.

7.2.12.Initial conditions and boundary conditions

The boundary conditions used are shown in Table 2. Where 𝑇𝑚 is the temperature of the thin steel plate supporting the
biomass particles (see figure 1). This is obtained experimentally by recording the plate surface temperature with a
pyrometer. 𝑇𝑚 is an polynomial function, dependent of time. In appendix C-14 is presented the polynomial functions
for all model conditions.

Table 2. Initial and boundary conditions.


t = 0 ∀ x [0, l] P = P0 vg = 0 𝑣𝑏 = 0 T(t 0 ) = T0 ρi = 0 εb ρb = εbo ρbo εg = εg0 ρ𝑤 = ρ𝑤𝑜

x = 0 ∀ t [0, tf] 𝛛𝐏 𝑣𝑏 = 𝟎 𝐓 = 𝐓𝐦 (t) 𝛛𝑴𝒊


=𝟎 =𝟎
𝛛𝐱 𝛛𝐱
x = l ∀ t [0, tf] 𝑵𝒊 = 𝐊𝐦𝐚𝐬𝐬 𝛒𝐠 (𝐲𝐢 − 𝐲𝐢∞ ) 𝐝𝐓 𝒚𝒊∞ = 𝟎
−𝐊 𝐞𝐟𝐟 = 𝐡𝐞𝐱𝐭 (𝐓 − 𝐓∞ )
𝐝𝐱

ρ𝑤 and ρ𝑤𝑜 are the moisture content by particle volume unit (kg/m3) at the beginning (t=0).In x=0, the mass balance
foe each specie i are represented by equation 32.

∂(ρi εgs 𝑣𝑔 ) ∂ ∂(ρi )


+ (Deffi εgs ) = Ni = 0
∂x ∂x ∂x (32)

For each species, x=l, in the gas phase, the net mass flow of out of the control volume must be equal the mass
transfer of component of x=l in the gas phase that surrounds the particle, as seen in equation 33.

∂(ρi εgs 𝑣𝑔 ) ∂(ρi εgl 𝑣𝑏 ) ∂ ∂(ρi )


− − + (Deffi εgs ) = K mass ρg (yi − yi∞ ) = Ni (33)
∂x ∂x ∂x ∂x

Where 𝐲𝐢∞ represents the accumulation of a gas species in the bulk gas phase surrounding the biomass. K mass is the
mass transfer coefficient on flat surface (See appendix C). Finally, the flow of energy out of the control volume at the
top of the biomass particle is equal to the convective heat transfer through the superficial area (Table 2). Where hext
is the external heat transfer coefficient that considers both convective and radiative heat transfer terms (see Appendix
C).

7.2.13.Simulation code

The pyrolysis model described in the previous section was implemented in the code-language of MatLab2009. The
code was discretized by special coordinates using a finite differences method with discretization up-wind for convection.
Each node also contains calculations for the physicochemical properties as a geometric mean with the neighboring
nodes. The integration over time was done with the built-in function “ODE 15s”, which integration method is time
implicitly and is useful for stiff systems of equations. Only a special discretization was made, creating a system of
nonlinear differential equations which were solved simultaneously. The physicochemical properties used for this model
are given in Table 3.
Chapter 7 221

Table 3. Physicochemical properties for the simulation.


Property Value Reference
Heat capacity biomass 𝐶𝑝𝑏 1387 J/kgK [28]
Heat capacity char 𝐶𝑝𝑐 1430 J/kgK [28]
Heat capacity stainless steel 𝐶𝑝𝑙 500 J/kgK www.engineeringtoolbox.com
Biomass thermal conductivity 𝐾𝑏 0.185 W/mK [28]
Thermal conductivity char 𝐾𝑐 0.41 W/mK [28]
Thermal conductivity gas 𝐾𝑔 W/mK [28]
Thermal conductivity Liquid phase 𝐾𝑙𝑖 0.518 W/mK [79]
Thermal conductivity stainless steel 𝐾𝑙 16 W/mK www.engineeringtoolbox.com
Density biomass 𝜌𝑏 700 kg/m3 [28]
Density char 𝜌𝑐 800 kg/m3 [28]
Density of liquid 𝜌𝑙 1000 [18]
Liquid viscosity 10-3 Pas [79]
Initial temperature 𝑇0 300 K -
Enthalpy of vaporization for water ∆𝐻𝑣𝑎𝑝 2200 kJ/kg www.engineeringtoolbox.com
Enthalpy of vaporization for Levo. ∆𝐻𝑣𝑎𝑝 92.2kJ/mol [56]
Biomass Permeability 10-14 m2 [28]
Char Permeability 5*10-12 m2 [28]

7.2.14.Model validation and verification

Model validation was performed with experimental data obtained in from the hot plate reactor described in Chapter 4.
Because of the limitations of the reactor and biomass thickness (< 2 mm), it was not possible to monitor the temperature
at several points inside the particle. Two sequences of experiments were carried out.
222
Chapter 7
A. Frontal view B. Top view

Figure 4. Biomass film configuration for external temperature measurement.(A).Frontal view of temperature
measurement on biomass surface to model validation.(B). Top view of set-up for temperature measurement.

The first sequence measured the external temperature of the bagasse thin-film directly with a thermocouple using the
reactor configuration shown in figure 4. To validate the temperature profile, surface temperature measurements were
performed using three biomass film thickness, 0.5 mm, 1 mm and 1.5 mm, at heating rates of 50 °C/s and 500 °C/s
and temperatures of 400 °C and 600 °C respectively. The methodology for preparing the sugarcane bagasse film and
execution of experiments is similar to the methodology described in Chapter 4. A data acquisition system was coupled
to the thermocouple to record the temperature values versus time for each experimental run. In this configuration, it
was necessary to leave a space in the central region of the plate (see Figure 4) to allow the pyrometer to record the
temperature of the steel plate surface and execute control actions. Direct measurement of temperature with the
pyrometer over biomass surface was not possible because of emissivity changes of the sample as devolatilized
progressed. Each measurement was performed in triplicate to ensure reproducibility of data.

A second sequence of experiments was carried out to validate they yields of solid residues, volatile condensable and
permanent. As with the first set of tests. Experiments were carried out at a heating rate of 500 °C/s at temperatures of
400 °C and 600 °C, respectively. Each experiment was carried out under atmospheric conditions with a hold time 10
s. Three sugarcane bagasse film thicknesses were again employed: 0.5 mm, 1 mm and 1.5 mm. The methodology for
measure the gas, char and volatile yield, are the same as described in chapter 5 and 6 for kinetics study.

7.3. Simulation results


The mesh independence was checked regarding spatial discretization, to ensure the stability of the solution of
differential equations. Figure 5 shows the temperature profile of the surface of the particle and first biomass node for N
= 10, N = 20; N = 50 nodes for the pyrolysis of 1mm biomass film at a final temperature of 600 °C and heating rate of
500 °C/s.
Chapter 7 223

Figure 5. Mesh independency for simulation verify with last node temperature of biomass particle.

The results show that for 10 nodes, the temperature profile has the same trend as was found using a larger number of
nodes, however the final temperature value is 50 degrees above the results found using 20 and 50 nodes. From this,
the simulations are independent of the number of nodes, when 20 or more nodes are employed. Because, as the
number of nodes increases, computational time increases, all other simulations in this chapter have been made with
20 nodes to minimize computation requirements.

7.3.1. Temperature profile

The surface temperature profiles predicted by the proposed mathematical model were validated experimentally.
Validation was limited to the surface because of experimental limitations that make verification of points inside the
particle very difficult (Thicknesses of biomass films <2 mm). Figure 6 shows the temperature profiles predicted by the
mathematical model versus the experimental values measured at the surface of bagasse film at different film
thicknesses (0.5 -1.5 mm), heating rates (50-500 oC/s) and temperatures (400-600 oC).

A B

D
C
224
Chapter 7
Figure 6. Comparison of experimental data vs. simulation for temperature profiles of the sugarcane bagasse film
surface at different thicknesses, heating rates and final reaction temperature. (A) Temperatures profiles (Simulated and
experimental) on biomass surface for a 0.5 mm particle thickness at different heating rates (50 °C/s and 500 °C/s) and
temperatures (400 °C - 600 °C). (B) Temperatures profiles (Simulated and experimental) on biomass surface for a
1mm particle thickness at different heating rates (50 °C/s and 500 °C/s) and temperatures (400 °C - 600 °C). (C)
Temperatures profiles (Simulated and experimental) on biomass surface for a 1.5mm particle thickness at different
heating rates (50 °C/s and 500 °C/s) and temperatures (400 °C – 600 °C). (D) Temperatures profiles (Simulated and
experimental) on biomass surface at 600 °C and 500 °C/s and particle thickness between 0.5 mm -1.5 mm.

Measured temperature profiles for the bagasse film surface fit reasonably well with profiles calculated by the proposed
model. In each figure an inflection point can be see for curves generated at for high heating rate tests, resulting in the
greatest deviation between experimental and theoretical data. These differences are attributed to thermal lag of the
thermocouple, which was not considered and was found to have insufficient sensitivity to accurately measure these
abrupt temperature changes over short periods of time. This temperature lag is reduced drastically for low heating rates
(50 °C / s) and at low temperatures (400 °C), especially for larger 1.5mm particles, were the thermocouple has sufficient
time to respond to temperature changes.

As shown in Figure 6-D, the difference in heating times for film thicknesses of 0.5 mm and 1.5 mm is small (a difference
of 0.54 s in time required to reach the set temperature), indicating that although particle thickness is tripled, heating
times remain small. This lag indicates that internal resistance to energy transfer is important (Bi> 0.1). The difference
in final temperature for the last node of the 0.5 mm and 1.5 mm particles is 50 °C. This difference is associated with
the change in external heat transfer coefficient (Grashof number proportional L3). The peak observed in all simulated
temperature profiles, specifically in the region of the inflection point of the curve, corresponds to over-heating of the
steel sheet due to thermal inertia and the actions of the PID control system configured for the reactor.

It is difficult to find reports in literature with experimental data for biomass particles pyrolysis heated on a hot surface.
Sepman et al. [80] showed the temperature profile of a 1 mm woody biomass sample heated in a hot plate reactor to a
final temperature of 600 oC using 25.5 W/m2-K. Figure 7 shows the comparison between the experimental results
reported by Sepman, et al. [80] and this model with an initial temperature of 127°C. By changing the boundary condition
for the steel plate according to equation 34 the energy transfer coefficient used in experiment can be modeled.

dT
−K = h(Tf − T) (34)
dx
Chapter 7 225
Figure 7. Verification of proposed model results vs. experimental data reported by Sepman et al [80] for 1mm
biomass particle pyrolysis in a Hot Plate Reactor at 600 °C final temperature.

The experimental data and model results fit quite well, showing only small differences during the first 0.5 s of heating,
possibly due to thermal lag of the thermocouple used in the measurement. The initial heating shows a linear behavior,
with an approximately heating rate of 400 °C/s, and slowing as the set temperature is approached. Based on these
results, the proposed model provides good prediction of the temperature profile of biomass particles under different
experimental conditions, and more specifically predicts well the behavior of a biomass this film in a hot plate reactor
similar to the reactor used in this thesis under similar reaction conditions.

