You are on page 1of 14

Energy Conversion and Management 267 (2022) 115892

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Kinetics mechanism of inert and oxidative torrefaction of biomass


Antonio Soria-Verdugo a, b, *, Eduardo Cano-Pleite a, Aidin Panahi b, Ahmed F. Ghoniem b
a
Carlos III University of Madrid (Spain), Thermal and Fluids Engineering Department, Avda. de la Universidad 30, 28911 Leganés, Madrid, Spain
b
Massachusetts Institute of Technology, Department of Mechanical Engineering, 77 Massachusetts Ave, 02139 Cambridge, MA, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Torrefaction of biomass is a promising pre-treatment process capable of substantially improving the properties of
Biomass torrefaction raw biomass for its use as a solid biofuel, among many other potential applications. Oxygen-lean torrefaction can
Extended mechanism effectively reduce the cost and complexity of inert torrefaction. In this work, torrefaction tests of crushed olive
Kinetics
stones were conducted in a thermogravimetric analyzer in various oxygen concentrations, under both non-
Olive stones
Oxygen concentration
isothermal and isothermal conditions. Non-isothermal torrefaction measurements were used to gain funda­
Oxygen-lean torrefaction mental knowledge on the inert and oxidative torrefaction processes by applying a model-fitting kinetics method.
Analysis of these measurements showed that an accurate description of inert torrefaction based on a two-step
reaction mechanism is possible, whereas a three-step reaction mechanism is necessary for oxidative torre­
faction. The new three-step mechanism was found to be accurate for describing the global mass loss during
isothermal torrefaction, obtaining average root mean squared errors between the model predictions and the TGA
measurements below 2.0 % for inert and oxidative torrefaction of olive stones. Furthermore, predictions using
the extended mechanism were in good agreement with the more fundamental torrefaction reactions derived from
the kinetics analysis of the non-isothermal torrefaction measurements.

design and control processes [61].


In contrast, olive stones generated in huge quantities as a residue of
1. Introduction the olive oil industry especially in Mediterranean countries, exhibit
favorable properties for thermochemical conversion. For instance, olive
Biomass has several advantages as a potential substitute to fossil stones have high mass and energy densities and adequate particle size
fuels such as zero net emissions of CO2, worldwide availability, and for biomass reactors [33], low sulfur and nitrogen contents that mini­
possibility of storage and availability on demand. Among available mize sulfur and nitrogen oxides emissions during combustion [23], and
biomass types, lignocellulosic biomass is preferred for technical and low chlorine, sodium, and potassium concentrations that prevent the
social reasons, such as lack of competition with food resources, its higher risk of reactor corrosion and soiling [35]. These promising properties
energy density, lower demand of fertilizers, water and pesticides, and can be further improved by subjecting olive stones to drying or torre­
faster growth [61]. Nevertheless, lignocellulosic biomass has some faction pre-treatment [39,40].
drawbacks, including its high moisture and oxygen contents, low energy Pre-treatment technologies, such as drying, densification and torre­
density, hydrophilic and fibrous nature, highly variable composition, faction, aim to improve biomass properties and circumvent some of the
and heterogenous and non-isotropic properties [8]. These inherent dis­ problems encountered during their conversion. Biomass moisture con­
advantages inhibit its extensive use at large scale [38]. The low energy tent can be drastically reduced by drying, however, the dried biomass
density of biomass makes its long-distance transport burdensome and may re-absorb water [12]. Densification processes like pelletizing and
uneconomical [43], whereas the high moisture content results in a high briquetting can effectively contribute to reducing handling and trans­
energy consumption during subsequent thermochemical conversion portation costs, nonetheless, due to the hydrophilic nature of biomass,
processes [50]. Hydrophilic solid fuels are likely to deteriorate biolog­ biological deterioration may occur during storage [22]. In contrast,
ically, hindering their long-term storage [65]. The fibrous nature of thermal degradation at 200–300 ◦ C, technically known as torrefaction,
lignocellulosic biomass leads to high energy demand for grinding [2], significantly improves its properties [38]. At this temperature range,
while its characteristic heterogenous properties complicates reactor

* Corresponding author at: Carlos III University of Madrid (Spain), Thermal and Fluids Engineering Department, Avda. de la Universidad 30, 28911 Leganés,
Madrid, Spain.
E-mail address: asoria@ing.uc3m.es (A. Soria-Verdugo).

