You are on page 1of 77

L IC E N T IAT E T H E S I S

Biocarbon for Fossil Coal Replacement

Aekjuthon Phounglamcheik

Energy Engineering
Licentiate Thesis

Biocarbon for Fossil Coal Replacement

Aekjuthon Phounglamcheik

December 2018

Luleå University of Technology


Department of Engineering Sciences and Mathematics
Division of Energy Science
Printed by Luleå University of Technology, Graphic Production 2018

ISSN 1402-1757
ISBN 978-91-7790-242-3 (print)
ISBN 978-91-7790-243-0 (pdf)
Luleå 2018
www.ltu.se
Abstract

This thesis aims to provide an overview of knowledge in charcoal production for fossil
coal replacement. Charcoal from biomass is a promising material to replace fossil coal in the
industrial sector. Charcoal with quality comparable to fossil coal can be produced by high-
temperature pyrolysis, but the efficiency of the production is relatively low due to low charcoal
yield at high pyrolysis temperature. Increasing charcoal yield by means of secondary char
formation in pyrolysis of thick wood particles, e.g., woodchips, is the primary method to
consider in this work. Secondary char formation can be promoted by increasing concentration
of volatiles during pyrolysis, and/or extending residence time of volatile inside the pore
structure of wood particles. This research has investigated how to increase the efficiency of
charcoal production by bio-oil recycling and CO2 purging. By applying bio-oil recycling, the
increase of charcoal yield was not only because of the increase in reactants, but also due to the
synergetic effect between bio-oil and woodchips upon physical contact. Use of CO2 as purging
gas lowered mass diffusion of volatiles inside the pore structure of woodchip resulting extra
charcoal. In addition, the effect of these techniques can be maximized by ensuring a good
contact of volatiles and solid surface, i.e., usage of thick particles and slow heating.
Characterization of charcoals implied negligible effect of these methods on charcoal properties
such as elemental composition, heating value, porosity, and chemical structure. Furthermore,
the reactivity of charcoal only increased slightly when these methods were applied. These
findings suggest that we can achieve high charcoal yields, both mass and energy, while
maintaining similar fuel properties in charcoal by bio-oil recycling and CO2 purging. In parallel,
a numerical model of pyrolysis in a rotary kiln reactor has been developed to increase the
understanding of how to implement the desired reaction parameters in pyrolysis reactors. The
simulation results showed the evolution of temperature profiles and products distribution along
the reactor length. Two important parameters have been studied, namely rotation speed and
feeding rate. The rotation speed was found to control the solid residence time, while the feeding
rate influences the heat capacity of holdup materials and product distribution. Nonetheless, to
deliver further detailed knowledge of charcoal production, further works are suggested.

iii | P a g e
List of appended papers

Paper A
Increasing efficiency of charcoal production with bio-oil recycling.
Phounglamcheik A., Wretborn T., Umeki K.
Energy & Fuels 32, 9650-9658, 2018.

The candidate carried out most of the macro-TG experiment as well as the ultimate analysis
and the measurement of higher heating value. Interpretation of the results was done by the
candidate under the supervision of the co-authors. Draft of manuscript and revisions were done
by the candidate according to the corrections from the co-authors.

Paper B
Effects of pyrolysis methods and conditions on properties and gasification reactivity of
charcoal from woodchip
Phounglamcheik A., Wang L., Romar H., Broström M., Ramser K., Skreiberg Ø., Umeki K.
Manuscript.

The candidate carried out the preparation of charcoals and the characterization of charcoals by
Raman spectroscopy. Interpretation of all results was done by the candidate under the
supervision of the co-authors. All authors participated in the discussion of the results. Draft of
the manuscript was done by the candidate according to corrections from the co-authors.

Paper C
Modeling and pilot plant runs of slow biomass pyrolysis in a rotary kiln
Babler M., Phounglamcheik A., Amovic M., Ljunggren R., Engvall K.
Applied Energy 207, 123-133, 2017.

The candidate derived the equations and developed the model together with the first author.
Simulation and interpretation of the results were done by the candidate under the supervision
of the first author. The candidate joined the pilot plant runs conducted by the co-authors. The
manuscript and revision were done by the first author with a fair effort from the candidate.

v|Page
List of articles not included in the thesis
Journal papers
Use of biomass in integrated steelmaking: Status quo, future needs and comparison to
other low-CO2 steel production technologies.
Suopajärvi H., Umeki K., Mousa E., Hedayati A., Romar H., Kemppainen A., Wang C.,
Phounglamcheik A., Tuomikoski S., Norberg N., Andefors A., Öhman M., Lassi U., Fabritius
T., Applied Energy 213, 384-407, 2018.

Slow pyrolysis of by-product lignin from wood-based ethanol production: A detailed


analysis of the produced chars.
Toloue N., Suopajärvi H., Mattila O., Umeki K., Phounglamcheik A., Romar H., Sulasalmi P.,
Fabritius T., Energy 164, 112-123, 2018.

Conference presentations
Pyrolysis of wood in a rotary kiln pyrolyzer: Modeling and pilot plant trails.
Phounglamcheik A., Babler M., Donag P., Amovic M., Ljunggren R., Engvall K.
Energy Procedia 105, 908-913, 2017.

Biomass pyrolysis with bio-oil recycle to increase energy recovery in biochar.


Phounglamcheik A., Wretborn T., Umeki K.
25th European Biomass Conference and Exhibition, (proceeding), 2017.

Effects of pyrolysis oil recycling and reaction gas atmosphere on the physical properties
and reactivity of charcoal from wood.
Phounglamcheik A., Wang L., Romar H., Broström M., Ramser K., Skreiberg Ø., Umeki K.
22nd International Symposium on Analytical and Applied Pyrolysis, (refereed), 2018.

Production of metallurgical charcoal from biomass pyrolysis: pilot-scale experiment.


Phounglamcheik A., Pitchot R., Andefors A., Norberg N., Umeki K.
22nd International Symposium on Analytical and Applied Pyrolysis, (refereed), 2018.

Change in size and density of a biomass char during heterogeneous reactions.


Phounglamcheik A., Umeki K.
25th International Conference on Chemical Reaction Engineering, (refereed), 2018.

vi | P a g e
Acknowledgment

First of all, this thesis would never exist under my name if Dr. Kentaro Umeki—
principle supervisor—did not give me this great opportunity to do a PhD in his group. I really
appreciate his supervision as leader, mentor, and coworker. He gave me guidance, advice, and
corrections. In the meantime, he also left free spaces to fill in my ideas. Thanks a lot for your
effort and all the opportunities you gave me during these first two years. I hope you are willing
to train me more for the upcoming years of my PhD and the future.
During these two years, I had the chance to work with many excellent researchers. I
wish to give my gratitude to Dr. Markus Broström, every single advice from him has been
meaningful to my work. I like to give my appreciation to Dr. Liang Wang, who did very hard
work on the TGA experiments in Paper B. Also, it has been my pleasure to work with Dr.
Henrik Romar, Dr. Kerstin Ramser, and Dr. Øyvind Skreiberg. In addition, Paper C would
never succeeded without Dr. Matthäus Bäbler, who worked with me from the beginning until
publishing, I really appreciate his effort and advice. Furthermore, I thank Mr. Thobias Wretborn
who gave an experimental initiative in Paper A.
I would never have done this research without the skills I have learned from my
colleagues in the thermochemical conversion group, Mr. Albert Bach-Oller, Mr. Angel David
Garcia Llamas, Ms. Thamali Rajika Jayawickrama, and Dr. Lu Ding. I like to thank my office
mate, Ms. Chunyan Ma, who I spends most of time during working hours and help me on
experiment. Moreover, my inspiration to do research has been substantially influenced by senior
researchers in the division of Energy Science, namely Dr. Marcus Öhman, Dr. Rikard Gebart,
Dr. Lars Westerlund, Dr. Xiaoyan Ji, Dr. Andrea Toffolo, Dr. Joakim Lundgren, Dr. Elisabeth
Wetterlund, Dr. Anna Krook-Riekkola, Dr. Erik Elfgren, and Dr. Mikael Risberg. Also, thanks
to all colleagues in the division to keep up the the spirit.
I like to express this sentence as my appreciation to my family, Phounglamcheik, and
Col. Narawadee Kirdjongrak for their solid supporting. At last, the ink would never spread
through this thesis without my effort. Thank you, Phai, you keep learning since the day you can
write the first alphabet until you can write thousands of words from your mind today. I hope
you will keep improving yourself while enjoying life.

vii | P a g e
Table of content

Abstract .....................................................................................................................................iii
List of appended papers............................................................................................................. v
List of articles not included in the thesis ................................................................................. vi
Acknowledgment ..................................................................................................................... vii
Chapter I Introduction .............................................................................................................. 1
1.1 Aim and objectives of work......................................................................................... 3
1.2 Research framework and outline of the thesis............................................................. 3
Chapter II Literature review ..................................................................................................... 5
2.1 Biomass pyrolysis........................................................................................................ 5
2.2 Reactivity of charcoal .................................................................................................. 8
2.3 Pyrolysis reactors....................................................................................................... 10
2.4 Summary of research questions ................................................................................. 13
Chapter III Research methods................................................................................................ 15
3.1 Pyrolysis experiment ................................................................................................. 16
3.2 Characterization of charcoal ...................................................................................... 18
3.3 Modeling of pyrolysis reactor.................................................................................... 21
Chapter IV Results and discussion ......................................................................................... 25
4.1 Effect of reaction parameters..................................................................................... 25
4.2 The high-efficiency charcoal production process...................................................... 36
4.3 Pyrolysis of woodchips in a rotary kiln ..................................................................... 41
Chapter V Conclusions and future works .............................................................................. 45
5.1 Conclusions ............................................................................................................... 45
5.2 Future works .............................................................................................................. 47
Bibliography ............................................................................................................................ 49
Appendix .................................................................................................................................. 65
Paper A .................................................................................................................................... 67
Paper B .................................................................................................................................... 85
Paper C .................................................................................................................................. 117

ix | P a g e
Chapter I
Introduction

Charcoal from biomass is a promising material to replace fossil coal in the industrial
sector. Slow pyrolysis, i.e., thermally driven degradation of biomass under an inert atmosphere
at low heating rate, is the typical method to produce charcoal. Carbon content, grindability, and
heating value of charcoal are higher than those of biomass, making it an attractive substitute for
fossil coal [1,2]. Industrial applications such as syngas production [3], co-firing power plant
[4,5], and metallurgical processes [2,6] are of great interest to alter fossil coal by charcoal.
Recently, the replacement of pulverized coal injection to blast furnaces with biomass products
has been suggested [2,6,7]. However, this application requires better characteristic of charcoal,
which is not able to achieve by low-temperature pyrolysis or torrefaction. As indicated in our
review article (not included in this thesis) [6], high-temperature pyrolysis is necessary to fulfill
the requirement of this application. Previous studies have shown that fuel characteristics of
charcoal, i.e., O/C and H/C ratios, heating value, and grindability, are comparable to pulverized
coal when it is produced at pyrolysis temperature of 500 °C or above [8,9,Paper A]. Meanwhile,
charcoal yield becomes lower at high pyrolysis temperature, e.g., charcoal yield decreased from
37 to 28% in pyrolysis of pine chips when temperature increased from 300 to 500 °C [9]. This
trade-off plays important role in the economic performance of charcoal production.
Secondary char formation, i.e., repolymerization of volatiles inside particle structure
during pyrolysis that yields extra charcoal, could be a way to increase the efficiency of charcoal
production while keeping charcoal quality high. This reaction shows significant appearance in
pyrolysis of thick particles, such as woodchip, rather than pyrolysis of biomass powder due to
the longer residence time of volatiles in the pore structure of large particle [10–13]. It has been
reported that the longer residence time of volatiles can be achieved by other techniques, for
example, pyrolysis in a closed system [14,15], high pressure [6,14,16–18], and low purging
1|Page
flow rate [19–21]. In addition, different purging gases—different heat and mass transfer
characteristics—might have consequences on other parameters, namely heating rate and
internal pressure. An idea to promote the secondary char formation by increasing volatiles
concentration inside the pore structure of thick biomass, could be an alternative method to
increase charcoal yield. To ensure the secondary char formation, recycling of bio-oil into
pyrolysis process is proposed in this thesis. The secondary reactions of volatiles from bio-oil
may progress even further at the surface of biomass structure by adsorbing bio-oil at the internal
pore of biomass, as shown in Huang et al [22]. It is important to quantify the effect of various
reaction parameters on synergetic effect in secondary char formation to give guidance for the
design of pyrolysis reactors and processes with bio-oil recycling.
The conversion rate or reactivity of charcoal is one important fuel characteristic that
needs considerations when charcoal is used in combustion, gasification, and metallurgical
processes. In general, reactivity is a parameter used for the design of burners and gasifiers
[23,24]. High reactivity generally reduces the required residence time of charcoal in reactors,
resulting in smaller reactor sizes. However, in order to replace coal in the industrial
applications, charcoal reactivity is preferred to be in the same level as coal to avoid major
process modification. Pyrolysis conditions and type of biomass strongly influence the reactivity
of charcoal as reported in the literature [25–28]. Promotion of secondary char formation may
affect the reactivity of charcoal, which is important to be elaborated. Furthermore, the reactivity
of charcoal produced from thick wood particles, while being a promising method to produce
charcoal for the industry, appears very limited in the literature.
In order to produce charcoal at a large capacity, pyrolysis must be studied at a reactor-
scale. Reactors such as rotary kilns and auger reactors are suitable reactors to convert
woodchips or biomass pellets into charcoal. These continuous reactors can produce charcoal
with homogenous quality [29]. In addition, these reactors have a possibility to control charcoal
yield and its quality simply by changing the operating conditions of the reactors, in contrast
with conventional kilns. As reported in the literature [29,30], rotary kilns can yield 19-38% of
charcoal, while auger reactors yield 11-45% of charcoal. Unfortunately, only a limited research
of these reactors are available for biomass pyrolysis due to the lack of commercial interests for
charcoal powder in the past [29]. However, increasing demand for fine charcoal in the industrial
sector has been attracting attentions to these types of reactor, corresponding to the needs of
scientific development.

2|Page
1.1 Aim and objectives of work
The aim of this research work is to contribute to the development of an efficient charcoal
production process. We intend to deliver a pyrolysis process that can produce charcoal for fossil
coal replacement, while keeping the process efficiency as high as conventional processes, if not
higher.
This work is provided to designate holistic knowledge of charcoal production from thick
particles. The specific objectives of this work are:
• to increase the knowledge of fuel conversion mechanisms in pyrolysis of thick particles;
• to clarify the effects of pyrolysis parameters on the yields (in mass and energy basis)
and the quality of charcoal;
• to suggest promising methods to increase the efficiency of charcoal production;
• to provide optimal pyrolysis conditions favored to increase yields of high-quality
charcoal; and
• to achieve an understanding of pyrolysis in the reactor scale, and know how to
implement desired pyrolysis conditions into the reactors.

Note that the mechanical properties of charcoal and economic analysis are beyond the
scope of this work.

1.2 Research framework and outline of the thesis


In order to accomplish the goals, the research framework has been developed as shown
in Figure 1.1. Three studied panels, namely pyrolysis reaction, process design, and reactor
design, are provided to cover all relevant knowledge of charcoal production. In the beginning,
the fundamental theory of biomass pyrolysis has been studied, including pyrolysis in thick
biomass and the role of pyrolysis parameters. The study led to formulate the methods those
have a potential to increase the efficiency of charcoal production, i.e., bio-oil recycling and CO2
purging. Experimental investigation on the proposed methods was carried out and the result has
been published in Paper A. Charcoals produced from the proposed methods were further
investigated for their properties and reactivity in particular to clarify the effects of the pyrolysis
conditions. This part has given an idea on the feasibility to use charcoal as an alternative
material to coal, which has been described in Paper B. In parallel, a reactor model has been
developed to describe pyrolysis of woodchips in a rotary kiln as published in Paper C. This
model provides an understanding of the design parameters in a rotary kiln pyrolyzer. In

3|Page
addition, the model is expected to be used as a tool to implement desired pyrolysis conditions
in technical reactors.