7.3.2. Validation of char and other species yield

In order to perform a complete validation of this model, a second sequence of experiments was performed using three
biomass thicknesses (0.5 mm, 1 mm, 1.5 mm), at 500 °C/s and two final reaction temperatures (600 °C-400 °C), under
atmospheric conditions with a hold time 10s. Predicted and experimental yields of CO, CO2, levoglucosan, furfuryl
alcohol, acetic acid, phenol, eugenol, vanillin, 5-HMF, aerosols (pyrolytic lignin + levoglucosan), water and solid
residues were compared. Figure 8 presents the comparison of these results for hot plate pyrolysis conducted at a
heating rate of 500 °C/s.
A- Acetic acid yield B-Furfuryl alcohol yield

C-5-Hidroxy methyl furfural yield D-Phenol yield

E-Eugenol yield F-Vanillin yield


226
Chapter 7
G-Carbon monoxide yield H-Carbon dioxide yield

Figure 8. Comparison of the experimental and theoretical values of chemical compounds released during sugarcane
bagasse pyrolysis at 500 °C/s using a 10 s of hold time and different particle thickness and temperatures.(A).Acetic
acid yield at different temperatures (400 °C - 500 °C) and particle thickness (0.5 mm-1.5 mm) at 500 °C/s. (B).Furfuryl
alcohol yield at different temperatures (400 °C- 500 °C) and particle thickness (0.5 mm -1.5 mm) at 500 °C/s. (C).5-
HMF yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at 500 °C/s. (D).Phenol
yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at 500 °C/s. (E). Eugenol
yield at different temperatures (400°C - 500 °C) and particle thickness (0.5 mm - 1.5 mm) at 500 °C/s. (F).Vanillin yield
at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at 500 °C/s. (G).Carbon monoxide
yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at 500 °C/s. (H). Carbon
dioxide yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at 500 °C/s.

Figure 8 shows that experimental values for phenol, vanillin, 5-HMF, and CO are slightly higher than estimated based
on the proposed model, especially at high temperatures. This deviation is most likely a result of important secondary
reactions between volatiles and the solid residue, or liquid phase reactions that were not taken into account in the
model. It is well known that during carbohydrate (levoglucosan, sucrose, cellobiosan) pyrolysis 5-HMF and CO are
produced as fragmentation products [46,81,82]. These reactions are favored by the presence of minerals, which
catalyze the fragmentation reactions [46]. Secondary reactions of levoglucosan that result in the formation of 5-HMF
were not considered. This omission is likely to contribute to higher than predicted yield obtained experimentally. These
differences are most noticeable at high temperatures (873 K-600°C), in which the intensity of these secondary reactions
becomes more important.

Similarly, the theoretical yield of vanillin predicted by this model is greater than was found experimentally. While
decomposition of lignin oligomers by secondary reactions results in the release of several monomer species, in the
proposed model, this decomposition is assumed to produce only vanillin, which is used as a reference compound
because of the difficulty in monitoring all the monomers that are formed. This assumption inherently over predicts
vanillin. For The experimental and theoretical yields for furfuryl alcohol, acetic acid, eugenol, and carbon dioxide \ are
quite close to each other, with the minor differences most likely related to experimental error. For these compounds,
the theoretical model has a good predictive ability, indicating that the mathematical models of the secondary reactions
that produce these compounds are appropriate for the description of their behavior. The Products released in the
greatest quantities during sugarcane bagasse pyrolysis are levoglucosan, water, char and CO2. For these compounds,
Figure 9 shows the comparison of the theoretical and experimental yield, the theoretical aerosol yield (population
balance) vs. experimental (lignin oligomers + levoglucosan) is also included.
Chapter 7 227
A-Char yield B-Water yield

C-Levoglucosan yield B-Aerosol yield

Figure 9. Comparison of the experimental and theoretical values for char, levoglucosan, water and aerosols released
during sugarcane bagasse pyrolysis at 500 °C/s with a 10 s hold time, and different particle thicknesses and
temperatures.(A).Char yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at
500 °C/s. (B).Water yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at
500°C/s. (C).Levoglucosan yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm)
at 500 °C/s. (D).Aerosol yield at different temperatures (400 °C – 500 °C) and particle thickness (0.5 mm - 1.5 mm) at
500 °C/s.

Trends in products yield are consistent with conventional understandings of pyrolysis [57], for example the solid residue
yield decreases with increasing temperature and is positively associated with particle thickness. in contrast, higher
levoglucosan yields are obtained at 500 °C and with a 0.5 mm particle thickness. For greater particle thicknesses,
levoglucosan secondary reactions are more intense, reducing the yield. The yield of aerosols obtained experimentally
is defined as the sum of lignin oligomers (lignin pyrolytic) and levoglucosan. It was not possible to measure heavier
sugar oligomers, due to experimental limitations. Good consistency was found for the quantity of aerosols obtained by
the population balance in the particle model and from experimental results. The differences in the yield are attributed
to the fact that this model does not include aerosol dynamics inside the porous medium. It is possible that, because of
diffusion, collision and interception mechanisms, the amount of aerosols retained aerosols in the particle is greater than
the amount predicted by this model.

Although there is little information regarding product yields from sugarcane bagasse pyrolysis in a hot plate reactor, the
model was verified with experimental results presented by Drummond, et al. [83] and Aiman, et al. [84] for 0.4 mm
sugarcane particles pyrolyzed in a wire mesh reactor at 1000 °C/s, with 1 s and 30 s hold times. Figure 10 compares
the solid residue yield predicted by this model to the experimental data obtained by Drummond et al. [83] and Aiman et
al [84].
228
Chapter 7

Figure 10. Verification of the solid residue yield predicted by the model using the experimental results reported by
Drummond et al. [83] and Aiman et al. [84] for 0.4 mm sugarcane bagasse particles pyrolyzed at 1000 °C/s, using 1 s
and 30 s hold times.

At higher holding times (30 s), the final biomass conversion is almost complete, even at low temperatures (377°C). With
a 1 s hold time, conversion is only completed for higher pyrolysis temperatures (550°C). The trends in the yield of solid
residues obtained from the model and experimental results are quite close, highlighting the predictive capabilities of
this mathematical model. The small differences in yields between the model predictions and those obtained
experimentally by Aiman et al. [84] and Drummond et al. [83] are, at least partially, due to variations in the chemical
composition of the sugarcane bagasse used in each experiment.

7.3.3. Intra-particle dynamics

Figure 11 shows the behavior over time of several key properties and species during the pyrolysis of sugarcane bagasse
particles. These parameters were evaluated using three particle sizes, with a heating rate of 1000 °C/s, a temperature
of 500 °C and a hold time of 3 s.
Chapter 7 229
A. Pressure B. Char density

C.CO2 density D. Levoglucosan density

E. Number of bubbles F. Bubble distribution (t=1.5s)

Figure 11. Intraparticle dynamics. (A). Overpressure and pressure gradient profiles at z=0 mm, z=0.5 mm and z=1 mm.
(B). Char density profile at z=0 mm, z=0.5 mm and z=1 mm. with volume shrinkage shown on the second axis. (C) CO2
density profile at z=0 mm, z=0.5 mm and z=1 mm. (D) Levoglucosan density profile at z=0 mm, z=0.5mm and z=1 mm.
(E) Bubble density profile at z=0 mm, z=0.5 mm and z=1 mm and cumulative bubbles density (second axis). (F). Bubble
size distribution at z=0 mm, z=0.5 mm and z=1 mm.

The maximum overpressure within the particle occurs at the lowest node (in contact with the heating plate), and is 1.33
times atmospheric pressure. The pressure gradient inside the particle is negative, indicating that the direction of gas
flow is toward the outside the biomass particle. The maximum overpressure occurs at 0.5 s, when the biomass surface
temperature is 477 °C and corresponds to the highest rate of gas release from this node. The highest pressure obtained
for the upper node of the particle occurs at 0.8 s. The permeability of gases inside the porous matrix is sufficiently high
230
Chapter 7
to allow the continuous outlet of gas and minimizes the overpressure inside the particle. Comparing the pressure profiles
with the char yield, shows that after 2 s all permanent gases (Figure 11C shows CO2) and volatiles (Figure 11D shows
levoglucosan) have by formed and released by the devolatilization reactions However, char formation in the last node
of the particle requires 3 s to stabilize, 1 s longer than for total gases production. This effect is the result of the way char
generation is calculated in this model, with total yield based on both the consumption of the initial biomass particle by
primary reactions as well as secondary reactions, especially those of lignin oligomers and levoglucosan reactions in
liquid phase,. In the last particle nodes there is a thermal lag, therefore these areas are heated more slowly, producing
an intermediate liquid that reacts more slowly, generating char as the main product.

The change in volume of the particle was approximately 40%, similar to the results reported by Shi et al. [30]. The rate
of maximum volume loss was found to match well to the maximum overpressure inside the particle, i.e. at the instant
the release of gas species reaches a maximum. Figure 11E shows both the bubble density versus time and the
cumulative bubble density versus time inside a model sugarcane bagasse particle at different positions. The total
number of bubbles in each node is delayed by 0.5 s, corresponding to the temperature of maximum gas formation and
intermediate liquid phase. The rate of bubble formation decreases considerably after 4s, except within the last node,
where the temperature is still low and is the liquid phase still remains. The total number of bubbles per unit volume
generated (sum of bubbles in all nodes and at each time) is ≈1.5*1010 bubbles/m3, all of which have been released after
6s. Bubble size distributions at different positions within the biomass particle (z=0.0 mm, z=0.5 mm and z=1.0 mm) after
1.5 s are shown in Figure 11F. The bubbles distribution are almost the same, with an average sizes of bubbles lose to
150 µm.The small differences can be explained by temperature gradients within the particle that modify the bubbles
nucleation, growth rate and other dynamics.

7.3.4. Bubbles dynamics

There is no information in the literature regarding the generation of bubbles inside the intermediate liquid phase during
pyrolysis of lignocellulosic materials. To overcome this, the model presented here shows quantitatively the dynamics of
bubble formation over time at each point inside the sugarcane bagasse particle. The distribution function used to
estimate the diameter of bubbles is lognormal, and was established experimentally by histograms of bubbles generated
in the liquid phase from lignin and sucrose (see Chapter 2-3). Figure 12A shows the distribution function of bubbles as
a function of pyrolysis temperature (400 °C – 600 °C) and heating rate (10 °C/s and1000 °C/s) at the midpoint of a 1
mm sugarcane bagasse particle at 50 % conversion.
Chapter 7 231
A. Z = 0.5 mm, t at x = 0.5 % w/w B. 1000 °C/s -500 °C in z = 0.5mm

rb (mm) rb (mm)
C. Cumulative bubbles (N bubbles/m3)

Figure 12. Bubbles dynamics at different heating rates and final temperatures. (A) Bubbles distribution at z=0.5 mm
and heating rates 1000 °C/s – 10 °C/s and 600 °C – 400 °C. (B). Bubbles distribution at z=0.5 mm and different
reaction times (0s, 1.5s, 5s). (C). Cumulative total bubbles density (Bubbles/m3) at different heating rates and final
temperature.

Figure 12A shows that the final bubble distribution depends on the heating rate and final reaction temperature
employed. At high heating rates (1000 °C/s) bubbles tend to be larger and more dispersed (standard deviation of the
distribution), with an average diameter of 200 μm. The sizes dispersion tends to increase with temperature. Because
bubble formation dynamics involves bubble nucleation, growth, birth and death rates, and each parameter has different
sensitivities to temperature (exponential-linear), it is expected that under certain conditions, some phenomena will
predominate over others, changing the overall bubble dynamics.

Bubbles generated at low heating rates were found to have average diameters of 50 µm and were also found to have
lower dispersion than bubbles formed at high heating rates. At higher heating rates, the yield of the intermediate liquid
phase is greater, and the generation of volatiles and permanent gases is almost instantaneous, promoting rapid bubble
generation and growth. In addition, the final reaction temperature is achieved in a very short time under these conditions,
indicating that nucleation, birth, death and growth dynamics are developed at the maximum temperature conditions.
Figure 12B presents the distribution function for bubbles generated at z = 0.5 mm and at different reaction times. At
time 0 the distribution is a horizontal line, indicating that bubbles are all of equal size or that there are no bubbles. As
232
Chapter 7
reaction time passes, the distribution becomes more dispersed because of the effect of temperature on growth rate and
bubbles nucleation. However, although distribution becomes more dispersed, the total number of bubbles tends to
reach a maximum and then decrease due to the consumption of intermediate liquid phase. Figure 12C shows the
cumulative bubble density (total bubbles) formed within the entire 1 mm particle under different reaction conditions. The
total density of bubbles is greater at higher heating rates and at higher final reaction temperature. At low heating rates,
total bubbles density is virtually insensitive to temperature (400 °C – 600 °C). Under these low heating conditions, the
reaction time is sufficiently large for the intermediate liquid phase to be completely formed and consumed.

