Вы находитесь на странице: 1из 86

Single particle ignition of pulverized solid biomass fuels:

experiments and modeling

Gonçalo Guedes de Andrade Costa Simões


Thesis to obtain the Master of Science Degree in

Mechanical Engineering

Supervisor: Prof. Mário Manuel Gonçalves Costa


Co-supervisor: MSc Miriam Estefânia Rodrigues Fernandes Rabaçal

Examination committee
Chairperson: Prof. Viriato Sérgio de Almeida Semião
Members of the Committee: Prof. Pedro Jorge Martins Coelho
MSc Miriam Estefânia Rodrigues Fernandes Rabaçal

May 2016

i
Acknowledgments

Agradeço aos meus orientadores Miriam Rabaçal e Professor Mário Costa pela orientação, amizade e
por tudo o que me ensinaram.
Agradeço ainda ao Duarte Magalhães, Afonso Ferreira e Sr. Pratas pelo apoio que me deram durante
a fase experimental do trabalho. Gostava ainda de agradecer ao Professor Viriato Semião por ter
disponibilizado a câmara de alta-velocidade fundamental para a execução deste trabalho.
Agradeço também aos meus colegas do Laboratório pela amizade e apoio durante este trabalho.
Finalmente agradeço à minha família e amigos pelo apoio contínuo ao longo da minha vida.

ii
Publications produced during the course of this thesis

G. Simões, D. Magalhães, M. Rabaçal, M. Costa, Single Particle Ignition Behavior of Several Biomass
Fuels under Different Atmospheres, Clean Air 2015 – 12th International Conference on Energy for a
Clean Environment, 5 - 9 July 2015, Lisbon, Portugal.


G. Simões, D. Magalhães, M. Rabaçal, M. Costa, Effect of gas temperature and oxygen concentration
on single particle ignition behavior of biomass fuels, accepted for oral presentation on the 36th
International Symposium on Combustion, Seoul, South Korea, 1-5 August 2016. Manuscript under
revision for publication in the Proceedings of the Combustion Institute.

iii
Resumo

O presente trabalho focou-se no estudo da ignição de partículas individuais de combustíveis sólidos


pulverizados. Os testes experimentais foram realizados num queimador de chama-plana McKenna
produzindo um escoamento confinado de produtos de combustão com composição e temperatura
especificas. Partículas de casca de pinheiro, palha, ramos de quivi, videira e de plátano na gama de
212 - 224 µm foram injetadas em diferentes condições de atmosfera. A ignição das partículas
individuais foi analisada para três condições com diferentes concentrações de oxigénio na zona de
ignição, nomeadamente 3.5 %, 5 % e 6.5 % (concentração seca). O modo de ignição e o tempo de
atraso à ignição foram avaliados através da análise de imagens obtidas através de uma câmara de
alta velocidade CMOS. De modo a compreender melhor os fenómenos envolvidos no processo de
ignição foi também feito um estudo numérico. Foi utilizada uma abordagem Lagrangiana-Euleriana
em que a fase gasosa foi modelada através das equações de Navier-Stokes e de transporte de
espécies químicas, e a fase sólida discreta seguiu uma user-defined-function (udf) que considera
sequencialmente as fases de: evaporação, volatilização, combustão do resíduo carbonoso. Para além
das condições estudadas neste trabalho, três condições adicionais (testadas num trabalho anterior
[42]) com diferentes temperaturas na zona de ignição, nomeadamente 1500 K, 1650 K e 1800 K,
foram utilizadas de modo a analisar numericamente o efeito da temperatura da atmosfera no
processo de ignição, Foi selecionado um caso de referência para validar o modelo, que
posteriormente foi utilizado para identificar dos fenómenos dominantes no tempo de atraso à ignição
de partículas sólidas de biomassa. Os resultados experimentais revelaram que (i) algumas partículas
de biomassa, nomeadamente as partículas de palha e de ramos de videira em algumas condições,
apesar do elevado conteúdo em voláteis apresentaram ignição à superfície da partícula, (ii) esta
ocorrência deste tipo de ignição não mostrou uma relação com a composição da partícula ou com a
sua forma/morfologia, no entanto a sua frequência aumentou com um aumento na concentração de
oxigénio (iii) apesar de os tempos de atraso à ignição terem decrescido ligeiramente com um
aumento da concentração de oxigénio, o efeito deste parâmetro é praticamente marginal. Para o caso
de referência selecionado, o modelo numérico mostrou um bom acordo com os resultados
experimentais, ilustrando que o modelo usado neste trabalho, é capaz de reproduzir o processo de
combustão de partículas sólidas isoladas, particularmente o tempo de atraso à ignição. Os tempos de
atraso à ignição calculados numericamente seguiram as principais tendências experimentais: (i) efeito
marginal da concentração de oxigénio, (ii) tempos de ignição menores para condições de temperatura
mais elevada. Finalmente, os resultados numéricos mostraram que o tempo de aquecimento e o
tempo de secagem dominam o tempo de atraso à ignição.

Palavras-chave: biomassa; partículas individuais; técnicas ópticas de alta velocidade; tempo de


atraso à ignição; modo de ignição; modelação numérica.

iv
Abstract

The current work focused on the study of the ignition of single-particles of pulverized solid fuels. The
experimental tests were carried out in an optical flat-flame McKenna burner that produced a confined
laminar flow of combustion products with a specific composition and temperature. Particles of pine
bark, wheat straw, vine, kiwi and sycamore branches were injected in different atmosphere conditions
in the size range of 212 - 224 µm. The ignition of single-particles was analyzed for three different
oxygen concentrations in the ignition zone, namely 3.5 %, 5 % and 6.5 % (dry concentration). The
ignition mode and the ignition delay time were evaluated through the analysis of images obtained by
means of a CMOS high-speed camera. In order to understand better the phenomena involved in the
ignition process an additional numeric study was done. A time-dependent Euler-Lagrange approach
was taken where the fluid phase is treated as a continuum by solving the Navier-Stokes, energy and
the species transport equations, and the discrete solid phase followed an user-defined-function (udf),
which considered sequentially the phases of: evaporation, devolatilization, char combustion. Besides
the conditions studied in this work, three other conditions (tested in a previous work [42]) with different
temperature in the ignition zone, namely 1500 K, 1650 K and 1800 K, were used in order to
numerically evaluate the effect of the temperature on the ignition process. A benchmark case was
selected to validate the model. The model was then used to identify the dominant phenomena on the
ignition delay time of the solid biomass particles.

The experimental results showed that (i) some biomass particles, namely the wheat straw and vine
branch particles under certain conditions, despite their high content of volatiles presented surface
ignition, (ii) this incidence did not seem to show a relation with particle composition or its
shape/morphology, however it increased with increasing oxygen concentrations, (iii) even though the
ignition delay times decreased slightly with increasing oxygen concentrations, the effect of this
parameter was marginal. For the selected benchmark case, the numeric results showed a good
agreement with the experimental data displaying that even a simple model, as the one used, is able to
reproduce in a reasonable way the combustion process of the single-particles, particularly the ignition
delay time. The numeric ignition times followed the same experimental trends: (i) marginal effect of
the oxygen concentration, (ii) decreasing ignition delay times with increasing temperatures. Finally, the
numerical results showed that heating and drying time dominate the particle ignition delay time.

Keywords: biomass; single particle; high-speed imaging techniques; ignition delay time; ignition
mode; numeric modeling.

v
Table of contents

Acknowledgments .................................................................................................................................... ii

Publications produced during the course of this thesis ........................................................................... iii

Resumo ................................................................................................................................................... iv

Abstract.....................................................................................................................................................v

Table of contents ..................................................................................................................................... vi

List of figures ......................................................................................................................................... viii

List of tables .............................................................................................................................................x

Nomenclature .......................................................................................................................................... xi

Subscripts ........................................................................................................................................ xii

Acronyms ........................................................................................................................................ xii

1. Introduction....................................................................................................................................... 1

1.1. Motivation ...................................................................................................................................... 1

1.2. Biomass ........................................................................................................................................ 4

1.3. Fundaments of biomass combustion and ignition ........................................................................ 6

1.4. Previous works.............................................................................................................................. 7

1.5. Objectives ................................................................................................................................... 17

1.6. Thesis outline .............................................................................................................................. 18

2. Experimental materials and methods ............................................................................................. 19

2.1 Fuel preparation and characterization ......................................................................................... 19

2.2 Experimental setup ...................................................................................................................... 19

2.3. Experimental techniques and uncertainties ................................................................................ 23

3. Mathematical model ....................................................................................................................... 28

3.1. Modeling approach ..................................................................................................................... 28

3.2. Gas-phase ................................................................................................................................. 28

3.3. Discrete-phase ............................................................................................................................ 31

3.3.1. Particle motion...................................................................................................................... 31

3.3.2. Heat and mass exchange .................................................................................................... 32

3.4. Volatile combustion ..................................................................................................................... 35

4. Numerics ........................................................................................................................................ 36

4.1. Geometry, mesh and boundary conditions ................................................................................. 36

vi
4.2. Solver options and discretization ................................................................................................ 37

5. Experimental results and discussion .............................................................................................. 39

5.1 Fuel characteristics ...................................................................................................................... 39

5.2. Test conditions ............................................................................................................................ 42

5.3. Ignition mode .............................................................................................................................. 44

5.4. Ignition delay time ....................................................................................................................... 46

6. Numeric results and discussion ......................................................................................................... 48

6.1. Benchmark case setup .............................................................................................................. 48

6.2. Ignition process ........................................................................................................................... 49

6.3. Definition of a numerical ignition criterion ................................................................................... 52

6.4. Comparison between experimental and predicted ignition delay times ..................................... 53

6.5. Analysis of the dominant phenomena on the ignition delay time ............................................... 55

7. Closure .............................................................................................................................................. 58

7.1. Conclusion .................................................................................................................................. 58

7.2. Outlook ........................................................................................................................................ 58

References ............................................................................................................................................ 60

Annex A – Thermocouple measurement correction .......................................................................... 64

Annex B – Image processing procedure ............................................................................................ 70

Annex C – Equipment used ............................................................................................................... 73

vii
List of figures

Figure 1.1 – Projections for (a) the consumption of primary energy split by fuel category and (b) for
the primary inputs into the power generation sector [6]. ......................................................................... 2
Figure 1.2 – Schematic of the carbon cycle resulting from the combustion of biomass [23]. ................ 4
Figure 1.3 – Schematic of the two alternative combustion processes of a solid fuel particle. ............... 6
Figure 1.4 – Comparison of the ration of different thermophysical properties and non-dimensional
groups for CO2 and N2 at 1200 K [35]. .................................................................................................. 15
Figure 2.1 – (a) Schematic representation and (b) photo of the experimental setup. .......................... 20
Figure 2.2 – Premixed flame sheet. ...................................................................................................... 21
Figure 2.3 – Particle feeding system: (1) syringe filled with particles, (2) air ejector, (3) electrical
vibrating motor, (4) feeding tube. .......................................................................................................... 22
Image acquisition system ................................................................................................................... 22
Figure 2.4 – Schematic representation of the temperature probe and thermocouple (values in mm
except when indicated). ......................................................................................................................... 24
Figure 2.5 – Gas sampling probe (values in mm). ............................................................................... 25
Figure 3.1 – Particle and gas phase coupling [49]. .............................................................................. 31
Figure 3.2 – Fuel constituents [50]. ...................................................................................................... 32
Figure 3.3 – Rate of surface reaction as a function of temperature of the solid particle [53]. .............. 33
Figure 4.1 – Geometry of the computational domain. .......................................................................... 36
Figure 4.2 – Mesh and labelled boundaries of the domain. ................................................................. 37
Figure 5.1 – Solid fuels after sieving: a) wheat straw, b) pine bark, c) vine branches, d) kiwi branches,
e) sycamore branches. .......................................................................................................................... 39
Figure 5.2 – SEM images of the solid fuels with x50 zoom factor. a) wheat straw, b) pine bark, c) vine
branches, d) kiwi branches, e) sycamore branches. ............................................................................. 40
Figure 5.3 – Aspect ratios of the biomass fuels computed from the SEM images. Pine bark particles’
aspect ratio is not depicted as these particles were considered quasi-spherical.................................. 40
Figure 5.4 – SEM images of a) kiwi branch particle (zoom x220); b) wheat straw particle (zoom x180);
c) vine branch particle (zoom x170); d) sycamore branch particle (zoom x250); e) pine bark particle
(zoom x250); and f) vine branch particle (zoom x170). ......................................................................... 41
Figure 5.6 – Mean dry O2 concentration profiles along the burner axis above the burner (see Table
5.3)......................................................................................................................................................... 44
Figure 5.7 – Selected images from high-speed cinematography of typical ignition events. a) Gas-
phase ignition of a kiwi branch particle (test 1); b) Surface ignition of a vine branch particle (test 3). 45
Figure 5.8 – Fraction of gas-phase/surface ignition for a) wheat straw particles and b) vine branch
particles. The remaining fuels experienced 100% of gas-phase ignitions over the three test conditions.
............................................................................................................................................................... 45
Figure 5.6 – Ignition delay times for the operating conditions O1-O3 for wheat straw (WS), kiwi branch
(KB), vine branch (VB), sycamore branch (SB) and pine bark (PB) particles in the range 212-224 µm.
Vertical bars represent 98 % confidence statistical error. ..................................................................... 47

viii
Figure 6.1 – Numeric and experimental (a) Dry oxygen mole fraction and (b) gas temperature along
the centerline for the conditions O1-O3 and T1-T3, respectively. ......................................................... 49
Figure 6.4 – Time-history of maximum OH mass fraction in the domain normalized by the maximum in
each condition. a) O1, b) O2, c) O3, d) T1, e) T2, f) T3. The dashed lines represent the experimental
ignition times. ......................................................................................................................................... 53
Figure 6.5 – Experimental and numeric ignition delay times for conditions a) O1-O3 and b) T1-T3. The
vertical bars represent 98 % confidence interval................................................................................... 54
Fig. 6.6 - Decomposition of ignition times in heat-up, drying, heat-up and devolatilization up to ignition
for test conditions a) O1-O3 and b) T1-T3. ........................................................................................... 55
Fig. 6.9 - Drying times as a function of the moisture content of the fuels for conditions T1-T3............ 56

ix
List of tables

Table 1.1 – Comparison of the elemental composition of biomass and coal (wt%, as received) [25]. ... 5
Table 1.2 – Summary of the previous works (continues). ....................................................................... 8
Table 1.3 – Drop tube reactor and entrained flow reactor main characteristics. .................................. 11
Table 1.4 – Main parameters influencing the ignition delay time of solid fuel particles. ....................... 16
Table 2.1 – Imaging settings. ................................................................................................................ 22
Table 4.1 – Summary of the boundary conditions. ............................................................................... 37
Table 5.1 – Properties of the solid fuels. ............................................................................................... 39
Table 5.2 – Summary of the main observations on the biomass fuels, which may affect ignition mode.
............................................................................................................................................................... 42
Table 5.3 – Operating conditions. ......................................................................................................... 42
Table 6.1 – Properties of pine bark [60]. ............................................................................................... 48
Table 6.2 – Assumed volatile matter composition of pine bark for the numerical simulations (%
weight). .................................................................................................................................................. 48
Table 6.3 – Test conditions. .................................................................................................................. 48
Table A.1 – Equipment used and respective working principle. ........................................................... 73

x
Nomenclature

𝐴 - pre-exponential factor (consistent units)


𝐴𝑝 - superficial area of the particle
𝐴𝑣 - pre-exponential factor for the devolatilization process
Bi - biot number
𝐶𝐷 - drag coefficient
𝐶𝑗,𝑟 - molar concentration of species 𝑗 in reaction 𝑟 (𝑘𝑚𝑜𝑙/𝑚3 )
𝑑𝑝 - diameter of the particle
𝐷 - diffusion constant
𝐷𝑖,𝑚 - mass diffusion coefficient of species i
𝐷𝑇,𝑖 - thermal diffusion coefficient
𝐸 - activation energy for the reaction (J/kmol)
𝐸𝑎,𝑣 - activation energy for the devolatilization process
𝑭𝑫 - drag force per unit mass
h - enthalpy
𝑘 - thermal conductivity
𝑘𝑓,𝑟 - forward rate constant for reaction 𝑟
𝑘𝑏,𝑟 - backward rate constant for reaction 𝑟
𝐾𝑣 - mass transfer coefficient
m - mass
𝑀𝑤 - molecular weight
ℳ𝑖 - symbol denoting species 𝑖
p - static pressure
𝑅 - universal gas constant (J/kmol-K)
𝑅̂ - arrhenius molar rate of reaction
𝑅𝑖 - net rate of species i created from chemical reactions
Re - relative Reynolds number of the particle
𝑆𝑖 - net rate of species i added from the dispersed phase
𝑆 - source term
t - time
T - temperature
𝑣 - velocity
𝑣′𝑖,𝑟 - stoichiometric coefficient for reactant 𝑖 in reaction 𝑟
𝑣′′𝑖,𝑟 - stoichiometric coefficient for product 𝑖 in reaction 𝑟
Y - local mass fraction of species i

𝛽 - temperature exponent

xi

𝜂𝑗,𝑟 - rate exponent for reactant species 𝑗 in reaction 𝑟
′′
𝜂𝑗,𝑟 - rate exponent for product species 𝑗 in reaction 𝑟
𝜇 - molecular viscosity
ρ - density
𝜏̅̅ - stress tensor

Subscripts

p - particle p
i - species i
r - reaction r
eff - effective
H2O - moisture fraction
v - volatile fraction
C - fixed carbon fraction
ash - ash fraction

Acronyms

1D - one dimensional
2D - two dimensional
CCS - carbon capture and storage
CFD - computational fluid dynamics
udf - user defined function
DTF - drop tube furnace
EFR - entrained flow reactor
CMOS - complementary metal oxide semiconductor
daf - dry, ash free
fps - frames per second
ID - internal diameter
pc - pulverized coal
wt - weight

xii
1. Introduction

1.1. Motivation

The planet is getting warmer. Presently, the effects of global warming are not just a distant
reality, but are happening all around the globe. The heat is melting glaciers and sea ice, rising
sea level and temperature; it is causing extreme weather phenomena, such as droughts,
floods, hurricanes and other storms; and it is compromising air quality and fresh water
supplies, affecting both health and agriculture. Averaged over all land and ocean surfaces,
temperatures rose approximately 0.85 degrees Celsius from 1880 to 2012 [1]. The
international community has accepted the limit increase of 2 degrees Celsius above pre-
industrial levels, as necessary to avoid the most severe and pervasive implications of climate
change [2]. Based on the available data, there is an overwhelming consensus that the main
cause of global warming is the increased level of greenhouse gases in the atmosphere [3].
Greenhouse gas emissions are largely dominated by carbon dioxide emissions. For instance,
in 2012 carbon dioxide emissions in the U.S. accounted for more than 80% of the total
emissions of greenhouse gases in that country [4]. Carbon dioxide emissions thus constitute
the main driver of global warming. CO2 emissions mostly result from the combustion of fossil
fuels for the energy and transportation sectors.