https://doi.org/10.1016/j.enconman.2022.115892
Received 14 April 2022; Received in revised form 12 June 2022; Accepted 13 June 2022
Available online 17 June 2022
0196-8904/© 2022 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

this 2-step first-order model has been extensively used in its original
Nomenclature form [48,24,37], and as the basis of more complex models to describe
the evolution of torrefaction volatiles and solid products [5] as well as
A Pre-exponential factor [s− 1] the thermochemistry of the process [6]. Multi-component models start
E Activation energy [kJ mol− 1] with the basic components: hemicellulose, cellulose, and lignin, and
k Reaction-rate constant [s− 1] follow their reactions during inert torrefaction. Because of the low
N Order of reaction [-] temperature of torrefaction, a two-component model was proposed by
N Number of reactions [-] Peng et al. [41]. For oxidative torrefaction, Senneca et al. [46] proposed
R Universal gas constant [J mol− 1 K− 1] two parallel pathways for the process, direct conversion of the solid and
T Time [s] a second combustion of the char and volatiles produced by pyrolysis.
T Temperature [◦ C] This model was complemented later by proposing a power law kinetic
X Remaining percentage of mass [%] model to describe oxidative pyrolysis [47]. However, these models were
XO2 Oxygen concentration [-] not developed specifically for the low temperature regime of torre­
W Relative contribution of a reaction [-] faction. In particular, for oxidative torrefaction, Uemura et al. [60]
wmin Minimum relative contribution of all reactions [-] proposed a two-component kinetic model, one for the decomposition of
α Conversion degree [-] hemicellulose and the other for the oxidation of the residue. Wang et al.
β Heating rate [K/s] [62], based on the model of Uemura et al. [60], assumed that the
oxidative torrefaction of biomass residues occurs by combination of
Abbreviations inert torrefaction and oxidation processes. Even though the literature
DTG Differential Thermogravimetric regarding kinetics of inert torrefaction is extensive, work focusing on
RMSE Root Mean Square Error kinetics of oxidative torrefaction is still scarce. However, torrefaction
TG Thermogravimetric under oxygen-lean conditions can lead to several advantages compared
TGA Thermogravimetric Analysis to inert torrefaction, thus, improving the knowledge of oxidative tor­
refaction is required.
The aim of this work is to further examine biomass torrefaction ki­
netics under both inert and oxidative conditions. Thermogravimetric
biomass releases most of its moisture and some of its volatiles [66]. Its
measurements of olive stones’ torrefaction in atmospheres containing
resulting energy density [26], reduces transportation costs. During tor­
various oxygen concentrations were carried out under non-isothermal
refaction, hydroxyl groups in the cell walls are reduced, preventing
and isothermal conditions. A model-fitting kinetics method was
water attachment to its surface by hydrogen bonding, significantly
applied to the non-isothermal measurements, conducted under 5 con­
improving its hydrophobicity [64] and its long-term storage potential.
stant heating rates, for a better understanding of the fundamentals of
Weakening the fibrous structure improves homogeneity and grindability
inert and oxidative torrefaction processes. Results of this analysis were
[30]. Torrefaction also reduces the O/C ratio [34] leading to better
used to develop an extended 2-step first-order kinetics model valid to
outcome of thermochemical processes, such as increasing the cold effi­
describe both inert and oxidative isothermal torrefaction. The extended
ciency of gasification [44] and reducing smoke generated during com­
mechanism was validated using isothermal torrefaction measurements
bustion [42].
conducted at various temperatures and oxygen concentrations and in­
Torrefaction typically occurs in a temperature range from 200 to
dependent non-isothermal torrefaction measurements varying the
300 ◦ C in oxygen-starved environment. Under these conditions, the
heating rates and the oxygen availability in the atmosphere.
governing parameters of the process are the reactor temperature, resi­
dence time, particle size, and moisture content [7]. The endothermic
2. Theoretical model
nature of inert torrefaction and the absence of oxygen results in a high
energy consumption. The capital cost of torrefaction reactors and the
2.1. Non-isothermal torrefaction kinetic models
complexity of the process can be substantially reduced by switching to
oxygen-lean torrefaction reactor [56]. In addition to cost reduction,
Non-isothermal measurements in thermogravimetric analyzers
Thengane et al. [57] found, applying a life cycle analysis, that the
(TGA) have been widely applied to biomass conversion to gain funda­
environmental impact of partially oxidative and high-temperature tor­
mental knowledge of the processes occurring during the thermal
refaction is lower than those of inert low-temperature torrefaction. In
degradation or chemical reaction of the feedstock. The kinetics of non-
this case, air partially oxidizes the gases released during pyrolysis pro­
isothermal conversion processes of solids are typically described as
ducing the heat required by the endothermic reactions, and leading to an
s1S1(s) → s2S2(s) + gG(g) [17], for which the rate of pyrolysis of the solid
overall autothermal process [27,28]. In oxygen-lean torrefaction, oxy­
reactant S1 into another solid S2 and gas G can be obtained by an overall
gen availability in the reactor can affect the process more strongly than
conversion reaction in the form:
other operating conditions [58]. Moreover, the method used for subse­
quent densification may also affect the properties of the final product dα
= k(1 − α)n , (1)
[4]. Numerous studies are available in the literature analyzing the effect dt
of the operating conditions, including oxygen concentration, on the
where t is time, n is the order of reaction, k is the reaction-rate
products of oxidative torrefaction using different biomass types as
constant, and α is the conversion degree defined as the ratio of mass
feedstocks [31,60,13,63,67,15].
loss at a specific time to the total mass loss during the complete process:
Inert torrefaction kinetics has been intensively studied. Models
describing inert torrefaction kinetics can be divided into single-step, α=
m0 − m
, (2)
multi-step, and multi-components models [14]. One-step models as­ m0 − mf
sume that biomass decomposes into a solid residue by releasing gas in a with m0 being the mass of the sample at the beginning of the torre­
single step, following either first- or nth-order kinetics [21]. Multi-step faction test, m the variable mass of the sample during the process, and mf
models consist of several steps, in which a solid residue releases gas the remaining mass at the end of torrefaction. An Arrhenius form is
generating a new solid that further decomposes. Among the multi-step assumed for the reaction-rate constant [3], thus, k depends on the
models, the 2-step first-order mechanism proposed by di Blasi and temperature T, the gas universal constant R, and the kinetics parameters,
Lanzetta [20] is preferred because of its simplicity and accuracy. In fact, i.e., pre-exponential factor A, and activation energy E:

2
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

( )
E
k = Aexp − . (3) dXB
RT = k1 XA − (k2 + kV2 )XB , (8)
dt
Kinetics of non-isothermal conversion processes are described by
dXC
either model-free or model-fitting methods, considering typically first- = k2 XB , (9)
order reactions, i.e., n = 1. In this case, the overall conversion rate, dt
Eq. (1), can be integrated to obtain the conversion degree as follows: dXV1
( ∫ ) = kV1 XA , (10)
A T − E/RT dt
α = 1 − exp − e dT . (4)
β 0 dXV2
= kV2 XB . (11)
The integral in the right-hand-side of Eq. (4) is the so-called tem­ dt
perature integral, which has no analytical solution. The system of differential equations, Eqs. (7)-(11), requires an initial
condition. At the beginning of the test, only fresh biomass is available.
2.1.1. Model-fitting kinetics Thus, the initial condition to solve the system of differential equations is:
Model-fitting kinetic methods derive the kinetics parameters of the {
conversion process by the fitting of the model prediction of α with t = 0→
XA = 100
(12)
temperature to the experimental measurements obtained from non- XB = XC = XV1 = XV2 = 0
isothermal TGA measurements. The evolution of α is obtained from Integrating the ODEs with the initial condition of Eq. (12), the time
the numerical integration of Eq. (4) [10]. In this model-fitting approach, evolution can be obtained, and the remaining percentage of solid mass is
the solid fuel is assumed to be composed of a finite number of pseudo- estimated as follows:
components, N, with each characterized by a weight, wj, over the
entire conversion process and a pair of kinetics parameters Aj and Ej. X = 100 − XV1 − XV2 (13)
Thus, the overall conversion degree can be determined as the sum of the
conversion degrees of each pseudo-component, yielding: 2.2.2. Extended 2-step first-order kinetics model for inert and oxidative
{ [ ( ( ))]} torrefaction
∑N Aj Ej Ej
α= wj 1 − exp − Te− Ej /RT − Ei , (5) In presence of oxygen, the 2-step model for inert torrefaction is now
β R RT extended as oxygen may also react with the solids A, B and C. These
j=1