Figure 1.1. Research framework and correlation among the papers.

This Licentiate thesis composes of 5 chapters, beginning with the introduction in this
chapter. Chapter II will briefly review relevant previous knowledge in the literature. After that,
chapter III will describe the research methods. Next, the main results will be discussed in
chapter IV. Finally, the conclusion will be given in chapter V as well as future plans.

4|Page
Chapter II
Literature review

Biomass mainly composes of carbon, hydrogen, and oxygen with a varied amount of
inorganic constituent depending on biomass feedstocks. In comparison to fossil coal, biomass
has lower carbon content, higher oxygen and volatile contents, lower heating value, and difficult
to grind. Pyrolysis is a well-known method to upgrade biomass into higher-quality solid fuel
namely charcoal. This chapter provides information on charcoal production processes including
the reactions occuring during biomass pyrolysis, the effects of pyrolysis conditions, and
pyrolysis reactors.

2.1 Biomass pyrolysis


Pyrolysis reaction is a thermal decomposition of organic materials in the absence of
oxygen environment. In biomass pyrolysis, thermal degradation of biomass produces a variety
of products, i.e., charcoal, bio-oil (or “tar”), and pyrolysis gas. Figure 2.1 represents a schematic
diagram of biomass pyrolysis, showing possible decomposition pathways of dried biomass
during the reaction [10].

Figure 2.1. Decomposition pathways of biomass pyrolysis (adapted from Neves et al.[10]).
5|Page
Biomass pyrolysis of wood particles can be divided into two stages: primary pyrolysis
and secondary pyrolysis. During primary pyrolysis, thermal decomposition induced by heat
transfer breaks chemical bonds in biomass, resulting in the releases of volatiles and
rearrangement of the carbon matrix of the solid residue [10]. The increase of temperature in this
step leads to increase the thermal stability of charcoal due to higher releases of volatiles [31–
34]. The products from this stage include primary char, bio-oil, and pyrolysis gas.
Primary volatiles can further react in the secondary stage of pyrolysis. Volatiles released
from biomass particles can undergo fragmentation and recombination, resulting in different
fraction of condensable and incondensable gases. In the meantime, large molecules of volatiles
inside pores of the particle can yield extra charcoal due to recombination reactions [35,36]. This
reaction refers to secondary char formation, and the char is generally called secondary charcoal
[10,37]. According to the mechanism of secondary pyrolysis, the reactions are not only
dominated by heat transfer, such as the primary pyrolysis, but also mass transfer of volatiles
within the pore structure of biomass particles. Therefore, the secondary char formation plays a
considerable role in pyrolysis of thick biomass particles, such as woodchips, compared with
pyrolysis of biomass powders due to longer residence time of volatiles in the pore structure [10–
13].
In charcoal production, slow pyrolysis, i.e., pyrolysis under low heating rate, is preferred
to achieve high charcoal yield. There is no clear definition existing in the definition of slow
pyrolysis with respect to heating rates. However, slow pyrolysis normally refers heating rates
up to 100 °C min-1, while fast pyrolysis normally refers heating rates higher than 1000 °C s-1
[29]. Figure 2.2. shows a summary of the effects of temperature, pressure, and heating rate on
charcoal yield from woody biomass in the literature. It shows that low temperature, high
pressure, and low heating rate result in high charcoal yields.

Figure 2.2. Influence of parameters on charcoal yield. (a) Effect of temperature; (b) Effect of
pressure; (c) Effect of heating rate [16,38–47].

6|Page
Among these parameters, temperature is the most influential parameters on charcoal
yields and properties [6,10,48,49]. In general, the increase of temperature gradually decreases
charcoal yield (see Figure 2.2a). Meanwhile, fuel quality of charcoal increases when
temperature increase. Charcoal from high-temperature pyrolysis contains high carbon content,
high heating value, low volatile and ash content [6,16,40,42,50,51]. For example, Figure 2.3
illustrates the effect of temperature on the main elemental content of charcoals. Carbon content
increases sharply at the temperature between 300 and 500 °C to reach about 80%, and then
keeps increasing at higher temperature with less sensitivity.

Figure 2.3. Influence of temperature on elemental compositions. (a) Carbon (b) Hydrogen
(c) Oxygen [16,39–43,47,52,53].

In order to increase charcoal yield while maintaining its quality, secondary char
formation is a promising solution. Several techniques have been reported to extend residence
time of volatiles inside the pore structure of wood particles in particular to yield extra charcoal.
This can be achieved by pyrolysis in a closed system [14,15], high pressure [6,14,16–18], and
low purging gas flow rate [19–21]. The type of purging gas may also influence residence time
of volatiles. As implied by Schonnenbeck et al. [54], the use of CO2 instead of N2 in pyrolysis
may inhibit the volatiles release due to the adsorption of CO2 through the microstructure of

7|Page
solid particle, resulting in higher charcoal yield. In addition, different carrier gases have
different heat and mass transfer characteristics, and they might have consequences on
temperature gradient and internal pressure especially in thick particles. However, the interaction
of these issues is not well elaborated in the literature.
An idea to promote the secondary char formation, by increasing volatiles concentration
inside the pore structure of thick biomass, could be another method to increase charcoal yield.
Hill and co-workers [55] have investigated that the charcoal yield from downstream of the fixed
bed was higher than that from upstream during pyrolysis of Aspen woodchips due to the
deposition of pyrolysis volatile on charcoal. Moreover, they have shown that deposition of
volatiles on biomass char during pyrolysis did not result in a negative impact on microporosity
and adsorption properties of charcoal [55]. Another way to ensure the secondary char formation
of volatiles is the recycling of bio-oil (large-molecule volatiles) into pyrolysis processes.
Secondary reactions of volatiles from bio-oil may progress even further at the surface of
biomass structure by adsorbing bio-oil at the internal pore of biomass, as reported in Huang et
al.[22]. Nevertheless, the usage of thick particles, which are commonly used in industrial
processes, may hinder adsorption of heavy oil as well as sufficient contact between volatiles
and pore surfaces.

2.2 Reactivity of charcoal


The conversion rate, or reactivity, of charcoal is a key parameter to design the capacity
of the reactor in thermochemical conversion processes, i.e., combustion and gasification
[23,24]. According to the agreeable discussion in the literature [25–28], reactivity of charcoal
is varied by three parameters: (i) content and type of inorganic compound, (ii) morphological
structure or availability of active surface area, (iii) content of functional groups and chemical
structure of charcoal. Inorganic matters in charcoal, alkali metals and alkaline earth (AAEM),
are well known as catalysts in oxidation and gasification [56–58] while other inorganic matters
such as Si, P, and Al may hinder the reactivity. Pore size and particle size of charcoal control
diffusion rate and local concentration of involved gases in the reactions [25], while surface area
represent availability of solid reactant. Meanwhile, chemical structures such as content of
carbon edges, defects, and functional groups highly influence the reaction rate of charcoal.
These parameters are strongly influenced by the type of biomass and pyrolysis conditions.
Biomass feedstocks significantly affect the content of inorganic constituents in charcoal
as most inorganic elements stays in charcoal. In woody biomass, Ca is the major ash forming

8|Page
element while it also contains Si, K, P, Mg, and Na [59,60]. In oxidation reactions, alkali metals
such as Na and K are active as catalysts [56]. The oxides and salts of alkali and alkaline earth
metals (AAEM) also have high catalytic activities for biomass gasification with steam [57].
Gasification reactivity of charcoal in CO2 is also reported to be dependent on the concentration
of K and Ca in wood [58]. Furthermore, alkali metals also act as catalysts in the polymerization
of volatile during pyrolysis, which raises charcoal yield and make charcoal less porous [28].
However, it has been reported that inorganic matters release when pyrolysis is conducted at
high heating rate (1000 °C s-1) and high temperature (800 °C) under high sweeping flows [61].
On the other hand, slow heating pyrolysis under temperature up to 700-800 °C have been
reported to have no significant release in AAEM [62–64].
Heating rate and temperature influence not only the release of inorganic matters during
biomass pyrolysis, but also give a dramatic effect on morphological structure and chemical
structure of charcoal. Porosity of charcoal highly depends on the heating rate during pyrolysis.
In pyrolysis under low heating rate, volatiles release gently from the pores in original biomass,
resulting no major change in char morphology compared to original biomass [52]. On contrary,
pyrolysis under high heating rate leads to charcoal with larger cavities and more open structure
due to the fast release of volatiles, overpressure, and coalescence of pores [65,66]. Hence,
charcoals from different heating rates compose of different portions of micropores (dpore < 2
nm), mesopores (2 nm < dpore < 50 nm), and macropores (dpore > 50 nm) [67]. This characteristic
directly indicates the availability of active surface area in charcoal, influencing the conversion
rate of charcoal. According to an analysis of the literature done by Di Blasi [23], surface area
developed by mesopores and macropores is the major contribution in oxidation and gasification,
while micropores barely participate in the reactions. Heating rate also affects elemental
composition and functional groups in charcoal. A larger amount of oxygen functional groups is
normally observed in charcoal produced under higher heating rate [52,65]. This is explained by
shorter residence time of volatile in the particle when pyrolysis takes place at high heating rate
[66,68].
Besides heating rate, high reaction temperature enhances surface area and chemical
structure of charcoal. High pyrolysis temperature leads to the increase in specific surface area
of charcoal [50,65,69,70]. However, the specific surface area of charcoal decreases at
temperature above 800 °C due to the increase of the carbon ordering and pore coalescence
known as thermal annealing [50,65,70]. Pyrolysis at higher temperate also leads to improving
structural order of charcoal, resulting larger aromatic rings cluster [67,71]. By using
characterization techniques such as XRD (X-ray diffraction), NMR (nuclear magnetic

9|Page
resonance), FTIR (Fourier-transform infrared spectroscopy), and Raman spectroscopy, the
conclusion has been drawn from the literature that increasing pyrolysis temperature promotes
the loss of functional groups and the formation of large aromatic rings system in charcoal
[24,67,72–76].
Pyrolysis conditions that would enhance secondary char formation can also potentially
affect the reactivity of charcoal. It has been reported that charcoal produced directly from pure
bio-oil has lower reactivity than charcoal from wood due to the absence of microporosity and
lower ash content in charcoal from bio-oil [77,78]. Anca-Couce and co-workers [79] conducted
pyrolysis experiments that enhanced secondary char formation by varying the initial mass (10
mg-100 g), particle size (0.2 mm-3 cm), and bed type (TGA and fixed-bed reactor). The result
showed significant low reactivity of secondary charcoal in comparison with primary charcoal.
Moreover, the effect was of a similar magnitude to the effect of thermal annealing obtained in
their work [79]. The authors have concluded the mechanism behind this reduction is the
deactivation of the active site due to secondary charcoal deposition or isolation of carbon edges
[79]. However, the effect of secondary char formation by the interaction between bio-oil and
woodchip on charcoal reactivity is not elaborated in the literature. Furthermore, the literature
on the effects of pyrolysis parameters on the reactivity of charcoals from thick particles, which
is a common method in charcoal production, is still limited while it may be significantly
different from charcoals from pulverized particles due to the presence of secondary reactions.

2.3 Pyrolysis reactors


Reactor design is a crucial procedure to designate desired pyrolysis conditions that can
produce charcoal with optimal efficiency. Pyrolysis reactors are usually classified based on
heating mechanisms and desired products, i.e., charcoal or bio-oil. For charcoal production,
slow and intermediate heating configuration are suitable to achieve high charcoal yield [29,80].
Fixed bed reactor such as kilns and retorts have been documented for wood carbonization
already in 18th century [29]. These reactors produce charcoal under slow heating rate in batch
and semi-batch processes [29]. However, the use and development of kilns made of metal or
earth have very little progress in the last century due to its serious disadvantages, e.g.,
inhomogeneous charcoal quality, low efficiency, difficulty in control, high release of pollutions,
and high labor demand [29,81]. Although retorts are the current technology to produce charcoal,
they were designed for the carbonization of wood logs. High capital cost and low efficiency are
the issues for their deployment today [29].

10 | P a g e
Presently, pyrolysis reactors using medium heating rate, so-called “converters”, are
getting higher attention to converting forestry and agricultural biomass in form of chips or
pellets into charcoal and bio-oil [29]. Converters do not require large volumes of circulating
gas or a heating carrier as the tradition reactors [29]. Furthermore, yields and characteristics of
products can be easily varied by operating conditions [30]. Auger reactor and rotary kilns are
popular examples of converters widely used in pyrolysis due to their robustness and mobile
ability. Converters have been lacking commercial interest due to little application of charcoal
powder in the past [29]. Nevertheless, growing interest of fine charcoal for soil amendment and
steel industry is leading to the recent development of converters [29]. Therefore, the demands
on scientific knowledge have been growing on the development of converters such as rotary
kilns and auger reactors for biomass pyrolysis.
Auger reactors, or screw reactors, are successfully used for the production of charcoal
and bio-oil. Figure 2.4 shows a schematic of an auger reactor adapted from commercial Pyreg
process [82]. In pyrolysis via this reactor, biomass is fed to the reactor by a hopper, and a screw
then carries biomass to hot zone of the reactor. Pyrolysis gas goes to a condenser, while charcoal
is obtained at the end of the screw [29]. Several research works have been using auger reactors
to study pyrolysis of various feedstock, such as agricultural materials [83,84], animal manure
[85], waste tires [86,87], sawdust [88–91], and wood particles [92–94]. This shows the
feasibility to use this reactor with different types of feedstock. According to the review paper
provided by Brassard et al. [30], pyrolysis of woody material in a auger reactor gives char yield
in the range of 11-45% (mass basis).

Figure 2.4. Schematic of an auger reactor (adapted from Pyreg process [82]).

11 | P a g e
Rotary kiln reactors have been widely used in cement production, pulp industry (lime
kiln), and mining industry (iron ore pellets production). This reactor has the ability to handle
widespread feedstock and broad particle size distribution. It is a reactor suitable for processing
materials whose physical properties change sharply [95]. The concept of rotary kilns was
basically developed from a tubular reactor, which includes inclination and rotation. Figure 2.5
represents schematic of the rotary kiln used in pyrolysis of woodchips, adapted from the
WoodRoll process, Cortus AB [96]. Woodchips or pellets fed into the reactor are indirectly
heated by flue gas or pyrolysis gas, and charcoal and bio-oil are then produced. It has been used
in many pyrolysis applications, for instance, slow pyrolysis of olive stones [97], sewage sludge
[98], and wood [99]. Rotary kilns can achieve charcoal yield in the range of 19-38% (mass
basis) [29].

Figure 2.5. Schematic of a rotary kiln reactor (adapted from WoodRoll process [96]).

Besides practical studies on these reactors, numerical simulation is a powerful tool to


increase understanding of pyrolysis in these reactors. Granular flows inside auger reactors have
been studied in the literature by using the Discrete Element Method, which can describe the
velocity of particles, mass flow rate, and flow pattern of granular bed [100–102]. Although few
researchers have developed models to describe pyrolysis in auger reactors [103–105], they are
focused on bio-oil production via fast pyrolysis. For pyrolysis in rotary kilns, only few
researchers developed numerical models to describe pyrolysis in this type of reactor. Some
publications have shown simulations of rotary kiln reactor by assuming it as plug flow reactors
[106,107]. Nevertheless, granular flow inside a rotary kiln can be accurately described by
Saeman’s expression [108]. This expression was implemented in the model of maize pyrolysis

12 | P a g e
in a rotary kiln done by Klose and Wiest [109]. Notwithstanding, the numerical model of rotary
kilns for pyrolysis of woodchips is very rare in the literature. By developing such models, the
correlation between reaction conditions and reactor parameters can be elaborated.