The obvious way to reduce CO2 emissions would then be to decrease the consumption of
fossil fuels, however with the growing world population and economy, primary energy demand
is expected to increase 40% between 2012 and 2035 [5,6]. This further aggravates the
problem: not only is it necessary to decrease carbon dioxide emissions, but also that has to
be achieved while the main source of carbon dioxide emissions is increasing.

Coal is the most polluting fossil fuel, and thus reducing the share of consumption of coal
would also be a path to reduce carbon dioxide emissions. Instead, coal consumption grew by
3% in 2013, making it the fastest-growing fossil fuel that year. Coal’s share of global primary
energy in that year reached 30.1 %, the highest since 1970 [7]. Due to its wide availability,
abundant reserves, and competitively low prices coal is and will continue to one of the major
energy resources in the long term, mainly in the power generation sector where it dominates.
The trends for the increasing consumption of coal can be seen in Fig. 1.1 [6]. Figure 1.1a)
shows data and the projections for the consumption of primary energy categorized by fuel
type, and Fig. 1.1b) the fraction of each fuel in the power generation sector. Even though,
coal looses some share of the power generation sector, it is still expected to dominate in this
sector in the next 20 years, and since the total input of primary energy into the power
generation sector increases the actual consumption of coal increases as well.

In this sense, the combustion of coal is and will continue to be one of the main players in what
concerns global warming. Besides greenhouse gases, CO2 and CH4, coal combustion also
results in other hazardous emissions, such as sulphur and nitrogen oxides, which cause acid
rain, ozone depletion and respiratory problems. Since coal is going to continue to be

1
massively used at least in the following decades, it becomes evident that it is necessary to
explore alternatives to conventional coal combustion in order to reduce its emissions. In this
context, technologies such as Carbon Capture and Storage (CCS) and/or co-firing of coal
with biomass are gaining attention due to their potential application in the short term.

(a) (b)

Figure 1.1 – Projections for (a) the consumption of primary energy split by fuel category and
(b) for the primary inputs into the power generation sector [6].

Carbon capture and storage (CCS) comprises compression of the CO2 rich combustion
product to a supercritical state, prior to its transportation and storage at geological depth. The
main CCS technologies are (i) scrubbing of the CO 2 gas for CO2 removal, (ii) integrated
gasification combined cycle with a shift reactor, and (iii) oxy-fuel combustion [8]. The latter
has significant advantages over the other two. In particular, it does not require CO2 capture
prior to compression and more importantly it is suitable for retrofit in existing power plants.
Oxy-fuel combustion was firstly proposed by Abraham and his co-workers [9] and consists in
replacing the air in combustion with oxygen diluted by a recycled flue gas in a conventional
boiler. This generates a CO2 rich stream, which is almost sequestration-ready. Reviews
[10,11] agree that O2/CO2 pulverized coal combustion is economically promising and
technically feasible with current technologies, reducing the risks associated with the
development of new technologies. In spite of this, successful implementation of oxy-fuel
combustion may not be direct and requires a detailed understanding of the effects of
replacing N2 with CO2 in the oxidizer stream. In previous studies [12,13] oxy-fuel conditions
were shown to modify important furnace operation parameters, such as heat transfer and gas
temperature, char burnout and hinder flame stability. Comprehensive knowledge of this
technology is crucial in the modeling of oxy-fuel combustion at an industrial scale.

2
Another interesting alternative to conventional coal combustion, with the objective of reducing
CO2 emissions, is co-firing of coal with biomass. This consists of burning mixtures of coal and
biomass instead of burning pure coal. Biomass is considered to be one of the key future
renewable resources, being the most important source of energy for three quarters of the
world’s population living in developing countries. Presently, the consumption of biomass
already corresponds to a share of 14% of the world’s primary energy [14]. However, the
properties of biomass fuels lift some problems in the combustion of biomass at an industrial
scale. Factors such as its low carbon content, low heating value, and high moisture content,
compared to those of coal, make biomass fuels difficult to ignite and can lead to problems in
flame stability. The combustion of pure biomass at a large scale would imply the construction
of new and expensive biomass specific power plants [15]. Co-firing arises as an alternative
that allows the use of biomass fuels in existing power plants. Since biomass is a renewable
fuel, co-firing coal with biomass results in a significantly smaller carbon footprint per amount
of energy released, as compared to the combustion of pure coal. Evidence also suggests that
there is a synergetic behavior in the combustion of coal and biomass. Overall, the low
pyrolysis temperature and high volatile content of biomass tend to improve the ignition and
combustion characteristics of the mixture [16,17]. Additionally, in most cases, emissions of
SO2 and NOx were reduced [18]. Successful tests and demonstrations have been carried out
[19], but there still exist several unresolved economical and technical challenges. For
instance, co-firing represents by far the cheapest means of renewable power generation in a
large fraction of the situations where it is feasible [20], however it is still more expensive than
neat coal firing. Technical issues include: fuel preparation, storage and delivery; Ash
deposition; Fouling; Corrosion; Pollutant formation; Impacts on SCR systems [20]. In addition,
most co-firing power plants use wood derived fuels, which puts pressure on forest resources.
Using alternative solid fuels such as agricultural residues can minimize this problem, however
these fuels pose a number of operational problems, including lower particle burnout and ash-
related impact [20].

In summary, carbon capture and biomass co-firing are interesting means to reduce the CO2
footprint associated with energy generation [4,21]. The current work focuses on the second
alternative, i.e., co-firing coal with biomass. Previous studies [12,22] have shown that both
technologies affect conditions inside the boiler, in particular ignition of the solid fuel particles,
which is a key feature of the combustion process, and has a critical influence over flame
stability. Specifically, the objective is to study the ignition behavior of single biomass particles
under different conditions.

3
1.2. Biomass

Biomass captures the energy of the sun and fixes carbon through the process of
photosynthesis. The biomass can then be burned to release the energy stored (as chemical
energy), releasing also in the process carbon to the atmosphere, generally as CO2 or CO.
This CO2 is again re-absorbed by the growing biomass, thus closing the carbon cycle, with no
net increase in atmospheric CO2. This is schematically represented in Fig. 1.2. Fossil fuels,
on the other hand, result from the transformation of biomass, at high temperature and
pressure, over several millions of years. They too can be burned releasing energy and carbon
that will eventually be recaptured in the next generation of fossil fuels. However, whilst the
period of the carbon cycle for fossil fuels is in the order of millions of years the same period
for biomass is only in the order of decades. In this sense, the carbon cycle for biomass is
neutral at a human time-scale and biomass is considered as a renewable fuel.

Figure 1.2 – Schematic of the carbon cycle resulting from the combustion of biomass [23].

Biomasses can be grouped according to their source [24]. The major groups are (i) woody,
which includes pine chips, an energy crop, willow, (ii) herbaceous, in which two energy crops
are listed - Miscanthus and Switch Grass, (iii) agricultural residues, including wheat straw,
rice husks, palm kernel expeller, bagasse (a sugar cane residue), olive residue or olive cake
(the waste from olive oil mills) and (iv) animal waste, such as cow dung. Based on this
categorization within each group biomass fuels may have extremely distinct properties.
Alternatively, biomass fuels can be grouped according to other parameters, such as their
structural composition.

Structurally, biomass consists cellulose, hemicellulose and lignite. The relative fraction of
each of these components depends on the biomass origin and stage of development and
obviously affects the combustion behavior of the fuel. Cellulose consists of structures of

4
macromolecules composed by semi-crystalline chains associated other molecules delimited
by hydrogen atoms. Hemicellulose is also a macromolecule composed of different sugars,
however, while cellulose is crystalline and strong, hemicellulose has a random amorphous
structure with little strength. Lignite, on the other hand, is composed by an aromatic polymer
synthesized from phenylpropanoid precursors [25]. In general, the thermal decomposition of
hemicellulose occurs at temperatures between 150 oC and 350 oC, and that of cellulose
occurs at temperatures between 300 oC and 350 oC. Lignite decomposes at higher
temperatures that can reach 500 oC [26,27]. Categorizing biomass fuels according to their
structural composition gives a better understanding about their combustion behavior than
simply grouping them according to their source.

As for coals, biomass fuels can also be characterized according to their ultimate and
proximate analysis. Table 1.1 presents the elemental analysis for some biomass fuels and
also, for comparison purposes, for coals from different ranks [25]. In terms of their ultimate
analysis, biomass fuels are mainly composed by carbon, hydrogen and oxygen, with their
relative fractions varying widely for different fuels. Other elements such as nitrogen, sulphur
and chlorine typically have mean dry weight fraction below 1% [28]. Carbon and hydrogen are
oxidized during the combustion reaction generating carbon dioxide and water vapor. The
oxygen, present in the organic bridges, is released during the thermal degradation process
and also contributes for the oxidation of the other elements.

Table 1.1 – Comparison of the elemental composition of biomass and coal (wt%, as received)
[25].

Herbaceous and Bituminous coal/sub-


Wood and woody
Fuel agricultural bituminous coal/ lignite
biomass
biomass coal
Moisture 4.7 – 62.9 4.7 – 62.9 0.4 – 20.2
Volatile Matter 30.4 – 79.7 41.5 – 76.6 12.2 – 44.5
Fixed Carbon 6.5 – 24.1 9.1 – 35.3 17.9 – 70.4
Ash 0.1 – 8.4 0.1 – 8.4 5 – 48.9
C 48.7 – 57.0 42.2 – 58.4 62.9 – 86.9
O 32.0 – 45.3 34.2 – 49.0 4.4 – 29.9
H 5.4 – 10.2 3.2 – 9.2 3.5 – 6.3
N 0.1 – 0.7 0.1 – 3.4 0.5 – 2.9
S 0.01 – 0.42 0.01 – 0.60 0.2 – 9.8

Concerning the proximate analysis, it can be seen from table 1.1 that biomass fuels normally
contain high volatile matter content and low fixed carbon content when compared with coal. In
addition to this, the volatile compounds released from biomass and coal are different.
Typically, the volatiles from biomass are lighter than those of coal [29]. There are also
differences in the chars of these fuels. It is generally accepted that biomass chars are more

5
reactive than coal chars due to their unordered structure [30]. The ash and moisture contents
can vary widely, but the moisture content is typically much higher than that of coal. The
proximate composition of the fuel affects the ignition and combustion behavior of the fuel. In
addition to the differences between biomasses and coals seen in table 1.1, other important
differences have been pointed by Riaza et al [31]. Namely they noted that (i) raw biomasses
have a highly fibrous nature, (ii) the biomass particles are less dense than coal particles,
therefore the total mass burnt for the same nominal particle size is lower, (iii) and that
biomass fuels have much lower heating values than coals.

1.3. Fundaments of biomass combustion and ignition

Biomass combustion involves a number of physical/chemical aspects of high complexity. The


combustion process of a solid fuel is schematically presented in Fig. 1.3.

Figure 1.3 – Schematic of the two alternative combustion processes of a solid fuel particle.

The process starts with an endothermic heat-up and drying phase, during which temperature
of the solid fuel particle increases and its moisture decays as water vapor is released to the
atmosphere. It is easy to understand that high moisture contents can delay/hinder the ignition
of the solid fuel. After this phase, generally a phase of devolatilization occurs. Further heating
leads to the thermal decomposition of the biomass fuel, resulting in the release of light gases
and heavier molecular fragments that either vaporize as tars or remain in the solid fuel
structure. Simultaneously, a fraction of the original fuel is converted into char. This process,
by which volatile compounds (light gases and tars) leave solid particle as it is transformed into
char, is an important step in the devolatilization phase [32]. Generally, for biomass fuels
ignition occurs in the gas phase from the reaction between the volatile gases and the oxidant
gas. The volatiles burn in a cloud around the particle, providing further heating to the particle.
The volatile flame is sustained until all of the volatile gases released from the particle have
been burnt. Upon extinction of the volatile flame, or in some cases before that (when oxygen
is able to diffuse to the surface of the particle), heterogeneous char oxidation occurs. In other
instances, depending on the conditions and on the properties of the solid fuel, heterogeneous

6
ignition can occur before a significant amount of volatiles has been released. In these cases,
upon the heterogeneous ignition, the volatiles and char burn simultaneously at the particle
surface.

The current work focuses mainly on the ignition of biomass fuels as this is a determining
feature of the combustion process, as discussed previously. Ignition is defined as the non-
stationary process in which radical formation starts taking place in a reactant mixture (initially
non-reactive), with subsequent heat release and rapid increase in temperature, originating a
self-sustained combustion process [33]. Ignition of solid fuel particles may occur
homogeneously, heterogeneously, or under a combination of both modes. Homogeneous
ignition results from the interaction between the volatiles released by the particle and the
oxidant both in the gas phase (hence the designation – homogeneous). Heterogeneous
ignition, on the other hand, results from surface oxidation, in a reaction between the solid
particle and the oxidant gas (hence the name – heterogeneous).

Howard and Essenhigh [34] studied the ignition behavior of single bituminous coal particles in
air exposed to high heating rates. In general, they observed homogeneous ignition for the
particles, however they saw that at high heating rates, the finite time lag before the release of
the volatiles became important; and if the particles reached a sufficiently high temperature
rapidly, it was possible that the particles ignited and burned by direct oxygen attack on the
solid surface before devolatilization set in. It became obvious to Howard and Essenhigh that
other parameters besides to the composition of the fuel, such as the operating conditions,
were important in the ignition process.

Several studies explore some of the parameters that influence the ignition of solid fuel
particles, namely the heating rate, oxygen concentration, coal type and temperature of the
ambient gas. An overview of a selected list of these works is presented in the following
section.

1.4. Previous works

As mentioned above, several studies have been conducted on the subject of ignition of solid
fuels (and the parameters which affect ignition). Table 1.2 lists a summary of a selection of
previous works, which includes reactor type, fuel injection type, fuel, and main results. Note
that not all the results are listed in this table, since several results from different works
overlapped.

Most works focus on the ignition of coal particles, since the combustion of biomass fuels at an
industrial scale is a recent subject. Ignition of biomass fuels is considerably different to that of
coal. Not only have these fuels different compositions, but also they are also distinct in terms
of structure, which affects the combustion process, and in particular the early stages of the
devolatilization process. As a consequence, the ignition behavior of these fuels can be
significantly different.