where the summation of the weights of all pseudo-component should gas–solid reactions are included to account for the effect of oxygen on
be
∑N the degradation process. For the extended 2-step mechanism, the orig­
j=1 wj = 1, and Ei is the exponential integral defined as follows:
inal terminology proposed by Di Blasi and Lanzetta [20] was used and
∫∞ −z
e the gases resulting from inert torrefaction reactions are designated as V1
Ei(z) = dz. (6)
z z and V2, whereas the gases released as a result of oxidative reactions are
called V1o, V2o and V3o, as shown in Fig. 2. Therefore, reaction-rate
2.2. Isothermal torrefaction kinetic models constants for the oxidation of each solid to release gases, kV1o, kV2o,
and kV3o, were included in the extended mechanism, which accounts for
2.2.1. 2-step first-order kinetic model both inert and oxidative torrefaction. Even though the original termi­
The 2-step first-order kinetic model was originally developed by Di nology of Di Blasi and Lanzetta [20] was used for the extended 2-step
Blasi and Lanzetta [20] and has been widely employed to describe inert mechanism, the presence of oxygen in the atmosphere during torre­
torrefaction of biomass during kinetically controlled conditions [5,45], faction may induce the existence of gas–gas chemical reactions affecting
obtaining robust and reliable predictions of the performance of different the composition of volatiles V1 and V2 released during inert torrefaction.
biomass samples during torrefaction and pyrolysis in the absence of Fig. 2 shows the mechanism proposed for the extended kinetics
oxygen. This model assumes that fresh biomass, A, decomposes into an model, representing the pyrolytic reactions in black and the oxidative
intermediate solid, B, by releasing moisture and volatiles, V1. The in­ contributions in blue. All the oxidative reactions will become zero with
termediate solid, B, is further decomposed into a solid residue, C, with the removal of oxygen from the atmosphere, and the 2-step mechanism,
more volatile matter, V2, released. The mechanism is shown schemati­ represented in Fig. 1 for inert torrefaction is recovered.
cally in Fig. 1. Similar to the mathematical procedure followed for the inert case,
The rate of variation of the mass percentage of each component, dXi/ the differential equations system for the extended 2-step kinetics
dt, can be calculated as a function of the rate constant of each reaction ki mechanism is derived.
and the percentage of mass of each component Xi:
dXA
= − (k1 + kV1 )XA , (7)
dt

Fig. 1. Schematic diagram of the 2-step first-order kinetics mechanism. Fig. 2. Schematic of the extended 2-step first-order kinetics mechanism.

3
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

dXA high volatile, carbon and hydrogen contents, resulting in a suitable


= − (kV1 + kAB + kV1o XO2 )XA , (14)
dt measured high heating value of 20.1 MJ/kg, comparable to that of coal.
Additionally, the contents in ash, nitrogen and sulfur are low, leading to
dXB
= kAB XA − (kV2 + kBC + kV2o XO2 )XB , (15) low potential pollutant emissions during conversion. Considering the
dt results, olive stones are a promising feedstock for thermochemical
dXC conversion applications.
= kBC ⋅XB − kV3o XO2 XC , (16)
dt
3.2. Torrefaction kinetics measurements in the TGA
dXV1
= kV1 XA , (17)
dt Torrefaction kinetics measurements, both non-isothermal and
isothermal, were conducted in the TGA Q500 from TA Instruments. A
dXV2
= kV2 XB , (18) mass of 5.0 ± 0.2 mg of olive stones with a particle size below 100 μm
dt was used in each test to avoid heat and mass transfer limitations inside
the samples [9]. All kinetics tests were preceded by a drying process, in
dXV1o
= kV1o XO2 XA , (19) which the sampled were heated from room temperature to 105 ◦ C at a
dt
rate of 20 ◦ C/min, followed by an isothermal process of 30 min to
dXV2o remove all moisture from the feedstock prior to the torrefaction
= kV2o XO2 XB , (20)
dt experiments.