Figure 13 presents the primary contributions to the bubble formation dynamics during the pyrolysis of 1 mm particles of
sugarcane bagasse using heating rates of 1000 °C/s and a final temperature of 500 °C. For illustrative purposes, only
shown the behavior of central node (z = 0.5 mm) is shown.
Chapter 7 233

Figure 13. Deconvolution of main bubble formation dynamics, during devolatilization of a 1mm sugarcane bagasse
particle at 1000 °C/s and 500 °C.

Figure 13 shows that in the central node (z=0.5 mm) bubble nucleation dynamics and bubble death by coalescence,
are the two most influential phenomena regarding total bubble density. This means that as soon as bubbles are formed,
there is a high probability of two bubbles colliding to form a new larger bubble, thereby reducing the total number of
bubbles in the distribution. Both phenomena occur at approximately the same order of magnitude. Growth rates do not
alter, implicitly, the total number of bubbles (only modify their distribution). However, it is possible that when bubbles
start to grow, the probability that they will collide increases and the death rate becomes more intense. According to the
234
Chapter 7
results obtained regarding bubble formation dynamics, the death rate and bubble nucleation decrease at the end of
pyrolysis due to consumption of the intermediate liquid phase. In this model it was considered that at each point in the
particle, all bubbles formed escape the particle by bursting and producing aerosols (a fraction of these aerosols escape
the vapor phase while part of them are retained). The rate of this aerosol ejection depends on the bubbles bursting rate,
as shown in Eq 20. The fraction of aerosols that are released is calculated by multiplying the ejection intensity, shown
in Eq. 13, with filtration efficiency described in Eq. 30. The integral, with respect to time, of the bubble bursting rate
results in a bubble density at each time t. A geometric relationship links the bubble diameter to the ejected aerosol
diameter and the number of ejected aerosol droplets with number of bubbles produced (see section 2.6).

For this model, two considerations were used for calculating the aerosol ejection intensity. The first was the relationship
𝑁
between the number of bubbles and number of aerosols ( 𝑏𝑢𝑏𝑏𝑙𝑒𝑠⁄𝑁 )and the second was the relationship
𝑎𝑒𝑟𝑜𝑠𝑜𝑙
𝐷
between the bubble diameter and the ejected aerosol diameter( 𝑏𝑢𝑏𝑏𝑙𝑒𝑠⁄𝐷 ). Both considerations were taken
𝑎𝑒𝑟𝑜𝑠𝑜𝑙
from literature reports examining bubbling and aerosol ejection from pure fluids. In chapters 3 and 4 of this work bubble
formation dynamics and aerosols distributions were studied, which serve to support the proposed considerations in the
model. The relationship between the average sizes of bubbles formed inside the liquid phase of organosolv lignin
(Chapter 3) and the size of captured aerosols at 900 mbar and 1200 °C/s (Chapter 4).

Daerosol
|
Dbubble lignin
≈ 0.3628.

Similarly, the relationship between the average bubble size formed during sucrose (model compound for cellulose)
pyrolysis and the average diameter of aerosols collected duing avicel cellulose pyrolysis aerosols at 900 mbar and 1200
°C/s was:
Daerosol
| ≈ 0.0923
Dbubblel cellulose

Unfortunately, this relationship cannot be obtained for biomass because, as mentioned in chapter 2 and 3, it is not
possible to visualize biomass directly during the initial phase of pyrolysis due to the high light absorption of the solid
matrix. However, since the ejected oligomers are compounds derived from lignin and cellulose, the above information
suggest that the size relationship between the bubble and aerosol diameter for sugarcane bagasse is between 0.3628
and 0.0923.
Daerosol
0.0923 ≤ | ≤ 0.3628
Dbubble biomass

An average ratio, based on the amount of lignin and cellulose present in raw biomass (see Table 1 Chapter 2) can be
taken:

Daerosol
| ≈ 0.09795
Dbubble biomass

The above result is almost equal to the value assumed for this model (0.1), and is equal to the value suggested by
Teixeira, et al. [18] for boiling reactive cellulose. Similarly, the relationship between the number of bubbles and number
of aerosols are estimated. In the case of organosolv lignin, this relationship is obtained by the following mathematical
relationship:
Chapter 7 235
Naerosol (Cumulative aerosol yield)ρaerosol 4.1 ∗ 1013 aerosol
| = t ≈ 14
≈ 0.1
Nbubbles lignin ∫ (Bursting rate)dt 3.87 ∗ 10 bubble
o

Naerosol 7.92 ∗ 1015 aerosol


| ≈ 14
≈ 48
Nbubbles cellulose 1.66 ∗ 10 bubble

On average, according to the chemical composition of the cellulose and lignin present in the original biomass,

Naerosol
| ≈ 15 aerosol⁄bubble
Nbubble biomass

In the model proposed in this chapter, this ratio was calculated with using the correlation proposed by Günther, et a.l
[85] and Zhang, et al. [65].

Although the ratio obtained experimentally is higher than the empirical correlation, it is a good first approximation (≈9
aerosol / bubble). There is no information about aerosol ejection viscous fluids at high temperatures (pyrolysis
conditions) or for lignocellulosic material or representative model compounds. In the simulation of a bubble collapse
within the liquid phase formed during cellulose pyrolysis Teixeira et al. [18] reported that up to 7 aerosols / bubble
collapse can be obtained. However this simulation did not taken into consideration aerosols release by a film/bursting
mechanism, which can contribute to the total quantity of aerosols recovered.

7.3.5. Effect of particle size on product yields

The effects of particle size on product yields from fast pyrolysis at 1000 °C/s and 500 °C are presented in Figure 14.
The behavior for particles sizes between 0.5 mm – 2 mm are evaluated.

0.5 mm, 1000 °C/s 500 °C

1 mm,1000 °C/s 500 °C


236
Chapter 7

2 mm, 1000 °C/s 500 °C

Figure 14. Effect of particle size on product distribution during fast pyrolysis of bagasse at 1000 °C/s and 500 °C.
Using particles of 0.5 mm, 1.0 mm and 2.0 mm.

Figure 14 clearly shows that the yield of aerosols, volatiles and levoglucosan (main products) are highest for the 0.5
mm particle and that these yields are severely reduced as particle size increases. In contrast to this, char yield are
higher for larger particles, especially secondary char formed by secondary reactions of lignin oligomers. An interesting
behavior is observed for intermediate liquid phase produced during pyrolysis of a 2 mm (figure 14). The intermediate
liquid phase (red line) has two peaks, suggesting a process in two steps. This behavior can be explained due to
temperature gradients within solid phase (see appendix B). For thick particles (2 mm) the nodes close to the plate reach
the final temperature quickly, so the liquid phase and aerosol ejection are intense. Nodes upwind not have enough
temperature to start the liquid phase formation. As the time goes, nodes upwind reach the final temperature, and
therefore the liquid phase formation and aerosol ejection starts to be important.
Chapter 7 237

Figure 15. Products global yield predicted by the model during pyrolysis with different particle thickness at 1000 °C/s
and 500 °C.

The summary of global product yields as a function of particle size is shown in Figure 15. There is a dramatic change
in product yields when the particle size is increased from 0.5 mm to 1 mm. These results show a consistent trend of
aerosols decreasing with particle size while char, gas and water formation increase. The fraction of retained aerosols
inside the biomass porous matrix also tends to increase with particle size; these retained aerosols react further to
generate secondary char, permanent gases and non-condensable volatile. During the pyrolysis of the 0.5 mm thick
particles, most aerosols formed can escape from the solid structure, minimizing secondary reactions to produce char,
water and gases, however this was not true of the larger particle sizes. The results found in this model are consistent
with the findings of other researchers [3,15,39,78].

7.3.6. Effect of final temperature on product yields

The effect of temperature on the products distribution during the pyrolysis of a 1mm sugarcane bagasse particle
pyrolysis at a heating rate 1000 °C/s and temperatures of 400 °C – 600 °C is presented in Figure 16. This is the
temperature range from which the highest bio-oil yields are reported for industrial level applications, and is also within
the temperature range for which the kinetic studies were performed made (see Chapter 6).
238
Chapter 7

1 mm,1000 °C/s 400 °C

1 mm, 1000°C /s 500 °C

1 mm, 1000 °C/s 600 °C


Chapter 7 239
Figure 16. Effect of temperature on the product distribution obtained from the fast pyrolysis of a 1 mm thick particle of
sugarcane bagasse at heating rates of 1000 °C/s.

The yield of oxygenated volatiles was found to increase with temperature due to an increase in reaction intensity. At
400 °C, the evolution of product species is slow, since the reaction rate is low at this temperature. At this temperature,
with a particle thickness of 1 mm, the biomass conversion is not quite complete X≈ 95%. Interesting behavior is
observed for levoglucosan, where the maximum yield is obtained from pyrolysis at 400 °C. At this temperature, the
decomposition reactions of levoglucosan are minimized, allowing levoglucosan to escape by evaporation or as aerosols.
This result is consistent with other reports on cellulose and biomass decomposition [86–89]. At higher temperatures,
levoglucosan’s secondary reactions become important, promoting char and water formation (primarily through cross-
linking and dehydration reactions). In the kinetic study developed in Chapter 6, levoglucosan yields remained constant
as temperatures were increased above 400 °C, because the particle thickness was very thin (60 μm) and a low reactor
pressure (150 mbar) was used to promote a rapid release of products, thus minimizing any potential decomposition
reactions inside the particle. A summary of the effects of temperature on the global products yield is shown in Figure
17.

Figure 17. Global product yields predicted by the proposed model during pyrolysis at different reaction temperatures,
with a heating rate of 1000 °C/s and 1 mm particle thickness.

Water and permanent gas yields increase with temperature due to the increased intensity of secondary reactions
involving condensable volatiles. Char yield decreased slightly at 600 °C, because char gasification reactions with CO2
and water vapor begin to become important. In addition, at 400 °C, the char yield is higher because a part of biomass
does not fully react, but is still considered a solid residue at the end of the experiment. Aerosol yields at temperatures
between 500 °C and 600 °C are practically the same, only a slight increase at 600 °C is noted. This is due to an
increase in the intensity of reactions that form the intermediate liquid phase and increase bubble nucleation rates.

At low temperatures, reaction pyrolysis reactions do not complete at the specified conditions, therefore not all volatile
have been released. At 600 °C conversion is 100% complete but part of the condensable volatiles have reacted through
secondary reactions to generate permanent gases and char. The completion of these effects results in the highest
yields of condensable volatilies being obtained at 500 °C, consistent with other studies [24,44,57,87,89,90].
240
Chapter 7

7.3.7. Effects of heating rate on product yields

One of the most important parameters to be taken into account during biomass pyrolysis studies is the heating rate.
Figure 18 shows the effect of the heating rate on the product distribution obtained from the pyrolysis of 1 mm particles
of sugarcane bagasse at 500 °C.

1 mm 50 °C/s 500 °C

1 mm 500 °C/s 500 °C

1 mm 1200 °C/s 500 °C


Chapter 7 241

Figure 18. The effect of heating rate on the product distribution obtained from the fast pyrolysis of 1 mm sugarcane
bagasse at 500 °C.