7
Table 1.2 – Summary of the previous works (continues).
Work Injection Reactor Fuel Studied parameters Main results
[35] The CO2 background gas delays ignition delay time,
Molina, Shaddix Atmosphere the compared to the N2 background
Ignition and devolatilization of High-volatile O2/N2 vs O2/CO2
Jet/spray EFR Increasing O2 concentration accelerates the ignition
pulverized bituminous coal particles bituminous coal Atmosphere
O2 concentration The CO2 background gas has no measurable effect
during oxygen/carbon dioxide coal
combustion on the duration of the volatile combustion
[13] High-volatile Atmosphere
Shaddix, Molina The CO2 background gas delays ignition delay time,
bituminous coal O2/N2 vs O2/CO2
Particle imaging of ignition and Single-particle EFR the compared to the N2 background
Sub- bituminous Atmosphere
devolatilization of pulverized coal Increasing O2 concentration accelerates the ignition
coal O2 concentration
during oxy-fuel combustion
Feeding rate Ignition delay times decreased with increasing
Atmosphere feeding rates until a minimum was reached increasing
[36] High-volatile temperature thereafter
Liu et al. bituminous coal Particle size The ignition delay time increased with increasing
Pulverized coal stream ignition delay Jet/spray EFR
Sub- bituminous Atmosphere particle size and decreasing atmosphere temperature
under conventional and oxy-fuel The effect of O2 concentration was weak
coal O2 concentration
combustion conditions
Atmosphere Replacing N2 with CO2 slightly increased the ignition
O2/N2 vs O2/CO2 temperature
The swelling bituminous coal consistently
experienced distinct volatile matter and residual char
[37] Lignite
combustion phases
Levendis et al. Sub- bituminous
Atmosphere The non-swelling but often fragmenting sub-
Combustion behavior in air of single Single-particle DTF coal
O2 concentration bituminous coal experienced variable behavior
particles from three different coal Bituminous coal The consistently fragmenting lignite coals did not
ranks and sugarcane bagasse
experience detached volatile flames
Bagasse experienced two phase combustion
8
Table 1.2 – Summary of the previous works (continues).
Work Injection Reactor Fuel Studied parameters Main results
Lignite CO2 background atmosphere and high O2
[38] concentration favor simultaneous combustion of volatiles
Khatami et al. Sub-bituminous Atmosphere
and char at the particle surface
Combustion behavior of single coal O2/N2 vs O2/CO2
Single-particle DTF Higher ranked coals required higher O2 concentration
particles from three different coal Bituminous coal Atmosphere
in the O2/CO2 atmosphere to match their temperature
ranks and sugarcane bagasse in Sugarcane O2 concentration
behavior in air, whilst sugarcane needed the least
O2/N2 and O2/CO2 atmospheres bagasse amount of O2
Semi-anthracite
High-volatile Ignition temperatures are higher and burnout lower
Atmosphere
[17] bituminous coal O2/N2 vs O2/CO2
for the O2/CO2 atmosphere
Riaza et al. Jet/spray EFR Semi-anthracite + Atmosphere
Ignition temperatures decreased and burnout
Oxy-fuel combustion of coal and Olive waste increased with increasing O2 concentration
O2 concentration
biomass blends High-volatile The addition of biomass to the coal decreased
Blend composition
bituminous anthracite + ignition temperature and increased burnout
Olive waste
Bituminous coal ignited homogeneously and lignite
[39] High-volatile
Atmosphere fragmented extensively, igniting heterogeneously
Khatami et al. O2/N2 vs O2/CO2 thereafter.
bituminous coal
Ignition characteristics of single coal Single-particle DTF Atmosphere Ignition delay times were longer in O2/CO2
Sub-bituminous atmosphere than for O2/N2 atmosphere
particles from three different ranks in O2 concentration
coal
O2/N2 and O2/CO2 atmospheres Increasing O2 concentration decreased ignition delay
times
High rank coals ignited heterogeneously, whilst
[40] Atmosphere bituminous coal ignited homogeneously
Riaza et al. Bituminous coal
O2/N2 vs O2/CO2 Replacing N2 with CO2 slightly increased the ignition
Single particle ignition and Single-particle DTF Semi- anthracite
Atmosphere temperature, decreased combustion temperatures and
combustion of anthracite, semi- Anthracite increased combustion times
O2 concentration
anthracite and bituminous coals in air
Increasing O2 concentration decreased ignition
and
temperature
9
Table 1.2 – Summary of the previous works.
Work Injection Reactor Fuel Studied parameters Main results
Bituminous coals showed lower ignition temperatures
than lignite coals
[41] Atmosphere
Heterogeneous ignition was observed for the lower
Yuan et al. temperature
Lignite temperatures and hetero-homogeneous was observed
The transition of heterogeneous- Jet/spray EFR Atmosphere
Bituminous coal for the higher temperatures
homogeneous ignition of dispersed O2 concentration
The increase in temperature triggered the transition
coal particles
from heterogeneous to hetero-homogeneous ignition
mode
Sugarcane
Increasing O2 concentration enhanced the
[31] bagasse Atmosphere
combustion intensity, decreased burnout times and
Riaza et al. Pine sawdust O2/N2 vs O2/CO2
increased the temperature of the burning char
Combustion of single biomass Single-particle DTF Torrefied pine Atmosphere
Burning times of the volatiles and the ignition delay
particles on air and oxy-fuel sawdust O2 concentration
times increased with the increase of particle density and
atmospheres Olive residue size
High-volatile
[42] bituminous coal Atmosphere The ignition delay time increased with increasing
Magalhães et al. Pine bark temperature particle size and decreasing atmosphere temperature
Single-particle EFR
High speed imaging of single particle Wheat straw Particle size Bituminous coal and pine bark ignited
ignition behavior of biomass fuels homogeneously and wheat straw heterogeneously
[43]
Zou et al. Atmosphere
Ignition behaviors of pulverized coal O2/N2 vs O2/H2O Higher H2O enhanced gasification and steam shift
particles in O2/N2 and O2/H2O Single-particle DTF Bituminous coal Atmosphere reactions, forming H2 and CO around the particle, which
mixtures in a drop tube furnace-using O2 concentration tends to reduce ignition delay times
flame monitoring techniques
10
Experimental studies on solid particle ignition are typically carried out either in drop tube
furnaces (DTF) [31,37-40,43] or in entrained flow reactors (EFR) [13,17,35,36,41,42]. In the
first type of reactor both convection and radiation play an important role. These reactors allow
an easy control of the combustion atmosphere, but do not permit an easy optical access to
the reaction zone, even though some recent works [37-40] have used DTF’s with optical
windows. In entrained flow reactors the dominant heat transfer mechanism is convection.
These reactors replicate better the conditions inside an industrial boiler in the sense that the
cold particles are injected in a cold stream into a very hot combustion product gas
atmosphere, enabling high heating rates. These reactors have the advantage of allowing an
easy optical access. Some of the characteristics of these reactors are listed in table 1.3.

Table 1.3 – Drop tube reactor and entrained flow reactor main characteristics.

Equipment
Characteristic
Drop tube reactor Entrained flow reactor
Dominant Heat transfer
Convection and Radiation Convection
mechanism
Maximum temperature (K) ~ 1300 ~ 1800
Maximum heating rate
~ 104 ~ 105
(K/s)
Gas composition Broad Limited
Optical accessibility Difficult Easy
Parameters control Easy Easy

Entrained flow reactors are particularly interesting when studying ignition, since most
experimental methods to identify the onset of ignition are based on optical diagnostic
techniques. Early studies of Howard and Essenhigh [34] defined the ignition event based on
the percentage of loss of carbon and volatiles during the early moments of combustion,
however, since then most works have used optical diagnostics to characterize this event.
Molina and Shaddix [35] defined the onset of ignition based on the CH* chemiluminescence
(although some interference of blackbody emission from hot soot and from the coal/char
particle was noted). In a more recent work, Shaddix and Molina [13], measured the ignition
delay time of individual coal particles by capturing the visible light signal emitted by igniting
particles, and defined the ignition onset as the point at which 60 % of the maximum luminosity
intensity was reached. The definition of ignition based on a fraction of the maximum of visible
light was also used by other authors [41, 42]. This technique is easy to use and the visible
light signal is strong and easy to capture. Although soot and char emission are also included
in the visible light signal captured this has proved to be a good indicator of ignition. Other
works [38,39] also use visible light to characterize ignition but simply define it as the time
between injection and first visible light signal. Riaza et al. [40] have applied high-speed

11
cinematography and three-color-pyrometry to evaluate the temperature-time history of coal
particles. In this work, the ignition delay time was defined as the time at which the maximum
particle temperature gradient was recorded.

In these works particles are injected either in jets [17,35,36,41] or as single particles
[13,31,37-40,42,43] (in both types of reactors). Jets have the advantage of being closer to
reality (in terms of industrial applications), however inter-particle effects become important
(for example radiation exchange between particles) which hampers modeling of the
combustion process. Sprays also enable injection of blends of fuels, which obviously is very
relevant in the context of co-firing coal with biomass. Studies on the ignition and combustion
properties as well as emissions of the blends can be performed. Nevertheless, to fully
understand and model the combustion behavior of blends or sprays of particles, firstly it is
essential to comprehend the combustion process of isolated particle. Single-particles studies
have a more fundamental nature. In these studies usually a certain parameter is isolated and
varied, while the other are kept, to study the influence on the ignition and combustion
processes of that particular parameter. These studies are very important in modeling the
combustion and ignition of solid fuel particles.

Parameters influencing the ignition of solid fuels

Several parameters influence the ignition process: the initial pressure and temperature of the
reactant mixture and its composition; the energy of the ignition source (as well as the spatial
and temporal distribution of this source, in the case of forced ignition); the velocity, and the
thermal and transport properties of the reactant mixture [36]. It is generally agreed that
parameters such as the heating rate, gas temperature and composition, and fuel type and
mean diameter, have significant effects on the ignition of solid fuels [22,41,44,45]. The main
parameters that influence the ignition of solid fuels are discussed in the following sections.

Effect of the surrounding gas temperature

Yuan et al. [41] burned bituminous coal particles streams in an entrained flow reactor
(Hencken burner) under three different ambience temperatures, from 1200 K to 1800 K. They
observed a general decrease in ignition delay time with increasing atmosphere temperature.
From 1200 K to 1500 K a steep reduction of the ignition delay was verified, which was
explained as a consequence of the transition of heterogeneous to hetero-homogeneous
ignition between those temperatures. The increase in temperature accelerated the
devolatilization process, which triggered the transition in ignition mode. Further increasing the
temperature from 1500 K to 1800 K resulted in an additional but smaller reduction of the
ignition delay. In this case, this was explained with the earlier release of volatiles as a
consequence of the higher temperature. The nature of the ignition mode was maintained,

12
however the devolatilization was enhanced by the increase of ambience temperature, thus
explaining the somewhat smaller effect of temperature in the second temperature step. Liu et
al. [36] studied diluted coal stream ignition and observed a comparable behavior with ignition
delay times increasing from 1320 K to 1230 K. Similarly, Magalhães [42] burned bituminous
coal and two biomass fuels, pine bark and wheat straw, under different temperature
conditions and observed the same behavior. Magalhães also observed that at high
temperatures the effect of the composition of the fuel seemed to loose importance and the
ignition delay times of different fuels were closer to one another.

Effect of the surrounding gas composition

Previous studies, focused on the effect of the atmosphere composition over the ignition of
solid fuels, typically assess the effect varying O2 concentration and the effect of changing the
background gas (from N2 to either CO2 or H2O).

The effect of the surrounding gas mixture composition can be explained by three main
aspects: (i) the reactivity of the mixture; (ii) the thermophysical and transport properties of the
mixture; and (iii) the chemical reactions that may occur due to the composition of the mixture.

Effect of O2 concentration

In several works [13,35-41] the effect of the oxygen concentration has been studied. In
general decreasing oxygen concentration was seen to increase ignition delay times, decrease
ignition temperature and favor homogeneous ignition.

Shaddix and Molina [13] explained the behavior of ignition delay time based on the adiabatic
thermal explosion theory, which for a one-step overall reaction, describes the auto-ignition
time as,

𝑐𝑣 (𝑇0 2 /𝑇𝑎 )
𝜏𝑖 = (1)
𝑞𝑐 𝑌𝐹,0 𝐴 𝑒𝑥𝑝(−𝑇𝑎 /𝑇0 )

where 𝑐𝑣 is the specific heat, 𝑇0 is the initial temperature of the local fuel/air mixture, 𝑞𝑐 is the
combustion heat release per mass of fuel, 𝑌𝐹,0 is the initial mass fraction of fuel and the
reactivity of the fuel/air mixture is given by 𝑘 = 𝐴 𝑒𝑥𝑝(−𝑇𝑎 /𝑇0 ). Since at high temperatures the
specific heats of oxygen and nitrogen are close, the effect of increasing O 2 concentration in
the O2/N2 atmosphere can only come from the characteristic reactivity of the local mixture.
Indeed, higher O2 concentrations increase the characteristic reactivity resulting in a decrease
of ignition delay time. Previous studies of bituminous coal ignition in entrained flow reactors
[13,35], and drop tube furnaces [37-40], have consistently explained this behavior as a
consequence of the higher reactivity of the fuel-oxidizer mixture. Khatami et al. [38] have also

13
observed that as O2 concentration in N2 increased from 20% to 100%, the ignition
temperatures decreased 100-150K. Riaza et al. [31] studied the oxy-fuel combustion of
several biomass fuels and observed a similar behavior to that of coal, i.e., decreasing ignition
delay time with increasing oxygen concentration.

O2 concentration was also observed to have an effect over the ignition mode. Khatami et al.
[39] reported that high oxygen molar concentrations induced heterogeneous ignition for sub-
bituminous coal. They stated that at oxygen mole fractions higher than 40 % fragmentation
occurred followed by heterogeneous ignition of the fragments.

Effect of background gas

In several works alternative atmospheres (to conventional combustion atmosphere) are


studied. In most cases, the nitrogen is replaced with carbon dioxide to simulate oxy-fuel
conditions. As mentioned in section 1.1, oxy-fuel conditions are very interesting in the context
of carbon capture and storage, which allows the reduction of the CO2 emissions and which
has consequently stimulated a growth spurt in the number of works on this subject. In some
recent works nitrogen is replaced with water vapor to study ignition and combustion of solid
fuels in the so-called oxy-steam conditions. Even though, neither oxy-fuel nor oxy-steam
conditions are studied in detail in the current work, the review of these works is still relevant to
fully understand the mechanisms by which changing the atmosphere can change the ignition
and combustion behavior of solid fuels.

In general, in the O2/CO2 atmosphere ignition times were delayed, and ignition temperatures
increased slightly in comparison with the O2/N2 atmosphere. In the O2/H2O atmosphere
ignition occurred sooner than in conventional atmosphere conditions.

Molina and Shaddix [35] studied the ignition and combustion behavior of bituminous coal
under both conventional and oxy-fuel conditions. To explain the effect of changing the
background atmosphere they looked at the phenomena that occur after injection of the coal
into the flow reactor. The first phenomenon is particle heat-up from room temperature to a
gas temperature at which volatiles begin to be released. They showed that for the case of a
non-reactive particle the only gas property that affects the initial particle heat-up is the thermal
conductivity. Comparing the thermophysical properties of N 2 and CO2 they observed that the
ratio of thermal conductivities was close to one (see Fig. 1.4), meaning during the initial
moments there should be no significant difference in particle heating rates for the two
different diluents. Therefore, the initial release of coal volatiles should occur at the same time
and at the same rate for the two diluents.

14
Figure 1.4 – Comparison of the ration of different thermophysical properties and non-
dimensional groups for CO2 and N2 at 1200 K [35].

After the heat-up phase, the devolatilization starts to take place. The volatiles released from
the particle then start to react to form radicals, which in turn react releasing heat. The
temperature starts to increase to generate a self-sustained reaction. This is the basis of the
ignition process. Different reacting mixtures react differently to the energy released by the first
radical reactions. The heat capacity (𝑐𝑝 𝜌) of the reacting mixture gives a measure of the
thermal sink for any heat that is chemically released. Molina and Shaddix [35] observed that
at high temperatures this product was 1.7 times larger for CO2 than for N2 (see Fig. 1.4),
indicating that in the O2/CO2 atmosphere more energy needs to be released to reach the self-
sustained combustion reaction. They used this argument to justify the delay of ignition verified
in O2/CO2 atmosphere compared to the O2/N2 atmosphere. Other authors [38,39] used the
same argument to explain this effect. Liu et al [36] and Riaza et al. [40] also observed a slight
increase in ignition temperature from conventional to oxy-fuel conditions.

In a later work, Shaddix and Molina [13] added that the O2/CO2 atmosphere had a tendency
to suppress radical formation, which also contributed to the delay of the ignition. This
influence of CO2 on ignition chemistry resulted from the reduction of the primary CO oxidation
reaction rate, 𝐶𝑂 + 𝑂𝐻 ↔ 𝐶𝑂2 + 𝐻, and enhancement of key recombination reactions, such

as 𝐻 + 𝑂2 + 𝑀 ↔ 𝐻𝑂2 + 𝑀. The efficiency of CO2 in these reactions is between 2 - 3 times

that of N2 [46].

More recently, Zou et al. [43] have investigated the ignition of pulverized coal particles in a
O2/H2O atmosphere. They observed a smaller ignition delay time in the presence of H 2O, and
higher particle temperature at the location of ignition in H2O than that in N2 at an identical
oxygen concentration, which they attributed mainly to the steam shift reaction, 𝐶𝑂 + 𝐻2 𝑂
↔ 𝐶𝑂 + 𝐻2 .

Effect of the solid fuel composition and properties

The composition of the solid fuel will obviously influence the ignition and combustion behavior
of the solid fuel. For instance, it is generally assumed that fuels with high volatile content/low

15
fixed carbon will ignite homogeneously. Despite this, Magalhães [42] saw in his work that
even biomass particles with high volatile content at high temperature conditions seemed to
ignite superficially. Some other works have assessed the potential catalytic effect of the ash
in the devolatilization process. In fact, in conditions with low temperatures/heating rates, the
presence of Na and K has been shown to accelerate the pyrolysis [48,49], and thus can
potentially reduce ignition delay times. Another important parameter is the moisture content.
As discussed in section 1.3, a high moisture content can be expected to delay ignition since
the drying phase is highly endothermic. In spite of this, few studies address the effect of the
moisture content. This is easy to understand taking into account that most works focus on
coals, which typically have low moisture contents. With the introduction of biomass fuels, rich
in moisture, this will become a more important issue.

In his work Magalhães [42] also analyzed the influence of the devolatilization kinetic
constants, the density and specific heat of the solid fuel with a sensitivity analysis. He saw
that even though the activation energy had an important weight the pre-exponential factor
presented a marginal effect. The specific heat showed a reduced effect on the ignition delay
time. The density, which can significantly vary in biomass fuels, had a much more
pronounced effect.

Summary of the parameters influencing the ignition delay time of


biomass and coal

Table 1.4 summarizes the effect of the main parameters on the ignition delay time of solid
fuels described above.

Table 1.4 – Main parameters influencing the ignition delay time of solid fuel particles.

Parameter Ignition delay time (tig)


Temperature of the atmosphere Temperature tig
Atmosphere O2 concentration O2 concentration tig
composition
Oxy-fuel conditions CO2 concentration tig
Oxy-steam conditions H2O concentration tig
Fuel type Coal coal rank tig
Biomass Insufficient data

Previous studies have addressed a broad range of parameters that can influence coal particle
ignition. However, most studies focused on coal are based in coal jets. In the case of
biomass, studies are scarce and none has fully addressed the influence of different
parameters on the ignition behavior of single biomass particles. In particular, the effect of the

16
moisture content, which is especially important in the context of biomass fuels, is almost
completely unaddressed in the literature.

1.5. Objectives

The present work focuses in the experimental and numerical study of the ignition behaviout of
single biomass particles. Specifically, two main objectives are listed:

1. Extend the previous work done [42], in order to include a broader diversity of biomass
fuels and the effect of atmosphere composition, more specifically the oxygen
concentration.

This was done by selecting three additional biomass fuels to the two previously
studied [42] and analyzing their ignition behavior in three new operating conditions
with different mean oxygen concentration in the ignition zone. These fifteen tests (3
conditions * 5 biomasses) added to the fifteen tests in [42] (5 conditions * 3 fuels),
and help to better characterize the behavior of single biomass particles. The effect of
particle composition, shape and surface morphology and oxygen concentration in the
atmosphere on the ignition mode and delay time was analyzed.

2. Validate a numeric strategy using a commercial CFD software to model the


combustion behavior, and more specifically ignition, of single biomass particles and
investigate the dominant phenomena on the ignition delay time.