dXV3o
= kV3o XO2 XC . (21) 3.2.1. Non-isothermal torrefaction measurements
dt The drying process was followed by a linear temperature rise starting
where XO2 is the oxygen concentration in the environment where the at 105 ◦ C. This was maintained until 400 ◦ C even though the analysis of
torrefaction process occurs. Initially, the percentage of mass of each non-isothermal torrefaction was performed only from 150 to 350 ◦ C.
component except fresh biomass A is zero: Raising the final temperature to 400 ◦ C guarantees that the TGA can
{ keep accurately a constant heating rate through 350 ◦ C, enhancing the
t = 0→
XA = 100
(22) reliability and accuracy of the results derived from the non-isothermal
XB = XC = XV1 = XV2 = XV1o = XV2o = XV3o = 0 temperature measurements. Five different constant heating rates of β
The time evolution of the percentage of mass of each component can = 1.00, 1.78, 3.16, 5.62, and 10.00 K/min were used to derive accurate
be derived by solving the system of differential equations, Eqs. (15)- values of the kinetics parameters of torrefaction according to Soria-
(21), subjected to the initial condition of Eq. (22). Then, the remaining Verdugo et al. [51]. The values of the heating rates were selected
percentage of solid mass can be estimated as follows: below 10 K/min, to keep the conditions in the kinetics regime, and
logarithmically spaced to improve visualization of the DTG curves. The
X = 100 − XV1 − XV2 − XV1o − XV2o − XV3o (23) concentration of oxygen was also varied by supplying the TGA furnace
Notice again that, for XO2 = 0, Eqs. (15)-(21) are the same as Eqs. (7)- with a flow rate of 60 ml/min of nitrogen, two different mixtures of
(11) for inert torrefaction. oxygen–nitrogen with concentrations of 5–95 and 10–90 vol%, and
synthetic air. Therefore, concentrations of oxygen, XO2, of 0, 0.05, 0.1,
3. Materials and methods and 0.21 were tested for each heating rate, similar to those used by
Senneca et al. [47] to study oxidative pyrolysis. The oxygen concen­
3.1. Biomass characterization tration in the atmosphere was measured in volumetric percentage,
however, it will be reported as dimensionless according to convention.
Crushed olive stones, obtained as a residue from the olive oil in­ The gases were specifically supplied by Abelló Linde for the torrefaction
dustry, were purchase from the company Olihueso (Córdoba, Spain) in experiments. Three replicates of each test were performed to check the
January 2019. The crushed olive stones were milled to a particle size replicability of the experimental process, obtaining deviations below
under 100 μm prior to the basic characterization and the torrefaction 2 % in all cases.
kinetics tests. The basic characterization of this feedstock, consisting in
an ultimate analysis, a thermogravimetric analysis, and a heating value 3.2.2. Isothermal torrefaction measurements
test, was performed and published in Soria-Verdugo et al. [55]. The For the isothermal torrefaction measurements, after the drying pro­
ultimate analysis was conducted in a Leco TruSpec CHN and S elemental cess, the temperature was increased rapidly from 105 ◦ C to the torre­
analyzer, whereas the heating value test was performed in an iso­ faction temperature of the test, using a heating rate of 200 K/min, and
peribolic calorimeter Parr 6300. The thermogravimetric analysis was the temperature was then kept constant for 180 min, while the sample
carried out in a Thermogravimetric Analyzer (TGA) Q500 from TA In­ was torrefied, mainly under isothermal conditions. To avoid any effect
struments, which was also used for the torrefaction kinetics tests. of possible variations of the sample moisture, the starting point of the
Further details of the equipment used for the basic characterization can torrefaction test was considered as the final point of the drying process.
be found in Soria-Verdugo et al. [52] and Soria-Verdugo et al. [53]. Thus, all percentages of mass shown in the results of the isothermal
Table 1 reports the results. The olive stones tested are characterized by torrefaction correspond to a dry basis.
During the tests, the temperature of the torrefaction process was
varied from 200 to 300 ◦ C in intervals of 25 ◦ C, to cover the typical range
Table 1 of temperature for torrefaction. The concentration of oxygen in the TGA
Basic characterization of the biomass samples studied. (VM: Volatile Matter, A: furnace during the isothermal torrefaction tests were the same as for the
Ash, C: Carbon, H: Hydrogen, N: Nitrogen, S: Sulfur, O: Oxygen, HHV: High
non-isothermal measurements, i.e., XO2 = 0, 0.5, 0.1, and 0.21.
Heating Value, db: dry basis, daf: dried ash free basis, *calculated by difference).
Values from three tests, attaining deviations below +/- 1.5 %.
4. Results and discussion
VM A C H N S O* HHV
[%db] [% [% [% [% [% [% [MJ/kg
db] daf] daf] daf] daf] daf] db] 4.1. Non-isothermal torrefaction process