2.4 Summary of research questions


According to the literature review explained in this chapter, research questions are
imposed as the following points
• What are possible methods to increase charcoal yield while maintaining its quality?
• What is the mechanism behind pyrolysis of woodchips with bio-oil embedding in the
particles? What are the effects of other conditions on the behavior of this method?
• Do different types of carrier gas affect yield and property of charcoal in pyrolysis of
thick particles? What is the mechanism behind it?
• What are the pyrolysis conditions required to increase the efficiency of charcoal
production?
• How do the pyrolysis conditions affect properties and reactivity of charcoal?
• What are important design parameters in pyrolysis reactor?
• How can we reflect the desired pyrolysis conditions in the design and operation of the
reactors?

13 | P a g e
Chapter III
Research methods

This chapter presents methods to pursue the objectives and answers the research
questions of this thesis. As discussed in the literature review, yield and fuel properties of
charcoal are mainly identified by pyrolysis temperature. In order to increase efficiency of
charcoal production process, enhancing secondary char formation is a promising solution to
increase charcoal yield without disturbing its quality. In this work, bio-oil recycling and CO2
purging are the proposed methods, applying to promote secondary char formation. The idea of
the charcoal production with the proposed methods is schematically represented in Figure 3.1.

Figure 3.1. Charcoal production process proposed in this thesis.

Several research methods have been applied to examine the validity of this process and
to investigate suitable operating conditions. Macro thermogravimetry is the method used to
study pyrolysis of thick wood particles. Gasification reactivity of charcoals produced from
various conditions was determined to investigate the effects of pyrolysis parameters on
reactivity and properties of charcoal. A numerical model was developed based on conservation
equations to describe pyrolysis of woodchips in a rotary kiln reactor, and to study effect of
operating parameters.

15 | P a g e
3.1 Pyrolysis experiment
In order to examine the performance of pyrolysis process with bio-oil recycling, an
experimental investigation has been done at laboratory scale. Bio-oil embedded woodchips
represented material with bio-oil recycling to compare with dried woodchip and bio-oil.
Charcoal yield is the primary measures to indicate the performance of pyrolysis methods in this
work.

3.1.1 Materials and preparation


Debarked chips of Norway spruce and birch were selected as representatives of
softwood and hardwood, respectively. Table 1 shows properties of woodchips. Moisture
contents were measured with MJ33 by METTLER TOLEDO, elemental composition with
EA3000 by Eurovector srl., and higher heating value with C200 by IKA. Dried woodchips were
prepared in an oven at 105 ºC for 24 hours and kept in a desiccator before the experiment. Fast
pyrolysis oil purchased from Fortum’s pyrolysis plant in Joensuu, Finland, was represented as
condensed bio-oil from pyrolysis process in this study. The bio-oil was selected due to its
stability over the long period to avoid the variety in its composition for each experiment. The
composition of bio-oil from actual pyrolysis process is most likely to be different from that in
the current study, but the general conclusions on the conversion behavior of bio-oil should be
applicable for the actual sample. The water content of bio-oil was measured by using WT-KF-
V100 Karl Fischer water content tester. Properties of bio-oil [110] are also shown in Table 3.1.

Table 3.1. Characteristic of raw biomasses.


Characteristics Spruce Birch Bio-oil*
Moisture content (original) wt.% 3.1 (± 0.15) 2.8 (± 0.48) 30.6 (± 0.55)
Moisture content (after drying) wt.% 0.4 (± 0.08) 0.1 (± 0.01) -
Ultimate analysis wt.%, dry
Carbon 49.5 47.1 55.4
Hydrogen 6.1 6.2 6.6
Nitrogen 0.25 0.27 0.14
Oxygen (by difference) 42.8 43.6 37.9
H/C molar ratio - 1.48 1.58 1.43
O/C molar ratio - 0.65 0.69 0.51
HHV MJ/kg 19.7 18.3 23.4
* Ultimate analyses and HHV of bio-oil are on wet basis due to its difficulty to separate water.

16 | P a g e
Bio-oil was embedded in woodchips at room temperature by brushing bio-oil on the
external surface of woodchip. The embedded sample was placed into the reactor 3 minutes after
embedding to ensure bio-oil to disperse into woodchips. In the base case, 20% (m/m) of bio-oil
was embedded at the surface of woodchip, i.e., 20% of bio-oil and 80% of woodchip on mass
basis. Meanwhile, 10% and 25% (m/m) of bio-oil on woodchip were also used for pyrolysis
experiment at selected conditions.

3.1.2 Pyrolysis in a macro-thermogravimetric reactor


Pyrolysis experiments of woodchips were carried out in a macro-thermogravimetric
(macro-TG) reactor. This method gives the accessibility to measure the mass decay of a large
particle during reactions. The detail of the macro-TG reactor is available in Paper A.
The experiment was divided into 2 distinctive temperature histories of the reactor during
the experiments: namely, isothermal condition and slow pyrolysis. Flow rate of the carrier gas,
either N2 or CO2, was 7 L min-1 at standard state for all temperature conditions. The mass of
sample and reactor temperature were recorded every 2 seconds with the precision of 1 mg and
3 °C, respectively.
Prior to the experiments under isothermal conditions, the reactor was heated to the
reaction temperature and purged with carrier gas. Single particle was manually lowered down
into the heating zone typically in 2–3 seconds. Therefore, the sample was rapidly heated by
surrounding gas flow inside the reactor, mimicking the heating rate of co-current configurations
in pyrolysis reactors. When the mass of sample became stable, the sample was moved to the
N2-purged cooling zone before being removed from the reactor. The reactor temperature ranged
between 300 and 700 °C with the precision of ± 3 °C. All the experiments under isothermal
condition had 3 repetitions. Mass yield of charcoal, yc, was calculated from the experimental
data as
𝑚𝑚𝑓𝑓
𝑦𝑦𝑐𝑐 = 𝑚𝑚 × 100% (1)
0

where 𝑚𝑚0 is the initial mass of sample, and 𝑚𝑚𝑓𝑓 is the final mass of sample. The final mass of
sample was defined as an interception point between major degradation line and post
degradation line to decrease random errors at the final stage of pyrolysis. The details are
available in Paper A.
Under the slow pyrolysis conditions, the sample was placed into the furnace at room
temperature with carrier gas flowing through the reactor. Then, the reactor was heated to the
reaction temperature at the heating rate of 3 °C min-1 with the precision of ± 3 °C. Ten seconds

17 | P a g e
after the reactor temperature reached the desired temperature, the sample was moved to the N2-
purged cooling zone and removed from the reactor.
Additional experiments of wood powder were carried out in PerkinElmer TGA 8000TM
Thermogravimetric Analyzer (TGA) to examine the pyrolysis behavior with negligible mass
diffusion resistance. Woodchips from the same origin were cut, and sawdust was collected as
wood powder. TGA experiments of wood powder, 0.9-1.1 mg, were conducted under the same
temperature profile as macro-TG experiment under slow pyrolysis condition (3 °C min-1) with
a carrier gas flow rate of 20 mL min-1 purged around the sample crucible.

3.2 Characterization of charcoal


Properties of charcoals, produced from the methods described in the previous section,
were investigated including elemental compositions, heating values, porosity, and chemical
structure. These properties were selected to describe the effects of pyrolysis conditions and the
methods on the quality of charcoals. Gasification reactivity of charcoals was also measured to
investigate the influences of pyrolysis conditions, and to give an idea about the possibility to
replace coal.

3.2.1 Elemental composition and heating values


Ultimate analysis of raw woodchips and char samples were carried out with EA3000, a
CHNS-O elemental analyzer from Eurovector srl. Higher heating value (HHV) of raw
woodchips and chars were measured by an oxygen bomb calorimeter, model C 200 from IKA.

3.2.2 N2 adsorption
The specific surface area, specific pore volume, and pore-size distribution of charcoal
were determined using an N2 adsorption method with a Micromeritics ASAP 2020 analyzer.
Prior to the measurements, about 200 mg of sample were degassed at low pressures (2 µm Hg)
and high temperatures (140 oC) for 180 minutes. Adsorption isotherms were obtained by
immersing sample tubes in liquid nitrogen (-197 oC) in order to obtain isothermal conditions.
Nitrogen was added to the samples in small steps and the resulting isotherms were obtained.
Specific surface areas and specific pore volume were calculated from adsorption isotherms
according to the Brunauer–Emmett–Teller (BET) method [111]. Pore-size distributions were
calculated using the Barrett-Joyner-Halenda (BJH) algorithm [112] and the Density Functional
Theory (DFT) function [113]. This method represents pore size into 3 different ranges, namely
micropore (< 2nm), mesopore (2-50 nm), and macropore (> 50 nm).

18 | P a g e
3.2.3 Raman spectroscopy
Raman spectroscopy was used to analyze aromatic rings clusters and chemical structure
of the charcoals. Raman spectra were collected using an inverted microscope (IX71, Olympus,
Japan) coupled to a spectrometer, Shamrock 303i (Andor Technology, Ireland). Spectra were
collected from 5 different spots with 120 seconds of exposure time. Detail of experimental setup
and processing data are available in Paper B.
The Raman spectra showed two main overlapping peaks with maximum intensities
around 1350 cm-1 and 1590 cm-1. These two peaks are commonly referred to the D and G bands,
respectively [114–116]. In the case of disordered and amorphous carbon such as biomass or
charcoal, the D and G bands are no longer sufficient to express the complexity of these materials
[117–121]. Therefore, interpretation of the Raman spectra in this work has been done by
deconvolution of Gaussian bands adapted from Smith et al. [72]. Detail of bands assignment
can be found in Paper B, while Figure 3.2 shows an example of deconvolution peaks.

Figure 3.2. An example of Raman spectra and deconvolution (spruce char from isothermal
condition at 700 °C).

The deconvolution method gives a sufficient idea of chemical structure contain in


charcoals. Nevertheless, not all chemical structure can be illustrated by the method due to high
amorphous carbon in the samples. The GT, summation of GG and GL, represents G band, which
mainly indicates small aromatic structures with 3-5 fused rings [74]. The DT, summation of Ds
and D, represents D band indicating large aromatic clusters, which consist of not less than 6
fused rings [74]. Uncertainty has remained in A band, A1 and A2, within the literature. This
valley region generally refers to irregular structures, which is the overlapping position for
cyclopentane, point defects, and heteroatoms [72,117,121]. Interpretation of Raman spectra was

19 | P a g e
done by the analyses of peak area ratios, by dividing the area of each peak with that of GT band.
The peak area ratio I(DT)/I(GT) has been introduced to compare the bonding structure of
amorphous carbon [72,114,122,123]. The I(DT)/I(GT) ratio of charcoal in this thesis increased
at high pyrolysis temperature, meaning more stable structure and large aromatic ring cluster as
widely reported in the literature [72,117]. Interpretations of peaks within the valley region, i.e.,
A1 and A2, showed no consistency in this thesis or in the literature [72,117,121]. Therefore,
interpretations of these bands are not considered in this thesis.

3.2.4 Gasification reactivity


The gasification reactivity of charcoals was measured by thermogravimetric analysis
with a TA instrument Q5000IR. Around 1 mg of powder sample was loaded and spread at the
bottom of a platinum pan. The sample and the furnace were purged with the mixture of N2 and
CO2 for 30 minutes with a volumetric ratio of 80% and 20%, respectively. Then, the sample
was heated up to 850 °C at a heating rate of 500 °C min-1 to mimic the condition in the industrial
processes. The sample was held at 850 °C and the change in sample mass was measured for 30
minutes. All charcoal samples were analyzed with 2 repetitions. The thermogravimetric curves
displayed mass loss due to both devolatilization and gasification, as shown in the supporting
information of Paper B. In order to determine gasification reactivity, initial mass (m0) of
charcoal was defined at the end of devolatilization (the detail is provided in the supporting
information of Paper B). Char conversion (X) was calculated by
𝑚𝑚0 − 𝑚𝑚
𝑋𝑋 = (2)
𝑚𝑚0
The reactivity of charcoals in this work is represented by initial rate constant (kini). The
data were fitted to the random pore model to determine the rate constant at the conversion up
to the minimum conversion appeared in the experimental data (Xmin=0.4). The integral form of
the random pore model [124] is read as:
2
� � ∙ ��1 − 𝜓𝜓 ∙ ln(1 − 𝑋𝑋) − 1� = 𝑘𝑘𝑖𝑖𝑖𝑖𝑖𝑖 ∙ 𝑡𝑡 (3)
𝜓𝜓
The structure parameter, 𝜓𝜓, was estimated from the experimental data under all
conditions using the reaction time at X = 0.4 by the following expression.
��1 − 𝜓𝜓 ∙ ln(1 − 𝑋𝑋) − 1� 𝑡𝑡
= (4)
��1 − 𝜓𝜓 ∙ ln(1 − 0.4) − 1� 𝑡𝑡𝑋𝑋=0.4

Hence, the initial rate constant of gasification, kini, can be determined. Examples of data
fitting and further details are available in Paper B.

20 | P a g e
3.3 Modeling of pyrolysis reactor
A numerical model was developed to describe pyrolysis in rotary kiln reactors. The
model is based on a set of conservation equations for mass and energy, coupled with submodels
for pyrolysis reaction, heat transfer, and granular flow inside the kiln. Figure 3.3 represents
schematic of the rotary kiln considered in the model.

Figure 3.3. Schematic of the rotary kiln pyrolyzer.