Given the typical lack of biomass material properties in the literature, a benchmark
test case was defined using the most reliable composition data available in order to
validate the CFD model. Properties such as specific heat or density have a heavy
weight on the numerical predictions, as seen in [42], and thus the uncertainty
associated with them strongly influences the numeric predictions. For these reasons,
pine bark, which was used in experimental and numerical studies in the past [50],
was selected as the benchmark case. The numeric predictions were compared
experimental measurements for a broad range of operating conditions in the
experimental setup of the current work. This included the three oxygen concentration
conditions and three temperature conditions from the previous work [42]. After being
validated, this model was then used to investigate the dominant phenomena on the
ignition delay time, by analyzing the relative contributions of heating, drying and
devolatilization times prior to ignition.

17
1.6. Thesis outline

This thesis is organized in seven chapters, of which the present one constitutes the
introduction that includes a literature survey and establishes the main objectives of the work.
Chapter 2 describes the experimental apparatus used throughout the study and the
experimental techniques used to collect the data reported in this thesis. Chapters 3 and 4
describe the theoretical mathematical and numeric models. Chapter 5 presents and
discusses the experimental results. Numerical results are discusses in section 6. Finally,
Chapter 7 lists the main conclusions of this study and provides guidelines for future work.

18
2. Experimental materials and methods

In the present section, the fuel preparation and characterization, the experimental setup and
experimental techniques are briefly described. All experiments were carried in the combustion
laboratory of the Mechanical Engineering department at Technical University of Lisbon.

2.1 Fuel preparation and characterization

Five different biomass fuels were used in this study. (i) Pine bark and (ii) wheat straw, which
were used in a previous work [42], were studied here for a new set of conditions. (iii) Vine
branches, (iv) kiwi branches, and (v) sycamore branches were additionally studied to include
a broader range of biomass compositions. All samples were finely sieved in a dry basis, in an
Electromagnetic Laboratory Sieve Shaker from Fritsch to the range of 212-224 µm. The
proximate and ultimate analysis and the heating values of the solid fuels used here were
carried out following the procedures specified in the standards ASTM‐ D‐ 3174, EN 14775
and EN 15403 for ash content, ASTM‐ D‐ 5865 and EN 14918 for high heating value and
ASTM‐ D‐ 3176 and ASTM‐ E‐ 870 for remaining parameters of the proximate and ultimate
analysis.

In order to analyze the morphology of the studied fuels, scanning electron microscope (SEM)
imaging was conducted. A quantitative analysis of the aspect ratio of each type of fuel was
performed using a minimum of 50 particles per sample. This factor was calculated simply as
the ratio between the largest length of the particle and the corresponding perpendicular
length.

Additionally from the SEM images a qualitative analysis of the particle surface morphology
was done. The surface morphology was split into three categories: scaled surface, surface
with superficial pores and surface with deep pores. For each fuel sample a minimum of 80
particles from the SEM images were considered in this analysis. In this categorization each
particle was observed individually and fitted into one of the three groups. The particles in
each surface morphology group were then counted to compute the fraction in each group for
the different biomass fuels.

2.2 Experimental setup

Figure 2.1 shows a schematic and a photograph of the experimental setup used in this work.
Globally, the setup consists of a feeding unit, an entrained flow reactor (flat-flame McKenna
burner and quartz tube), and an image acquisition system. Measuring of the temperature and
species was done with a thermocouple probe, and a gas species probe, respectively.

19
Figure 2.1 – (a) Schematic representation and (b) photo of the experimental setup.

Flat flame McKenna burner

Experiments were performed on a flat flame optical McKenna burner. The burner consists of
stainless-steel cylinder enveloping a water-cooled bronze porous sintered matrix of 60 mm in
diameter (see fig. 2.1). Two flow meters, respectively Tylan General model noFC280SA and
Omega model noFMA-A2317, allowed the control of the methane and primary air flow-rates
to the burner. Both flow meters had an operational range of 0-50 dm3/min and an uncertainty
of +/- 2 %. Methane was supplied by a 200 bar industrial bottle, and primary air by an Atlas
Copco GX7ff compressor, with a 10 bar capacity. More details in Annex C. In addition,
cooling water was fed to the burner through copper pipes (1 and 2 in Fig. 2.1a). The
fuel/oxidizer mixture was introduced in such a way that it was evenly distributed through the
sintered matrix, enabling a stabilized premixed flame for a range of air/fuel ratios. Figure 2.2
illustrates the pre-mixed flame sheet. This particular burner was fitted with a hole in the
center, with I.D. 1.55 mm, through which particles of solid fuels could be injected. Above the
burner a high-grade fused quartz tube of I.D. 70 mm, height of 500 mm and thickness of 2
mm confined the flow and avoided the entrainment of ambient air, while providing optical
access.

Even though, the flow produced by this this burner is laminar, it enables ultra-high heating
rates, and in this sense some of the conditions of an industrial boiler can be replicated. In
particular, conditions are much closer to industrial applications than those of drop tube
furnaces. The ultra-high heating rates provided by this device, in the order of magnitude of
105 K/s [42], can reproduce the conditions similar to those in real applications, in which
typically a cold flow of particles is suddenly exposed to extremely high heating rates.

20
Figure 2.2 – Premixed flame sheet.

Particle feeding system

The particle-feeding unit in Fig. 2.3 was used to fulfill the requirements of particle-by-particle
injection. The system consisted of a rotameter (ABB model noA6131B033INA0AAS), a 10 mL
syringe, and a vibrating electric motor. The rotameter, calibrated for nitrogen, allowed
measuring in the range of 0-0.34 dm3/min and had the uncertainty of +/- 2 %. In the
experiments, the rotameter controlled the transport air flow rate. The biomass particles stored
in the syringe were fed by gravitational force into the stream of transport air and transported,
and injected through the central hole of the burner into the ignition zone. The vibrating electric
motor avoided the clogging of the syringe hole and ensured a low feeding rate of dispersed
particles.

21
Figure 2.3 – Particle feeding system: (1) syringe filled with particles, (2) air ejector, (3)
electrical vibrating motor, (4) feeding tube.

Image acquisition system

The image acquisition system consisted of a backlight and a high-speed camera


complemented with two lenses: a 2x tele-converter and an AF Micro Nikkor f/2.8 60 mm. A
led backlight was used to visualize the particle before ignition. Led was chosen over other
light sources because it provided a fairly continuous light intensity with time, important for the
success of the image processing tools used. An Optronis CamRecord CR600x2 CMOS high-
speed camera was used to record the ignition events. No filter was used and therefore the
visualized signal corresponds to the visible light intensity, including both black body emission,
and chemiluminescence emission. The camera was positioned with its optical axis
perpendicular to the axis of the burner and focused on the ignition zone at the burner axis.
The table 2.1 shows the imaging settings used.

Table 2.1 – Imaging settings.

Settings Value
Aperture 2.8-5.6
Exposure time [s] 1/15000
FPS 3500
Resolution [pix] 250*452

22
As mentioned above, a quartz tube was used to confine the flow and provide optical access.
Whilst some authors [35] prefer to use square cross section quartz chimneys when applying
laser techniques, others [38, 31,42] observe the visible light signal with no laser inducement
and use circular section quartz tubes with similar thickness and inner diameter to that of the
present work, and do not observe any image distortion effects.

2.3. Experimental techniques and uncertainties

In combustion problems, diagnostics can mainly be accomplished through two different


approaches: probe sampling techniques and optical techniques. Generally, probes result in
directly local physical property measurements, but may cause perturbations in the flow and
induce heat sinks. These effects depend on the size and the material of the probe. Additional
impact on flow patterns appears if the suction of the probe is not isokinetic. On the other
hand, optical techniques have no interference with the system, but are generally an
expensive solution.

In the current work, measuring of gas temperature and gas species was done using probe
sampling techniques and observation of the ignition events was done using optical
techniques.

Gas temperature

In order to characterize the ignition zone in terms of gas temperature, local mean temperature
measurements were performed using uncoated 76‐μm‐diameter fine‐wire platinum/platinum:
13% rhodium (type‐R) thermocouples. The hot junction is supported on 350 μm wires of the
same material, located in a twin‐bore alumina sheath with an external diameter of 6 mm.
These thermocouples present a range of operation from -50 ºC to 1768 ºC, sufficient for the
measurements undertaken. The thermocouple probe (see Fig. 2.4) is mounted on a traverse
mechanism, which allowed axial movements. No thermal distortion of the probe was
observed, and its positioning was accurate to within ± 1 mm. The analog outputs of the
thermocouple were transmitted via an A/D board to a computer where the signals are
processed and the mean values computed. The acquisition time for each measured point is
30 s.

23
Figure 2.4 – Schematic representation of the temperature probe and thermocouple (values in
mm except when indicated).

Errors associated with the measured gas temperature have two different natures, (i) one
associated with the repeatability of the tests and (ii) another associated with the heat transfer
phenomena about a thermocouple bead.

In the measurements done in the McKenna burner, there is a strong radial temperature
gradient close to the burner, because of the injection of cold air into the center hole into a
region of high temperature combustion products gas. This results in a temperature
measurement that is dependent on the proper alignment of the thermocouple, especially
close to the surface of the burner. To cope with this error the presented temperatures
measured in the current work result from a repetition of three measurements with respective
thermocouple alignments.

The values of the temperature computed by the acquisition system represent the temperature
of the thermocouple bead and not the temperature of the gas around it. The computed
temperatures were corrected as explained in Annex A and similarly to what was done by
Shaddix [48] to yield an estimate of the gas temperature.

Inserting the probe inside the flow of hot gases may also cause local perturbations of the flow.
In the present case, due to the laminar nature of the flow, these were considered to have
minor influence.

Gas species concentration

Gas species concentrations were measured to characterize the flow of combustion products.
A water-cooled probe (Fig. 2.5), mounted in a traverse mechanism, was used to collect the
gas sample. The sample was sucked through the probe with the aid of an oil-free diaphragm
pump. A condenser removed the main particulate matter excess and the main condensate. A
filter and a dryer removed any residual particles and moisture, so that a constant supply of
clean dry gaseous mixture was delivered to the analyzers. The measured species were
obtained under dry basis conditions.

24
Figure 2.5 – Gas sampling probe (values in mm).

The analytical instrumentation comprised a magnetic pressure analyzer for O2


measurements, a non-dispersive infrared gas analyzer for CO2 and CO measurements, a
flame ionization detector for HC measurements and finally a chemiluminescent analyzer for
NOx measurements. The analogical outputs of the analyzers were transmitted via A/D boards
to a computer where the signals were processed and the mean values computed. Zero and
span calibrations with standard mixtures were performed before and after each measurement
session. O2, CO, CO2, hydrocarbons (HC), and NO x concentrations were set to zero by
injecting a sample of pure N2 to the analyzers. O2 concentration was adjusted according with
the volumetric percentage of oxygen in the atmospheric air (20.9% vol.). The adjustments for
the remaining species were performed via the injection of a standard mixture of gases (CO,
CO2, HC, and N2). The uncertainties associated with these measuring techniques result
mainly from the aerodynamic perturbations in the flow. Water condensed on the surface of
the water-cooled probe, and ran downwards by gravitational force along the probe. To avoid
this water from falling to the surface of the burner the suction had to be such that it resulted in
a non-isokinetic flow, which means that the measured quantities did not correspond solely to
the centerline of the burner, but to an average over the volume flow aspirated. This effect was
only noticeable in a close vicinity of the burner in which the exhaust gas flow and the
transport air flow were not fully mixed. In practice, close to the burner the concentration of O2
should be slightly higher than the one measured due to the aspiration of the surrounding gas.
Following the same reasoning CO2, CO, and HC concentrations in the centerline should be
lower than the measured ones. As in the temperature measurements the presented species
concentrations result from a repetition of three tests.

Particle imaging

The ignition criterion used herein is based on the visible light signal, similarly to what was
done by Yuan [41]. This technique is simple to use and the visible light signal is strong and
easy to capture. Although soot and char emission are included in the visible light signal
captured this has proved to be a good indicator of ignition [13,41]. In this work, ignition was

25
considered to occur when 15% of the maximum pixel luminosity was reached for either
volatile or surface ignition. Since the high speed videos focused on the ignition events of
single particles instead of a time average (long exposure time) of a stream of particles as in
Yuan et al. [41] a broader criterion was needed. The light intensity signal of the high speed
images, used to compute the ignition delay time, was normalized by the maximum luminosity
intensity of each video. Videos of different particles exhibited different ratios of maximum
luminosity to the baseline luminosity. A more inclusive criterion than the 10 % criterion used in
[41] was needed and the 15% showed a better agreement with the ignition events observed.

A Matlab® routine was used to calculate the ignition delay time. The code reads the data from
the .tif files, stores it in matrices, eliminates the background noise, and identifies the frame in
which the pixel value was closest to 15% of the maximum luminosity intensity. The code also
“tracks” the position of the particle, and uses the frame-rate to compute the frame-to-frame
velocity of the particle. This procedure is explained in Annex B. Finally, these velocities and
the position of the particle at the ignition frame are used to calculate the ignition delay time.

In order to capture the ignition of the particle in the frame, the frame was lifted from the top of
the burner, i.e., the frame did not capture the whole trajectory of the particle from injection to
ignition. In the blind spot from injection to the first frames where the particle was in frame, the
velocity of the particle was assumed to vary linearly. The velocity of the particle at injection
was assumed to be equal to the gas velocity at injection. This velocity then decayed linearly
to the first velocity of the particle measured.

The method used to determine the ignition delay time introduces some error. Even in the
range of 212-224 µm particles do not have all the same size, but even if they did have,
different particles have different compositions and structures, and the 15% luminosity
corresponds to slightly different stages of the combustion process in different particles. In
practice, even though this criterion is simple, the ignition frame was always in an interval of
+/- 5 frames around the first visual signs of luminosity (from visual observation). Even if this
criterion corresponded to the exact ignition instant, the ignition delay calculation is based on
the determination of the position of the particle and its velocity, which are also subject to
uncertainties. To bridge this problem a minimum of 20 single particle events was recorded for
each experimental condition (to achieve statistical significance).

Determination of the ignition mode was evaluated by direct visual observation of the recorded
images. It should be noted that by observing only the luminosity at the particle surface it is
impossible to determine whether the ignition occurred homogeneously or heterogeneously. In
this case it is more correct to talk about gas-phase and surface ignition. Ignition is considered
to take place in the gas-phase when a bright area, wider than the particle, and surrounding it,
is discernible from the dark particle in the center. Surface ignition occurs when the particle
surface becomes bright, whether it is the whole particle, or just discrete spots on its surface.
Figure 2.6 illustrates these two distinct types of ignition events. On the left a, gas phase

26
ignition occurs, resulting from the interaction of the volatile gases released from the particle
and the oxidizing gas. On the right b, direct surface oxidation occurs.

Figure 2.6 – Ignition events of a kiwi branch particle: a) gas-phase ignition b) surface
ignition.

When analyzing individual particles it is particularly important to guarantee that the observed
particles are within the selected size range. Despite the accuracy of the sieving process, the
high directionality of biomass can allow some particles of large dimensions to pass through
fine sieves. Ignition events of particles out of the range of interest were not considered in the
analysis of the results. This was valid for particles smaller than the desired range, that might
get trapped with the other during the sieving process or originated afterwards, and also for the
larger particles that passed through the sieve.

27
3. Mathematical model

In this section, the mathematical model used to study the ignition and combustion of the
biomass particles is briefly described.

3.1. Modeling approach

The CFD software Fluent v.14.5 was used to model the ignition and combustion of the single-
particles in the McKenna burner. A time-dependent Euler-Lagrange approach was taken. In
this approach, the fluid phase is treated as a continuum by solving the Navier-Stokes, energy
and the species transport equations. The dispersed phase is solved by tracking the particles
through the calculated flow field, and solving mass and energy balances.

3.2. Gas-phase

The reactive flow is described by the Navier-Stokes, energy and species transport equations.
These conservation principles are briefly presented below.

The Navier-Stokes equations state the conservation of mass and momentum respectively as
follow:

𝜕𝜌
+ ∇ ∙ (𝜌𝑣
⃗⃗⃗ ) = 𝑆𝑚 (3.1)
𝜕𝑡
𝜕(𝜌𝑣)
+ ∇ ∙ (𝜌𝑣 ⃗⃗⃗ + 𝑆𝑚,𝑝
⃗⃗⃗ ) = −∇𝑝 + ∇ ∙ (𝜏̅̅ ) + 𝜌𝑔
⃗⃗⃗ 𝑣 (3.2)
𝜕𝑡

where 𝑆𝑚 is the mass source added to the continuous phase from the dispersed phase. 𝑝 is
the static pressure, 𝜏̅̅ is the stress tensor (described below), and 𝜌𝑔
⃗⃗⃗ is the gravitational body
force, and 𝑆𝑚,𝑝 is the momentum sink/source from the solid particles.

The stress tensor 𝜏̅̅ is given by:

2
𝜏̅̅ = 𝜇 [(∇𝑣 ⃗⃗⃗ 𝑇 ) − ∇ ∙ 𝑣
⃗⃗⃗ + ∇𝑣 ⃗⃗⃗ 𝐼] (3.3)
3

where 𝜇 is the molecular viscosity, 𝐼 is the unit tensor, and the second term on the right hand
side is the effect of volume dilatation.

The energy equation employed in this work uses as the main variable the sum of internal
energy and kinetic energy:

𝑝 𝑣2
𝐸 =ℎ− + (3.4)
𝜌 2

28
where sensible enthalpy ℎ is defined for ideal gases as:

𝑇
ℎ = ∑ 𝑌𝑗 (∫ 𝑐𝑝,𝑗 𝑑𝑇 ) (3.5)
𝑗 𝑇𝑟𝑒𝑓

where 𝑇𝑟𝑒𝑓 is 298.15 K.

The conserved energy equation is on the following form:

𝜕(𝜌𝐸)
⃗⃗⃗ (𝜌𝐸 + 𝑝)) = ∇ ∙ (𝑘𝑒𝑓𝑓 ∇T − ∑ ℎ𝑗 ⃗⃗𝐽𝑗 + (𝜏̅̅ 𝑒𝑓𝑓 ∙ 𝑣
+ ∇ ∙ (𝜌𝑣 ⃗⃗⃗ )) + 𝑆ℎ + 𝑆𝑝 (3.6)
𝜕𝑡
𝑗

where 𝑘𝑒𝑓𝑓 is the effective conductivity, and ⃗⃗𝐽𝑗 is the diffusion flux of species j. The first three
terms on the right-hand side represent the energy transfer due to conduction, species
diffusion, and viscous dissipation, respectively. The source terms 𝑆ℎ and 𝑆𝑝 include the heat
of chemical reaction, and heat exchanged with the solid particles.