77.0 0.7 52.4 6.1 0.9 0.1 40.5 20.1


The conversion degree α was obtained from the non-isothermal

4
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

torrefaction measurements run in the TGA for a temperature range of of considering a variable number of reactions occurring during
150 – 350 ◦ C. The DTG curves for each heating rate and oxygen con­ torrefaction.
centration were plotted in Fig. 3 to analyze the effect of the heating rate Predictions of the evolution of the conversion degree α with tem­
β and oxygen concentration XO2 on the evolution of the rate of conver­ perature T obtained from Eq. (5) were fitted to the experimental con­
sion dα/dt, which is represented as a function of temperature T. As ex­ version curves obtained for the five heating rates tested under each
pected, an increase of the heating rate leads to higher rates of conversion oxygen concentration. The fitting procedure was based on a non-linear
for all values of the oxygen concentration tested. The DTG curves for least squared technique implemented in Matlab®, considering a vari­
inert torrefaction, XO2 = 0, are characterized by two clear peaks, whose able number of reactions or pseudo-components N for each case,
rate curves overlap, indicating the existence of two primary reactions in obtaining as fitting parameters the relative contribution of each reaction
this process, as stated by the typical 2-step mechanism proposed by Di wj to the global conversion process and the corresponding kinetics pa­
Blasi and Lanzetta [20]. The peak temperatures shift slowly to higher rameters Aj and Ej of each reaction. The fitting was subjected to the
values as β increases [19,54]. For inert torrefaction, the first peak is condition that the sum of all relative contributions should be unity,

obtained at temperatures between 230 and 260 ◦ C depending on the j wj = 1 and performed independently for the data of each oxygen
heating rate, whereas the second peak appears in a temperature range concentration.
from 300 and 340 ◦ C. According to the literature [18,25,68], a compensation effect exists
The rate of conversion at low temperatures, i.e., the first peak, is not when the fitting is performed for each data set, i.e., each oxygen con­
much affected by the presence of oxygen. In contrast, when XO2 greater centration, independently. The compensation effect states that the
than 0, the second peak is reduced significantly to levels closer to the fitting of an experimental data set can be performed with the same
first peak. Oxidative torrefaction enhances the reduction of the second goodness of fitting for different pairs of Aj and Ej, since there is a linear
peak height, which appears at lower temperatures between 275 and relation between the logarithm of the pre-exponential factor and the
325 ◦ C, and a shoulder, i.e., an overlapping smaller peak at slightly activation energy. Therefore, the same value of the pre-exponential
higher temperature around 330–340 ◦ C, can be observed, which may be factor of each reaction was set as an initial condition for all the oxy­
attributed to an extra reaction occurring when oxygen is available gen concentrations, and the fitting procedure varied the activation en­
during torrefaction. This extra reaction is more visible for high heating ergy and the weight of each reaction to reach a proper fitting, allowing a
rates in an air atmosphere. direct comparison of the results in terms of activation energies and
Analysis of the non-isothermal measurements was used to gain weights of each reaction.
fundamental understanding of the torrefaction process, including the The non-linear square fitting method requires an initial condition for
effect of the oxygen concentration. Non-isothermal torrefaction mea­ the values of wj, Aj and Ej, which serves as a starting point for the iter­
surements were analyzed applying a model-fitting kinetics method since ation. Based on the values reported in the literature, the initial condi­
these methods allow the selection of the number of pseudo-components tions for each parameter were selected in the ranges wj,0 = 1/N, Aj,0 =
considered for the calculation and permit the quantification of the effect 106 – 1011 s− 1 and Ej,0 = 100 – 300 kJ/mol.

Fig. 3. DTG curves of non-isothermal torrefaction for various heating rates and oxygen concentrations.