The effect of heating rate on the final distribution of products has been little studied. In previous figures, it is has been
shown that at low heating rates, the yield of most light oxygenated volatile species are higher, with the exception of
levoglucosan. At low heating rates, volatiles and heavy oligomers stay inside the particle longer, taking part in additional
secondary reactions, and producing a variety of lower molecular weight compounds (monomers, permanent gases,
char and water). The opposite trend is observed for Levoglucosan, a primary reaction product, which shows yield
increases with heating rate; due to the rapid ejection of primary liquid phase products. At a heating rate of 50 °C/s the
particle reaches the set temperature of Tf = 500 °C in the nodes nearest to the hot plate within 10 s, while the nodes
furthest from the plate are found to have lower temperatures at this time, indicating that the pyrolysis process was not
complete. . Differences in the yield of levoglucosan at heating rates of 500 °C/s, 1000 °C/s and 1200 °C/s are small
(2% w/w). Because biomass has a low thermal conductivity, it imposes significant resistance to energy transfer,
regardless of the external energy transfer rate. This high internal heat transfer resistance means that the Biot number
is sensitive to both the particle size and the external energy transfer coefficient (Bi = h*d / k).

Figure 19. Global products yields as predicted by the proposed model from the pyrolysis of a 1 mm sugarcane
bagasse particle at 500 °C using different heating rates.
242
Chapter 7

The condensable volatile yield is lower at low heating rates. There is a 5 % difference compared to the yield obtained
at higher heating rates, which are largely stable at heating rates above 500 °C/s). Aerosol yields at 50 °C/s are very
low; most of liquid phase formed reacts to generate secondary char, water, and volatiles (modeled here as vanillin). A
slight increase in gas yield occurs at high heating rates because all nodes inside the particle reach the equilibrium
temperature quickly, increasing the intensity of secondary reactions at all points. As outlined in section 7.4, there are
differences between the yields obtained experimentally for volatiles and permanent gases and those obtained by
simulation, indicating that not all secondary reactions necessary to precisely explain the pyrolysis process are
accurately represent in the model. This includes possible errors associated with the known catalytic effect of mineral
material on secondary reactions. It is quite possible that homogenous phase reactions considered in this chapter do
not fully account for this effect, resulting in errors in the calculated intensity of cracking reactions.

7.3.8. Sensitivity analysis

In previous sections, the effect of several parameters on the global product distribution and bubble formation dynamics
were discussed. In this section, a sensitivity analysis was performed to estimate which transport properties are the most
influential on the temperature profile, mass loss and total number of bubbles that evolve during biomass pyrolysis. The
proposed model contains many parameters from which the 9 most important parameters for the mass and energy
transfer equations, as well as and bubble dynamics were selected. These parameters are: biomass thermal conductivity
(Kb), biomass heat capacity (Cp), the external energy transfer coefficient (Hext), the energy transfer coefficient form
the hot Plate to biomass (hi), enthalpy of reaction (DH), liquid phase viscosity (Vis), liquid phase surface tension
(TenSurf), liquid phase vapor pressure (PVapo) and the bubble growth constant (K). Using the subroutine “sens_sys”
in MatLab R2014 [91], the sensitivity matrix was calculated with biomass temperature, bubbles density as the
dependent property, equation 35:

∂X i
Sij = (35)
∂Pj

Where 𝑺𝒊𝒋 represents the sensitivity of the variable Xi (temperature, bubbles density) with respect to the parameter Pj
(thermal conductivity, permeability, etc.). Figure 20 shows the results obtained from the sensitivity analysis for
temperature and bubble density.

Figure 20. Sensivity analysis of main parameters over temperature (model node) and number of bubbles.
Chapter 7 243
The sensitivity parameters are evaluated at each instant in time, however in order to obtain a representative value over
the entire time interval, and to provide easily distinguishable global information, global sensitivities were calculated by
integrating the sensitivity function of each parameter with respect to time, equation 36:

t
∂X i
Sensitivity = ∫ dt (36)
∂Pj
0

For both Figures 20A and B, the sensitivity scale is logarithmic, so the larger the bar in the negative region of the graph,
the lower the sensitivity. Contrary to this, in the area in the positive region, the higher the sensitivity. These graphs were
made using a logarithmic scale for visualization purposes.

Figure 20A shows that biomass thermal conductivity is the most influential parameter on the overall temperature profile
inside the biomass particle. For a particle of constant size and overall energy transfer coefficient (hi) constant, the Biot
number (Bi) decreases as thermal conductivity increases, and therefore the internal resistance to energy transfer
becomes more negligible (positive effect). Following this line of analysis, the overall energy transfer coefficient (hi) also
has a positive influence on the temperature profile inside the particle, however, the effect is not as intense as was found
biomass thermal conductivity. This result is interesting because it has been found that, under certain, conditions (particle
size and temperature), regardless of the degree to which the external energy transfer coefficient increases, the
resistance to internal energy transfer and the global product distribution do not change, because the biomass thermal
conductivity is so low that it limits the dynamics of energy transfer (Bi> 1) [24,49,92]. The external energy transfer
coefficient, enthalpy of reaction and biomass heat capacity each show a negative correlation to the overall temperature
profile. These results are expected since a higher heat capacity demands more energy to increase the temperature of
the particle. Similarly, when the external energy transfer coefficient increases, the energy transfer rate from the particle
surface to the surroundings increases, “cooling” the internal nodes. The enthalpy of reaction (DH) has an almost
negligible effect on temperature profile inside the particle. These results are consistent with those reported by Lede, et
al [93].

Figure 20B shows the sensitivity of bubble number density (aerosols ejection intensity proportional to bubble density)
to several parameters. The results show that this parameter is most sensitive to the viscosity and surface tension of the
liquid phase with both parameters having a positive influence. It is known in research into gas bubbling dynamics that
when the viscosity and surface tension increase, the probability of bubble coalescence decreases, increasing the bubble
number density [21,94]. The overall energy transfer coefficient (hi) was also found to have a positive impact on the
bubble density. This result was shown in section 3.4 by simulations at heating rates of 10 °C/s and 1000 °C/s. At high
heating rates (high hi) the intermediate liquid phase yield is higher, and therefore the bubble nucleation rate
(proportional to the amount of liquid phase formed) also increases. The liquid phase vapor pressure and bubble growth
constant have been found to have a negative effect on the bubble number density. An increase in the bubbles growth
rate promotes bubble collision and coalescence (collisions of bigger bubbles), thereby decreasing the bubble number
density. Likewise, when the vapor pressure increases, bubble density decreases (both the minimum Gibbs free energy
of bubble formation and the frequency of bubble collisions increase). However, this effect is minimal, seen by the
negative log value of the sensitivity factor.

7.4. Conclusions
A new single particle model has been proposed that couples physical phenomena, such as bubble dynamics inside
intermediate liquid phase, with chemical kinetics of the main species released during sugarcane bagasse pyrolysis.
244
Chapter 7
This model is able to predict the aerosol ejection intensity caused by bubbles bursting inside the intermediate liquid
phase. The temperature profiles predicted by the model closely follow experimentally determined temperature profiles.
The model predictions for the yields of aerosols, char, light oxygenated compounds and permanent gases are close to
experimental results obtained by pyrolysis in a hot plate reactor at different reaction conditions. The most sensitive
variable for aerosol yield is the particle size, showing a yield close to 0 with a 2 mm particle.

Temperature has been found to have a positive effect on the overall aerosol yield due to intensification of chemical
reactions leading to the intermediate liquid phase and bubble dynamics (nucleation). Heating rate was also found to
have a positive effect on the bubbles bursting rate and the rate of aerosol ejection, but this effect was minor in
comparison to particle size. Thermal behavior was found to be most sensitive to the biomass thermal conductivity and
global energy transfer coefficient from the hot plate to the biomass. The heat capacity and enthalpy of reaction were
found to have only minor effects on the temperature profile inside the particle.

Liquid properties, specifically the viscosity and surface tension are the most important parameters regarding bubble
dynamics based on a sensitivity analysis. Both parameters are related to bubble nucleation. At high viscosities, and at
high surface tension, bubbles have a reduced probability to collide and coalesce, meaning that the total number of
bubbles remains high. The global heat transfer coefficient (proportional to heating rate) was found to have a positive
effect on the rate at which bubbles bursts. These results are consistent with the results of other researchers. The bubble
growth rate coefficient and vapor pressure were found to have a negative contribution. While increases in these
parameters result in larger bubbles, the collision probability also increases reducing the bubble number density. Further,
at high vapor pressures, the minimum Gibbs energy required to form a bubble increases as does the collision frequency,
with both effects reducing the bubble number density.
Chapter 7 245