We consider a rotary kiln of length L and inner diameter D rotating with an angular
velocity ω. Inside the kiln, a moving bed of granular solid flows with a volumetric flow rate Q
in axial direction, whereas Q refers to the bulk volume of granular solid per unit time. A small
stream of nitrogen is fed at the entrance of the kiln. The kiln is indirectly heated from outside
the wall by flue gas flowing in axial direction along the tube wall, co-currently to the biomass
flow.
The model considers two separated domains, namely a solid bed domain that occupies
the lower part of the rotating kiln, and a gas phase domain that occupies the upper part of the
rotating drum. Product gas generated in the solid bed domain is treated to transfer to the gas
phase domain as a source term. Both domains are assumed to be radially well mixed and to
describe a plug flow in axial direction (no axial dispersion) [125], with the cross-section area
of both domains evolving along the kiln axis (see Figure 3.2b). Heat and mass transfer inside
the wood chips are assumed to be fast with respect to the pyrolysis reaction which allows for
treating the particles as homogeneous and lumping the intra-particle heat transfer into the
overall heat transfer from the kiln wall to the granular bed [109,126]. This assumption is
reasonable considering the moderate heating rate experienced by the particles in the kiln. The
model considers several biomass species and pyrolysis products that exist either in the solid bed
domain or in the gas phase domain. Reactions are assumed to take place solely in the solid bed
domain, i.e., once a species that formed in the solid phase domain transfers to the gas phase
domain, it will not undergo any further reactions. Secondary reactions of volatile products occur

21 | P a g e
only for species trapped in the solid bed domain.
Within this framework, the mass balance of species j in the solid bed domain reads as:
𝜕𝜕 𝜕𝜕𝐹𝐹𝑗𝑗
(1 − 𝜀𝜀) �𝜌𝜌𝑗𝑗 𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 � + = (1 − 𝜀𝜀)𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 𝑟𝑟𝑗𝑗 (5)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
where 𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 is cross-sectional area of the solid bed, 𝜌𝜌𝑗𝑗 and 𝐹𝐹𝑗𝑗 = (1 − 𝜀𝜀)𝜌𝜌𝑗𝑗 𝑄𝑄 are the density and
the density and the mass flow rate of species 𝑗𝑗, respectively, and 𝑟𝑟𝑗𝑗 is rate of formation of
species 𝑗𝑗. Further symbols are explained in the nomenclature. The mass balance for species 𝑘𝑘
in the gas phase domain read as:
𝜕𝜕 𝜕𝜕𝐹𝐹𝑘𝑘
�𝜌𝜌 𝑆𝑆 � + = (1 − 𝜀𝜀)𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 𝑟𝑟𝑘𝑘 (6)
𝜕𝜕𝜕𝜕 𝑘𝑘 𝑔𝑔𝑔𝑔𝑔𝑔 𝜕𝜕𝜕𝜕
where 𝑆𝑆𝑔𝑔𝑔𝑔𝑔𝑔 is the cross-sectional area of the gas phase and 𝐹𝐹𝑘𝑘 = 𝑆𝑆𝑔𝑔𝑔𝑔𝑔𝑔 𝑢𝑢𝑔𝑔 𝜌𝜌𝑘𝑘 is the mass flow rate
of species k. The right-hand side of Eq. (6) describes the formation of species k that takes place
in the solid bed domain. The transfer rate from the solid bed to the gas phase, which is relevant
for tar, is incorporated into the rate formation 𝑟𝑟𝑘𝑘 as described in Paper C.
Energy balances for the two domains are expressed in a similar form. The energy
balance for the solid bed domain reads as:
𝜕𝜕𝑇𝑇𝑏𝑏 𝜕𝜕𝑇𝑇
�����
(1 − 𝜀𝜀)𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 (𝜌𝜌𝐶𝐶𝑝𝑝 )𝑏𝑏
�����𝑝𝑝 )𝑏𝑏 𝑏𝑏 = (1 − 𝜀𝜀)𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 𝐺𝐺 + 𝑞𝑞𝑏𝑏𝑏𝑏𝑏𝑏
+ (𝐹𝐹𝐶𝐶 (7)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
where Tb is solid bed temperature, 𝐺𝐺 = ∑𝑛𝑛𝑖𝑖=1 𝑟𝑟𝑖𝑖 (−∆ℎ𝑖𝑖 ) is the heat production due to the
reactions, with ∆ℎ𝑖𝑖 the reaction enthalpy of reaction i (defined in the conventional way with
∆ℎ𝑖𝑖 > 0 for an endothermic reaction). The terms 𝑞𝑞𝑏𝑏𝑏𝑏𝑏𝑏 composes of the heat flux from the wall
to bed (𝑞𝑞𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤−𝑏𝑏𝑏𝑏𝑏𝑏 ) and the gas phase to the bed (𝑞𝑞𝑔𝑔𝑔𝑔𝑔𝑔−𝑏𝑏𝑏𝑏𝑏𝑏 ).
The energy balance of the gas phase domain reads as:
𝑛𝑛𝑘𝑘
𝜕𝜕𝑇𝑇𝑔𝑔 𝜕𝜕𝑇𝑇𝑔𝑔 𝑇𝑇𝑔𝑔
�����
𝑆𝑆𝑔𝑔𝑔𝑔𝑔𝑔 (𝜌𝜌𝐶𝐶 )
𝑝𝑝 𝑔𝑔 + �����
(𝐹𝐹𝐶𝐶 )
𝑝𝑝 𝑔𝑔 = −(1 − 𝜀𝜀)𝑆𝑆𝑏𝑏𝑏𝑏𝑏𝑏 �� 𝑟𝑟𝑘𝑘 � 𝐶𝐶𝑝𝑝,𝑘𝑘 𝑑𝑑𝑑𝑑� + 𝑞𝑞𝑔𝑔𝑔𝑔𝑔𝑔 (8)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇𝑏𝑏
𝑘𝑘=1

where Tg is gas phase temperature, 𝑞𝑞𝑏𝑏𝑏𝑏𝑏𝑏 composes of the heat flux from the wall to gas
(𝑞𝑞𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤−𝑔𝑔𝑔𝑔𝑔𝑔 ) and the gas phase to the bed (𝑞𝑞𝑔𝑔𝑔𝑔𝑔𝑔−𝑏𝑏𝑏𝑏𝑏𝑏 ). The first term on the right-hand side of Eq.
(8) describes the heat that is needed to bring species k from the bed temperature, i.e., the
temperature where it is formed, to the gas phase temperature.
The temperature of the kiln wall is controlled by the flue gas flowing along the wall.
Assuming negligible heat capacity of the wall material and fast heat transfer from the flue gas
to the wall, we can write the following energy balance for the wall temperature Tw:
𝜕𝜕𝑇𝑇𝑤𝑤 𝜕𝜕𝑇𝑇𝑤𝑤
(𝑆𝑆𝑆𝑆𝐶𝐶𝑝𝑝 )𝑓𝑓𝑓𝑓 + (𝐹𝐹𝐶𝐶𝑝𝑝 )𝑓𝑓𝑓𝑓 = −𝑞𝑞𝑘𝑘𝑘𝑘𝑘𝑘𝑘𝑘 − 𝑞𝑞𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙 (9)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
22 | P a g e
where “fg” stands for flue gas, 𝑞𝑞𝑘𝑘𝑘𝑘𝑘𝑘𝑘𝑘 is the heat fluxes from the wall to the bed and the gas phase,
and 𝑞𝑞𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙 is the heat fluxes from the wall to the surroundings.
The five balance equations given by Eqs. (5)-(9) form the basis of the rotary kiln model
developed in this work. Assuming steady state allows for reducing them to a set of ordinary
differential equations whose solution comprises the density profiles of the different biomass
species and the temperature profile in the different domains of the kiln. The governing equations
are coupled with submodels for pyrolysis reactions, heat transfer, and granular flow inside the
kiln. The reaction adopted in this model is based on the two-step mechanism proposed by Park
et al. [127]. The reaction assumes that dry wood decomposes in a set of primary reactions into
gas, tar, and an intermediate solid that subsequently decomposes exothermically into char. The
heat transfer model accounts conduction, convection, and radiation, which was derived from
the balance between three components namely solid bed, gas, and reactor wall. The evolution
of granular bed inside the rotary kiln pyrolyzer is described by the well-known Saeman model
[108]. Detail information of the submodels and boundary conditions are available in Paper C.
Calculations were performed in Matlab using standard ODE-solvers for stiff systems.

23 | P a g e
Chapter IV
Results and discussion

This chapter discusses the results obtained from this research. Only main outcomes are
provided in this chapter, more results and discussion are available in Paper A, B, and C.

4.1 Effect of reaction parameters


As discussed briefly in the literature review, temperature plays an important role on
yields and property of charcoal. Secondary char formation in pyrolysis of thick particles could
be a key to increase charcoal yield without negative impacts on the property of charcoal. Bio-
oil recycling and purging with CO2 are suggested in this thesis to increase the secondary char
formation. This section firstly describes mechanisms behind pyrolysis of thick particles with
the proposed techniques. Then, the effects of these reaction parameters on the properties of
charcoal will be explained.

4.1.1 Increase of charcoal yield by promoting secondary char formation

Bio-oil recycling
As explored in Paper A, bio-oil recycling has the potential to increase charcoal yield.
The first supporting reason is additional bio-oil in embedded samples simply added charcoal
mass from pyrolysis of bio-oil origin. Therefore, mass and energy of bio-oil recycle stream are
recovered based on process engineering aspect.
Another and important reason is the synergetic effect due to the physical contact
between bio-oil and woodchips. Volatile compounds from bio-oil can be shifted toward
charcoal by contact with the solid surface of woodchip/char or by high concentration of volatiles
inside the internal pore of the particle. To examine the degree of synergetic effect, the
experimental data of bio-oil embedded biomass was compared with the expected charcoal yield

25 | P a g e
from interpolation between pure wood chips and bio-oil, as shown in Figure 4.1. The char yield
in the figure was calculated by dividing the mass of charcoal with total mass of reactants (bio-
oil + wood particle). Higher experimental values than the interpolation values mean the
presence of synergetic effects. At low-temperature range (300 and 340 °C), experimental results
clearly showed higher char yield than interpolation data. This synergetic effect was apparent
regardless of the amount of bio-oil embedding. The synergetic effect between bio-oil and
woodchip seems to have diminished at higher pyrolysis temperature (400 and 500 °C) as the
experimental and interpolation data showed no significant differences (statistical analysis is
available in Paper A).

Figure 4.1. Examination of synergetic effects between woodchips (spruce) and bio-oil by
comparing char yield of bio-oil embedded woodchips with interpolations between char yields
of bio-oil and wood chips. The error bars are standard deviation obtained from 3 repetitions
for each experimental point. The result of birch is available in Paper A.

26 | P a g e
The most likely reason of diminishing synergetic effect at high pyrolysis temperature is
due to the loss of physical contact between volatiles and solid surface. At higher temperature in
isothermal conditions, larger temperature differences between the reactor and the particle
surface gives higher heating rate and causes more intensive devolatilization. The latter leads to
accumulation of volatiles inside the particle, increasing internal pressure and mechanical stress.
This effect was observed by fragmentation and cracking in charcoal after pyrolysis at high
temperature as shown in Figure 4.2. These cavities or fragmentation in charcoal allowed
volatiles generated during pyrolysis to be released from the particle easier, hence limiting the
progress of secondary char formation.

Figure 4.2. Differences in spruce char morphology at different pyrolysis temperature under
isothermal conditions. The result of birch is available in Paper A.

To isolate the effect of fragmentation and cavities caused by high heating rate,
experiments were carried out at slow heating rate. Figure 4.3 shows the results from room
temperature to 500 °C at the heating rate of 3 °C min-1 (reproducibility of the TG curves is
illustrated in Paper A). The slow pyrolysis results showed well-known pyrolysis sequences. Up
to about 150 °C, moisture in woodchip (and part of bio-oil) evaporated, and then major
degradation took place. Mostly hemicellulose fraction degraded at a temperature range between
150 to 280 °C, which can be observed as the shoulder of the derivative thermogravimetric
(DTG) curves. Then, degradation of cellulose took place up to the temperature of ca. 340 °C.
The final region with slow degradation rate corresponds to the decomposition of lignin.

27 | P a g e
Figure 4.3. Thermogravimetric curve of woodchips in macro-TG at slow heating rate (3 °C
min-1) under N2 flow. The result of birch is available in Paper A.

When considering the bio-oil embedded woodchips, the experimental results showed
higher charcoal yield than interpolation results along with all the temperature range.
Furthermore, charcoal obtained after slow pyrolysis did not appear cavities or breakage. The
results revealed that the charcoal yield of bio-oil embedded woodchips was higher than that
when bio-oil and woodchips were pyrolyzed independently. Hence, keeping enough contact
between woodchips and bio-oil is important to ensure the profound effect of bio-oil recycling.

Purging with CO2


Effect of purging gases was examined under slow pyrolysis to isolate the transport
effects from intrinsic chemical effects (the results under isothermal conditions are available in
Paper A). Thermogravimetric analysis was carried out at heating rate of 3 °C min-1 in CO2 and
N2 using both woodchip (with macro-TG) and powder (with TGA), as shown in Figure 4.4. The
effect of heat transfer and particle breakage was absent for slow pyrolysis of woodchips, while
heat and mass transfer as well as secondary pyrolysis had minimal effect on slow pyrolysis of
wood powder.

28 | P a g e
Figure 4.4. Thermogravimetric curve of woodchips in macro-TG and wood powder in TGA
as a comparison between N2 and CO2 flows. The result of birch is available in Paper A.

For slow pyrolysis of woodchip, the pyrolysis reactions can be divided into three regions
with respect to the similarity and difference between pyrolysis under CO2 and N2. At the
temperature below 400 °C, which corresponds to the degradation of cellulose and
hemicellulose, CO2 and N2 showed no visible difference. However, the pyrolysis of woodchip
in CO2 showed higher charcoal yield than in N2 at the temperature between 400 °C to 700 °C,
corresponding to the degradation of lignin. At T=700 °C, charcoal yield of spruce chip in CO2
pyrolysis was 27.6%, compared with 23.0% in N2 pyrolysis. At the temperature higher than 700
°C, residual mass in CO2 atmosphere decreased sharply due to gasification reaction. In pyrolysis
of wood powder using TGA, the occurrence of secondary reactions was minimized so that only
primary decomposition takes place [127,128]. Wood powder showed no significant difference
in mass degradation path between CO2 and N2 at the temperature below 700 °C, confirming the
absence of the effect of CO2 on intrinsic pyrolysis reaction, often discussed in the literature
[129,130]. Charcoal yield of wood powder was much lower than woodchips due to the minimal
occurrences of secondary char formation in pyrolysis of wood powder.
For the pyrolysis under slow heating, heat transfer effect is negligible. Therefore, high
yield of charcoal from woodchips under CO2 atmosphere at T=400-700 °C is the effect of mass
diffusion, and subsequent change in reaction pathways. The absence of the CO2 effects in the
TGA results with wood powder sample (minimal mass transfer limitation) indicates that it was
low mass diffusivity in CO2 that have influenced the pyrolysis reactions (calculation of mass
diffusivity is available in Paper A). Furthermore, the increase in mass yield can be explained

29 | P a g e
either by the promotion of secondary reactions, the change in the reaction pathways or merely
adsorption of CO2 at the active sites of the char matrix. If adsorption of CO2 is the case, the CO2
exist as a condensed phase and could increase the charcoal yield. At the same time, CO2 also
affects the elemental composition of charcoal due to its high oxygen content. However,
adsorption of CO2 is unlikely to be the reason in this study since carbon content of charcoal was
not affected (Figure 4.4). Promotion of secondary reactions is plausible because low diffusivity
of volatile through CO2 means high internal pressure or long residence time of volatiles. Both
effects are favorable for the secondary char formation. On the other hand, it is still possible that
the reaction pathways of primary pyrolysis were modified. The previous study showed that even
short contact with steam can increase the size of aromatic ring cluster in charcoal [131].

4.1.2 Charcoal properties affected by promoting secondary char formation

Unlike charcoal yield, bio-oil recycling does not affect the elemental composition of
charcoal significantly. Figure 4.5 represents van Krevelen diagram as a comparison of charcoal
from original woodchips with bio-oil embedded woodchips. Elemental composition of
charcoals was close to that of pulverized coal, but it was slightly rich in oxygen and lean in
hydrogen than pulverized coal. Bio-oil recycling showed negligible influence on elemental
composition.

Figure 4.5. Van Krevelen diagram of char from dried woodchip and bio-oil embedded
woodchips under isothermal conditions. The result of birch is available in Paper A.

30 | P a g e
Figure 4.6 represents higher heating value (HHV) of charcoal. Bio-oil recycling had no
significant effect on HHV of charcoal (p-values by 2-way ANOVA was 0.76 *), but it increased
with reaction temperature. HHV reached stable at the temperature above 500 °C, and the value
was close to that of pulverized coal reported by Wang et al. (34.4 MJ kg-1) [7].

Figure 4.6. Higher heating value of char with and without bio-oil embedding under isothermal
conditions. The error bars are standard deviation obtained from 3 repetitions for each experimental
point. The result of birch is available in Paper A.

Based on the measurement of charcoal gasification reactivity, Figure 4.7 shows rate
constants (kini) indicated the reactivity of charcoals obtained from dried woodchip and bio-oil
embedded woodchips. Charcoals obtained from bio-oil embedded woodchips showed
significantly higher reactivity than charcoals from dried woodchips (p-values are shown in the
figure). The rate constants showed around 1.4 times higher than woodchips when charcoal
produced from bio-oil embedding woodchips.

*
p-values less than 0.05 were considered significant with 95% confidence.