The conservation equations for the chemical species takes the general form for species 𝑖:

𝜕(𝜌𝑌𝑖 )
⃗⃗⃗ 𝑌𝑖 ) = −∇ ∙ ⃗𝐽𝑖 + 𝑅𝑖 + 𝑆𝑖
+ ∇ ∙ (𝜌𝑣 (3.7)
𝜕𝑡

where 𝑌𝑖 is the local mass fraction of species 𝑖, 𝑅𝑖 is the net rate of production of species 𝑖 by
chemical reactions and 𝑆𝑖 is the rate of creation by addition from the dispersed phase. ⃗𝐽𝑖 is the
diffusion flux of species 𝑖, which arises due to gradients of concentration and temperature. To
model the mass diffusion due to concentration gradients, the following law is used:

⃗𝐽𝑖 = −𝜌𝐷𝑖,𝑚 ∇ 𝑌𝑖 − 𝐷𝑇,𝑖 ∇𝑇 (3.8)


𝑇

Here 𝐷𝑖,𝑚 is the mass diffusion coefficient for species 𝑖 in the mixture and 𝐷𝑇,𝑖 is the thermal
(Soret) diffusion coefficient.

The reaction rates appear in equation (3.7) as a source term. In a laminar flow, a laminar
finite-rate model is used, in which reaction rates are determined by Arrhenius kinetic
expressions.

The net source of chemical species 𝑖 due to reaction is computed as the sum of the Arrhenius
reaction sources over the 𝑁𝑅 reactions in which the species participates in:
𝑁𝑅

𝑅𝑖 = 𝑀𝑤,𝑖 ∑ 𝑅̂𝑖,𝑟 (3.9)


𝑟=1

29
where 𝑀𝑤,𝑖 is the molecular weight of species 𝑖 and 𝑅̂𝑖,𝑟 is the Arrhenius molar rate of
creation/destruction of species 𝑖 in reaction 𝑟.

A general chemical reaction 𝑟 can be written in the general form as follows:

𝑁 𝑘𝑓,𝑟 𝑁

∑ 𝑣′𝑖,𝑟 ℳ𝑖 ∑ 𝑣′′𝑖,𝑟 ℳ𝑖 (3.10)

𝑖=1 𝑘𝑏,𝑟 𝑖=1

where,

𝑁 - number of chemical species in the system


𝑣′𝑖,𝑟 - stoichiometric coefficient for reactant 𝑖 in reaction 𝑟
𝑣′′𝑖,𝑟 - stoichiometric coefficient for product 𝑖 in reaction 𝑟
ℳ𝑖 - symbol denoting species 𝑖
𝑘𝑓,𝑟 - forward rate constant for reaction 𝑟
𝑘𝑏,𝑟 - backward rate constant for reaction 𝑟

For the general reversible chemical reaction 𝑟 the molar rate of creation/destruction of
species 𝑖 is given by:

𝑁 𝑁
′ ′′
𝜂𝑗,𝑟 𝜂𝑗,𝑟
𝑅̂𝑖,𝑟 = (𝑣′𝑖,𝑟 − 𝑣′′𝑖,𝑟 ) (𝑘𝑓,𝑟 ∏[𝐶𝑗,𝑟 ] − 𝑘𝑏,𝑟 ∏[𝐶𝑗,𝑟 ] ) (3.11)
𝑗=1 𝑗=1

where,

𝐶𝑗,𝑟 - molar concentration of species 𝑗 in reaction 𝑟 (𝑘𝑚𝑜𝑙/𝑚3 )



𝜂𝑗,𝑟 - rate exponent for reactant species 𝑗 in reaction 𝑟
′′
𝜂𝑗,𝑟 - rate exponent for product species 𝑗 in reaction 𝑟

The rate constants in equation (3.11) are computed using an Arrhenius expression:

−𝐸𝑟
𝑘𝑟 = 𝐴𝑟 𝑇𝛽𝑟 𝑒𝑥𝑝 ( ) (3.11)
𝑅𝑇
where,

𝐴𝑟 - pre-exponential factor (consistent units)


𝛽𝑟 - temperature exponent
𝐸𝑟 - activation energy for the reaction (J/kmol)
𝑅 - universal gas constant (J/kmol-K)

30
3.3. Discrete-phase

As mentioned above, each particle was transported in a Lagrangian frame of reference where
its position, velocity, mass, and diameter are evolved.

In the model, some simplifications were considered, namely:

Bi << 1 (internal temperature gradients neglected);


Spherical particles;
Negligible catalytic effects.

3.3.1. Particle motion

The dispersed phase can exchange momentum, mass and energy with the fluid phase, as
schematically represented in Fig. 3.1.

Figure 3.1 – Particle and gas phase coupling [49].

The position of the particle in the continuous phase is computed by integrating the force
balance on the particle, which is written in a Lagrangian reference frame:

𝜕𝑣
⃗⃗⃗ 𝑝 ⃗⃗⃗ (𝜌𝑝 − 𝜌)
𝑔
= 𝐹𝐷 (𝑣
⃗⃗⃗ − 𝑣
⃗⃗⃗ 𝑝 ) + (3.12)
𝜕𝑡 𝜌𝑝

where 𝑣
⃗⃗⃗ 𝑝 , 𝜌𝑝 and 𝑣
⃗⃗⃗ , 𝜌 represent the respectively velocity and density of the particle and gas
phases. 𝑔
⃗⃗⃗ is the gravity acceleration and 𝐹𝐷 (𝑣 ⃗⃗⃗ 𝑝 ) is the drag force per unit mass where:
⃗⃗⃗ − 𝑣

18𝜇 𝐶𝐷 𝑅𝑒
𝐹𝐷 = (3.13)
𝜌𝑝 𝑑𝑝2 24

𝜇 is the molecular viscosity of the fluid, 𝑅𝑒 is the relative Reynolds number, which is defined
as:

𝜌𝑑𝑝 |𝑣
⃗⃗⃗ − 𝑣
⃗⃗⃗ 𝑝 |
𝑅𝑒 ≡ (3.14)
𝜇

31
The drag coefficient for the spherical particle is given by Morsi and Alexander [49] as,

𝑎2 𝑎3
𝐶𝐷 = 𝑎1 + + 2 (3.15)
𝑅𝑒 𝑅𝑒

where 𝑎1 , 𝑎2 and 𝑎3 are constants that apply over several ranges of Re.

3.3.2. Heat and mass exchange

To model the heat and mass transfer to and from the particle a user-defined-function was
developed considering the evaporation, devolatilization and surface oxidation phases. Initially,
an evaporation phase, in which the particle loses mass as it releases water vapor to the
environment, takes place. When all of the moisture content of the particle has evaporated the
devolatilization phase begins. Similarly, when all the volatiles have been released from the
particle, the surface oxidation begins.

In the model these phases occur sequentially and therefore the model is not able to predict
the occurrence of heterogeneous ignition. This is a strong limitation, however ignition of
biomass particles generally occurred in the gas-phase, and thus this model can be interesting
as a first approach.

In the proposed model the particle consists of moisture, volatiles, char and ash:

𝑚𝑝 = 𝑚𝐻2𝑂 + 𝑚𝑣 + 𝑚𝑐 + 𝑚𝑎𝑠ℎ (3.16)

where the subscripts 𝑝, 𝐻2 𝑂, 𝑣, 𝑐, and 𝑎𝑠ℎ correspond respectively to particle, moisture,


volatile, char and ash.

Fig. 3.2 depicts the solid fuel constituents and the mass models.

Figure 3.2 – Fuel constituents [50].

Evaporation

The moisture content evolution is given by:

32
𝑑𝑚𝐻2𝑂 𝑝𝐻 𝑂,𝑠𝑎𝑡 𝑝𝐻2𝑂
= 𝑘𝑣 ( 2 − ) 𝐴𝑝 𝑀𝑤,𝐻2𝑂 (3.17)
𝑑𝑡 𝑅𝑇𝑝 𝑅𝑇

where 𝑘𝑣 is the mass transfer coefficient of steam into air, 𝑝𝐻2𝑂,𝑠𝑎𝑡 is the saturation pressure
of water at particle temperature, and 𝑝𝐻2𝑂 is the partial pressure of water in the gas phase.

The latent heat of vaporization for water is calculated from the Watson relation [51].

Devolatilization

The devolatilization was modeled based on a single kinetic rate law [52]:

𝑑𝑚𝑣 −𝐸𝑎,𝑣
= 𝐴𝑣 𝑒𝑥𝑝 ( ) 𝑚𝑣 (3.18)
𝑑𝑡 𝑅𝑇𝑝

where 𝐴𝑣 and 𝐸𝑎,𝑣 are the pre-exponential factor and activation energy.

Char combustion

The char combustion model has no influence over the ignition (in the approach followed
herein). For this reason, and since the main focus of this work is to model the ignition
process, a simple diffusion-limited model was used. The diffusion-limited surface reaction rate
model assumes that the surface reaction proceeds at a rate determined by the diffusion of the
gaseous oxidant to the surface of the particle, neglecting the effect of chemistry. This is valid
at high temperatures, where the reaction at the particle surface becomes very intense, and
where the all oxygen is consumed therein [33], see Fig. 3.3.

Figure 3.3 – Rate of surface reaction as a function of temperature of the solid particle [53].

33
For the bulk diffusion limitation the evolution of the mass of char is given by,

𝑑𝑚𝑐 𝑀𝑤,𝐶
= −2𝜋𝑑𝑝 𝐷 𝑌𝑖 𝜌𝑔 (3.19)
𝑑𝑡 𝑀𝑤,𝑖

where 𝑑𝑝 is the particle diameter, 𝐷 is the diffusion constant, as in [54], 𝑌𝑖 is the mass fraction
of the chemical species in the gas, 𝜌𝑔 is the gas density, and 𝑀𝑤,𝑖 is the molecular weight of
species 𝑖.

If greater focus is put upon the char combustion phase a more detailed model should be
used. A more complex model, based on the random pore model [55], was used by Kajitani et
al. [56] and Watanabe and Otaka [53], to include the chemical kinetic effects. In this model,
however, the reaction rate greatly depends on the fuel type and gasifying agent. Therefore,
kinetic rate parameters specific to the test conditions are required.

Heat exchange

The particle exchanges heat with the surrounding gas due to convection, conduction and
radiation. The energy balance across the surface of a particle, assuming that the temperature
within a particle is uniform and a transparent media, is given by,

𝑑𝑇𝑝 −𝐴𝑝 (3.20)


= [ℎ (𝑇 − 𝑇𝑔 ) + 𝜀𝜎(𝑇𝑝4 − 𝑇𝑤4 )] + S𝑟
𝑑𝑡 𝑚𝑝 𝑐𝑝,𝑝 𝑐 𝑝
where 𝑇𝑝 , 𝑇𝑤 and 𝑇𝑔 are the particle, wall, and gas temperatures respectively. 𝑐𝑝,𝑝 , 𝑚𝑝 , 𝐴𝑝 and
𝜀 are the particle heat capacity, mass, surface area and emissivity respectively, 𝜎 is the
Stefan–Boltzmann constant, ℎ𝑐 = 𝑁𝑢 𝑘/𝑑𝑝 is the convective heat transfer coefficient with
𝑁𝑢 = 2 + 0.6𝑅𝑒 1/2 𝑃𝑟1/3 , and S𝑟 is the temperature source term due to vaporization, S𝑟,𝑒𝑣𝑎𝑝 ,
and heterogeneous reactions, S𝑟,ℎ𝑒𝑡 ,

S𝑟 = S𝑟,𝑒𝑣𝑎𝑝 + S𝑟,ℎ𝑒𝑡 (3.21)

The heat source die to vaporization is given by,

1 𝑑𝑚𝐻2𝑂 (3.22)
S𝑟,𝑒𝑣𝑎𝑝 = 𝜆𝐻2𝑂 (− )
𝑚𝑝 𝑐𝑝,𝑝 𝑑𝑡

where 𝜆𝐻2𝑂 is the water latent heat of vaporization. S𝑟,ℎ𝑒𝑡 is defined as in [50].

34
3.4. Volatile combustion

Due to the nature of the model and the assumptions made, the ignition will invariably occur in
the gas phase.

This is a limitation of the model, however, as the experimental results will show, the ignition of
the biomass fuels, in particular of the pine bark particles, occurred almost exclusively in the
gas phase.

Volatile composition and chemical mechanism

The volatile gases are made up of light gaseous species such as: CO, CO2, H2, H2O, CH4,
C2H4, C2H2, and heavy gaseous species such as tar.

Considering all of the volatile gases separately involves the transport of a large number of
species and becomes computationally expensive. In some works, this difficulty is overcome
by grouping the volatile gases in a single postulated substance C aHbOcNd. In this study,
however, due to the laminar nature of the flow the chemical reaction time-scale becomes
important, especially, when modeling ignition.

The composition of the volatiles was unknown and thus had to be assumed. For bituminous
coal particles Goshayeshi and Sutherland [50] considered the decomposition of volatiles in
CO, H2, generally accepted as lighter species, and C 2H2 to represent the heavier species. For
the biomass fuels however, due to the high content in oxygen, the volatile matter constituent
species will be lighter with a larger fraction of CO. The decomposition of the volatiles
considered here is then given by,

𝑎 (𝑏 − 1) 𝑐
𝐶𝐻𝑎 𝑂𝑏 𝑁𝑐 → 1𝐶𝑂 + 𝐻2 + 𝑂2 + 𝑁2 (3.23)
2 2 2

where a, b and c are calculated from the ultimate and proximate analysis of the fuel as in [57].

To model the ignition process of the volatiles a detailed chemistry mechanism was used. The
DRM mechanism [58] with 103 reactions among 22 chemical species was chosen. This
mechanism results from a truncation of the original GRI-mechanism [59] (53 species and 325
reactions) with the objective of developing a smaller set of reactions (i.e., still a detailed
mechanism but with a smaller number of variables) to reproduce closely the main combustion
characteristics predicted by the full mechanism. Vascellari et al. [52] used this mechanism to
study the ignition and combustion of single-coal-particles and obtained a good match
between the experimental and numeric results.

The GRI-mechanism could be used for a more complete analysis and a more accurate
simulation, however this mechanism implies heavy computational costs. Therefore, the DRM
mechanism is considered a compromise between accuracy and computational cost.

35
4. Numerics

The combustion and ignition of the biomass particles was investigated using the commercial
CFD software ANSYS Fluent v.14.5. In the following sub-sections some of the options taken
will be discussed, namely concerning the geometry of the domain and its mesh, the boundary
conditions, discretization methods and solver options.

4.1. Geometry, mesh and boundary conditions

The physical domain was represented by a 2D-axy-simmetric computational domain. The


dimensions of the domain followed the dimensions of the burner. Figure 4.1 shows the
geometry of the domain, where the x-axis corresponds to the vertical direction. The top 80
mm above the burner were modeled.

Figure 4.1 – Geometry of the computational domain.

Figure 4.2 shows the mesh used with labels for the boundaries of the domain. The
computational grid consisted of approximately 8000 rectangular cells and was refined so that
a grid-independent unsteady solution was obtained. The mesh was structured and non-
uniform, with a larger cell density close to the axis of the burner, where the reactions
occurred. The flat flame sheet was not modeled in this work. Instead, a flow of hot
combustion products gas was injected into the domain from the top of the burner (boundary 3
in Fig. 4.2). Some simplifications were made in boundaries 2 and 4, respectively
corresponding to the transport air tube wall and quartz tube wall, namely by considering that
these walls were at a constant temperature (close to the neighbouring gas temperature).
These boundaries are far from the region where the reaction will occur and have a marginal
influence on the final result. Finally, Table 4.1 summarizes the boundary conditions (BC)
considered.

36
Figure 4.2 – Mesh and labelled boundaries of the domain.

Table 4.1 – Summary of the boundary conditions.

Boundary Region Type of BC Thermal BC

1 Transport air inlet Velocity-inlet Fixed temperature

2 Transport air tube wall Wall Fixed temperature

3 Hot combustion product gas inlet Velocity-inlet Fixed temperature

4 Quartz tube wall Wall Fixed temperature

Null temperature
5 Top border of the domain Pressure-outlet
gradient

6 Axis Axis Null heat flux

4.2. Solver options and discretization

The simulations were carried in two steps. Firstly, a steady-state simulation was performed
for the gas phase to set the environment conditions in which the particle was injected. After
that, the particle was introduced in the domain in a transient simulation. A time step of 1e-6
was used with 100 iterations per time step.

The coupled algorithm was used for pressure-velocity coupling with second-order spatial and
temporal discretization. This solver belongs to the family of pressure-based solvers, which
use a pressure equation (derived from the continuity and momentum equations) to ensure the
conservation of mass.

When simulating combustion it can be difficult to reach a converged solution. The reactions
lead to a large heat release and subsequent density changes and large accelerations in the
flow. Additionally, the reaction does not occur until an ignition temperature is reached, but

37
then proceeds very quickly until the reactants are consumed. Some reactions have very fast
time scales, in order of 10-10 s (reaction time scales are much faster than the convection and
diffusion time scales), while others have much slower time scales, on the order of 1 s. This
time-scale disparity results in numerical stiffness, which means that an extensive
computational work is required to integrate the chemical source terms. To overcome these
difficulties a fractional step algorithm is used. In the first fractional step, the chemistry in each
cell is reacted at constant pressure for the flow-time-step [49]. In the second step, the
convection and diffusion terms are treated just as in a non-reacting simulation.

38
5. Experimental results and discussion

5.1 Fuel characteristics

Figure 5.1 shows the solid fuels, after the sieving process and Table 5.1 below presents the
proximate and ultimate analysis, and low heating value, for the studied fuels. All the biomass
fuels tested present a high volatile content, ranging from 58.9 % for the pine bark to 73.5 %
for sycamore. The fuels presented consistent values of fixed carbon, with the exception of
pine bark in which this content reached 25.9 %. The fuels presented a moisture content
between 8 and 14%, much higher than that typical of coal. In terms of the ash content wheat
straw stands out from the other fuels because of its exceptionally high fraction of 14.7%,
when all the other fuels present an ash content below than 2.4 %.