7.5. References
[1] T. Damartzis, A. Zabaniotou, Thermochemical conversion of biomass to second generation biofuels through
integrated process design—A review, Renew. Sustain. Energy Rev. 15 (2011) 366–378.
[2] C. Di Blasi, Modelling the fast pyrolysis of cellulosic particles in fluid-bed reactors, Chem. Eng. Sci. 55 (2000)
5999–6013.
[3] J.I. Montoya, C. Valdés, F. Chejne, C.A. Gómez, A. Blanco, G. Marrugo, J. Osorio, E. Castillo, J. Aristóbulo, J.
Acero, Bio-oil production from Colombian bagasse by fast pyrolysis in a fluidized bed: An experimental study,
J. Anal. Appl. Pyrolysis. 112 (2015) 379–387.
[4] R.H. Venderbosch, W. Prins, Fast pyrolysis technology development, Biofuels, Bioprod. Biorefining. 4 (2010)
178–208.
[5] J.I. Montoya Arbeláez, F. Chejne Janna, M. Garcia-Pérez, Fast pyrolysis of biomass: A review of relevant
aspects. Part I: Parametric study, DYNA. 82 (2015) 239–248.
[6] M. Ringer, V. Putsche, J. Scahill, Large-Scale Pyrolysis Oil Production: A Technology Assessment and
Economic Analysis, National Renewable Energy Laboratory (NREL), Golden, CO., US, 2006.
[7] J.. Diebold, T.A. Milne, S. Czernik, A. Osamaa, A.V. Bridgwater, A. Cuevas, S. Gust, D. Hufman, J. Piskorz,
Proposed specification for various grades of pyrolysis oils. In: Bridgwater, A.V., Boocock, 1997.
[8] A.. Bridgwater, Fast Pyrolysis of Biomass: A Handbook, 3rd ed., PyNe, 2008.
[9] P.J. Dauenhauer, D.G. Vlachos, M.S. Mettler, Top ten fundamental challenges of biomass pyrolysis for biofuels,
Energy Environ. Sci. 5 (2012) 7797.
[10] M.S. Mettler, S.H. Mushrif, A.D. Paulsen, A.D. Javadekar, D.G. Vlachos, P.J. Dauenhauer, Revealing pyrolysis
chemistry for biofuels production: Conversion of cellulose to furans and small oxygenates, Energy Environ. Sci.
5 (2012) 5414–5424.
[11] S. Kersten, M. Garcia-Perez, Recent developments in fast pyrolysis of ligno-cellulosic materials, Curr. Opin.
Biotechnol. 24 (2013) 414–420.
[12] H.Z. Li, J. Villermaux, H. Martin, J. Lédé, Fusion-like behaviour of wood pyrolysis, J. Anal. Appl. Pyrolysis. 10
(1987) 291–308.
[13] M. Le Bras, J. Yvon, S. Bourbigot, V. Mamleev, The facts and hypotheses relating to the phenomenological
model of cellulose pyrolysis, J. Anal. Appl. Pyrolysis. 84 (2009) 1–17.
[14] P.J Dauenhauer, Joshua L Colby, Christine M Balonek, Wieslaw J Suszynski, Reactive boiling of cellulose for
integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[15] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Effects of particle size on the fast
pyrolysis of oil mallee woody biomass., Fuel. 88 (2009) 1810–1817.
[16] J. Lédé, J., Diebold, J.P., Peacocke, G.V.C. and Piskorz, The nature and properties of intermediate and
unvaporized biomass pyrolysis materials., Dev. Thermochem. Biomass Convers. (1997) 27–42.
[17] S. Saka, M. Asmadi, H. Kawamoto, Gas- and solid/liquid-phase reactions during pyrolysis of softwood and
hardwood lignins, J. Anal. Appl. Pyrolysis. 92 (2011) 417–425.
[18] Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt, D.P. and
Dauenhauer, Aerosol generation by reactive boiling ejection of molten cellulose, Energy Environ. Sci. 4 (2011)
4306.
[19] G. Zhu, X. Zhu, Z. Xiao, F. Yi, Study of cellulose pyrolysis using an in situ visualization technique and
thermogravimetric analyzer, J. Anal. Appl. Pyrolysis. 94 (2012) 126–130.
[20] F. Yi, G. Zhu, X. Zhu, R. Zhou, Z. Xiao, Pyrolysis characteristics of bean dregs and in situ visualization of
pyrolysis transformation, Waste Manag. (2012).
[21] K. Butler, A numerical model for combustion of bubbling theroplastic materials in microgravity, (2002) 70.
http://fire.nist.gov/bfrlpubs/fire02/PDF/f02002.pdf.
[22] F.J. Resch, P. De Toulon, D. Sciences, J.S. Darrozes, Marine Liquid Aerosol Production From Bursting of Air
Bubbles, 91 (1986) 1019–1029.
[23] J.S. Lee, B.M. Weon, S.J. Park, J.H. Je, K. Fezzaa, W.-K. Lee, Size limits the formation of liquid jets during
bubble bursting., Nat. Commun. 2 (2011) 367.
[24] Kersten, X. Wang, W. Prins, W.P.M. van Swaaij, Biomass Pyrolysis in a Fluidized Bed Reactor. Part 1:
Literature Review and Model Simulations, Ind. Eng. Chem. Res. 44 (2005) 8773–8785.
[25] C. Di Blasi, Modeling Intra- and Extra-Particle Processes of Wood Fast Pyrolysis, AIChE J. 48 (2002) 2386–
246
Chapter 7
2397.
[26] R. Bounaceur, B. Ouartassi, A. Dufour, A. Zoulalian, Modelling intra-particle phenomena of biomass pyrolysis,
Chem. Eng. Res. Des. 89 (2011) 2136–2146.
[27] J. Ratte, F. Marias, J. Vaxelaire, P. Bernada, Mathematical modelling of slow pyrolysis of a particle of treated
wood waste., J. Hazard. Mater. 170 (2009) 1023–1040.
[28] A. Sharma, V. Pareek, S. Wang, Z. Zhang, H. Yang, D. Zhang, A phenomenological model of the mechanism
of lignocellulosic biomass pyrolysis processes, Comput. Chem. Eng. 60 (2014) 231–241.
[29] Michel Bellais, Modelling of the pyrolysis of large wood particles, KTH - Royal Institute of Technology
Stockholm, 2007.
[30] X. Shi, F. Ronsse, J.G. Pieters, Finite element modeling of intraparticle heterogeneous tar conversion during
pyrolysis of woody biomass particles, Fuel Process. Technol. 148 (2016) 302–316.
[31] F. Patuzzi, S. Ciuta, M.J. Castaldi, M. Baratieri, Intraparticle gas sampling during wood particle pyrolysis:
Methodology assessment by means of thermofluidynamic modeling, J. Anal. Appl. Pyrolysis. 113 (2015) 638–
645.
[32] E. Leonardi, G. de Vahl Davis, K. Yuen, G.H. Yeoh, Modelling the pyrolysis of wet wood – I. Three-dimensional
formulation and analysis, Int. J. Heat Mass Transf. 50 (2007) 4371–4386.
[33] A.H. Mahmoudi, F. Hoffmann, B. Peters, Semi-resolved modeling of heat-up, drying and pyrolysis of biomass
solid particles as a new feature in XDEM, Appl. Therm. Eng. 93 (2016) 1091–1104.
[34] Myongsook S., Oh William A., Peters Jack B. Howard, An experimental and modeling study of softening coal
pyrolysis, AIChE J. 35 (1989).
[35] P.N. Ciesielski, M.F. Crowley, M.R. Nimlos, A.W. Sanders, G.M. Wiggins, D. Robichaud, B.S. Donohoe, T.D.
Foust, Biomass particle models with realistic morphology and resolved microstructure for simulations of
intraparticle transport phenomena, Energy and Fuels. 29 (2015) 242–254.
[36] C. Di Blasi, Physico-chemical processes occurring inside a degrading two-dimensional anisotropic porous
medium, Int. J. Heat Mass Transf. 41 (1998) 4139–4150.
[37] J. Eitelberger, K. Hofstetter, Prediction of transport properties of wood below the fiber saturation point - A
multiscale homogenization approach and its experimental validation. Part I: Thermal conductivity, Compos. Sci.
Technol. 71 (2011) 134–144.
[38] X. Bai, K.H. Kim, R.C. Brown, E. Dalluge, C. Hutchinson, Y.J. Lee, D. Dalluge, Formation of phenolic oligomers
during fast pyrolysis of lignin, Fuel. 128 (2014) 170–179.
[39] S. Zhou, M. Garcia-Perez, B. Pecha, A.G. McDonald, R.J.M. Westerhof, Effect of particle size on the
composition of lignin derived oligomers obtained by fast pyrolysis of beech wood, Fuel. 125 (2014) 15–19.
[40] M. LI, J.-G. Dong, Z.-X. Huang, L. LI, W. Gao, H.-Q. Nian, Z. Fu, P. Cheng, Z. Zhou, Analysis of Cigarette
Smoke Aerosol by Single Particle Aerosol Mass Spectrometer, Chinese J. Anal. Chem. 40 (2012) 936–939.
[41] R. Hall, L. Murdoch, R. Falta, B. Looney, B. Riha, Evaluation of liquid aerosol transport through porous media,
J. Contam. Hydrol. 190 (2016) 15–28.
[42] T.M. Bucher, H.V. Tafreshi, G.C. Tepper, Modeling performance of thin fibrous coatings with orthogonally
layered nanofibers for improved aerosol filtration, Powder Technol. 249 (2013) 43–53.
[43] S. Marre, J. Palmeri, A. Larbot, M. Bertrand, Modeling of submicrometer aerosol penetration through sintered
granular membrane filters, J. Colloid Interface Sci. 274 (2004) 167–182.
[44] A. Anca-Couce, Reaction mechanisms and multi-scale modelling of lignocellulosic biomass pyrolysis, Prog.
Energy Combust. Sci. 53 (2016) 41–79.
[45] J. Lédé, M. Ferrer, O. Boutin, Radiant flash pyrolysis of cellulose—Evidence for the formation of short life time
intermediate liquid species, J. Anal. Appl. Pyrolysis. 47 (1998) 13–31.
[46] P.R. Patwardhan, Understanding the product distribution from biomass fast pyrolysis, Iowa State University,
2010. Doctoral Thesis.
[47] J. Diebold, J. Scahill, Ablative pyrolysis of macroparticles of biomass, in: Proceedings of Specialists’ Workshop
on fast Pyrolysis of Biomass, Copper Mountain, in: R.P. Overend, T.A. Milne, L.K. Mudge (Eds.), Fundam.
Thermochem. Biomass Convers., Springer Netherlands, Copper Mountain, 1985: pp. 539–555.
[48] J. Diebold, J. Scahill, Ablative pyrolysis of biomass in solid-convective heat transfer environments, in: R. P.
Overend, T. A. Milne, L. K. Mudge (Eds.), Fundam. Thermochem. Biomass Convers., Springer Netherlands,
1985: pp. 539–555.
[49] C. Di Blasi, Heat transfer mechanisms and multi-step kinetics in the ablative pyrolysis of cellulose, Chem. Eng.
Chapter 7 247
Sci. 51 (1996) 2211–2220.
[50] N. Bech, P.A. Jensen, K. Dam-Johansen, M.B. Larsen, Modelling solid-convective flash pyrolysis of straw and
wood in the Pyrolysis Centrifuge Reactor, Biomass and Bioenergy. 33 (2009) 999–1011.
[51] S. Aramideh, Q. Xiong, S.-C. Kong, R.C. Brown, Numerical simulation of biomass fast pyrolysis in an auger
reactor, Fuel. 156 (2015) 234–242.
[52] O. Onay, O. Mete Koçkar, Fixed-bed pyrolysis of rapeseed (Brassica napus L.), Biomass and Bioenergy. 26
(2004) 289–299.
[53] P. Mellin, E. Kantarelis, W. Yang, Computational fluid dynamics modeling of biomass fast pyrolysis in a fluidized
bed reactor, using a comprehensive chemistry scheme, Fuel. 117 (2014) 704–715.
[54] D. Barisano, G. Canneto, F. Nanna, A. Villone, E. Alvino, M. Carnevale, G. Pinto, Production of gaseous carriers
via biomass gasification for energy purposes, in: Energy Procedia, Elsevier BV, 2014: pp. 2–11.
[55] Y. Su, Y. Luo, Y. Chen, W. Wu, Y. Zhang, Experimental and numerical investigation of tar destruction under
partial oxidation environment, Fuel Process. Technol. 92 (2011) 1513–1524.
[56] http://www.nist.gov/.
[57] R.J.. Westerhof, Refining fast pyrolysis of biomass, Thesis Degree of Doctor of Science, University of Twente,
2011.
[58] P. Hasler, T. Nussbaumer, P. Morf, Mechanisms and kinetics of homogeneous secondary reactions of tar from
continuous pyrolysis of wood chips, Fuel. 81 (2002) 843–853.
[59] Weerawut Chaiwat, Isao Hasegawa, Takaaki Tani, Kenshi Sunagawa, Kazuhiro Mae, Analysis of cross-
linking behavior during pyrolysis of cellulose for elucidating reaction pathway, Energy Fuels. 23 (2009) 5765–
5772.
[60] S. Whitaker, Simultaneous Heat, Mass, and Momentum Transfer in Porous Media: A Theory of Drying, Adv.
Heat Transf. 13 (1977) 119–203.
[61] Morten Grunnar Gronli, A theorical and experimental study of the thermal degradation of biomass, Norwegian
University of Science and Technology, 1996.
[62] S.S. Alves, J.L. Figueiredo, A model for pyrolysis of wet wood, Chem. Eng. Sci. 44 (1989) 2861–2869.
[63] Y. Lin, J. Cho, J.M. Davis, G.W. Huber, Reaction-transport model for the pyrolysis of shrinking cellulose
particles, Chem. Eng. Sci. 74 (2012) 160–171.
[64] C. Di Blasi, Heat, momentum and mass transport through a shrinking biomass particle exposed to thermal
radiation, Chem. Eng. Sci. 51 (1996) 1121–1132.
[65] J. Zhang, J.J.J. Chen, N. Zhou, Characteristics of jet droplet produced by bubble bursting on the free liquid
surface, Chem. Eng. Sci. 68 (2012) 151–156.
[66] E. Ghabache, A. Antkowiak, C. Josserand, T. Séon, On the physics of fizziness: How bubble bursting controls
droplets ejection, Phys. Fluids. 26 (2014).
[67] Montoya .J, B. Pecha, F. Chejne Janna, Micro-Explosion of Liquid Intermediates During the Fast Pyrolysis of
Sucrose and Organosolv Lignin, Submitt. to J. Anal. Appl. Pyrolysis. (n.d.).
[68] D. Li, Coalescence between small bubbles or drops, Huagong Xuebao/Journal Chem. Ind. Eng. 45 (1994) 38–
44.
[69] L. Duchemin, S. Popinet, C. Josserand, S. Zaleski, Jet formation in bubbles bursting at a free surface, Phys.
Fluids. 14 (2002) 3000–3008.
[70] S.C. Georgescu, J.L. Achard, É. Canot, Jet drops ejection in bursting gas bubble processes, Eur. J. Mech.
B/Fluids. 21 (2002) 265–280.
[71] Shen, P.P.G. Debenedetti, A kinetic theory of homogeneous bubble nucleation, J. Chem. Phys. 768 (2003)
768.
[72] A. Trubetskaya, P.A. Jensen, A.D. Jensen, M. Steibel, H. Spliethoff, P. Glarborg, Influence of fast pyrolysis
conditions on yield and structural transformation of biomass chars, Fuel Process. Technol. 140 (2015) 205–
214.
[73] L. Sun, P. Fu, S. Hu, S. Su, J. Wang, J. Xiang, Evaluation of the porous structure development of chars from
pyrolysis of rice straw: Effects of pyrolysis temperature and heating rate, J. Anal. Appl. Pyrolysis. 98 (2012)
177–183.
[74] L. Ghazaryan, Aerosol dynamics in porous media, University of Twentee, 2000.
[75] H. Fallah, H.B. Fathi, H. Mohammadi, The Mathematical Model for Particle Suspension Flow through Porous
Medium, Geomaterials. 2 (2012) 57–62.
248
Chapter 7
[76] J.R. Serrano, H. Climent, P. Piqueras, E. Angiolini, Filtration modelling in wall-flow particulate filters of low soot
penetration thickness, Energy. 112 (2016) 883–898.
[77] C. Yue, Q. Zhang, Z. Zhai, Numerical simulation of the filtration process in fibrous filters using CFD-DEM
method, J. Aerosol Sci. 101 (2016) 174–187.
[78] Westerhof, H.S. Nygård, W.P.M. van Swaaij, Kersten, Brilman, Effect of Particle Geometry and Microstructure
on Fast Pyrolysis of Beech Wood, Energy & Fuels. 26 (2012) 2274–2280.
[79] J. Lédé, Comparison of contact and radiant ablative pyrolysis of biomass, J. Anal. Appl. Pyrolysis. 70 (2003)
601–618.
[80] A.V. Sepman, de Goey, Plate reactor as an analysis tool for rapid pyrolysis of biomass, Biomass and Bioenergy.
35 (2011) 2903–2909.
[81] R. Vinu, Linda J. Broadbelt., A mechanistic model of fast pyrolysis of glucose-based carbohydrates to predict
bio-oil composition., Energy Environ. Sci. 5 (2012) 9808–9826.
[82] Matthew S.Mettler, Alex D Paulsen, Dionisios G Vlachos, Paul J. Dauenhauer, The chain length effect in
pyrolysis: bridging the gap between glucose and cellulose, Green Chem. 14 (2012) 1284–1294.
[83] Drummond, I.W. Drummond, Pyrolysis of Sugar Cane Bagasse in a Wire-Mesh Reactor, Ind. Eng. Chem. Res.
35 (1996) 1263–1268.
[84] J.F. Stubington, S. Aiman, Pyrolysis kinetics of bagasse at high heating rates, Energy & Fuels. 8 (1994) 194–
203.
[85] A. Günther, S. Wälchli, P.R. von Rohr, Droplet production from disintegrating bubbles at water surfaces. Single
vs. multiple bubbles, Int. J. Multiph. Flow. 29 (2003) 795–811.
[86] X. Gong, Y. Yu, X. Gao, Y. Qiao, M. Xu, H. Wu, Formation of anhydro-sugars in the primary volatiles and solid
residues from cellulose fast pyrolysis in a wire-mesh reactor, Energy and Fuels. 28 (2014) 5204–5211.
[87] M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W.J. Lee, C.Z. Li, Effects of temperature on the formation of
lignin-derived oligomers during the fast pyrolysis of Mallee woody biomass, Energy and Fuels. 22 (2008) 2022–
2032.
[88] A.D. Paulsen, M.S. Mettler, P.J. Dauenhauer, The role of sample dimension and temperature in cellulose
pyrolysis, in: Energy and Fuels, 2013: pp. 2126–2134.
[89] C. Branca, C. Di Blasi, A. Galgano, Chemical characterization of volatile products of biomass pyrolysis under
significant reaction-induced overheating, J. Anal. Appl. Pyrolysis. (2015).
[90] S.S. Liaw, S. Zhou, H. Wu, M. Garcia-Perez, Effect of pretreatment temperature on the yield and properties of
bio-oils obtained from the auger pyrolysis of Douglas fir wood, Fuel. 103 (2013) 672–682.
[91] V. M. García Mollá R. Gómez Padilla, Description of the MATLAB functions SENS_SYS and SENS_IND.,
Meliana (Valencia), SPAIN, 2002.
[92] J. Lédé, O. Authier, Temperature and heating rate of solid particles undergoing a thermal decomposition. Which
criteria for characterizing fast pyrolysis?, J. Anal. Appl. Pyrolysis. 113 (2015) 1–14. d
[93] J. Lédé, Reaction temperature of solid particles undergoing an endothermal volatilization. Application to the
fast pyrolysis of biomass, Biomass and Bioenergy. 7 (1994) 49–60.
[94] S. Orvalho, M.C. Ruzicka, G. Olivieri, A. Marzocchella, Bubble coalescence: Effect of bubble approach velocity
and liquid viscosity, Chem. Eng. Sci. 134 (2015) 205–216.
249
Conclusions