31 | P a g e
Figure 4.7. Rate constant in the comparison between charcoals from bio-oil embedded
woodchips and dried woodchips. Condition 1: isothermal condition, 300 °C, N2; condition: is
isothermal condition, 700 °C, N2; condition 3: is slow heating, 700 °C, N2. The error bars are
standard deviation obtained from 2 repetitions for each experimental point.

Chemical structure of charcoal interpreted by Raman spectra ratios, i.e., (I(DT)/I(GT)),


are represented in Figure 4.8. The results between charcoals from dried woodchips and bio-oil
embedded woodchips did not show significant differences (p-values are 0.16-0.20). I(DT)/I(GT)
ratios of charcoals from bio-oil embedded materials at 700 °C (condition 2 and 3 in Figure 4.8)
were slightly lower than dried woodchips, meaning slightly smaller size of aromatic structure
cluster in charcoals from bio-oil embedded woodchips. The result showed the opposite effect
in the charcoals produced at 300 °C. According to the figure, chemical structure of charcoals
from bio-oil embedded woodchips was likely affected by charcoal from pure bio-oil. Charcoals
produced from pure bio-oil at pyrolysis temperature of 700 °C under both isothermal condition
and slow heating rate showed significant lower in I(DT)/I(GT) ratios than charcoals from
woodchips (p-values are 0.0002 and 0.0007, respectively), which means it contains smaller
aromatic cluster than charcoals from woodchips. In contrast, pyrolysis at 300 °C showed no
significant difference in I(DT)/I(GT) ratio among the materials, i.e., dried woodchips, bio-oil
embedded woodchips, and bio-oil. Therefore, this may imply that the growth of aromatic ring
cluster is less significant for charcoal from bio-oil, which may increase the reactivity of
charcoals from bio-oil embedded woodchips.

32 | P a g e
Figure 4.8. I(DT)/I(GT) ratios of charcoals from dried woodchip, bio-oil embedded
woodchips, and pure bio-oil. Condition 1: isothermal condition, 300 °C, N2; condition 2:
isothermal condition, 700 °C, N2; condition 3: slow heating, 700 °C, N2. The error bars are
standard deviation obtained from 5 repetitions for each experimental point.

In comparison of carrier gases explained in the previous section, charcoal yield from
CO2 pyrolysis (27.3%) was significantly higher than N2 pyrolysis (23.0%), but charcoals from
two purging gases contain equivalent in carbon content at 91.4% and 92.1%, respectively. The
reactivity was examined for charcoals from both carrier gases. Figure 4.9 represents the rate
constant of charcoals produced under N2 and CO2 flows. These charcoals were obtained at the
pyrolysis temperature of 700 °C under slow heating rate of 3 °C min-1. The rate constant of
charcoal obtained from pyrolysis under CO2 seemed slightly higher than charcoal from N2, but
the difference was not significant (p-value is 0.38).

Figure 4.9. Rate constant of charcoals produced from different carrier gases. The error bars
are standard deviation obtained from 2 repetitions for each experimental point.

33 | P a g e
The specific pore volumes of charcoals from the two carrier gases are displayed in
Figure 4.10. The total specific pore volume of charcoal obtained from CO2 pyrolysis (0.158 cm3
g-1) was slightly higher than charcoal from N2 pyrolysis (0.151 cm3 g-1). Furthermore, charcoal
from CO2 pyrolysis contained higher specific pore volume in every pore-sized category, i.e.,
micropores (dpore < 2 nm), mesopores (2 nm < dpore < 50 nm), and macropores (dpore > 50 nm).
According to the literature [23], mesopores and macropores are the major contributions in
gasification reaction. Therefore, higher specific mesopore and macro volumes in charcoal
obtained from CO2 pyrolysis may result in higher char reactivity than charcoal from N2
pyrolysis.

Figure 4.10. Pore-size distribution of charcoals from pyrolysis with different carrier gases.

According to the interpretation of the results from Raman spectroscopy, Figure 4.11
shows a comparison of I(DT)/I(GT) ratios between charcoals obtained from pyrolysis under N2
and CO2. Charcoal obtained from CO2 pyrolysis showed slightly lower I(DT)/I(GT) ratio than
charcoal from N2 pyrolysis, but the difference was not significant (p-value is 0.20). However,
the result can imply that charcoal from CO2 pyrolysis likely contains smaller aromatic ring
structure than charcoal from N2 pyrolysis. This behavior may be due to the lower diffusivity of

34 | P a g e
volatiles during the pyrolysis of CO2, resulting in more secondary charcoal with smaller
aromatic structure. Weaker structure, i.e., smaller aromatic cluster, in charcoal from CO2 can
also support the result of higher reactivity in comparison to charcoal obtained from N2 pyrolysis.

Figure 4.11. I(DT)/I(GT) ratios of charcoals from different reaction gases. The error bars
standard deviation obtained from 5 repetitions for each experimental point.

4.1.3 Summary of the effects of pyrolysis parameters


According to the literature review and works have been done in Paper A and B, the
general conclusion on the effects of pyrolysis parameters can be drawn. The pyrolysis
parameters studied in this thesis includes temperature, heating rate, biomass species, bio-oil
recycling, and purging with CO2. Table 4.1 represents an indication of the pyrolysis parameters
on charcoal yield, carbon content, porosity, chemical structure, and reactivity of charcoal. The
effects of parameters on the variables are indicated by the magnitude of impact, i.e., high,
medium, and low effects, and the correlation between the parameters and the variables, i.e., +
and – signs indicating positive and negative correlation, respectively.
Table 4.1. An indication of pyrolysis conditions on the yield and properties of charcoal.
Size of
Pyrolysis condition Char yield Carbon content Porosity aromatic Reactivity
cluster
Temperature High effect High effect High effect High effect High effect
(–) (+) (+) (+) (–)
Heating rate Medium effect Low effect High effect Medium effect High effect
(–) (–) (–) (+)
Inorganic content Medium effect Low effect Low effect Low effect Medium effect
(+) (+)
Bio-oil recycling Medium effect Low effect Low effect Low effect Medium effect
(+) (+)
CO2 Medium effect Low effect Low effect Low effect Low effect
(+)
Note; + refers to positive correlations between parameter and variable
– refers to negative correlations between parameter and variable.

35 | P a g e
As mentioned in the literature review and investigation done in Paper A, temperature
obviously gives major impacts on all the variables. Increasing temperature dramatically
decreases charcoal yield, but significantly increases carbon content of charcoal. Based on the
results in Paper B, charcoals obtained at higher pyrolysis temperatures contain higher porosity
and larger aromatic structure. These properties affect the reactivity of charcoals. However,
among porous charcoals obtained in this work, aromatic structure plays more important role in
charcoal reactivity. Therefore, charcoals produced from higher temperatures generally have
lower reactivity.
The other parameters considerably affect charcoal yield, but do not give a major impact
on carbon content. This implies that parameters such as heating rate, type of biomass, bio-oil
recycling, and CO2 can be considered to optimize carbon yield and efficiency of charcoal
production process. Pyrolysis of woodchips under slow heating rate gives higher charcoal yield
than at high heating rate. While carbon content of charcoal does not change by heating rate, the
reactivity of charcoal decreased significantly by lowering the heating rate due to dramatic
changes in porosity and chemical structure of charcoal (the detail is available in Paper B).
Different type of biomass also gives different level of charcoal yield due to various amount of
cellulose, hemicellulose, and lignin. Inorganic constituent in biomass significantly increases
reactivity of charcoal as reported in the literature review and Paper B. Bio-oil recycling and
purging with CO2 increase charcoal yield, but do not give high impact on other properties.
Although the bio-oil recycling produce charcoal with higher reactivity than normal conditions,
the effect is only minor in comparison with the other parameters.

4.2 The high-efficiency charcoal production process


According to the effects of pyrolysis parameters described previously, the efficiency of
pyrolysis processes is discussed in this section.

4.2.1 Comparisons of yields


To measure the performance of pyrolysis methods, process charcoal yield, i.e., mass
ratio of final charcoal to dried woodchip, was defined. Figure 4.12 shows the process charcoal
yields on mass basis with and without bio-oil recycling, which was prepared by embedding 20%
(mass/mass) of bio-oil on woodchip. With bio-oil recycling, charcoal yield increased for all the
reaction temperature. In pyrolysis of spruce, char yield increased by 5.4% with bio-oil recycling
on average.

36 | P a g e
Figure 4.12. Process charcoal yield of bio-oil embedded woodchips under isothermal
conditions in comparison with those of woodchips and bio-oil. The error bars are standard
deviation obtained from 3 repetitions for each experimental point. The result of birch is
available in Paper A.

Figure 4.13 shows a comparison of carbon yields among all the pyrolysis conditions
examined in this study (calculation can be found in Paper A). As well discussed in the literature,
higher temperature ended up lower carbon recovery although charcoal had higher quality
[6,16,40,42,50]. However, other reaction parameters also affected carbon yields, where the
magnitude of the combined effect can be equivalent to that of reaction temperature by more
than 300 °C. Bio-oil recycling and the use of CO2 as carrier gas instead of N2 are recommended
when designing pyrolysis processes. However, it is important to apply these modifications with
thick particles at slow heating rate in order to ensure the effect by taking advantages of good
contact of volatile and particle surface. Slow pyrolysis of a thick particle, in contrast with
isothermal conditions, produces charcoal without major change in fiber structure by cavities
and breakages. This behavior is more favorable for secondary char formation and contact
between internal solid surface and volatiles, resulting in higher charcoal yield. Moreover,
particles under slow heating experience uniform internal temperature and long reaction time.
Therefore, charcoal from slow pyrolysis contains higher carbon content (lower functional
groups) than that from isothermal conditions. As a consequence, the benefit from bio-oil
recycling and the use of CO2 as carrier gas is more apparent under slow pyrolysis.

37 | P a g e
Figure 4.13. Carbon yield in char from various pyrolysis conditions. The result of birch is
available in Paper A.

All the measures combined, carbon yield from slow pyrolysis under CO2 flow with bio-
oil recycling at 700 °C showed as high carbon yield at that of isothermal conditions under N2
flow at 340 °C. Meanwhile, the carbon content of charcoal was significantly higher; the carbon
content in charcoal was ca. 90% instead of ca. 75%. When comparing the same pyrolysis
temperature, on the other hand, charcoal quality remained similar value while the yield
increased from 16.2% to 26.7% for spruce.
Figure 4.14 shows the expected energy balance of pyrolysis units in different
configurations. We calculated the energy balance by assuming that there was no heat loss and
using experimental data at the temperature of 700 ˚C (the calculation is available in appendix
A1). For the pyrolysis under isothermal condition without bio-oil recycling, a large share of
energy from biomass is converted into sensible and chemical heat of volatiles (gas and bio-oil)
that are leaving the process (Figure 4.14a). By using slow heating (Figure 4.14b), the energy
yield of charcoal can be increased from 30% to 39%. With bio-oil recycling under the optimized
conditions, i.e., slow pyrolysis of woodchips under CO2 flow (Figure 4.14c), the energy yield
of charcoal could further increase to 49%.

38 | P a g e
Figure 4.14. Energy balance of pyrolysis units in different configurations at 700 ° a)
isothermal condition under N2, b) slow pyrolysis under N2, c) slow pyrolysis with bio-oil
recycling under CO2.

39 | P a g e
4.2.2 Comparison of the charcoal reactivity
Besides the efficiency of charcoal production, the reactivity of charcoals can be used to
evaluate pyrolysis process regarding the requirements of applications. Figure 4.15 shows the
comparison of the rate constants among charcoals produced from several processes. The figure
clearly indicates that pyrolysis under isothermal condition produced charcoal with highest
reactivity. Pyrolysis under slow heating rate at 3 °C min-1 produced charcoal with much less
reactivity than isothermal condition (by factor 2.4). Promoting secondary reaction by means of
bio-oil recycling and CO2 purging seems to increase reactivity of charcoal. In slow pyrolysis,
20% bio-oil recycling increased the reactivity of charcoal by 22% from pyrolysis without bio-
oil recycling. Utilization of CO2 in slow pyrolysis also increased the reactivity slightly (by 2.5%
from slow pyrolysis under N2). However, because of the dominant effect of slow heating rate,
the reactivity of charcoal from the process with improved yields (section 4.2.1) is still much
lower than charcoal from the process without these measures.

Figure 4.15. Comparison of rate constants between charcoals from pyrolysis process with and
without optimized conditions (slow pyrolysis with bio-oil recycling under CO2).

40 | P a g e
4.3 Pyrolysis of woodchips in a rotary kiln
4.3.1 Axial profiles
According to the model of pyrolysis in a rotary kiln, we can visualize the axial profiles
of various quantities along the kiln axis, as represented in Figure 4.16. The simulation result in
the figure is for a kiln with a diameter of 0.4 m and a length of 2 m, operated at an inclination
of 3°. The height of the dam, i.e., the rim at the outlet of the rotary kiln, is 2.5 cm. In the base
case scenario, the rotary kiln is run at a rotation speed of 2 rpm and at biomass feeding rate of
30 kg h-1. Other parameters are given in Paper C.

Figure 4.16. Simulation result representing profiles along the kiln axis of a) bed height and
bed velocity, b) bed, gas and wall temperature, c) pyrolysis products, and d) net heat fluxes.

At the given conditions, the bed height (solid line in Figure 4.16a) remains at about the
inlet level for half of the kiln length before going a smooth decrease to reach the dam height at
the kiln exit. This decrease causes an increase of the bed velocity when reaching the outlet
(dashed line in Figure 4.16a).

41 | P a g e
The temperature profiles are shown in Figure 4.16b. At the beginning part of the reactor,
the biomass bed increases to reach the pyrolysis temperature at around 0.5 m. The heating rate
in this first 0.5 m decreases from around 250 °C min-1 to less than 30 °C min-1 when reaching
temperature of 360 °C. This heating profile with the given conditions, i.e., co-current heating,
is analogous to the isothermal conditions applied in the pyrolysis of single particle done in the
macro-TG experiment. At around 0.5 m into the reactor, the pyrolysis process sets in which
affects the bed temperature in two ways: the endothermic decay of dry wood into primary
products causes a retardation of the increase of Tb, followed by an increase due to the exothermic
decomposition of the primary products. The exothermic decomposition results in a maximum
of the bed temperature that for the studied conditions exceeds the wall temperature. To highlight
the effect of the reaction enthalpy we run simulations where the heat production due to the
reaction is set to zero (i.e., G = 0 in Eq. (7)). The resulting bed temperature in this case, shown
by the dashed curve in Figure 4.16b, describes a smooth relaxation towards the wall
temperature. In contrast to the bed temperature, the gas temperature describes a very sharp
increase to reach a maximum close to the kiln entrance beyond which it shows a slow decay.
This sharp increase of Tg is caused by the small amount of gas that is present at this early stage.
Figure 4.16c shows the evolution of the pyrolysis products. The decay of the dry wood
sets in at around 0.4 m (Tb ≈ 330 °C) at which point the primary products (gas, inert tar, reactive
tar, and intermediate solid) are formed. The given rotation speed and the reaction parameters
result in a high production of inert tar compared with reactive tar. At around 1.2 m the primary
reactions are completed, and the gaseous products reached their final values. Beyond 1.2 m, the
decomposition of the intermediate solid into charcoal takes place. Charcoal yield obtained at
the outlet of the rotary kiln is 22%.
Figure 4.16d shows the net heat flux to the solid bed (qbed), the gas phase (qgas) and the
surroundings (qloss). The bed, whose temperature is lowest in the first half of the kiln, has a
positive net heat flux up to ca. 1 m. Later, a small negative heat flux is observed which is caused
by the exothermic decay of the primary pyrolysis products. The gas phase has a positive net
heat flux only at the very beginning and around 0.6 m. In between, the heat that is transferred
from the wall to the gas is delivered to the bed. The heat loss to the surroundings is roughly
constant along the kiln at ca. 1.3 kW m-1. A global energy balance reveals that despite this small
number, ca. 30% of the heat provided to the process is lost to the surroundings.