Figure 5.1 – Solid fuels after sieving: a) wheat straw, b) pine bark, c) vine branches, d) kiwi
branches, e) sycamore branches.

Table 5.1 – Properties of the solid fuels.

Parameter Pine Bark Wheat Straw Sycamore Vine Kiwi


Proximate Analysis (wt%, as received)
Volatile Matter 58.9 64.9 73.5 70.2 71.7
Fixed Carbon 25.9 11.5 16.0 17.2 17.6
Moisture 13.9 8.9 9.3 10.5 8.3
Ash 1.3 14.7 1.2 2.1 2.4
Ultimate Analysis (wt%, as received)
Carbon 47.8 39.4 47.6 49.3 48.2
Hydrogen 4.3 5.2 5.98 6.19 6.4
Nitrogen 0.3 0.5 0.54 0.67 0.72
Sulfur < 0.02 < 0.02 < 0.16 < 0.15 < 0.16
Oxygen 32.4 31.3 45.72 43.69 44.52
Low heating value (MJ/kg) 17.1 18.8 18.1 17.7 17.9

39
Figure 5.2 shows typical SEM group images of the five biomass fuels under a x50 zoom
magnification and Fig. 5.3 shows the estimated aspect ratio for each fuel.

Figure 5.2 – SEM images of the solid fuels with x50 zoom factor. a) wheat straw, b) pine
bark, c) vine branches, d) kiwi branches, e) sycamore branches.

Figure 5.3 – Aspect ratios of the biomass fuels computed from the SEM images. Pine bark
particles’ aspect ratio is not depicted as these particles were considered quasi-spherical.

From this analysis, it was observed that pine bark particles had a quasi-spherical shape
(aspect ratio <2) and also that this sample had a fairly homogeneous size distribution. The
remaining biomass fuels were typically more elongated with values of aspect ratios close to 4,
in general, and close to 5 for sycamore branch particles. The samples with higher values of

40
aspect ratio also presented higher heterogeneity in the size distribution in comparison with
the samples with more spherical particles.

In addition to this shape analysis the particle morphology was further studied by doing a
qualitative analysis of the SEM images. This analysis illustrated that wheat straw particles
had a similar scaled surface to that of kiwi branch particles, and vine branch particles
presented two distinct particle types, one resembling the scales seen in wheat straw and kiwi,
and the other highly porous closer to the aspect of sycamore branch and pine bark particles.
However, these porous particles were not completely alike. Whilst in the sycamore branch
particles the pores were superficial, both pine bark and kiwi branch particles presented
deeper pores. This is illustrated in Fig. 5.4. Fig. 5.4a, b and c show the scaled appearance of
kiwi branch, wheat straw and vine branch particles, respectively, and Fig 5.4d, e, and f show
the porous structure of a sycamore branch, pine bark and vine branch particles, respectively.

Figure 5.4 – SEM images of a) kiwi branch particle (zoom x220); b) wheat straw particle
(zoom x180); c) vine branch particle (zoom x170); d) sycamore branch particle (zoom x250);
e) pine bark particle (zoom x250); and f) vine branch particle (zoom x170).

In this qualitative analysis the particles were split into three categories according to their
appearance: superficial pores, deep pores and scaled surface. Table 5.2 contains a summary
with the main observations concerning the biomass fuels, including the aspect ratio, the
fraction of particles within each surface morphology group, and the proximate analysis.

41
Table 5.2 – Summary of the main observations on the biomass fuels, which may affect
ignition mode.

Aspect Ratio Surface Morphology Proximate Analysis


Wheat straw 3.5
Elongated
Vine branches 3.5
Elongated
Pine bark <2
Quasi-spherical
Sycamore 4.8
branches Highly elongated
Kiwi branches 4.2
Elongated

It is important to note that table 5.2 only considers some factors that may affect the ignition
mode. Other parameters, such as the thermodynamic properties of the particles and the
volatile constituent species, are likely to play an important role in this process and should be
further explored in future studies.

5.2. Test conditions

Table 5.3 shows the test conditions of this work. The solid fuels were burned under three
conditions. The fuel and oxidizer flow rates were varied to yield different equivalence ratios,
and hence different oxygen concentrations for the same temperature. The resulting mean dry
oxygen concentrations for tests 1, 2 and 3 were 3.5 %, 5.1 % and 6.5 %, respectively, which
correspond to concentrations that a particle might actually face in an industrial boiler.
Transport air was kept constant and at the lowest flow rate that would ensure particle feeding.
The reported values of temperature and O2 concentration are the ones measured at the
ignition zone.

Table 5.3 – Operating conditions.

Parameter Test 1 Test 2 Test 3


Thermal input [kW] 1.0 1.1 1.4
Methane [dm3/min] 1.70 1.87 2.40
Primary Air [dm3/min] 20.5 25.0 34.0
Transport Air [dm 3/min] 0.14
Equivalence ratio ϕ [-] 0.79 0.71 0.67
Gas temperature [K] 1670 1680 1690
O2 concentration [dry vol %] 3.5 5.1 6.5

42
Figures 5.5 and 5.6 show the temperature and O2 concentration profiles above the burner
along its axis, respectively. Temperature near the injection hole in the burner is low as a
consequence of the cold transport-air flow. Upwards from the injection hole, the temperature
increases rapidly because of the mixture between the transport air and the combustion
products. It then reaches a plateau and heat losses through the quartz have a marginal
impact at the height of interest above the burner. Figure 5.5a corresponds to the temperature
of the thermocouple bead as measured. The temperature profile in Fig. 5.5b is an estimate of
the gas temperature around the thermocouple, based on heat transfer about the
thermocouple bead, as explained in annex B.

Figure 5.5 – Mean gas temperature profiles along the burner axis above the burner (see
Table 5.3).

Close to the surface of the burner and to the injection hole the radial temperature gradients
are strong, and thus a small change in the alignment of the thermocouple probe results in
significant variations of the measured temperature. This can be seen in in the error bars in
Fig. 5.5a. The estimated gas temperatures differ the most from the measured thermocouple
temperatures far from the burner where the radiation losses to the environment play a
meaningful role.

43
Figure 5.6 – Mean dry O2 concentration profiles along the burner axis above the burner (see
Table 5.3).

Oxygen concentration decreases from the injection point where air is injected and stabilizes
after the transport-air flow is fully mixed with the combustion products. Condition 3 was the
leanest of the tested conditions and stabilized at a slightly higher position as compared with
the other conditions, with O 2 concentration stabilizing further away from the burner.

Again the repeatability error in the measurements was higher close to the burner, for the
same reasons as before. Values below 4mm were not measured. The flame sheet was
established at a height of approximately 3-4 mm, and presented a thickness of 1-2 mm (see
Fig. 2.2).

5.3. Ignition mode

Figure 5.7 shows selected images of the typical ignition events captured. Figure 5.7a shows
the gas-phase ignition of a kiwi branch particle and Fig. 5.7b the surface ignition of a vine
branch particle under test conditions 1 and 3, respectively. Additionally, in Fig. 5.7a it is
possible to identify different steps of the combustion process. The first image depicts the
particle before ignition, during the particle heat-up phase. At t = 16 ms the ignition event, as
defined by the 15 % of maximum luminosity criterion, can be observed. Then, 6 ms later it is
still possible to observe the volatile cloud burning and at t = 22 ms the volatile flame is close
to extinction. Afterwards at t = 25 ms it is possible to observe the char oxidation. In some
cases it was possible to identify the onset of char oxidation before the extinction of the volatile
cloud. A similar analysis can be made for Fig. 5.7b. The first and second images show again
the particle heat-up and the onset of ignition at t = 15 ms, now on the particle surface. It can
be seen that the particle has an elongated shape and that the ignition occurs at its edges.
This is a common characteristic in most of the surface ignition events captured. After ignition,
from t= 18 ms to t = 21 ms, the flame spreads to the rest of the particle surface. In this case,
the oxidation occurs at the particle surface with no clear volatile cloud.

Ignition of the biomass particles occurred mostly in the gas-phase over the three test
conditions. In fact, surface ignition was only observed for wheat straw and vine branch
particles. Figure 5.8 shows the fraction of gas-phase and surface ignitions over the three
oxygen concentrations for wheat straw and vine branch particles. In these fuels, the fraction
of the particles that ignited at the particle surface increased with an increase in oxygen
concentration, suggesting that higher oxygen concentrations favour surface ignition. With
higher oxygen partial pressure oxygen will more easily reach the particle surface and be
available react in the heterogeneous reaction, thus explaining this effect.

44
Figure 5.7 – Selected images from high-speed cinematography of typical ignition events. a)
Gas-phase ignition of a kiwi branch particle (test 1); b) Surface ignition of a vine branch
particle (test 3).

The surface ignition in these fuels, particularly in the case of the vine branch particle, does
not appear to be correlated with the proximate or ultimate analysis, as all biomass have
similar composition, as shown in table 5.2. Furthermore, the high volatile content of these
fuels, around 65 % for wheat straw and 70 % for vine branches (table 5.1), might suggest a
tendency for gas-phase ignition. Parameters such as shape, thermodynamic properties, and
volatile constituent species may have an important effect on the ignition mode.

a) Wheat Straw b) Vine


Homogeneous Heterogeneous Homogeneous Heterogeneous
100 100

80 80
Frequency [%]

Frequency [%]

60 60

40 40

20 20

0 0
3.5 5.5 6.5 3.5 5.1 6.5
O2 concentration [dry vol %] O2 concentration [dry vol %]

Figure 5.8 – Fraction of gas-phase/surface ignition for a) wheat straw particles and b) vine
branch particles. The remaining fuels experienced 100% of gas-phase ignitions over the three
test conditions.

45
As mentioned above, surface ignition generally occurred at the edges of elongated particles
and then spread to the whole surface. At these sharp edges the surface-to-volume ratio is
very high and temperature increases very rapidly promoting the oxidation at the particle
surface before a significant amount of volatiles has even been released. Interestingly, it can
be observed in Fig. 5.7 that the particles that ignited through hot spots were the ones with
lower aspect ratios (apart from the quasi-spherical pine bark particles), contrarily to what
might be expected. Even though, the shape of the particle may play a role in the ignition
mode, it seems to have an inferior weight when compared to other parameters.

From observation of table 5.2 it can be seen that what differentiates the wheat straw from the
other fuels, in particular from the kiwi branch particles, is its exceptionally high ash content.
This might be indicative of a catalytic behavior that promotes reactions at the surface of the
particle before a significant amount of volatiles has been released. This hypothesis should be
studied in future works, however if this is the case, then for the wheat straw particles the main
mechanism (from the ones considered herein) responsible for the fraction of surface ignition
events observed was likely this catalytic behavior, since in the other parameters considered
this fuel is very similar to kiwi branches. The surface ignition events of vine particles are more
difficult to explain with the observations in table 5.2. The deep pores in these particles might
trap air and hence favour diffusion of oxidizer inside the particle. This feature however is
shared with the pine bark particles that had an even lower volatile content, but ignited 100 %
in the gas-phase. In the case of the vine branch particles the surface ignition is more likely
related to the factors not considered in this work.

Other parameters such as the thermodynamic properties, and volatile constituent species
may have an important effect on the ignition mode, and should be studied in the future.

5.4. Ignition delay time

Figure 5.6 shows the evolution of ignition delay time with oxygen concentration for all tested
conditions. In general a slight decrease in the ignition delay time was witnessed for higher
oxygen concentrations. The higher oxygen partial pressure facilitates the diffusion of oxygen,
either to the volatiles emitted from the particle or to the surface of the particle, and
accelerates the ignition process. In the current work, however, the oxygen concentration
range was limited (3.5 % to 6.5 %) and in these conditions it had a marginal effect over the
ignition delay time. Additionally, it can be seen in Fig. 5.6 that the different fuels presented
similar ignition delay times. This is consistent with the results of Riaza et al. [31] who also
observed that unlike coal the effect of biomass type was weak and the behavior of the
biomass appeared to be more unified. These results are not surprising based on the
composition of the fuels. From table 2.3 it is possible to observe that all tested fuels had
similar proximate analysis results, and so a similar ignition behavior might be expected.

46
Despite this, in phenomena as delicate as ignition even small differences in composition
could have a strong effect. The small difference in ignition delay times can also be explained
with the results from Magalhães [42] who observed that at high temperatures the effect of the
composition of the biomass fuel seemed to lose importance, and different fuels exhibited
similar ignition delay times. In his work, he saw this phenomenon for particles in a similar size
range than that of the current work (224-250 µm) at temperatures higher than 1650K, again
close to the operating temperature in this work (see table 4.1). Therefore, not only are the
compositions of the tested fuels similar, but also their effect on the ignition delay time might
be further dampened by the high temperature operating condition, and so it is easy to
understand the similar ignition delay times presented by the tested fuels.

30
Ignition delay time (ms)

O1 O2 O3
25
20
15
10
5
0
WS KB VB SY PB
Biomass type
Figure 5.6 – Ignition delay times for the operating conditions O1-O3 for wheat straw (WS),
kiwi branch (KB), vine branch (VB), sycamore branch (SB) and pine bark (PB) particles in the
range 212-224 µm. Vertical bars represent 98 % confidence statistical error.

47
6. Numeric results and discussion

6.1. Benchmark case setup

Pine bark was selected as the benchmark fuel. The relevant properties of this fuel, namely
density, specific heat, devolatilization kinetic constants, were taken from a previous work [60]
and are presented in Table 6.1.

Table 6.1 – Properties of pine bark [60].

Fuel 𝛒𝐩 (𝐤𝐠/𝒎𝟑 ) 𝐜𝐩,𝐩 (𝐉/𝐤𝐠 𝐊) 𝐀𝐯 (𝐬 −𝟏 ) 𝐄𝐯 (𝐤𝐉/𝐦𝐨𝐥)


Pine bark 900 1500 15 × 103 50

The volatile matter of the pine bark particles was assumed to be composed of CO, H 2, O2 and
N2 as described in section 3.4, and is presented in Table 6.2.

Table 6.2 – Assumed volatile matter composition of pine bark for the numerical simulations
(% weight).

Volatile matter composition


CO 0.87
H2 0.07
O2 0.05
N2 ~0

Six operating conditions were selected. Besides the three experimental conditions with
different oxygen concentrations tested in this work, three additional conditions with different
gas temperatures in the ignition zone (from a previous work [42]) were considered. These
conditions are presented in table 6.3.

Table 6.3 – Test conditions.

Parameter O1 O2 O3 T1 T2 T3
Thermal input [kW] 1.0 1.1 1.4 0.6 1.0 2.0
Equivalence ratio ϕ [-] 0.79 0.71 0.67 0.71
O2 concentration [dry vol. %] 3.5 5.1 6.5 7.6
Gas temperature [K] 1680 1500 1650 1800

Figure 6.1 shows the axial profiles along the axis of the burner of oxygen concentration and
temperature for the conditions presented in table 6.3. Figure 6.1a shows the three distinct
oxygen conditions (O1-O3), and Fig. 6.1b shows the three different temperature conditions
(T1-T3). Predictions show that despite all the simplifications considered the numeric profiles

48
and experimental results show a reasonable agreement. The main differences appear close
to the burner arise from the difficulty of measuring these quantities close to the burner,
especially the species concentrations (as described in section2.3).

Figure 6.1 – Numeric and experimental (a) Dry oxygen mole fraction and (b) gas temperature
along the centerline for the conditions O1-O3 and T1-T3, respectively.

6.2. Ignition process

Figure 6.2 shows the mass and temperature time histories of a typical pine bark particle
(specifically for test condition O3).

Figure 6.2 – Particle temperature and non-dimensional mass depletion. Zones I, II, and II
show the drying, devolatilization and char combustion phases.

In Fig. 6.2 three distinct phases can be identified, namely: (I) evaporation, (II) devolatilization,
and (III) char oxidation. In phase (I), the mass decays quickly as the temperature of the
particle increases and water vapor is released into the domain. This is a highly endothermic

49
process, and during this phase the temperature of the particle is limited by the latent heat of
the water in the particle. After the drying phase, the temperature of the particle stars to
increase more quickly since no heat is required to vaporize the water. The mass starts to
decay again when the particle reaches a temperature close to 500 K, and as volatiles gases
CO and H2 start to be released to the surrounding gas. The heating process is further
accelerated in the devolatilization phase. As the mass decays, so does the product 𝑚 ∙ 𝑐𝑝 ,
implying that for the same amount of heat the temperature of the particle is more easily
changed. In the conditions tested, and in particular for the condition O3 illustrated in Fig. 6.2,
ignition occurred just after the onset of devolatilization. In this case approximately for t = 23
ms. This is a consequence of the high temperatures of the surrounding atmosphere, at which
the reaction rates are very high. Although the ignition occurred (according to the criterion
discussed in section 6.4. - next section) the temperature of the gas phase around the particle
does not immediately shoot. In these first moments after its onset, the devolatilization occurs
slowly and the amount of volatile gases that are injected into the gas phase is very little (as
seen in Fig. 6.2). Even though the reaction of these volatiles is fast to occur, the amount of
energy released by them is not sufficient to increase the temperature significantly. After all the
volatiles have been released the char oxidation (III) begins. In this phase the particle
continues to lose mass now at a rate at a much slower rate than that of the devolatilization
phase, consistently to what was described in [33].

As the devolatilization starts approximately at t = 20 ms the volatile gases CO and H 2 start to


be injected into the domain. In the first stages of combustion, these stable species dissociate
to form radicals. A radical is an atom, molecule or ion that has unpaired valence electrons,
e.g. CH* or OH*, thus making it highly reactive. The dissociation of reactants into radicals is
strongly endothermic and quite slow in the beginning. After being formed radicals will react
rapidly in one of two types of reaction: either reacting in a chain branching propagation or
chain termination reaction. Chain propagation occurs when the number of radicals in the
reaction products is equal or greater than the number of radicals in the reactants. A particular
case is the chain branching where the number of radicals increases. These radicals in turn
react quickly releasing heat that contributes for the dissociation reactions and to further
radical propagation reactions.