5
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

The effect of the number of pseudo-components considered for the Table 2


fitting was analyzed based on the minimum relative contribution of all Kinetics parameters and relative weight of each reaction obtained from the
reactions for each case. Fig. 4 shows the minimum relative contribution fitting-model considering a total of 3 reactions.
of all reactions considered, wmin, as a function of the total number of XO2 0 0.05 0.1 0.21
reactions or pseudo-components taken into account, N, for each oxygen w1 [-] 0.362 0.377 0.366 0.355
concentration XO2. Clearly, when only one reaction is considered, the A1 [s− 1] 1010 1011 1011 1011
only relative contribution is w = 1, and thus, wmin = 1 for N = 1 for all E1 [kJ/mol] 126.6 126.0 125.7 125.3
values of XO2. As the number of reactions is increased, the minimum w2 [-] 0.620 0.520 0.516 0.506
A2 [s− 1] 1010 1011 1011 1011
relative contribution of the reactions decreases, obtaining a negligible
E2 [kJ/mol] 152.1 148.1 147.1 146.2
contribution for N ≥ 4 for all oxygen concentrations. Differences be­ w3 [-] 0.018 0.103 0.118 0.139
tween inert and oxidative torrefaction were obtained when considering A3 [s− 1] 1010 1011 1011 1011
3 reactions. In this case, the minimum relative contribution of the re­ E3 [kJ/mol] 167.9 161.5 161.1 160.3
actions for inert torrefaction is close to zero, which indicates that 2 re­
actions could properly describe inert torrefaction. In contrast, the
a slight shift of the peak to lower temperature for higher oxygen
minimum relative contribution of the third reaction is higher for
concentrations.
oxidative torrefaction, thus, the third reaction should not be neglected
The contribution of the second reaction differs for inert and oxidative
for oxidative torrefaction, for which considering only 2 reactions may
torrefaction. A higher peak at higher temperature was obtained for inert
lead to inaccurate predictions of the conversion degree.
torrefaction, whereas the peaks of oxidative torrefaction were shorter
The relative contribution and kinetics parameters of each reaction
and occurred at lower temperature, as an effect of the reaction of the
for inert and oxidative torrefaction of olive stones considering 3 re­
solid biomass with the oxygen available in the reactor. Similar to the
actions are reported in Table 2. The non-linear fitting process followed
first reaction, a slight displacement of the peak corresponding to the
for all the heating rates of each oxygen concentration was performed
second reaction of oxidative torrefaction to lower temperatures was
using the same values for the pre-exponential factors of each reaction Aj,
obtained for increasing values of the oxygen concentration. Regarding
so that, the effect of the oxygen concentration can be directly observed
the contribution of the third reaction, it occurs at higher temperatures
in the values of the activation energy Ej and the relative contribution of
with a higher contribution when the oxygen concentration in the
each reaction wj. Both the relative contribution and the activation en­
furnace increases. The negligible influence of the third reaction to inert
ergy of the first reaction for inert torrefaction and oxidative torrefaction
torrefaction can be observed in Fig. 5 b).
with the 3 different oxygen concentrations tested are similar. In contrast,
The model-fitting kinetics analysis of the non-isothermal torrefaction
both the relative contribution and the activation energy of the second
shows that inert torrefaction of olive stones can be properly described by
reaction decreased when the concentration of oxygen available in the
2 first-order reactions, confirming the validity of the 2-step first-order
atmosphere increases. In the case of the third reactions, the differences
mechanism proposed by di Blasi and Lanzetta [20].
between inert and oxidative torrefaction are bigger. The relative weight
of the third reaction is below 2 % for the case of inert torrefaction,
whereas for oxidative torrefaction the third reaction accounts for more 4.2. Description of the isothermal torrefaction process by the extended 2-
than 10 % of the global conversion. The activation energy of the third step model
reaction of oxidative torrefaction slightly decreases with the oxygen
concentration. This result is in good agreement with the previous find­ The evolution of the remaining percentage of mass, X, during
ings of Anca-Couce et al. [1]. isothermal torrefaction in the TGA is shown in Fig. 6 for the various
For better understanding of the contribution of each reaction, Fig. 5 torrefaction temperatures and oxygen concentrations tested. In all cases,
shows the experimental DTG curves for β = 3.16 K/min and all oxygen the beginning of the torrefaction was defined at the end of the pre­
concentrations, and the relative contribution of each reaction according liminary drying process to allow direct comparison of the results ob­
to the model-fitting estimations, considering the fitting parameters tained for different operating conditions. Fig. 6 shows that for all oxygen
included in Table 2. The relative contribution of the first reaction (solid concentrations, increasing the torrefaction temperature results in a
lines in Fig. 5 b)) is similar for both inert and oxidative torrefaction, with faster and more intense reduction of the remaining percentage of mass, i.
e., a more vigorous release of gases in the process (notice the different y-
axis scale for the torrefaction temperatures of 200 and 225 ◦ C in Fig. 6),
in agreement with the results of Lu et al. [31], Uemura et al. [60], and
Zhang et al. [67]. Regarding the effect of the oxygen concentration,
increasing the amount of oxygen leads to a higher and faster reduction of
the remaining mass of olive stones for all cases. However, the effect of
the oxygen concentration is stronger at moderate torrefaction temper­
atures of 225 and 250 ◦ C. Comparing the results of the isothermal tor­
refaction process in an inert atmosphere, XO2 = 0, to those conducted
with XO2 = 0.05, 0.1 and 0.21, reveals that the presence of oxygen ac­
celerates the torrefaction of biomass. For the case of high torrefaction
temperatures, T = 275 and 300 ◦ C, the presence of oxygen contributes to
oxidize the produced solid residue, resulting in a slow reduction of the
remaining percentage of mass, X, at the end of the torrefaction test. In
contrast, in the case of torrefaction in the absence of oxygen, XO2 = 0, an
almost uniform value for X is attained at the end of the isothermal tor­
refaction experiment. For all torrefaction temperatures, the remaining
percentage of mass is reduced when oxygen is available, attaining values
of around 20 % for torrefaction at 300 ◦ C, a result previously reported by
Kung et al. [29].
Fig. 4. Minimum contribution as a function of the number of reactions The kinetics parameters, Ai and Ei, of each reaction of the extended 2-
considered for the fitting-model. step mechanism, shown schematically in Fig. 2, were determined by

6
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

Fig. 5. a) DTG curves for β = 3.16 K/min obtained from the TGA analysis, b) contribution of each reaction considered to the DTG curves for β = 3.16 K/min
according to the model-fitting method.