42 | P a g e
4.4.2 Influence of feeding rate and rotation speed
The two parameters of highest relevance for operating a rotary kiln are the feeding rate
and the rotation speed. While the feeding rate controls the productivity, the rotation speed
influences the solid residence time inside the kiln.
Figure 4.17 shows the influence of the feeding rate on the rotary kiln pyrolyzer. The
solid mean residence time (MRT) and the bed height at the kiln entrance as a function of the
feeding rate are shown in Figure 4.17a. The MRT is slightly influenced by the feeding rate
which is typical for rotary kilns [125,132]. The observed increase of the bed height reflects the
higher solid hold up that is needed for compensation of the higher feeding rate. Figure 4.17b
shows the outlet temperatures of the gas, the bed, and the wall. Increasing the feeding rate
decrease the outlet temperatures because of the higher heat capacity of the kiln contents. The
distribution of the final products is shown in Figure 4.17c. For a feeding rate below 35 kg h-1,
the product distribution is slightly influenced by the feeding flow. In this region, the flue gas
delivers enough heat for the pyrolysis process to complete. The situation changes for a feeding
rate larger than 35 kg h-1 where an increased amount of reactive tar, intermediate solid and
unconverted wood is observed. In this region, the heat provided by the flue gas is not enough
to maintain the pyrolyzer at a high enough temperature required to complete the pyrolysis
process. However, the result may behave differently in the case of counter-current
configuration, which can be explored in the future work.

Figure 4.17. Influence of the feeding rate.


43 | P a g e
The effect of the rotation speed is shown in Figure 4.18. Increasing the rotation speed
decreases the MRT and the bed height (Figure 4.18a). Despite the shorter residence time, the
temperatures at the kiln exit (Figure 4.18b) and the final products (Figure 4.18c) are only little
affected by the increase of the rotation speed. This minor influence on these quantities is the
result of two counter-acting effects, i.e., the increase in heat transfer with increased rotation
speed which compensates for the decrease in residence time. An effect of the rotation speed on
the product distribution is only observed for high values of where some intermediate solid
remains in the product, while the improved bed agitation allows for the tars to escape into the
gas phase. The decrease in the amount of solids observed in Figure 4.18c is caused by an
increase in heating rate. For example, increasing the rotation speed from 2 to 4 rpm increases
the maximum heating rate in the first section of the kiln from 250 to 470 °C min-1. This high
heating rate leads to produce high tar, resulting in a slight decrease in the solid mass.

Figure 4.18. Influence of the rotation speed.

44 | P a g e
Chapter V
Conclusions and future works

5.1 Conclusions
Due to the trade-off between quality and yield of charcoal given by temperature,
secondary char formation is the key solution to increase charcoal yield in pyrolysis of thick
particles. This thesis has examined mechanisms affecting charcoal yield by bio-oil recycling
and CO2 purging. Bio-oil recycling, represented by bio-oil embedded woodchips, increased
charcoal yield by adding mass from bio-oil recycling and synergetic effect due to the contact
between bio-oil and biomass during secondary char formation. Bio-oil embedded on woodchip
increased the concentration of volatile inside the pore structure of the particles, resulting
extension of the secondary char formation. This synergetic effect was diminished by the
physical disintegration of particles at high heating rate and high temperature while it remained
effective for slow pyrolysis even at high temperature. Pyrolysis under CO2 also results in higher
charcoal yield. CO2 seems to have no direct effects on primary pyrolysis reactions but via intra-
particle heat and mass transfer. As a consequence of low mass diffusivity, volatiles and solid
surface had better contact. This led to higher charcoal yield due to enhanced secondary char
formation.
Elemental composition and heating value of charcoal were mainly affected by the
reaction temperature, and bio-oil recycling and CO2 purging had negligible effects. Carbon
content and heating value of charcoal obtained at the pyrolysis temperature above 500 °C were
comparable to pulverized coal. However, charcoals produced from pyrolysis with bio-oil
recycling and CO2 purging had gasification reactivity slightly higher than charcoals from
pyrolysis without these measures. This higher in reactivity is because charcoal from bio-oil and
secondary char formation contained less stable structure (i.e., smaller aromatic cluster).

45 | P a g e
Pyrolysis under CO2 also modified porosity of charcoal resulting slightly more porous charcoal
than pyrolysis under N2, which affected the reactivity. Nevertheless, the impacts of increase in
charcoal reactivity by bio-oil recycling and CO2 purging were very small in comparison with
other parameters, i.e., temperature, heating rate, and inorganic content.
On contrary to the significant increase in charcoal yield by bio-oil recycling and CO2
purging, carbon content and heating value had negligible effect. As a consequence, these
methods significantly increase carbon yield and energy yield of charcoal production process.
The findings from this work suggest that bio-oil recycling, CO2 purging, and slow heating rate
are key process parameters to achieve higher process efficiency. With all of these measures
applied, the optimized conditions in this study showed significantly higher efficiency of the
charcoal production process (26.7% of carbon yield, and 49% of energy yield) than the process
without these measures (16.2% of carbon yield, and 30% of energy yield). However, the
reactivity of charcoal produced from the conditions optimized yields is lower than charcoal
from the process without these measures by factor ca. 2.
A numerical model of a rotary kiln pyrolyzer was used to simulate pyrolysis of
woodchips under a given condition. The heating history inside the rotary kiln with co-current
heating was analogous to the isothermal condition in the macro-TG experiments. The
simulation results reveal the effects of dimension and operating conditions on the solid
residence time, materials temperature, and biomass conversion. Feeding rate of biomass highly
influences product distribution in the kiln, although it gives very low impact on the solids
residence time. Increasing the feeding rate increased heat capacity of material in the kiln,
resulting in lower temperature and unconverted wood. Rotation speed controlled the residence
time of solid inside the kiln. Increasing the rotation speed decreased the residence time, but it
increased the degree of solid mixing resulting in faster heating of biomass. Therefore,
increasing the rotation speed can shorten production time with low effects on product quality.
Nevertheless, further investigation on dimension and operating conditions can be made by using
this reactor model in order to implement the desired pyrolysis parameters favored to maximize
efficiency of charcoal production.

46 | P a g e
5.2 Future works
The works presented in this thesis contributed in the suggesting of a charcoal
production process with high efficiency. Paper A revealed the benefit of secondary char
formation by pyrolysis using thick particles rather than wood powder. In the coming years, the
candidate will extend the investigation of pyrolysis in multiple particles to investigate the inter-
particle transport of volatile products. Further laboratory experiments will investigate the
pyrolysis of woodchips bed, e.g., fixed bed and granular bed. This knowledge would give useful
information that can be applied further to pilot-scale pyrolysis. Pyrolysis of woodchips in a
pilot-scale auger reactor can be studied incorporation with industrial partners, namely Harads
Arctic Heat AB, Sweden.
In addition, the candidate intends to study more on the model of pyrolysis in a rotary
kiln that has been developed in Paper C. As explained earlier, dimension and operating
conditions of rotary kilns are flexible to vary pyrolysis parameters in the reactor. Therefore,
desired pyrolysis conditions suggested from the investigation in the macro-TG experiment
(single particle and/or multiple particles) and simulation of a particle model (an existing model),
such as final temperature, heating rate, CO2 purging, and other relevant parameters can be
implemented into the reactor model. For example, low heating rate that was favored for high
charcoal yield can be obtained in a rotary kiln with counter-current heating. As a consequence,
this reactor model can suggest configurations, dimensions, and operating conditions that could
maximize efficiency of charcoal production. Nevertheless, due to the assumption of perfect
mixing in radial direction and no dispersion in axial direction in the model, the model is limited
for understanding more detailed mechanisms of pyrolysis in a rotary kiln. Therefore, an
extension of the model could bring the model get closer to the reality. Moreover, validation of
the reactor model should be done by experimental investigation.
Possibility to use charcoal as fossil coal replacement has been evaluated by the mean
of gasification reactivity and properties of charcoal in Paper B. According to the plan, the
candidate will study on changes in size and apparent density of single charcoal particle in
combustion and gasification environments. The result from this work is expected to increase
understanding in uses of charcoal in the industrial applications. The investigation could be done
by experiments performed in a single particle converter using a flat flame burner, where the
reaction conditions can be controlled. A CCD or a high-speed camera can be applied to capture
the change of particle size and temperature. In addition, the candidate will develop a particle
model to simulate changes of a charcoal particle during heterogeneous reactions.

47 | P a g e
According to the plan, also shown in Figure 5.1, the PhD thesis is expected to deliver
holistic knowledge of a charcoal production technology.

Figure 5.1. Future plan toward the PhD thesis. Blue lines indicate future works.

48 | P a g e
Bibliography

[1] J. Koppejan, S. Sokhansanj, S. Melin, S. Madrali, Status overview of torrefaction


technologies, IEA Bioenergy Task 32. (2012) 1–54.

[2] E. Mousa, C. Wang, J. Riesbeck, M. Larsson, Biomass applications in iron and steel
industry: An overview of challenges and opportunities, Renew. Sustain. Energy Rev.
65 (2016) 1247–1266. doi:10.1016/j.rser.2016.07.061.

[3] A. Molino, V. Larocca, S. Chianese, D. Musmarra, Biofuels production by biomass


gasification: A review, Energies. 11 (2018) 1–31. doi:10.3390/en11040811.

[4] J. Dai, S. Sokhansanj, J.R. Grace, X. Bi, C.J. Lim, S. Melin, Overview and some issues
related to co-firing biomass and coal, Can. J. Chem. Eng. 86 (2008) 367–386.
doi:10.1002/cjce.20052.

[5] Blackwood technology, Co-firing of biomass, Blackwood Technol. (2017).


http://www.blackwood-technology.com/applications/co-firing/ (accessed November 7,
2018).

[6] H. Suopajärvi, K. Umeki, E. Mousa, A. Hedayati, H. Romar, A. Kemppainen, C.


Wang, A. Phounglamcheik, S. Tuomikoski, N. Norberg, A. Andefors, M. Öhman, U.
Lassi, T. Fabritius, Use of biomass in integrated steelmaking – Status quo, future needs
and comparison to other low-CO2steel production technologies, Appl. Energy. 213
(2018) 384–407. doi:10.1016/j.apenergy.2018.01.060.

[7] C. Wang, P. Mellin, J. Lövgren, L. Nilsson, W. Yang, H. Salman, A. Hultgren, M.


Larsson, Biomass as blast furnace injectant - Considering availability, pretreatment and
deployment in the Swedish steel industry, Energy Convers. Manag. 102 (2015) 217–
226. doi:10.1016/j.enconman.2015.04.013.

[8] N.T. Farrokh, H. Suopajärvi, O. Mattila, K. Umeki, A. Phounglamcheik, H. Romar, P.


Sulasalmi, T. Fabritius, Slow pyrolysis of by-product lignin from wood-based ethanol
production– A detailed analysis of the produced chars, Energy. 164 (2018) 112–123.
doi:10.1016/J.ENERGY.2018.08.161.
49 | P a g e
[9] S. Şensöz, M. Can, Pyrolysis of Pine ( Pinus Brutia Ten.) Chips: 1. Effect of Pyrolysis
Temperature and Heating Rate on the Product Yields, Energy Sources. 24 (2002) 347–
355. doi:10.1080/00908310252888727.

[10] D. Neves, H. Thunman, A. Matos, L. Tarelho, A. Gómez-Barea, Characterization and


prediction of biomass pyrolysis products, Prog. Energy Combust. Sci. 37 (2011) 611–
630. doi:10.1016/j.pecs.2011.01.001.

[11] H. Thunman, F. Niklasson, F. Johnsson, B. Leckner, Composition of volatile gases and


thermochemical properties of wood for modeling of fixed or fluidized beds, Energy and
Fuels. 15 (2001) 1488–1497. doi:10.1021/ef010097q.

[12] D.S. Scott, J. Piskorz, CONTINUOUS FLASH PYROLYSIS OF BIOMASS., Can. J.


Chem. Eng. 62 (1984) 404–412. doi:10.1002/cjce.5450620319.

[13] M. Nik-Azar, M.R. Hajaligol, M. Sohrabi, B. Dabir, Effects of heating rate and particle
size on the products yields from rapid pyrolysis of beech-wood, Fuel Sci. Technol. Int.
14 (1996) 479–502. doi:10.1080/08843759608947593.

[14] L. Basile, A. Tugnoli, C. Stramigioli, V. Cozzani, Thermal effects during biomass


pyrolysis, Thermochim. Acta. 636 (2016) 63–70. doi:10.1016/j.tca.2016.05.002.

[15] L. Wang, M. Trninic, Ø. Skreiberg, M. Gronli, R. Considine, M.J. Antal, Is Elevated


Pressure Required To Achieve a High Fixed-Carbon Yield of Charcoal from Biomass?
Part 1: Round-Robin Results for Three Different Corncob Materials, Energy & Fuels.
25 (2011) 3251–3265. doi:10.1021/ef200450h.

[16] J. Recari, C. Berrueco, S. Abelló, D. Montané, X. Farriol, Effect of temperature and


pressure on characteristics and reactivity of biomass-derived chars, Bioresour. Technol.
170 (2014) 204–210. doi:10.1016/j.biortech.2014.07.080.

[17] P. Rousset, C. Figueiredo, M. De Souza, W. Quirino, Pressure effect on the quality of


eucalyptus wood charcoal for the steel industry: A statistical analysis approach, Fuel
Process. Technol. 92 (2011) 1890–1897. doi:10.1016/j.fuproc.2011.05.005.

[18] L. Basile, A. Tugnoli, C. Stramigioli, V. Cozzani, Influence of pressure on the heat of


biomass pyrolysis, Fuel. 137 (2014) 277–284. doi:10.1016/j.fuel.2014.07.071.

50 | P a g e
[19] S. Yorgun, D. Yildiz, Slow pyrolysis of paulownia wood: Effects of pyrolysis
parameters on product yields and bio-oil characterization, J. Anal. Appl. Pyrolysis. 114
(2015) 68–78. doi:10.1016/j.jaap.2015.05.003.

[20] İ. Demiral, S. Şensöz, Fixed-Bed Pyrolysis of Hazelnut ( Corylus Avellana L.)


Bagasse: Influence of Pyrolysis Parameters on Product Yields, Energy Sources, Part A
Recover. Util. Environ. Eff. 28 (2006) 1149–1158. doi:10.1080/009083190966126.

[21] A.E. Pütün, N. Özbay, E.P. Önal, E. Pütün, Fixed-bed pyrolysis of cotton stalk for
liquid and solid products, Fuel Process. Technol. 86 (2005) 1207–1219.
doi:10.1016/j.fuproc.2004.12.006.

[22] Y. Huang, S. Kudo, K. Norinaga, M. Amaike, J.I. Hayashi, Selective production of


light oil by biomass pyrolysis with feedstock-mediated recycling of heavy oil, Energy
and Fuels. 26 (2012) 256–264. doi:10.1021/ef2011673.

[23] C. Di Blasi, Combustion and gasification rates of lignocellulosic chars, Prog. Energy
Combust. Sci. 35 (2009) 121–140. doi:10.1016/j.pecs.2008.08.001.

[24] Z. Bouraoui, M. Jeguirim, C. Guizani, L. Limousy, C. Dupont, R. Gadiou,


Thermogravimetric study on the influence of structural, textural and chemical
properties of biomass chars on CO2 gasification reactivity, Energy. 88 (2015) 703–710.
doi:10.1016/j.energy.2015.05.100.