The theory of thermal explosion is based upon a very simple concept. A thermal explosion
occurs when a chemical system undergoes an exothermic reaction during which insufficient
heat is removed from the system so that the reaction process becomes self-heating. Since
the rate of reaction, and hence the rate of heat release, increases exponentially with
temperature, the reaction rapidly runs away; that is, the system explodes. When the opposite
occurs thermal ignition is impossible. Chain termination reactions occur when radicals react to
give either a molecular species or a radical of lower activity that cannot propagate the chain.
These reactions are associated with the extinction of the combustion process.

50
Thus, looking at the concentration of radical species is a typical way of measuring the ignition,
both experimentally and numerically. Figure 6.3 shows the contours of temperature and OH
mass fraction around the particle between t = 21 ms and t = 30 ms (close to ignition) and
images of an ignition event from a pine-bark particle under test condition 3.

Figure 6.3 – Contour plots of temperature and normalized OH mass fraction in the vicinity of
the injection hole and close to the axis of the burner between t = 21 ms and t = 30 ms.

51
In Fig. 6.3 it is possible to see the relatively slow dissociation reactions by looking at the
slowly increasing concentration of OH, which then shoots and quickly reaches OH
concentrations close to the maximum, with a simultaneous increase in gas temperature
around the particle. In the later frames, a region of low OH mass fraction can be identified
close to the volatiles source (where the particle would be). Again this is a sign of the relatively
slow dissociation of the stable species close to the source of volatiles.

6.3. Definition of a numerical ignition criterion

Computationally, it is not obvious how to determine the ignition point. For example, threshold
values of temperature or species mass fractions, or the inflection point in the particle
temperature history can be used, as suggested in [50].

CH* chemiluminescence has proved to be a good experimental indicator of the onset of


ignition [35]. Hence the mass fraction of CH, as a CH* representative, as discussed in the
previous section, would be the obvious choice as an indicator of ignition. However, to
compute the mass fraction of CH a large chemical mechanism, such as the GRI3.0, would be
necessary. Since the DRM mechanism was used, the OH radical was considered instead in
the numerical simulations to predict the ignition onset in a similar way to that of CH. Vascellari
et al. [52] observed in their work that the onset of CH and OH happened almost
simultaneously, but for slightly different mixture fraction values, thus showing that this is a
good criterion to determine the ignition delay time. In their work they obtained a good
agreement between the experimental results of Molina and Shaddix [35], and the numerical
results with OH as indicator of ignition.

A similar approach was considered in this work, by taking a fraction of the maximum OH
mass fraction in the domain as an indicator for the onset of ignition. Figure 6.4 shows the
profiles of maximum OH mass fractions in the domain normalized by the maximum OH mass
fraction in each test, for test conditions O1-O3 and T1-T3, with the respective experimental
ignition times. In the studied cases of the pine-bark particles the maximum fraction of OH
increases steeply just after the onset of devolatilization, showing that at the high temperatures
the reaction is very fast to begin after some volatile gases have been injected into the gas
phase. The plots in Fig. 6.4 exhibit a hump corresponding to period of the volatile-cloud
flame. Figure 6.4c can be compared to Fig. 6.2 to show that the maximum OH mass fraction
shoots just after the onset of devolatilization when ignition occurs, in this case approximately
for t = 23 ms, and starts to decay slightly after all volatiles in the particles have been released,
corresponding to the extinction of the volatile flame, approximately for t = 42 ms.

52
Figure 6.4 – Time-history of maximum OH mass fraction in the domain normalized by the
maximum in each condition. a) O1, b) O2, c) O3, d) T1, e) T2, f) T3. The dashed lines
represent the experimental ignition times.

Furthermore, it can be seen from Fig. 6.4a to c and from d to f that this hump becomes
narrower, showing that both increasing the oxygen concentration and gas temperature tend to
reduce the duration of the volatile-flame. In order to be consistent with the experimental
analysis, the numeric ignition was considered to occur when 15 % of the maximum OH mass
fraction was reached.

6.4. Comparison between experimental and predicted ignition delay


times

Figure 6.5 shows the experimental and computational ignition delay times, as computed by
the criterion above stated. As stated above, at the high temperatures tested the numeric
ignition occurred just briefly after the onset of devolatilization, not varying significantly with the
different oxygen concentrations tested. This reduced influence of the oxygen concentration
over the ignition delay time can be observed in Fig. 6.5a. An increase in the gas temperature,
on the other hand, results in an anticipation of the onset of the devolatilization, which in turn
reflects in a shorter ignition delay time. Thus, in the simulations the surrounding gas
temperature shows a much more intense effect on the ignition time (see Fig. 6.5b).

53
Figure 6.5 – Experimental and numeric ignition delay times for conditions a) O1-O3 and b)
T1-T3. The vertical bars represent 98 % confidence interval.

From Fig. 6.5 it can also be concluded that the ignition delay times were, generally, over-
predicted. This is similar to what was observed by Vascellari et al. [52] in a similar unresolved
simulation of the ignition of bituminous coal particles. Despite this, in general, a good
agreement between the experimental and numeric results was achieved, especially taking
into account the complexity of the studied phenomena. An exception occurred for the test
condition T3. In this case the deviation between experimental and numeric ignition delay is 53
%. At these higher temperatures the model fails to reproduce the experimental ignition times.
Possibly other parameters not captured by the model, such as catalytic effects, become
important. Also, it should be noted that the kinetic rate constants of devolatilization used (see
table 6.1) were computed in a drop tube furnace for temperatures much lower than that of
condition T3 [60], which may further contribute to the large deviation between experimental
and numeric results. At higher temperatures devolatilization occurs sooner and faster. This
effect is not considered when using the kinetic constants derived at a lower temperature
condition.

These simulations show that the model used herein, can predict the ignition delay time of
solid fuel particles, given that adequate kinetic rate constants of devolatilization are used.
This is valid for the cases where ignition occurred in the gas-phase. The model used was
limited and did not have the ability to predict the heterogeneous ignition.

54
6.5. Analysis of the dominant phenomena on the ignition delay time

The ignition process up to ignition was discussed in section 1.3. As mentioned there, three
phases occur, namely heat-up, evaporation and devolatilization phases. Through the analysis
of the numeric results, it is possible to understand the weight of each of these phases in the
ignition delay time. Figure 6.6 shows the ignition delay times decomposed by their constituent
phases: sequentially heat-up, drying, heat-up, and devolatilization up to ignition. These were
obtained from the mass and temperature profiles of the particles for the tested conditions.

Fig. 6.6 - Decomposition of ignition times in heat-up, drying, heat-up and devolatilization up to
ignition for test conditions a) O1-O3 and b) T1-T3.

There is a clear contrast between Fig. 6.6a and b. While the oxygen concentration has
practically no effect in any of the phases, the temperature of the atmosphere seems to affect
all of them. This can be seen in the decreasing duration of these phases with increasing gas
temperature.

55
In particular it can be seen that the heat-up and drying phases dominate the ignition delay
time, consisting of up to 80-90% of the total ignition delay time in the tested cases. Properties
of the particle, which have a strong influence during these phases such as the density,
specific heat and moisture content of the particle, will thus have a strong impact on the
ignition delay time. As discussed in section 1.3, the effect of density and specific heat has
already been investigated in a previous work [42]. The specific heat was shown to have a
reduced effect on the ignition delay time, but the density, which can significantly vary in
biomass fuels, presented a much more pronounced effect. It then becomes interesting to
analyse the effect of the moisture content.

The effect of this parameter was analysed by considering a particle with the same properties
of pine bark, namely the specific heat and density and with a variable moisture content. This
study was done only for the three different temperature test conditions since, as seen in Fig.
6.6, the oxygen concentration had a marginal effect either in the drying or ignition delay times.
Figure 6.9 shows the drying times as a function of the particle moisture content for tests T1-
T3.

Fig. 6.9 - Drying times as a function of the moisture content of the fuels for conditions T1-T3.

In any of the three conditions it can be seen that, as expected, drying times increase with the
moisture content. This growth is very rapid for low moisture fractions, but quickly reaches a
slower steadier pace, which is seen in the higher moisture percentages.

The drying phase is governed by the moisture concentration gradient between the surface of
the particle and the surrounding atmosphere, which promotes the transport of this water
through diffusion (the model is able to ensure that high evaporations do not occur in saturated
environments). On the other hand, the moisture concentration at the particle surface
increases with increasing particle temperature, exhibiting a exponential behavior. In the case
of particles with a lower moisture content, given the low quantity of water to evaporate, the
drying phase occurs in a short period in which the particle does not have time to heat-up
significantly. In this case the moisture content at the particle surface does not reach high

56
values. In particles with higher moisture contents the drying phase occurs in a longer period,
due to the additional mass to evaporate. In this case the particles have more time to heat-up
and the drying phase will occur at higher temperatures, when compared with lower moisture
particles. In these conditions, there is an increase in the drying rate due to the increasing
moisture concentration at the particle surface.

57
7. Closure

7.1. Conclusion

The current work focused on the experimental and numerical study of the ignition of single-
particles of pulverized solid fuels. The experimental study extended previous work, in order to
include a broader diversity of biomass fuels and the effect of atmosphere composition, more
specifically the oxygen concentration. Furthermore, a numeric strategy using a commercial
CFD software to model the combustion behavior, and more specifically ignition, of single
biomass particles was validated. Subsequently, the dominant phenomena on the ignition
delay time were numerically investigated. The main conclusions of the experimental study
are:

Some biomass particles, namely the wheat straw and vine branch particles under
certain conditions, despite their high content of volatiles presented surface
ignition.
This incidence did not seem to show a relation with particle composition or its
shape/morphology, however it increased with increasing oxygen concentrations.
Even though the ignition delay times decreased slightly with increasing oxygen
concentrations, the effect of this parameter was marginal.

The main conclusions of the numeric work are:

For the selected benchmark case, the numeric results showed a good agreement
with the experimental data showing that the model is able to reproduce in a
reasonable way the combustion process of the single-particles, particularly the
ignition delay time.
The numeric ignition times followed the same experimental trends: (i) marginal
effect of the oxygen concentration, (ii) decreasing ignition delay times with
increasing temperatures.
The model was then used to investigate the effect of the fuel moisture content. As
expected, drying times increased with increasing moisture and decreasing
atmosphere temperature. The drying phase was seen to represent a significant
fraction of the ignition delay, showing that an increase in the moisture is likely to
be reflected in greater ignition delay times.

7.2. Outlook

One interesting observation that resulted from this study is that the ignition mode of biomass
fuels is not as linear as one might have thought. High volatile composition does not always

58
mean gas-phase ignition, in the same way that surface ignition does not always correlate with
higher aspect ratios of the particles (sharp/elongated particles). In this context, it would be
interesting to do a more detailed analysis of the dominant parameters behind the ignition
mode, including ash composition, porosity, volatile constituent species, thermodynamic
properties, etc. Some of these could be achieved simply by having a more complete
characterization of the fuels tested, others such as the ash composition could be studied by
either washing or enriching the particles with certain elements.

It would also be interesting to investigate the effect of the heat transfer mechanism on the
ignition of these fuels, using for example a drop tube furnace, where the dominant
mechanism is radiation, in comparison to the McKenna burner where convection is dominant.
This could be achieved by burning the same biomass fuels tested in this work in similar
temperature and oxygen conditions but in a drop tube furnace with an optical windoe.

Finally, it should be mentioned that the numeric analysis included in this work is fairly simple,
and that progressing to a more complete and detailed model would also be an interesting
step.

59
References

[1] U.S. Global Change Research Program, 2014 National Climate Assessment. [Consult. 28
August 2015]. Available on http://nca2014.globalchange.gov

[2] IPCC Fifth Assessment report, Working Group I - Climate Change 2013: The Physical
Science Basis. [Consult. 28 August 2015]. Available on

http://www.climatechange2013.org/images/report/WG1AR5_SPM_FINAL.pdf

[3] J.P. Smart, R. Patel, G.S. Riley, Oxy-fuel combustion of coal and biomass – the effect on
radiative and convective heat transfer and burnout, Combustion and Flame, 157 (12) (2010),
pp. 2230–2240.

[4] United States Environmental Protection Agency, Overview of green-house gases.


[Consult. 28 August 2015]. Available on

http://www.epa.gov/climatechange/ghgemissions/gases.html

[5] International Energy Agency, World Energy Outlook 2014 – Executive Summary.

[6] Bp Energy Outlook 2035, January 2014. Available on bp.com/energyoutlook

[7] BP Statistical Review of the World Energy June 2014.

[8] T.F. Wall, Combustion processes for carbon capture, Proceedings of the Combustion
Institute 31 (1)(2007), pp. 31-47.

[9] B. Abraham, J.G. Masburry, E.P. Lynch, A.P.S. Teotia, Coal oxygen process provides CO 2
for enhanced oil recovery, Oil and Gas 80(1982), pp. 68-75.

[10] B. Buhre, L. Elliott, C. Sheng, R. Gupta, T. Wall, Oxy-fuel combustion technology for
coal-fired power generation, Progress in Energy and Combustion Science, 31, (4), (2005),
283–307.

[11] R. Tan, G. Corragio, S. Santos, Report No. IFRF Doc. No. G 23/y/1, International Flame
Research Foundation, (2005).

[12] M. Ditaranto, J. Hals, Combustion instabilities in sudden expansion oxy-fuel flames,


Combust. Flame, vol. 146, pp. 493-512,2006.

[13] C.R. Shaddix, A. Molina, Particle Imaging of ignition and devolatilization of pulverized
coal during oxy-fuel combustion, Proc. Combust. Inst., vol 32, pp. 2091-2098,2009.

[14] F. Al-Mansour, J. Zuwala, An evaluation of bio-mass co-firing in Europe, Biomass and


Bioenergy, 34, (2010), 620-629.

[15] A. Williams, J. Jones, M. Mall, M. Pourkashanian, Combustion of pulverised coal and


biomass, Progress in Energy and Combustion Science, 27, (2001), 587-610.

[16] M. Sami, K. Annamalai, M.Wooldridge, Co-firing of coal and biomass fuel blends, Prog
Energy Combust. Sci., vol 24, pp. 232-240,2001.

60
[17] J. Riaza, M.V. Gil, L. Álvarez, C. Pevida, J.J. Pis, F. Rubiera, Oxy-fuel combustion of coal
and biomass blends, Energy, vol 41, pp. 429-435,2012.

[18] F. Kazanc, R. Khatami, P.M. Crnkovic, Y.A. Levendis, Emissions of NOx and SO2 from
Coals of Various Ranks, Bagasse, and Coal-Bagasse Blends Burning in O2/N2 and O2/CO2
Environments, Energy Fuels, 25 (7) (2011), pp. 2850–2861.
[19] D.A. Tillman, Biomass cofiring: the technology, the experience, the combustion
consequences, Biomass Bioenergy, 19 (6) (2000),pp. 365–384.

[20] L. Baxter, Biomass-coal co-combustion: opportunity for affordable renewable energy,


Fuel, 84 (10) (2005), pp. 1295–1302.

[21] R. Saidur, E. Abdelaziz, A. Demirbas, M. Hossain, S. Mekhilef, A review on biomass as a


fuel for boilers, Renewable Sustainable Energy Review, 15, (5), (2011).

[22] T.F. Wall, R. Gupta, V. Gururajan, D. Zhang, The ignition of coal particles, Fuel, 70, (9),
(1991), 1011–1016.

[23] Integro – Earth Fuels, The Benefits of Biomass Power. [Consult. 28 August 2015].
Available on http://www.integrofuels.com/sustainability/biomass-benefits/

[24] J. Koppejan, S. van Loo, The Handbook of Biomass Combustion and Co-firing, (2012).

[25] S. Vassilev, D. Baxter, L. Andersen, C. Vassileva, An overview of the chemical


composition of biomass, Fuel, 89, (2010), 913–933.

[26] M.J. Antal, Biomass pyrolysis: a review of the literature – Part I – carbonhydrate
pyrolysis, Advances in Solar Energy Technology 11 (1983), pp. 61 – 111.

[27] K.G. Mansaray, A.E. Ghaly, Thermal degradation of rice husks in nitrogen atmosphere,
Bioresource Technology 65 (1998), pp. 13 – 20.

[28] J. Dias, Utilização de biomassa: avaliação dos resíduos e utilização de pellets em


caldeiras domesticas, Tese de Mestrado, Instituto Superior Técnico, Universidade Técnica de
Lisboa (2002)

[29] B.V. Babu, A.S. Chaurasia, Heat transfer and kinetics in the pyrolysis of shrinking
biomass particle, Chemical Engineering Science 59 (10) (2004), pp. 1999-2012.

[30] F. Costa, Características de combustão de biomassa torrada num reactor de queda livre,
Tese de Mestrado, Instituto Superior Técnico, Universidade Técnica de Lisboa (2014)

[31] J. Riaza, R. Khatami, Y. Levendis, L. Álvarez, M. Gil, C. Pevida, F. Rubiera, J. Pis,


Combustion of single biomass particles in air and in oxy-fuel conditions, Biomass and
Bioenergy, 64, (2014), 162-174

[32] T.H. Fletcher et al., A chemical percolation model for devolatilization: Summary

[33] P. Coelho, M. Costa, Combustão, Edições Orion, 2ª Edição, (2012).

61
[34] J. Howard, R. Essenhigh, The mechanism of ignition of pulverized coal, Combustion and
Flame, 9, (3), (1965), 337-339.

[35] A. Molina, C. Shaddix, Ignition and devolatilization of pulverized bituminous coal particles
during oxygen/carbon dioxide coal combustion, Proceedings of the Combustion Institute, 31,
(2), (2007), 1905–1912.

[36] Y. Liu, M. Geier, A. Molina, C. Shaddix, Pulverized coal stream ignition delay under
conventional and oxy-fuel combustion conditions, International Journal of Greenhouse Gas
Control, 5S, (2011), S36-S46.

[37] Y.A. Levendis, K. Joshi, R. Khatami, A.F. Sarofim, Combustion behavior in air of single
particles from three different coal ranks and from sugarcane bagasse, Combustion and Flame
158 (2011), 452–465.