8. Chapter 8. Conclusions
The development of this research has shown significant conclusions regarding new experimental methodologies to
understand the main physical events and chemicals that occur during devolatilization of lignocellulosic material, and
the integration of these phenomena to a sigle particle model, to theoretically predict the dynamics of bubble and ejection
of aerosols. A summary of the main conclusions of this work are shown below:

1. It is well known that char produced from lignocellulosic biomass retains its general structure during pyrolysis.
In this work, visualization studies on cellulose, organosolv lignin, milled wood lignin, xylan, de-ashed xylan,
sugarcane bagasse, and de-ashed sugarcane bagasse were performed. The lignin melted, bubbled
dramatically, but shrank again. The cellulose melted and shrank. Ash-free xylan melted completely and shrank
into a pool, then carbonized. However, when ash content was high in the xylan, the xylan foamed dramatically
and grew before carbonizing due to the catalytic effects of ash on gas formation. In sugarcane bagasse,
removing the ash appears to have enhanced the formation of a liquid intermediate; SEM of the char indicates
that more of the biomass melted (likely the hemicellulose and lignin. Future work needs to be done to better
understand the effect of metal content on the formation of products responsible for char retaining the structure
of the original biomass.

2. A new methodology has been proposed to describe of the dynamics of bubble formation in organosolv lignin
and sucrose that combines fast speed visualization (125 fps) with mathematical modeling. The model uses a
population balance to theoretically predict overall rates of bubble birth and death, bubble bursting, and aerosol
ejection. Experimentally, it was observed that gas bubbles follow a lognormal distribution versus bubble size
within the liquid intermediate phase for both materials. This distribution function changes over time due to
increased viscosity from resolidification reactions that generate char. The model predicts aerosol ejection
yields of 21.18 % w/w from organosolv lignin and 17.40 % w/w from sucrose during pyrolysis.

3. For the first time a simple methodology based on SEM visualization for morphological characterization of
ejected aerosols during biomass and model compounds pyrolysis is proposed. Under vacuum (150 mbar) and
high heating rates (1200 °C/s) conditions aerosol ejection intensity for all studied materials is favored, by
favoring oligomers evaporation, bubbles bursting and micro-explosions at the intermediate liquid phase. For
all studied materials, relative aerosol yields are favored at high heating rates and vacuum conditions. Higher
relative yield was obtained for avicel cellulose at 150 mbar and 1200 °C/s (71.78 % Wt.).

4. In this work a new reactor and methodology was presented to study sugarcane bagasse with direct
impregnation of a biomass sample on a steel sheet. The film used for pyrolysis studies had an average
thickness of 60 µm, uniform over the entire surface distribution. This allowed for the same reaction conditions
each point on the film. Parametric studies across temperature and heating rate were performed with sugarcane
bagasse. The results showed that beyond 100 °C/s, no significant changes are observed in the yields of the
products due to the low thermal conductivity of the biomass. At high heating rates (> 100 °C/s), the highest
yields of oligomers of sugars and lignin and lowest production of char were observed. The bio-oil obtained at
low heating rates were high in water, organic acids, aldehydes, and ketones with a lower concentration of
oligomers compared recovered with high heating rates. Overall yields obtained at high heating rates are on
the same order of magnitude as those previously reported for bagasse in a fluidized bed reactor.
250
Conclusions

5. Good fittings of experimental data were achieved for main species during bagasse devolatilization with an
order one distribution of activation energies model (DAEM). Kinetic parameters were in the order of 106 - 107
s-1, activation energies between 170 - 220 kJ/mol and standard deviation between 15 – 21 kJ/mol. It was
analyzed and proposed a one stage reaction for pyrolytic lignin and levoglucosan behavior as model
compounds for oligomers secondary reactions during biomass pyrolysis. These reactions produce water,
carbon dioxide and char as main products.

6. Was proposed a new single particle model that couples physical phenomena as bubbles dynamics inside
intermediate liquid phase, with chemical phenomena like kinetics of main species released during sugarcane
pyrolysis. The model is able to predict the aerosol ejection intensity produced by bubbles bursting within the
intermediate liquid phase. The variable most sensitive to aerosol yield is the particle size, showing yields close
to 0 at 2 mm. The temperature has a positive effect on aerosol yield by intensification of chemical reactions
and bubble dynamics (nucleation). Also heating rate has positive effect on bubbles bursting and aerosol
ejection but with minor intensity in comparison with particle size. Was found, the most sensitive parameters
for thermal behavior is the biomass thermal conductivity and global energy transfer coefficient from the hot
plate to the biomass. Also the liquid properties as viscosity and surface tension are the most sensitive
parameters. Both parameters are related with the bubbles nucleation.

8.1. Future works


Throughout the development of this research there have been many research topics that can be studied in more detail
to improve the understanding of the biomass pyrolysis process, and explore this as a ''bio-refinery'' of which it can be
taken advantage of the assortment crew of compounds for pharmaceutical purposes, food, or applications in the
manufacturing sector. An important step towards this goal, is the fundamental scale knowledge of the physical and
chemical phenomena and their interactions in a real operation that will allow the 'Chemical-Tunning' 'to a desired
product at the lowest possible costs.
The work carried out in this thesis, opened a first door to clarify some of these phenomena, by implementing
experimental new techniques and coupling with theoretical models. However, it takes a lifetime to answer all questions
in the field of pyrolysis. In view of the above global outlook, future work that may contribute to this purpose they are
presented below:

 It is important to deepen in the study of reactivity of the main compounds involved within the intermediate liquid
phase and its interactions with the mineral material. It is well known that lignocellulosic biomass, decompose at
different temperature ranges [1], due to the hemicellulose, lignin and cellulose content. Each of these compounds
forms a liquid phase intermediate, which can interact with each other chemically. The importance of mass and
energy transfer within this liquid phase must be explored for a better understanding of the dynamics of bubble
formation and aerosol ejection.

 New visualization techniques, including thermography visualization are important to the study of micro-explosions
and bubbles dynamics within the liquid phase intermediate. These should be enough sensitive to capture of short
duration events (< 1 ms) and sufficient resolution to track the ejected aerosol (size 1 - 10 μm). These methodologies
allow build the bridge between dynamics of bubble formation and ejection of aerosols.
251
Conclusions

 Understanding the bubbles dynamics within the liquid phase, should be proposed new experimental strategies as
the use of blowing agents used in polymer science [2], to promote micro- explosion by bubbles bursting. The
recovered products are oligomers of sugars (levoglucosan, cellotriosan, cellobiosan) with high potential in the
pharmaceutical sector.

 It is important to advance in develop of quantification techniques and rapid characterization of oligomers belongs
to lignin and cellulose. This will be useful to estimate the physicochemical properties of the intermediate liquid
phase also to bring new clues to explain the reaction mechanisms.