[25] N.M. Laurendeau, Heterogeneous kinetics of coal char gasification and combustion,
Prog. Energy Combust. Sci. 4 (1978) 221–270. doi:10.1016/0360-1285(78)90008-4.

[26] F. Mermoud, S. Salvador, L. Van de Steene, F. Golfier, Influence of the pyrolysis


heating rate on the steam gasification rate of large wood char particles, Fuel. 85 (2006)
1473–1482. doi:10.1016/j.fuel.2005.12.004.

[27] M. Asadullah, S. Zhang, Z. Min, P. Yimsiri, C.Z. Li, Effects of biomass char structure
on its gasification reactivity, Bioresour. Technol. 101 (2010) 7935–7943.
doi:10.1016/j.biortech.2010.05.048.

[28] K. Raveendran, A. Ganesh, Adsorption characteristics and pore-development of


biomass-pyrolysis char, Fuel. 77 (1998) 769–781. doi:10.1016/S0016-2361(97)00246-
9.

51 | P a g e
[29] J.A. Garcia-Nunez, M.R. Pelaez-Samaniego, M.E. Garcia-Perez, I. Fonts, J. Abrego,
R.J.M. Westerhof, M. Garcia-Perez, Historical Developments of Pyrolysis Reactors: A
Review, Energy and Fuels. 31 (2017) 5751–5775.
doi:10.1021/acs.energyfuels.7b00641.

[30] P. Brassard, S. Godbout, V. Raghavan, Pyrolysis in auger reactors for biochar and bio-
oil production: A review, Biosyst. Eng. 161 (2017) 80–92.
doi:10.1016/j.biosystemseng.2017.06.020.

[31] L. Axelsson, M. Franzén, M. Ostwald, G. Berndes, G. Lakshmi, N.H. Ravindranath,


Perspective: Jatropha cultivation in southern India: Assessing farmers’ experiences,
Biofuels, Bioprod. Biorefining. 6 (2012) 246–256. doi:10.1002/bbb.

[32] A. V Bridgwater, Upgrading biomass fast pyrolysis liquids., Environ. Prog. Sustain.
Energy. 31 (2012) 261–268. doi:doi.org/10.1002/ep.11635.

[33] N. Sathitsuksanoh, A. George, Y.H.P. Zhang, New lignocellulose pretreatments using


cellulose solvents: A review, J. Chem. Technol. Biotechnol. 88 (2013) 169–180.
doi:10.1002/jctb.3959.

[34] V. Chaturvedi, P. Verma, An overview of key pretreatment processes employed for


bioconversion of lignocellulosic biomass into biofuels and value added products, 3
Biotech. 3 (2013) 415–431. doi:10.1007/s13205-013-0167-8.

[35] P. McKendry, Energy production from biomass (part 3): Gasification technologies,
Bioresour. Technol. 83 (2002) 55–63. doi:10.1016/S0960-8524(01)00120-1.

[36] C. Laine, Structures of hemicelluloses and pectins in wood wnd pulp, Helsinki
University of Technology, 2005. http://lib.tkk.fi/Diss/2005/isbn9512276909/.

[37] 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.
doi:10.1016/j.rser.2014.06.013.

[38] S. Şensöz, M. Can, Pyrolysis of pine (Pinus Brutia Ten.) chips: 1. Effect of pyrolysis
temperature and heating rate on the product yields, Energy Sources. 24 (2002) 347–
355. doi:10.1080/00908310252888727.

52 | P a g e
[39] T.X. Xuebin Wang, Weigang Xu, Limeng Zhang, Zhongfa Hu, Houzhang Tan,
Yingying Xiong, Char Characteristics from the Pyrolysis of Straw, Wood and Coal at
High Temperatures, J. Biobased Mater. Bioenergy. 7 (2013) 675–683.
doi:10.1166/jbmb.2013.1380.

[40] A. Li, H.L. Liu, H. Wang, H. Bin Xu, L.F. Jin, J.L. Liu, J.H. Hu, Effects of temperature
and heating rate on the characteristics of molded Bio-Char, BioResources. 11 (2016)
3259–3274. doi:10.15376/biores.11.2.3259-3274.

[41] V. Chiodo, G. Zafarana, S. Maisano, S. Freni, F. Urbani, Pyrolysis of different


biomass: Direct comparison among Posidonia Oceanica, Lacustrine Alga and White-
Pine, Fuel. 164 (2016) 220–227. doi:10.1016/j.fuel.2015.09.093.

[42] H. Yang, L. Huang, S. Liu, K. Sun, Y. Sun, Pyrolysis Process and Characteristics of
Products from Sawdust Briquettes, BioResources. 11 (2016) 2438–2456.
doi:10.15376/biores.11.1.2438-2456.

[43] A. Phounglamcheik, T. Wretborn, K. Umeki, Biomass pyrolysis with bio-oil recycle to


increase energy recovery in biochar, in: 25th Eur. Biomass Conf. Exibition, Stockholm,
2017: pp. 1388–1392. doi:10.5071/25thEUBCE2017-3CV.2.6.

[44] L. Wang, Ø. Skreiberg, S. Van Wesenbeeck, M.G. Gronli, M.J. Antal, Experimental
Study on Charcoal Production from Woody Biomass, Energy & Fuels. 30 (2016)
7994–8008. doi:10.1021/acs.energyfuels.6b01039.

[45] P.T. Williams, S. Besler, The Influence of Temperature and Heating Rate on the Slow
Pyrolysis of Biomass, Renew. Energy. 7 (1996) 233–250. doi:10.1016/0960-
1481(96)00006-7.

[46] D. Chen, J. Zhou, Q. Zhang, Effects of heating rate on slow pyrolysis behavior, kinetic
parameters and products properties of moso bamboo, Bioresour. Technol. 169 (2014)
313–319. doi:10.1016/j.biortech.2014.07.009.

[47] A. Dufour, P. Girods, E. Masson, Y. Rogaume, A. Zoulalian, Synthesis gas production


by biomass pyrolysis: Effect of reactor temperature on product distribution, Int. J.
Hydrogen Energy. 34 (2009) 1726–1734. doi:10.1016/j.ijhydene.2008.11.075.

53 | P a g e
[48] T. Kan, V. Strezov, T.J. Evans, Lignocellulosic biomass pyrolysis: A review of product
properties and effects of pyrolysis parameters, Renew. Sustain. Energy Rev. 57 (2016)
1126–1140. doi:10.1016/j.rser.2015.12.185.

[49] E.S. Noumi, P. Rousset, A. De Cassia Oliveira Carneiro, J. Blin, Upgrading of carbon-
based reductants from biomass pyrolysis under pressure, J. Anal. Appl. Pyrolysis. 118
(2016) 278–285. doi:10.1016/j.jaap.2016.02.011.

[50] M.J. Antal, M. Grønli, The Art, Science, and Technology of Charcoal Production, Ind.
Eng. Chem. Res. 42 (2003) 1619–1640. doi:10.1021/ie0207919.

[51] S.B. Saleh, J.P. Flensborg, T.K. Shoulaifar, Z. Sárossy, B.B. Hansen, H. Egsgaard, N.
Demartini, P.A. Jensen, P. Glarborg, K. Dam-Johansen, Release of chlorine and sulfur
during biomass torrefaction and pyrolysis, Energy and Fuels. 28 (2014) 3738–3746.
doi:10.1021/ef4021262.

[52] P.A. Della Rocca, E.G. Cerrella, P.R. Bonelli, A.L. Cukierman, Pyrolysis of
hardwoods residues: On kinetics and chars characterization, Biomass and Bioenergy.
16 (1999) 79–88. doi:10.1016/S0961-9534(98)00067-1.

[53] H. Abdullah, H. Wu, Biochar as a fuel: 1. Properties and grindability of biochars


produced from the pyrolysis of mallee wood under slow-heating conditions, Energy
and Fuels. 23 (2009) 4174–4181. doi:10.1021/ef900494t.

[54] S. Zellagui, C. Schönnenbeck, N. Zouaoui-Mahzoul, G. Leyssens, O. Authier, E.


Thunin, L. Porcheron, J.F. Brilhac, Pyrolysis of coal and woody biomass under N2 and
CO2 atmospheres using a drop tube furnace - Experimental study and kinetic modeling,
Fuel Process. Technol. 148 (2016) 99–109. doi:10.1016/j.fuproc.2016.02.007.

[55] A. Veksha, H. McLaughlin, D.B. Layzell, J.M. Hill, Pyrolysis of wood to biochar:
Increasing yield while maintaining microporosity, Bioresour. Technol. 153 (2014)
173–179. doi:10.1016/j.biortech.2013.11.082.

[56] A. Zolin, A. Jensen, P.A. Jensen, F. Frandsen, K. Dam-Johansen, The influence of


inorganic materials on the thermal deactivation of fuel chars, Energy and Fuels. 15
(2001) 1110–1122. doi:10.1021/ef000288d.

54 | P a g e
[57] J.M. Encinar, J.F. González, J.J. Rodríguez, M.J. Ramiro, Catalysed and uncatalysed
steam gasification of eucalyptus char: Influence of variables and kinetic study, Fuel. 80
(2001) 2025–2036. doi:10.1016/S0016-2361(01)00073-4.

[58] M.P. Kannan, G.N. Richards, Gasification of biomass chars in carbon dioxide:
dependence of gasification rate on the indigenous metal content, Fuel. 69 (1990) 747–
753. doi:10.1016/0016-2361(90)90041-N.

[59] S. V. Vassilev, D. Baxter, L.K. Andersen, C.G. Vassileva, An overview of the


composition and application of biomass ash. Part 1. Phase-mineral and chemical
composition and classification, Fuel. 105 (2013) 40–76.
doi:10.1016/j.fuel.2012.09.041.

[60] P. Thy, C. Yu, B.M. Jenkins, C.E. Lesher, Inorganic composition and environmental
impact of biomass feedstock, Energy and Fuels. 27 (2013) 3969–3987.
doi:10.1021/ef400660u.

[61] T. Okuno, N. Sonoyama, J.I. Hayashi, C.Z. Li, C. Sathe, T. Chiba, Primary release of
alkali and alkaline earth metallic species during the pyrolysis of pulverized biomass,
Energy and Fuels. 19 (2005) 2164–2171. doi:10.1021/ef050002a.

[62] J.G. Olsson, U. Jäglid, J.B.C. Pettersson, P. Hald, Alkali metal emission during
pyrolysis of biomass, Energy and Fuels. 11 (1997) 779–784. doi:10.1021/ef960096b.

[63] P.A. Jensen, F.J. Frandsen, K. Dam-Johansen, B. Sander, Experimental Investigation of


the Transformation and Release to Gas Phase of Potassium and Chlorine during Straw
Pyrolysis, Energy and Fuels. 14 (2000) 1280–1285. doi:10.1021/ef000104v.

[64] K.O. Davidsson, B.J. Stojkova, J.B.C. Pettersson, Alkali emission from birchwood
particles during rapid pyrolysis, Energy and Fuels. 16 (2002) 1033–1039.
doi:10.1021/ef010257y.

[65] M. Guerrero, M.P. Ruiz, M.U. Alzueta, R. Bilbao, A. Millera, Pyrolysis of eucalyptus
at different heating rates: Studies of char characterization and oxidative reactivity, J.
Anal. Appl. Pyrolysis. 74 (2005) 307–314. doi:10.1016/j.jaap.2004.12.008.

55 | P a g e
[66] C. Fushimi, K. Araki, Y. Yamaguchi, A. Tsutsumi, Effect of heating rate on steam
gasification of biomass. 1. Reactivity of char, Ind. Eng. Chem. Res. 42 (2003) 3922–
3928. doi:10.1021/ie030056c.

[67] M. Morin, S. Pécate, M. Hémati, Y. Kara, Pyrolysis of biomass in a batch fluidized bed
reactor: Effect of the pyrolysis conditions and the nature of the biomass on the
physicochemical properties and the reactivity of char, J. Anal. Appl. Pyrolysis. 122
(2016) 511–523. doi:https://doi.org/10.1016/j.jaap.2016.10.002.

[68] F. Kurosaki, K. Ishimaru, T. Hata, P. Bronsveld, E. Kobayashi, Y. Imamura,


Microstructure of wood charcoal prepared by flash heating, Carbon N. Y. 41 (2003)
3057–3062. doi:10.1016/S0008-6223(03)00434-2.

[69] R.A. Brown, A.K. Kercher, T.H. Nguyen, D.C. Nagle, W.P. Ball, Production and
characterization of synthetic wood chars for use as surrogates for natural sorbents, Org.
Geochem. 37 (2006) 321–333. doi:10.1016/j.orggeochem.2005.10.008.

[70] M. Keiluweit, P.S. Nico, M. Johnson, M. Kleber, Dynamic molecular structure of plant
biomass-derived black carbon (biochar), Environ. Sci. Technol. 44 (2010) 1247–1253.
doi:10.1021/es9031419.

[71] E.S. Noumi, J. Blin, J. Valette, P. Rousset, Combined Effect of Pyrolysis Pressure and
Temperature on the Yield and CO2 Gasification Reactivity of Acacia Wood in macro-
TG, Energy and Fuels. 29 (2015) 7301–7308. doi:10.1021/acs.energyfuels.5b01454.

[72] M.W. Smith, I. Dallmeyer, T.J. Johnson, C.S. Brauer, J.S. McEwen, J.F. Espinal, M.
Garcia-Perez, Structural analysis of char by Raman spectroscopy: Improving band
assignments through computational calculations from first principles, Carbon N. Y. 100
(2016) 678–692. doi:10.1016/j.carbon.2016.01.031.

[73] J. Robertson, Diamond-like amorphous carbon, Mater. Sci. Eng. R Reports. 37 (2002)
129–281. doi:10.1016/S0927-796X(02)00005-0.

[74] M. Asadullah, S. Zhang, C.-Z. Li, Evaluation of structural features of chars from
pyrolysis of biomass of different particle sizes, Fuel Process. Technol. 91 (2010) 877–
881. doi:10.1016/j.fuproc.2009.08.008.

56 | P a g e
[75] R.K. Sharma, J.B. Wooten, V.L. Baliga, M.R. Hajaligol, Characterization of chars from
biomass-derived materials: Pectin chars, Fuel. 80 (2001) 1825–1836.
doi:10.1016/S0016-2361(01)00066-7.

[76] R.K. Sharma, J.B. Wooten, V.L. Baliga, X. Lin, W.G. Chan, M.R. Hajaligol,
Characterization of chars from pyrolysis of lignin, Fuel. 84 (2004) 1469–1482.
doi:10.1016/j.fuel.2003.11.015.

[77] C. Branca, C. Di Blasi, Combustion kinetics of secondary biomass chars in the kinetic
regime, Energy and Fuels. 24 (2010) 5741–5750. doi:10.1021/ef100952x.

[78] C. Branca, C. Di Blasi, R. Elefante, Devolatilization and heterogeneous combustion of


wood fast pyrolysis oils, Ind. Eng. Chem. Res. 44 (2005) 799–810.
doi:10.1021/ie049419e.

[79] A. Anca-Couce, A. Dieguez-Alonso, N. Zobel, A. Berger, N. Kienzl, F. Behrendt,


Influence of Heterogeneous Secondary Reactions during Slow Pyrolysis on Char
Oxidation Reactivity of Woody Biomass, Energy and Fuels. 31 (2017) 2335–2344.
doi:10.1021/acs.energyfuels.6b02350.

[80] W. Emrich, Handbook of Charcoal Making: The Traditional and Industrial Methods,
1985.