[38] R. Khatami, C. Stivers, K. Joshi, Y. Levendis, A. Sarofim, Combustion behavior of single


particles from three different coal ranks and from sugar cane bagasse in O2/N2 and O2/CO2
atmospheres, Combustion and Flame, 159, (2012), 1253-1271.

[39] R. Khatami, C. Stivers, Y. Levendis, Ignition characteristics of single coal particles from
three different ranks in O2/N2 and O2/CO2 atmospheres, Combustion and Flame, 159, (2012),
3554–3568.

[40] J. Riaza, R. Khatami, Y. Levendis, L. Álvarez, M. Gil, C. Pevida, F. Rubiera, J. Pis, Single
particle ignition and combustion of anthracite, semi-anthracite and bituminous coals in air and
simulated oxy-fuel conditions, Combustion and Flame, 161, (2014), 1096-1108.

[41] Y. Yuan, S. Li, G. Li, N. Wu, Q. Yao, The transition of heterogeneous–homogeneous


ignitions of dispersed coal particle streams, Combustion and Flame, 161, (9), (2014), 2458-
2468.

[42] D. Magalhães, Ignition behavior of single biomass and coal particles, Msc Thesis,
Instituto Superior Técnico, University of Lisbon, (2015).

[43] C. Zou, L. Cai, D. Wu, Y. Liu, S. Liu, C. Zheng, Ignition behaviors of pulverized coal
particles in O2/N2 and O2/H2O mixtures in a drop tube furnace using flame monitoring
techniques, Proceedings of the Combustion Institute, 35, (3), (2014) 3629-3636.

[44] G. Soete, Ignition of coal particles: a review, Rev. Inst. Franc. Petr., 37, (1981), 403–530.

[45] X. Du, K. Annamalai, The transient ignition of isolated coal particle, Combustion and
Flame, 97, (1994), 339–354.

[46] S. G. Davis, A.V. Joshi, H. Wang, F. Egolfopoulos, An optimized kinetic model of H2/CO
combustion, Proceedings of the Combustion Institute, 30, (2005) 3147-3155.

[47] J. B. Howard, R.H. Essenhigh, Mechanism of solid-partical combustion with simultaneous


gas phase volatiles combustion, Proceedings of the Combustion Institute, 11, (1967) 399-
408.

62
[48] J C.R. Shaddix, in: 33rd National Heat Transfer Conference NHTC'99, American Society
of Mechanical Engineers, Albuquerque, NM, USA, 1999, p. 1150.

[49] ANSYS, Inc., ANSYS Fluent Theory Guide, Release 15.0, November 2013.

[50] B. Goshayeshi, J. Sutherland, A comparison of various models in predicting ignition delay


in single-particle coal combustion, Combustion and Flame, 161, (7), (2014), 1900-1910.

[51] K.K. Watson, Ind. Eng. Chem. 35 (4)(1943)398-406

[52] M. Vascellari, H. Xu, C. Hasse, Flamelet modeling of coal particle Ignition, Proceedings
of the Combustion Institute, 34, (2), (2013), 2445-2452.

[53] H. Watanabe, M. Otaka, Fuel 85 (2006) 1935-1943.

[54] B.E. Poling, J.M. Prausnitz, J.P. O'Connel, The properties of gases and liquids, McGraw
Hill, 5th Edition, (2000;

[55] S.K. Bhatia, D.D. Perlmutter, AIChE J 1980;26:379.

[56] S. Kajitani, S.Hara, H. Matsuda, Fuel 2002;81:539.

[57] B. Franchetti, Large Eddy Simulation of Air and Oxy-Coal Combustion, PhD Thesis,
Imperial College London, (2013).

[58] A. Kazakov, M. Frenklach, Reduced Reaction sets based on gri1.2, 1994, available at

http://www.me.berkeley.edu/drm/
[59] Gregory P. Smith, David M. Golden, Michael Frenklach, Nigel W. Moriarty, Boris
Eiteneer, Mikhail Goldenberg, C. Thomas Bowman, Ronald K. Hanson, Soonho Song,
William C. Gardiner, Jr., Vitali V. Lissianski, and Zhiwei Qin
available at http://www.me.berkeley.edu/gri_mech/
[60] G. Wang, (Co‐)combustion of solid fuels: experiments and modeling, PhD Thesis,
Instituto Superior Técnico, University of Lisbon, (2014).

63
Annex A – Thermocouple measurement correction

A.1 - Working principle

A thermocouple is a device made by two different wires joined at one end, called junction end
or measuring end. The other end of the thermocouple is called tail end or reference end.
Figure A1.1 shows a schematic of a general thermocouple configuration.

Figure A1.1 – Schematic of thermocouple configuration.

The working principle of a thermocouple is based on the Seebeck and Peltier effects. These
effects state that when two different or unlike metals are joined together at two junctions, an
electromotive force (emf) is generated within the circuit because of the differences in
temperature between the junctions. Thus a voltage difference can be measured between the
two thermo-elements at the tail end, i.e., the thermocouple is a temperature-voltage
transducer.

A.2 - Heat transfer about a thermocouple

Figure A1.2 shows a schematic of a Mckenna burner with a thermocouple as mounted in the
tests.

Figure A1.2 – Montage of the thermocouple positioning for the tests in the McKenna burner.

64
In this situation the thermocouple exchanges heat through convection, conduction, radiation
and catalytic reaction. Figure A1.3 presents a schematic of the thermocouple head with these
modes of heat transfer.

Figure A1.3 – Schematic of the thermocouple head [A1].

In the most general case an energy balance on the thermocouple takes the following form,

𝑑𝑇𝑡𝑐 (A1)
𝑄̇𝑐𝑎𝑡 + 𝑄̇𝑐𝑜𝑛𝑣 + 𝑄̇𝑟𝑎𝑑 + 𝑄̇𝑐𝑜𝑛𝑑 = 𝜌𝑐𝑝 𝑉
𝑑𝑡

where 𝑄̇𝑐𝑎𝑡 is the heat transfer associated with surface induced catalytic reactions, 𝑄̇𝑐𝑜𝑛𝑣 is
the heat transfer by convection between the gases and the thermocouple wires, 𝑄̇𝑟𝑎𝑑 is the
radiant heat transfer between the thermocouple and its surroundings, 𝑄̇𝑐𝑜𝑛𝑑 is the heat
transfer by conduction along the thermocouple wires.

Since the object of a thermocouple is to determine the local gas temperature, it is desirable
for the convective heat transfer term to dominate over all others associated with the
thermocouple wire and junction.

In practice the catalytic effects are very difficult to quantify, so attempts are usually made to
minimize these terms (typically by applying a non-catalytic coating on the thermocouple wire
and junction, when required. For these reasons, the catalytic term is not considered in this
analysis.

The temperature measurements are taken when the thermocouple reaches a steady-state,
i.e., its temperature was constant in time. In this case equation A1 reduces to,

𝑄̇𝑐𝑜𝑛𝑣 + 𝑄̇𝑟𝑎𝑑 + 𝑄̇𝑐𝑜𝑛𝑑 = 0 (A2)

65
Convective term

Figure A1.4 shows the convective heat transfer from the hot combustion product gas to the
thermocouple.

Figure A1.4 – Schematic of the convection heat transfer to the thermocouple head.

The convective term is given by,

𝑄̇𝑐𝑜𝑛𝑣 = 𝐴𝑏 ℎ(𝑇𝑔 − 𝑇𝑡𝑐 ) (A3)

where 𝑇𝑔 and 𝑇𝑡𝑐 are the temperatures of the gas and of the thermocouple, and 𝐴𝑏 is the
surface area of the thermocouple bead. Shaddix [A1] suggest that the bead diameters are
typically 2-5 times larger than the wire diameter. Thus, the thermocouple bead was assumed
to have the shape of a cylinder with base diameter, and height equal to 3.5 times the
diameter of the thermocouple wires, and as a result a surface area of 𝐴𝑏 = 𝜋(3.5)2 𝑑𝑤𝑖𝑟𝑒 . ℎ in
equation A3 is the convection coefficient,

𝑁𝑢 𝑘𝑔 (A4)
ℎ=
𝑑𝑏

where 𝑁𝑢 is the Nusselt number, 𝑘𝑔 is the gas thermal conductivity, and 𝑑𝑏 is the bead
diameter. In his work, Shaddix reviewed several correlations for the Nusselt number in this
type of thermocouple and recommended a widely the following correlation by Kramers [A1],

𝑁𝑢𝑑,𝑐𝑦𝑙 = 0.42 𝑃𝑟 0.42 + 0.57 𝑃𝑟1/3 𝑅𝑒𝑑 1/2 (A5)

66
which extends over 0.01<𝑅𝑒𝑑 <10000, with the gas properties evaluated at the "film-
temperature", Tm.

Conduction term

Figure A1.5 shows the conduction heat transfer along the thermocouple wires.

Figure A1.5 – Schematic of the conduction heat transfer to the thermocouple head.

The conduction term is hard to calculate, however a rough estimate can be made by
considering some simplifications.

The conduction term is given by,

𝑑 𝑇𝑡𝑐 (A6)
𝑄̇𝑐𝑜𝑛𝑑 = −𝐴𝑐𝑟𝑜𝑠𝑠 𝑘𝑏
𝑑𝑥

where 𝐴𝑐𝑟𝑜𝑠𝑠 is the cross section of the thermocouple wires, and 𝑘𝑏 is the thermal
conductivity of the bead.

The thermal conductivity of the bead is taken as an average of the conductivities of the two
wires. The experimental correlations for these properties are given by [A2],

𝑘𝑃𝑡 = 0.0198 𝑇 + 64.141 (A7)

𝑘𝑃𝑡−10%𝑅ℎ = 0.006 𝑇 + 28.385

𝑑 𝑇𝑡𝑐 Δ𝑇
The term can be approximated by , which can be estimated by assuming that at a
𝑑𝑥 Δ𝑥

certain temperature at a certain distance from the bead. Assuming that at a distance equal to
two times the diameter of the hole, through which particles are injected, the temperature is

67
equal to the plateau temperature far away from the injection hole (see section 4.1 - Test
conditions) the order of magnitude of this term can be estimated.

Radiative term

Figure A1.6 shows the radiation heat transfer from the hot flame sheet to the thermocouple,
and from the thermocouple to the cold ambient environment. Near the burner the
thermocouple gains heat by radiation from the hot flame, but far from the burner the net
radiation exchange results in a heat loss.

Figure A1.6 – Schematic of the radiation heat transfer to the thermocouple head.

Taking into account that the current calculation is only a rough estimate of the error the
assumptions described below were taken.

The atmosphere gas was assumed to be transparent to radiation, which is a reasonable


consideration taking into account that this gas is mainly constituted by di-atomic gases, N2
and O2.

The thermocouple was assumed to have a gray, diffuse surface. The thermocouple bead was
split into two surfaces, the top surface, which emits radiation to a distant large black surface
at ambient temperature, and the bottom surface which partially emits radiation to the cooler
"environment surface" and absorbs radiation emitted by a hot circular black surface
(emulating the hot flame).

In these conditions, the radiation term is then given by,

𝑄̇𝑟𝑎𝑑 = (𝐴𝑏 /2) 𝜎 𝜀 (𝑇𝑡𝑐 4 − 𝑇𝑎𝑚𝑏𝑖𝑒𝑛𝑡 4 ) + (𝐴𝑏 /2) 𝜎𝜀 [𝐹12 (𝑇𝑓𝑙𝑎𝑚𝑒 4 − 𝑇𝑡𝑐 4 ) + 𝐹13 (𝑇𝑡𝑐 4 − (A8)

68
𝑇𝑎𝑚𝑏𝑖𝑒𝑛𝑡 4 )]

where 𝜎 is the Stefan-Boltzmann constant, 𝑇𝑎𝑚𝑏𝑖𝑒𝑛𝑡 and 𝑇𝑓𝑙𝑎𝑚𝑒 are the ambient temperature
and the temperature of the flame (assumed to be the adiabatic temperature of the flame) and
𝐹𝑖𝑗 are the surface view factors of surfaces. Surface 1 is the bottom surface of the
thermocouple, surface 2 is the hot surface emulating the flame, and surface 3 is the distant
surface at ambient temperature emulating the ambient atmosphere. The view factor 𝐹12 is
calculated as in [A2], and 𝐹13 is given by 𝐹13 = 1 − 𝐹12 .

The emissivity of the bead is taken as an average of the emissivities of the two wires. The
experimental correlations for these properties are given by [A1],

𝜀𝑃𝑡 = 0.136 𝑙𝑛(𝑇) − 0.8047 (A9)

𝜀𝑃𝑡−10%𝑅ℎ = 0.1357 𝑙𝑛(𝑇) − 0.7887

A.3 – References

[A1] C.R. Shaddix, Correcting thermocouple measurements for radiation loss: a critical
review, Prodeedings of the 33rd National Heat Transfer Conference, Albuquerque, New
Mexico, August 15–17, 1999.

[A2] Hindasageri, R. P. Vedula, and S. V. Prabhu, Thermocouple error correction for


measuring the flame temperature with determination of emissivity and heat transfer
coefficient, Review of Scientific 84, 024902 (2013)

[A3] T.L Bergman, A.S. Lavine, F.P. Incropera, D.P. Dewitt, Fundamentals of heat and mass
transfer, 7th edition, John Wiley & Sons,

69
Annex B – Image processing procedure

In this section the image processing procedure is briefly explained. This procedure was
developed in order to ‘’track’’ the particle, i.e., determine the particle position in each frame.
The position of the particle was then used to compute the velocity between every two frames
up to the frame corresponding to ignition. Finally, this velocity, along with the position of the
particle, was used to determine the experimental ignition delay time.

In order to see the particle before ignition a backlight was used. This backlight was
very dim so not to outshine the light from the volatile cloud. The particle before ignition was
seen as a dark blob. In the beginning, since the backlight is approximately uniform, the blob
corresponding to the particle is darker than the whole background. In some cases this is true
up until ignition. In others, however, this does not happen as a consequence of the ignition
criterion taken. In this work, ignition was considered to occur when 15% of the maximum pixel
luminosity was reached, which means that in some cases some faint luminosity can be
spotted around the particle even before ignition. In this scenario, the dark blob corresponding
to the particle can be darker than the whole background, or it can just be a local minimum of
luminosity, i.e., it can be darker than the dim luminosity around the particle, but still brighter
than the background.

Considering these constraints and the noise in the images a matlab® routine was
developed to compute the velocity of the particle from frame to frame. In this section the
procedure is briefly explained. The main idea of the algorithm is to locate a region with
luminosity close to the maximum frame luminosity and in that region look for the particle. In
the first frames, far from the ignition frame, pixel luminosity is more or less uniform and the
region close to the maximum of the frame corresponds roughly to the whole frame, with the
exception of the dark blob of the particle and some local minimums due to the noise. Close to
ignition, in the cases where the luminosity cloud can be identified this cloud corresponds to
the maximum region where the particle is.

The steps of the algorithm are listed below:

1 – Read the image and normalize it with the initial frame, in which only the backlight is
in frame.

2 – Define the region close to the maximum. This region is defined as the pixels with
luminosity intensity above a certain threshold. This base level is defined as the top x%
luminosity intensities, where x is a function of the maximum pixel intensity. (In the frames
close to ignition the range of intensities is much broader and to get the region of interest
where the particle is contained x has to increase accordingly.)

3 – Clean small local maximum and minimum blobs.

4 – In the resulting region, look in the original image for a region close to the minimum.

70
5 – Calculate the centroid of the resulting blob, and use its position to compute the
velocity of the particle.

In Table B1 some of the steps stated above are shown for the case where a luminosity
cloud can be identified.

71
Table B1 – Examples of the post-processing of the high-speed images.
Clean small local
In the resulting region Find the centroid of
Define the region minimums (3) to
Normalize the image Clean small local in the original image the minimum region
Original Image Read the image (1) close to the maximum define the region
(1) maximums (3) define a region close to define the position
luminosity (2) (yellow) in which the
to the minimum (4) of the particle (5)
particle is contained.
Frame far from ignition
Frame close to ignition
72
Annex C – Equipment used

Table A.1 shows photographic or schematic images and the respective working principles of the
equipment used in the experimental analysis.

Table A.1 – Equipment used and respective working principle.


Equipment Photograph/Schematic Working principle

Rotameters are also known as gravity type


flow-meters because they are based on
the opposition between the downward
ABB
force of gravity and upward drag force
Rotameter
from the fluid. When the flow is constant
the rotameter stays in one position that
can be related to the volumetric flow rate.

The two used flow-meters use the same


working principle.
Aalborg
In order to sense the flow in the sensor
Mass-flow
tube, heat flux is introduced at two
meter GFM
sections of the sensor tube by means of
37
precision wound heater coils. Heat is
transferred through the thin wall to the gas
flow inside. Heat is carried by he gas
stream from the upstream coil to the
downstream coil windings. The resultant
temperature dependent resistance
differential is detected by the electronic
control circuit. An output is generated that
Tylan mass- is a function of the amount of heat carried
flow controller by the gases to indicate mass-molecular
FC 280 SA based flow rates.

Because of the temperature difference


between the junction and the tail end a
voltage difference can be measured the
Thermocouple
two thermoelements at the tail end, due to
the Seebeck effect: so the thermocouple is
a temperature-voltage transducer.

73
O2: Paramagnetic
CO: Non dispersive infra-red
CO2: Non dispersive infra-red
HC: Flame ionization
Species
analyzer NOx: Chemiluminescence

For a more detailed description read the


manual.

CMOS stands for complementary


metal oxide semiconductor.
The sensors are pixilated and convert
light into an electric charge and
process it into electronic signals. There
are millions of photosensitive diodes
on the surface of these chips, each of
them capturing one pixel of the image.
High-speed When the light hits the sensor, these
camera series of pixels collect photons and
record the brightness of the light that
falls on it by accumulating photons.
When the exposure to light rays is
completed, this charge is converted
into a digital value. In a CMOS sensor
this conversion takes place in each
pixel instead of in a common output
structure.

74

Вам также может понравиться