 Although the mathematical model proposed in this research is quite detailed, even should incorporate the dynamics
of aerosols within the pororous matrix. Especially can build more accurate models including a realistic
representation of the internal structure of biomass as suggested Ciesielinki et al [3], and aerosol transport through
these structures, as proposed by Gazaryan [4].

 At reactor scale, many researchers have suggested that the bed agglomeration of inert material is due to collisions
of aerosols ejected during devolatilization that acts like glue with near particles [5,6]. The bed hydrodynamics in
fluidized beds can be seriously affected by the agglomeration of these solids. New models at reactor and plant
scale, involving the aerosol ejection and the collisions of these with inert material within the bed, allow more realistic
designs with improved performance.

 The understanding of aerosol ejection, also its chemical composition needs to be explored in deep. The synergism
between physical-chemical properties and solid surfaces (new materials) should be taken into account to develop
new selective condensers based on wettability. These new designs promote high bio-oil recover minimizing
upgrading and refining cost.
252
Conclusions

8.2. References
[1] H. Reith, P. de Wild, E. Heeres, Biomass pyrolysis for chemicals, Biofuels. 2 (2011) 185–208.
[2] Ruiz, M. Vincent, J.F. Agassant, T. Sadik, C. Pillon, C. Carrot, Polymer foaming with chemical blowing agents:
Experiment and modeling, Polym. Eng. Sci. 55 (2015) 2018–2029.
[3] P.N. Ciesielski, M.F. Crowley, M.R. Nimlos, A.W. Sanders, G.M. Wiggins, D. Robichaud, B.S. Donohoe, T.D.
Foust, Biomass particle models with realistic morphology and resolved microstructure for simulations of
intraparticle transport phenomena, Energy and Fuels. 29 (2015) 242–254.
[4] L. Ghazaryan, Aerosol dynamics in porous media, University of Twentee, 2000.
[5] Roel Johannes Maria {Westerhof, Refining fast pyrolysis of biomass, Universtity of Twentee, 2011.
[6] T. Iwasaki, S. Suzuki, T. Kojima, Influence of Biomass Pyrolysis Temperature, Heating Rate and Type of
Biomass on Produced Char in a Fluidized Bed Reactor, Energy Environ. Res. 4 (2014) 64–72.
253
Appendix

9. Appendix
Appendix A. Videos (Chapter 2 – 5)
 Video A. Organosolv lignin
 Video B. Ball mill lignin
 Video C. Raw xylan
 Video D. Deashed xylan
 Video E. Avicel cellulose
 Video F. Raw sugarcane bagasse
 Video G. Acid washed sugarcane bagasse
 Video H. Sucrose
 Video J. Organosolv lignin in vacuum
 Video K. Sucrose in vacuum

Appendix B. Chromatograms (Chapter 5 – 6) and Temperature profile at


1000°C/s and Tf=500C.

B1-GC-MS Spectrogram from bio-oil obtained at 1200°C/s and 500°C.


254
Appendix

Retention time
Peak Compound name m/z
min
1 Water 1.045 18
2 Furan 1.23 39
3 Formic Acid 1.25 29
4 Acetic Acid 1.50 43
5 Glycoaldehyde 1.95 31
6 Acetoin 2.02 45
7 Furfuryl Alcohol 3.5 98
8 Phenol 5.2 94
9 Naphthalene 8.4 128
10 Pyrocatechol 8.5 110
11 Benzofuran 8.8 118
12 5-Hydroxymethyl-Furfural 9.0 97
13 Eugenol 10.8 164
14 Vanillin 11.3 151
15 Levoglucosan 12.5 60
16 1,2-Benzene dicarboxilic acid 22.0 104
17 1-Octacosanol 27.0 57
18 Unknown compound 18.0 ----
19 N-Hexadecanoic acid 17.6 43,73
20 1-Pentadecamine,N,N-Dimethyl 16.35 58
21 4-((1E)-3-Hydroxy-1-propenyl)-2-methoxyphenol 15.5 137
22 1-Butene,3-Methyl-3-(1-ethoxyethoxy) 14.14 69,41
23 1,3-Propaniedol-2(Hidroxymethyl)-2-nitro 11.71 57,29
24 2-Methoxy-4-Vinyl phenol 10.27 135,150
25 2(5H) Furanone 4.5 55
26 O-Xylene 4.0 91
27 1,2-Epoxy-3-Propyl Acetate 2.5 43,45

B2. µGC Chromatogram for permanent gases.


255
Appendix

Molecular weigth
Peak Compound Retention time (min) Formula
(g/mol)
1 Hydrogen 0.22 H2 2
2 Carbon monoxide 1.125 CO 28
3 Methane 1.575 CH4 17
4 Carbon dioxide 0.44 CO2 44

B3. HPLC Spectrogram for sugars


256
Appendix

Peak Compound Retention time Formula Molecular weigth


(min) (g/mol)
1 Unknown 7 ------- --------
2 Cellotriosan 10.948 C18H30O15 486.420
3 Cellobiosan 13.855 C12H20O10 324.280
4 1,6 Anhydro β-D-Glucofuranose 16.248 C6H10O5 162.140
5 Unknown 18 ------- --------
6 1,4:3,6 Dianhydro α-D-Gluco Pyranose 20.005 C6H8O4 144.130
7 Levoglucosenone 21.865 C6H6O3 126.110
8 Unknown 25 ------- --------
9 Unknown 28 ------- --------

B4. Temperature profiles for diferent biomass thickness at 1000°C/s

2mm-1000°C/s 1mm-1000°C/s

0.5mm-1000°C/s 0.1mm-1000°C/s
257
Appendix
258
Appendix

Appendix C. Auxiliary equations for modeling (Chapter 7)


C.1. Effective Thermal Conductivity
The effective thermal conductivity of the biomass has two contributions: molecular contribution and radiation.

K eff = Kcond + Krad

With Kcond is the molecular contribution by conduction, and Krad is the pore radiation contribution

Krad = 4σϵdpore T 3

dpore Is the pore diameter. It is calculated by linear interpolation between the initial pore diameter (raw material) and
the pore diameter of final char. Pores diameter are take of Ciesielinki et al [1].

dpore = (1 − X)db + Xdc

X is the biomass conversion, db , dc are the diameter pores for raw biomass and char respectively. Using again linear
interpolation, the thermal conductivity in each time are calculated with the below expression

Kcond = (1 − X)Kb + XKc

Kb y Kc are the thermal conductivities for raw biomass and char both presented in table 3 chapter 7.

C.2. Effective diffusivity


Transport of species i by diffusion within the particle are calculated using the Chapman-Enskoog method. The binary
diffusivity of each species and subsequently diffusivity in the mixture are described as.

T 3/2 Mwi + Mwj


Dij = 1.86 ∗ 10−3 √
Pσ2ij ω Mwi Mwj

Dij is the diffusivity of species i in species j, σ2ij, ω are the binary interaction parameters. The diffusivity of species i in
all mixture are calculated like linear combination of binaries diffusivities:

Di/mix = ∑ yi Dij
i

Finally, in order to taken into account, the tortuosity of the porous matrix, an correction factor is included, Where τ is
the tortuosity coeffiecient.

Di/min
Deffi =
τ
259
Appendix

C.3. Gas Viscosity


The dynamic viscosity of the whole gas phase is determined similar to the way diffusivity, using Chapma-Enskog
method for each species i.

(Mwi T)1/2
μi = 2.6693 ∗ 10−6
ε2 ωT ∗

μi , is the dynamic viscosity of species i, T ∗ is the reduced temperature in gas phase, ω y ε are the integral and diameter
of collisions respectively. Finally, using addition law, the viscoity in gas mixture is estimated by.

μmix ∑ yi μi
i

C.4. Calorific capacities


Like other properties, the solid heat capacity was estimated by linear interpolation according to the conversion between
the virgin biomass and char.

Cps = (1 − X)Cpb + XCpc

Cps , Cpb , Cpc are the solid heat capacities for raw biomass and char. For the gases, is estimated the individual heat
capacities of each species as temperature polynomial

Cpi = a + bT + cT 2

All polymomic functions are founded in NIST. http://webbook.nist.gov/

Finally, the heat capacity of the gas mixture is calculated as a linear combination of the heat capacity of each species
in the mixture.

Cpi = ∑ yi Cpi
i

The heat capacity of the liquid phase, was taken the levoglucosan as model compound

www.uni-kassel.de/upress/online/frei/978-3-89958-708-1.volltext.frei.pdf.

C.5. External heat transfer coefficient


The external heat transfer coefficient hext took into account the radiative and convective contribution:

hext = hconv + hrad


260
Appendix

With hrad calculated using the expression:

hrad =∈ ϑ(T 3 − T∞
3)

∈, ϑ are the surface emissivity and Stephan-Boltzmann constant. The convective coefficient is estimated for a flat
surface with the hottest surface in the bottom by expression:

K l Ra1/4
hconv =
L

L, Ra son are the plate length and Ra is the Raley number. K l .is the stainless Steel termal conductivity

C.6. External mass transfer coefficient


The external mass transfer coefficient for a flat surface based on the Sherwood number (Sh) was calculated by:

ShDeffN2
K masa =
L

Sh =0.664(Re1/2 Sc1/3 )

Re, Sc are Reynolds and Schmidt dimensionless numbers of the external gas (N2).

C.7. Porous permeability


The permeability of gases through the porous matrix was calculated knowing the permeability of the virgin biomass and
char respectively by the following a linear interpolation.

Ug = (1 − X)Ub + XUc

Ub , Uc , are the permeability of the raw biomass and char taken from Sharma et al [2].

C.8. Gas bubbles rising velocity


For small diameter bubbles as those formed within the liquid phase (0.01-1mm), bubbles are completely spherical [3,4],
and can be applied Stokes equation for calculating the bubles rising velocity

gDbubble (ρl − ρg )
v𝑏 =
18ul

With g gravity, Dbubble bubble diameter, ρl , ρg , ul are the densities of liquid phase and gas (bubbles).

C.9. Bubbles nucleation


261
Appendix

According to the classical theory of nucleation, the rate of bubble formation may represented by Arrhenius law.

∆𝐄𝐜𝐫𝐢

𝐉̇𝐧𝐮𝐜 = 𝐍𝐨 𝐞 𝐊𝐛𝐓

With 𝐉̇𝐧𝐮𝐜 It represents the rate of bubble nucleation (# bubbles / s m3) by particle volume. A reasonable estimate for
No is calculating the number of liquid molecules formed, and assuming that each molecule when it decomposes is a
seed for a new bubble

Navogadro 1bubble
No = (εl ρl )
Mwl 1molecule

∆Ecri , is the mínimum gibbs energy to form a new bubble:

16π γ3
∆Ecri = ( )
3 (Pv − P)

With γ, Pv like superficial tension and vapor pressure of liquid compound. The value of the surface tension was taken
from Paul et al [5] and the vapor pressure was taken Dufour et at and Suuberg et al, [6] for cellulose oligomers and
from [7] for lignin oligomers.

C.10. Enthalpy of reaction


The enthalpy of reaction was calculated by the heats of formation of products and reagents:

DHrxn = DHprod − DHreact

DHprod , DHreact are the enthalpies of products and reagents

DHprod = ∑ hfio + ∫ Cpi dT


i To

The formation enthalpy for the raw biomass are calculted following the method descrited in [8].