[81] Y. Seboka, Charcoal production: opportunities and barriers for improving efficiency
and sustainability., in: Bio-Carbon Oppor. East. South. Africa Harnessing Carbon
Financ. to Promot. Sustain. For. Agro-Forestry Bio-Energy, 2009: pp. 102–126.
doi:10.1111/j.1365-2141.2009.07705.x.

[82] Pyreg, Biomass Pyreg Process, (2018). http://www.pyreg.de/wp-


content/uploads/03_biomass_pyreg_process.pdf (accessed November 7, 2018).

[83] Y. Yu, Y. Yang, Z. Cheng, P.H. Blanco, R. Liu, A. V. Bridgwater, J. Cai, Pyrolysis of
Rice Husk and Corn Stalk in Auger Reactor. 1. Characterization of Char and Gas at
Various Temperatures, Energy and Fuels. 30 (2016) 10568–10574.
doi:10.1021/acs.energyfuels.6b02276.

57 | P a g e
[84] S. Liang, Y. Han, L. Wei, A.G. McDonald, Production and characterization of bio-oil
and bio-char from pyrolysis of potato peel wastes, Biomass Convers. Biorefinery. 5
(2015) 237–246. doi:10.1007/s13399-014-0130-x.

[85] M.I. Schnitzer, C.M. Monreal, G.A. Facey, P.B. Fransham, The conversion of chicken
manure to biooil by fast pyrolysis I. Analyses of chicken manure, biooils and char by
13C and 1H NMR and FTIR spectrophotometry, J. Environ. Sci. Heal. Part B. 42
(2007) 71–77. doi:10.1080/03601230601020894.

[86] E. Aylón, A. Fernández-Colino, R. Murillo, G. Grasa, M. V. Navarro, T. García, A.M.


Mastral, Waste tyre pyrolysis: Modelling of a moving bed reactor, Waste Manag. 30
(2010) 2530–2536. doi:10.1016/j.wasman.2010.04.018.

[87] J.D. Martínez, R. Murillo, T. García, A. Veses, Demonstration of the waste tire
pyrolysis process on pilot scale in a continuous auger reactor, J. Hazard. Mater. 261
(2013) 637–645. doi:10.1016/j.jhazmat.2013.07.077.

[88] J. Solar, I. de Marco, B.M. Caballero, A. Lopez-Urionabarrenechea, N. Rodriguez, I.


Agirre, A. Adrados, Influence of temperature and residence time in the pyrolysis of
woody biomass waste in a continuous screw reactor, Biomass and Bioenergy. 95
(2016) 416–423. doi:10.1016/j.biombioe.2016.07.004.

[89] É. Le Roux, M. Chaouch, P.N. Diouf, T. Stevanovic, Impact of a pressurized hot water
treatment on the quality of bio-oil produced from aspen, Biomass and Bioenergy. 81
(2015) 202–209. doi:10.1016/j.biombioe.2015.07.005.

[90] D. Mohan, C.U. Pittman, M. Bricka, F. Smith, B. Yancey, J. Mohammad, P.H. Steele,
M.F. Alexandre-Franco, V. Gómez-Serrano, H. Gong, Sorption of arsenic, cadmium,
and lead by chars produced from fast pyrolysis of wood and bark during bio-oil
production, J. Colloid Interface Sci. 310 (2007) 57–73. doi:10.1016/j.jcis.2007.01.020.

[91] L. Ingram, D. Mohan, M. Bricka, P. Steele, D. Strobel, D. Crocker, B. Mitchell, J.


Mohammad, K. Cantrell, C.U. Pittman, Pyrolysis of wood and bark in an auger reactor:
Physical properties and chemical analysis of the produced bio-oils, Energy and Fuels.
22 (2008) 614–625. doi:10.1021/ef700335k.

58 | P a g e
[92] M. Garcia-Perez, T.T. Adams, J.W. Goodrum, D. Geller, K.C. Das, Production and fuel
properties of pine chip bio-oil/biodiesel blends, Energy and Fuels. 21 (2007) 2363–
2372. doi:10.1021/ef060533e.

[93] P. Bhattacharya, P.H. Steele, E.B.M. Hassan, B. Mitchell, L. Ingram, C.U. Pittman,
Wood/plastic copyrolysis in an auger reactor: Chemical and physical analysis of the
products, Fuel. 88 (2009) 1251–1260. doi:https://doi.org/10.1016/j.fuel.2009.01.009.

[94] N. Puy, R. Murillo, M. V. Navarro, J.M. López, J. Rieradevall, G. Fowler, I.


Aranguren, T. García, J. Bartrolí, A.M. Mastral, Valorisation of forestry waste by
pyrolysis in an auger reactor, Waste Manag. 31 (2011) 1339–1349.
doi:10.1016/j.wasman.2011.01.020.

[95] P. V. Barr, J.K. Brimacombe, A.P. Watkinson, A heat-transfer model for the rotary
kiln: Part I. pilot kiln trials, Metall. Trans. B. 20 (1989) 391–402.
doi:10.1007/BF02696991.

[96] Cortus AB, WoodRoll, (2018). http://www.cortus.se/technology.html#tech1 (accessed


November 7, 2018).

[97] P. Sanginés, M.P. Domínguez, F. Sánchez, G. San Miguel, Slow pyrolysis of olive
stones in a rotary kiln: Chemical and energy characterization of solid, gas, and
condensable products, J. Renew. Sustain. Energy. 7 (2015) 043103.
doi:10.1063/1.4923442.

[98] Y.N. Chun, S.C. Kim, K. Yoshikawa, Pyrolysis gasification of dried sewage sludge in
a combined screw and rotary kiln gasifier, Appl. Energy. 88 (2011) 1105–1112.
doi:10.1016/j.apenergy.2010.10.038.

[99] Y. Mei, R. Liu, Q. Yang, H. Yang, J. Shao, C. Draper, S. Zhang, H. Chen, Torrefaction
of cedarwood in a pilot scale rotary kiln and the influence of industrial flue gas,
Bioresour. Technol. 177 (2015) 355–360. doi:10.1016/j.biortech.2014.10.113.

[100] P.A. Shimizu, Y.; Cundall, Three-Dimensional DEM Simulations of Bulk Handling by
Screw Conveyors, J. Eng. Mech. 127 (2001) 864–872. doi:Doi 10.1061/(Asce)0733-
9399(2001)127:9(864).

59 | P a g e
[101] P.J. Owen, P.W. Cleary, Prediction of screw conveyor performance using the Discrete
Element Method (DEM), Powder Technol. 193 (2009) 274–288.
doi:10.1016/j.powtec.2009.03.012.

[102] P.W. Cleary, Large scale industrial DEM modelling, Eng. Comput. 21 (2004) 169–204.
doi:10.1108/02644400410519730.

[103] 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.
doi:10.1016/j.fuel.2015.04.038.

[104] F. Codignole Luz, S. Cordiner, A. Manni, V. Mulone, V. Rocco, Biomass fast


pyrolysis in a shaftless screw reactor: A 1-D numerical model, Energy. 157 (2018)
792–805. doi:10.1016/j.energy.2018.05.166.

[105] A. Funke, R. Grandl, M. Ernst, N. Dahmen, Modelling and improvement of heat


transfer coefficient in auger type reactors for fast pyrolysis application, Chem. Eng.
Process. - Process Intensif. 130 (2018) 67–75. doi:10.1016/j.cep.2018.05.023.

[106] F. Marias, H. Roustan, A. Pichat, Modelling of a rotary kiln for the pyrolysis of
aluminium waste, Chem. Eng. Sci. 60 (2005) 4609–4622.
doi:10.1016/j.ces.2005.03.025.

[107] E. Benanti, C. Freda, V. Lorefice, G. Braccio, V.K. Sharma, Simulation of olive pits
pyrolysis in a rotary kiln plant, Therm. Sci. 15 (2011) 145–158.
doi:10.2298/TSCI090901073B.

[108] W. Saeman, Passage of solids through rotary kilns: Factors affecting time of passage,
Chem. Eng. Prog. 47 (1951) 508–514.

[109] W. Klose, W. Wiest, Experiments and mathematical modeling of maize pyrolysis in a


rotary kiln, Fuel. 78 (1999) 65–72. doi:10.1016/S0016-2361(98)00124-0.

[110] A. Bach-Oller, K. Kirtania, E. Furusjö, K. Umeki, Co-gasification of black liquor and


pyrolysis oil at high temperature: Part 1. Fate of alkali elements, Fuel. 202 (2017) 46–
55. doi:10.1016/j.fuel.2017.04.013.

[111] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of Gases in Multimolecular Layers, J.


Am. Chem. Soc. 60 (1938) 309–319. doi:10.1021/ja01269a023.

60 | P a g e
[112] E.P. Barrett, L.G. Joyner, P.P. Halenda, The Determination of Pore Volume and Area
Distributions in Porous Substances. I. Computations from Nitrogen Isotherms, J. Am.
Chem. Soc. 73 (1951) 373–380. doi:10.1021/ja01145a126.

[113] N. a. Seaton, J.P.R.B. Walton, N. Quirke, A new analysis method for the determination
of the pore size distribution of porous carbons from nitrogen adsorption measurements,
Carbon N. Y. 27 (1989) 853–861. doi:10.1016/0008-6223(89)90035-3.

[114] F. Tuinstra, J.L. Koenig, Raman Spectrum of Graphite, J. Chem. Phys. 53 (1970)
1126–1130. doi:10.1063/1.1674108.

[115] E. Bar-Ziv, A. Zaida, P. Salatino, O. Senneca, Diagnostics of carbon gasification by


Raman microprobe spectroscopy, Proc. Combust. Inst. 28 (2000) 2369–2374.
doi:10.1016/S0082-0784(00)80649-9.

[116] A. Sadezky, H. Muckenhuber, H. Grothe, R. Niessner, U. Pöschl, Raman


microspectroscopy of soot and related carbonaceous materials: Spectral analysis and
structural information, Carbon N. Y. 43 (2005) 1731–1742.
doi:10.1016/j.carbon.2005.02.018.

[117] J. McDonald-Wharry, M. Manley-Harris, K. Pickering, Carbonisation of biomass-


derived chars and the thermal reduction of a graphene oxide sample studied using
Raman spectroscopy, Carbon N. Y. 59 (2013) 383–405.
doi:10.1016/j.carbon.2013.03.033.

[118] X. Li, J. ichiro Hayashi, C.Z. Li, Volatilisation and catalytic effects of alkali and
alkaline earth metallic species during the pyrolysis and gasification of Victorian brown
coal. Part VII. Raman spectroscopic study on the changes in char structure during the
catalytic gasification in air, Fuel. 85 (2006) 1509–1517.
doi:10.1016/j.fuel.2006.01.011.

[119] A.C. Ferrari, J. Robertson, Raman spectroscopy of amorphous, nanostructured,


diamond-like carbon, and nanodiamond, Philos. Trans. R. Soc. A Math. Phys. Eng. Sci.
362 (2004) 2477–2512. doi:10.1098/rsta.2004.1452.

[120] A.C. Ferrari, J. Robertson, Interpretation of Raman spectra of disordered and


amorphous carbon, Phys. Rev. B. 61 (2000) 14095–14107.
doi:10.1103/PhysRevB.61.14095.

61 | P a g e
[121] C. Hu, S. Sedghi, A. Silvestre-Albero, G.G. Andersson, A. Sharma, P. Pendleton, F.
Rodríguez-Reinoso, K. Kaneko, M.J. Biggs, Raman spectroscopy study of the
transformation of the carbonaceous skeleton of a polymer-based nanoporous carbon
along the thermal annealing pathway, Carbon N. Y. 85 (2015) 147–158.
doi:10.1016/j.carbon.2014.12.098.

[122] C. Guizani, M. Jeguirim, S. Valin, L. Limousy, S. Salvador, Biomass chars: The effects
of pyrolysis conditions on their morphology, structure, chemical properties and
reactivity, Energies. 10 (2017) 796. doi:10.3390/en10060796.

[123] D.M. Keown, X. Li, J.I. Hayashi, C.Z. Li, Characterization of the structural features of
char from the pyrolysis of cane trash using Fourier transform-Raman spectroscopy,
Energy and Fuels. 21 (2007) 1816–1821. doi:10.1021/ef070049r.

[124] L. Lin, M. Strand, Investigation of the intrinsic CO2gasification kinetics of biomass


char at medium to high temperatures, Appl. Energy. 109 (2013) 220–228.
doi:10.1016/j.apenergy.2013.04.027.

[125] B. Colin, J.L. Dirion, P. Arlabosse, S. Salvador, Wood chips flow in a rotary kiln:
Experiments and modeling, Chem. Eng. Res. Des. 98 (2015) 179–187.
doi:10.1016/j.cherd.2015.04.017.

[126] Z. Ma, N. Gao, L. Zhang, A. Li, Modeling and simulation of oil sludge pyrolysis in a
rotary kiln with a solid heat carrier: Considering the particle motion and reaction
kinetics, Energy and Fuels. 28 (2014) 6029–6037. doi:10.1021/ef501263m.

[127] W.C. Park, A. Atreya, H.R. Baum, Experimental and theoretical investigation of heat
and mass transfer processes during wood pyrolysis, Combust. Flame. 157 (2010) 481–
494. doi:10.1016/j.combustflame.2009.10.006.

[128] C. Di Blasi, Modeling intra- and extra-particle processes of wood fast pyrolysis, AIChE
J. 48 (2002) 2386–2397. doi:10.1002/aic.690481028.

[129] O. Senneca, F. Cerciello, S. Heuer, P. Ammendola, Slow pyrolysis of walnut shells in


nitrogen and carbon dioxide, Fuel. 225 (2018) 419–425.
doi:10.1016/j.fuel.2018.03.094.

62 | P a g e
[130] G. Wang, J. Zhang, W. Chang, R. Li, Y. Li, C. Wang, Structural features and
gasification reactivity of biomass chars pyrolyzed in different atmospheres at high
temperature, Energy. 147 (2018) 25–35. doi:10.1016/j.energy.2018.01.025.

[131] D.M. Keown, J.I. Hayashi, C.Z. Li, Drastic changes in biomass char structure and
reactivity upon contact with steam, Fuel. 87 (2008) 1127–1132.
doi:10.1016/j.fuel.2007.05.057.

[132] X.Y. Liu, E. Specht, Mean residence time and hold-up of solids in rotary kilns, Chem.
Eng. Sci. 61 (2006) 5176–5181. doi:10.1016/j.ces.2006.03.054.

63 | P a g e
Appendix

A1. Calculation of energy balance (Figure 4.14)


Energy yield was calculated by equation A1

𝐻𝐻𝐻𝐻𝐻𝐻𝑖𝑖
𝑦𝑦𝑖𝑖𝐸𝐸 = 𝑦𝑦𝑖𝑖𝑀𝑀 (A1)
𝐻𝐻𝐻𝐻𝐻𝐻𝑤𝑤

where 𝑦𝑦𝑖𝑖𝐸𝐸 and 𝑦𝑦𝑖𝑖𝑀𝑀 energy and mass yields, respectively, and HHV is higher heating value
(kJ/kg). Subscript i is material species (i.e. charcoal), and w is raw wood.
Minimum heat requirement, regarding assumptions of no heat loss and no reaction heat,
that supply energy to heat biomass up to 700 ˚C was calculated by equation A2

700 ˚C
𝑄𝑄𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 = 𝑚𝑚̇𝑤𝑤 ∫25 ˚C 𝐶𝐶𝑝𝑝 𝑑𝑑𝑑𝑑 (A2)

where 𝐶𝐶𝑝𝑝 = 0.1031 + 0.003867𝑇𝑇, (kJ/kg.K) ‡.

Simpson W, TenWolde a. Physical properties and moisture relations of wood. Wood Handb Wood as an Eng

Mater 1999:3.1-3.24. doi:10.1007/s13398-014-0173-7.2.

65 | P a g e

You might also like