C.11. Population balances


The general Population balance to describe the bubbles withing the liquid phase is the follow:

𝛛(𝐅(𝐭,𝐫𝐛 ,𝐱)) 𝛛(𝐅(𝐭,𝐫𝐛 ,𝐱)v𝑏 ) 𝛛 ̇ − 𝐃𝐞𝐚𝐭𝐡


̇ ) + 𝐉̇𝐧𝐮𝐜𝐥𝐞 𝛅(𝐫𝐛 − 𝐫 ∗ )=0
𝛛𝐭
+ 𝛛𝐱
+ 𝛛𝐫 (𝐫𝐛̇ 𝐅(𝐭, 𝐫𝐛 , 𝐱)) + (𝐁𝐢𝐫𝐭𝐡
𝐛

rb Is the bubble radius, F(t, rb , x) is the distribution function of the number of bubbles per unit volume of particle.
Mathematical expressions proposals for the birth and death of the bubbles are presented below:
262
Appendix
rb
1 r
̇ = φ ∫ F (t, b − r1 , x) F(t, r1 , x) dr1
Birth
2 α
0


̇ = φF(t, rb , x) ∫ F(t, r1 , x) dr1
Death
0

The above expressions represent rates per unit volume of particle for the birth and death of bubbles. Substituting in the
original equation of Population balance

rb
𝛛(𝐅(𝐭, 𝐫𝐛 , 𝐱)) 𝛛(𝐅(𝐭, 𝐫𝐛 , 𝐱)v𝑏 ) 𝛛 1 rb
+ + (𝐫𝐛̇ 𝐅(𝐭, 𝐫𝐛 , 𝐱)) + ( φ ∫ F (t, − r1 , x) F(t, r1 , x) dr1 )
𝛛𝐭 𝛛𝐱 𝛛𝐫𝐛 2 α
0

− (φF(t, rb , x) ∫ F(t, r1 , x) dr1 ) + 𝐉̇𝐧𝐮𝐜𝐥𝐞 𝛅(𝐫𝐛 − 𝐫 ∗ ) = 𝟎


0

Multiplying the derivative within by rbn and remembering that by definition the moment is:


∂Mn ∂(rbn F(t, rb , x))
=∫ drb
∂t ∂t
0

∞ ∞ ∞
𝛛(𝐫𝐛𝐧 𝐅(𝐭, 𝐫𝐛 , 𝐱)) 𝛛(𝐫𝐛𝐧 𝐅(𝐭, 𝐫𝐛 , 𝐱)v𝑏 ) 𝛛
∫ 𝐝𝐫𝐛 + ∫ 𝐝𝐫𝐛 + ∫ (𝐫 𝐧 𝐫 ̇ 𝐅(𝐭, 𝐫𝐛 , 𝐱)) 𝐝𝐫𝐛
𝛛𝐭 𝛛𝐱 𝛛𝐫𝐛 𝐛 𝐛
𝟎 𝟎 𝟎
∞ rb
1 rb
+ ∫ ( φ ∫ 𝐫𝐛𝐧 F (t, − r1 , x) F(t, r1 , x) dr1 ) 𝐝𝐫𝐛
2 α
𝟎 0
∞ ∞ ∞

− (∫ φF(t, rb , x) ∫ 𝐫𝐛𝐧 F(t, r1 , x) dr1 ) 𝐝𝐫𝐛 + ∫ 𝐫𝐛𝐧 𝐉̇𝐧𝐮𝐜𝐥𝐞 𝛅(𝐫𝐛 − 𝐫 ∗ ) 𝐝𝐫𝐛 = 𝟎


0 0 𝟎

Replacing according to the definition of moment


∂Mn ∂
+ 〈v𝑏 〉 (Mn ) + n〈rḃ〉 ∫ rbn−1 F(t, rb , x)drb
∂t ∂x
0
∞ rb
1 rb
+ ∫ ( φ ∫ 𝐫𝐛𝐧 F (t, − r1 , x) 𝐫𝐛𝐧 F(t, r1 , x) dr1 ) 𝐝𝐫𝐛 − 𝛗𝐌𝟎 𝐌𝟏 + 𝐫𝐛𝐧 𝐉̇𝐧𝐮𝐜𝐥𝐞
2 α
𝟎 0

Now the term with the double integral represents the coalescence can be replaced by a Taylor expansion to eliminate
double integral
263
Appendix
n
1 n
Coa = αn+1 φ ∑ ( ) Mn−i Mi
2 i
i=0

Substituting n = 1,2,3 obtain the equations of the first three moments:

∂(M0 ) 1 ∂
− J̇nuc − φαM02 + φM02 + (v𝑏 M0 )
∂t 2 ∂x

∂(M1 ) ∂
+ 〈Ḋ〉M0 −
̇ Jnuc rb − φα2 M0 M1 + φM0 M1 + (v𝑏 M1 )
∂t ∂x

∂(M2 ) ∂
+ 2〈Ḋ〉M1 −
̇ Jnuc rb 2 − φα3 M0 M2 − α3 φM12 + φM0 M2 + (v𝑏 M1 )
∂t ∂x

The distribution function is constructed from the values of the moments:

1 (log(rb ) − r̅b )2
F(t, rb , x) = exp (− )
√2πσ 2σ2

With:
M1
r̅ bg =
M0

M1 M0 M2
σg = √( 2 − 1)
M0 M1

r̅ bg
r̅b = log
σ 2
√1 + ( g )
r̅ bg
( )

2
σg
σ = √log (1 + ( ) )
r̅ bg

C.12. Breakage and coalescence probability

Parameters φ, α are the coalescence and breakage probability coefficients. These parameter were calculated
according to correlations available for gas-liquid bubbles dynamics.

The breakage probability of one bubble with size rb(k+k+1)to produce 2 bubbles with sizes rb(k) and rb(k+1) are
presented in the below equation [9].
264
Appendix

3 2 3 2
30 rb(k) rb(k)
Breakage probablity (rb(k)/ rb(k+1) ) = 〈( )( 3 ) (1 − 3 ) 〉
rb(k) rb(k+1) rb(k+1)

Characteristic time for breakage (1/s) [10]:

1/2
ρl R eq 3
τb = ( )
16σ

Where ρl , σ are the liquid phase density and surface tension respectively

req + ϑmin
R eq =
2

ϑmin is the minimum eddy size to break a bubble:

ϑmin ≈ 0.084(req )

And

rb(k) + rb(k+)
req = 〈 〉
2

2
Coalescence probability (rb(k)/ rb(k+1) ) = 〈4πρg (rb(k) + rb(k+1) ) (rb(k) + rb(k+1) )〉

taken from Oh, et al [11].

Characteristic time for coalescence (1/s): Taken from Hengel, et al [10]

1/2
ρl req 3
τc = ( )
16σ

Where req is the bubble equivalent size, computed according the mathematical expression

φ = τc (Coalescence probability (rb(k)/ rb(k+1) ))

α = τb (Breakage probability (rb(k)/ rb(k+1) ))

C.13. Bubbles growth rate


265
Appendix

The bubbles growth rate 𝐷̇ or 𝑑𝑟⁄𝑑𝑡 can be written according the follow mathematical equation:
1/2
𝑑𝑟 2 (𝑃𝑣 − 𝑃)
𝐷̇ = ={ }
𝑑𝑡 3 𝜌𝑙

Where 𝑃𝑣 , 𝑃 represents the vapor pressure of the liquid phase, and external pressure (pressure estimated within the
particle by ideal gas law). 𝜌𝑙 is the liquid phase density. The vapor pressure of liquid phase was taken from Dufour et
al [6] and Suuberg et al [7] for the levoglucosan and heavy tars. The above equation represent the growth rate due
mainly to evaporation of molecules within liquid phase, specially levoglucosan (the molten phase have high quantity of
levoglucosan). Is not taken into account effect of surface tension, because at high temperature (temperature of pyrolysis
> 300°C), it have a low magnitude (< 0.01 mN/m2). Also is considered themal equilibrium, then the bubble and liquid
phase have the same temperature then, does not exist thermal effects on bubble growth. It is known that gas bubbles
can grow by chemical reactions within the liquid phase; however this process is not included in the present model,
mainly by the complexity associated with coalescence, breackage and mass transfer restrictions within the molten
phase.

C.14. Polynomials for plate temperature as boundary condition (Tm)

The plate temperature Tm, measured experimentally, is expressed as polynomial function dependent of time,
according the expression:
20

𝑇𝑚 = ∑ 𝑎(𝑖) 𝑡 𝑖
𝑖=0

The values of each parameter for different experimental conditions are obtained fitting the experimental data in a
polynom of order 20, using matlab polyfit function. The values are presented in the underneath table.
Experimental
1000 °C/s 1000 °C/s 1000 °C/s 1200 °C/s 500 °C/s 50 °C/s
condition
400 °C 500 °C 600 °C 500 °C 500 °C 500 °C
Parameter
a(0) 0 0 0 0 0 0
a(1) 0 0 0 0 0 0
a(2) 0 0 0 0 0 0
a(3) 0 0 0 0 0 0
a(4) 0 0 0 0 0 0
a(5) 0 0 0 0 0 0
a(6) 0 -4 -10 0 5 -0.2
a(7) -30 30 70 -20 -45 1.8
a(8) 200 -188 -490 150 293 -11
a(9) -1080 920 2540 -870 -1480 53.7
a(10) 4670 -3467 -10300 3900 5823 -204.5
a(11) -15880 9902 32450 -13770 -17656 603.3
a(12) 42080 -20820 -78410 37870 40637 -1364.7
a(13) -85580 30411 142450 -80040 -69287 2327.9
a(14) 130600 -26646 -188680 127150 84279 -2921.9
a(15) -14463 5673 173570 -146810 -68936 2605.8
a(16) 11017 15457 -102260 116820 34436 -1569.5
a(17) -52440 -17501 33290 -58220 -8942 591.2
a(18) 12380 6496 -4860 14460 810 -122.9
a(19) -120 188 120 -180 529 61.1
a(20) 50 46 30 50 22 24.8
266
Appendix

References

[1] P.N. Ciesielski, M.F. Crowley, M.R. Nimlos, A.W. Sanders, G.M. Wiggins, D. Robichaud, B.S. Donohoe, T.D.
Foust, Biomass particle models with realistic morphology and resolved microstructure for simulations of
intraparticle transport phenomena, Energy and Fuels. 29 (2015) 242–254.
[2] A. Sharma, V. Pareek, S. Wang, Z. Zhang, H. Yang, D. Zhang, A phenomenological model of the mechanism
of lignocellulosic biomass pyrolysis processes, Comput. Chem. Eng. 60 (2014) 231–241.
[3] M.K. Tripathi, K.C. Sahu, R. Govindarajan, Dynamics of an initially spherical bubble rising in quiescent liquid,
Nat. Commun. 6 (2015) 6268.
[4] J. Lambert, I. Cantat, R. Delannay, R. Mokso, P. Cloetens, J.A. Glazier, F. Graner, Experimental growth law for
bubbles in a moderately “wet” 3D liquid foam, Phys. Rev. Lett. 99 (2007).
[5] Paul J Dauenhauer, Joshua L Colby, Christine M Balonek, Wieslaw J Suszynski, Reactive boiling of cellulose
for integrated catalysis through an intermediate liquid, Green Chem. 11 (2009) 1555–1561.
[6] R. Bounaceur, B. Ouartassi, A. Dufour, A. Zoulalian, Modelling intra-particle phenomena of biomass pyrolysis,
Chem. Eng. Res. Des. 89 (2011) 2136–2146.
[7] E. Suuberg, V. Oja, Vapor pressures and heats of vaporization of primary coal tars, J. Chem. Thermodyn. 12
(1997) 243–248.
[8] G. Boissonnet, Des Transformations, de la Biomasse, 2003.
[9] L.E. Patruno, C.A. Dorao, P.M. Dupuy, H.F. Svendsen, H.A. Jakobsen, Identification of droplet breakage kernel
for population balance modelling, Chem. Eng. Sci. 64 (2009) 638–645.
[10] E.I. V Van Den Hengel, N.G. Deen, J.A.M. Kuipers, Application of coalescence and breakup models in a
discrete bubble model for bubble columns, Ind. Eng. Chem. Res. 44 (2005) 5233–5245. doi:10.1021/ie0492449.
[11] Myongsook S., Oh William A., Peters Jack B. Howard, An experimental and modeling study of softening coal
pyrolysis, AIChE J. 35 (1989).
267
Appendix

You might also like