Fig. 6. TG curves for various temperatures and oxygen concentrations (Notice the different y-axis scale for temperatures of 200 and 225 ◦ C).

fitting the remaining percentage of mass, Eq. (23), to the experimental whereas the activation energies range from 125 to 330 kJ/mol. Among
measurements shown in Fig. 6. A non-linear least squared technique, the oxidative reactions, the rate constant of V2o is several orders of
implemented in Matlab®, was used, considering the experimental magnitude lower than those of the rest of oxidative and pyrolytic re­
measurements conducted for all torrefaction temperatures and oxygen actions, and thus, it can be neglected. The values obtained for the ki­
concentrations for the fitting at the same time. netics parameters of the isothermal torrefaction in Table 3 are similar to
The optimal values of the kinetics parameters are reported in Table 3. those derived from the non-isothermal measurements in Table 2.
The pre-exponential factors obtained for the pyrolytic reactions range The kinetic parameters reported in Table 3 were used to construct an
from 106 to 107 s− 1, and the corresponding activation energies vary from Arrhenius plot, where ln(ki) is represented as a function of 1/T for a
90 to 125 kJ/mol. These values are similar to those obtained by Martín- temperature range from 200 to 300 ◦ C. The Arrhenius plot is shown in
Lara et al. [32] and Cano-Pleite et al. [10]. In contrast, the range of Fig. 7, depicting the curves of the rate constant corresponding to pyro­
variation of A and E is wider for the oxidative reactions, obtaining values lytic reactions in black and those related to oxidative reactions in blue.
for the pre-exponential factors varying between 106 and 107 s− 1, In the case of the oxidative reactions, the rate constant ki would be

7
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

Table 3 temperature and oxygen concentration. The values of RMSE, are 1.7,
Pre-exponential factors and activation energies of each reaction obtained from 1.5, 1.9, and 2.0 % for the cases of oxygen concentration of XO2 = 0,
the fitting of the extended 2-step first-order model estimations to the TGA 0.05, 0.1 and 0.21, respectively. Similar values of the average RMSE
measurements of olive stones for all torrefaction temperatures and oxygen were found for all the different temperatures tested above 200 ◦ C,
concentrations. confirming the robustness of the fitting method for the whole range of
Pyrolytic reactions Oxidative reactions torrefaction temperatures.
Rate A E [kJ/ Rate A E [kJ/
constant [s¡1] mol] constant [s¡1] mol] 4.2.1. Contribution of each reaction to the global mass loss
kAB 6.4⋅106 98.8 kV1o 3.6⋅106 124.9 The time evolution of the solid composition during torrefaction of
kV1 3.6⋅106 120.8 kV2o 3.0⋅106 328.8 olive stones at different torrefaction temperatures and oxygen concen­
kBC 9.1⋅106 92.2 kV3o 1.1⋅106 199.7 trations is depicted in Fig. 9. The composition of the solid, i.e., the dis­
kV2 2.0⋅106 106.9 tribution of the solid components A, B, and C, is not strongly affected by
oxygen availability. In contrast, the torrefaction temperature signifi­
cantly influences the composition of the solid, in good agreement with
the results present in the literature for inert torrefaction of lignocellu­
losic biomass [49,11,16]. For torrefaction at 200 ◦ C, the amount of
unreacted solid A decreases slowly, until a percentage above 30 % after
180 min for all the oxygen concentrations tested, whereas the rest of the
solid corresponds to the intermediate solid B, with a negligible per­
centage of the final residue C, in agreement with Chen et al. [16]. When
increasing the torrefaction temperature to 250 ◦ C the fresh solid A is
completely consumed after around 90 min and the final solid C is
generated in higher quantities; the final residue after 180 min is a
mixture of solids B and C, in line with the results previously reported by
Chen et al. [16]. At torrefaction temperature of 300 ◦ C, the fresh solid A
is consumed fast, after around 10 min, and the intermediate solid B is
also consumed at this high torrefaction temperature in approximately
120 min; the final solid after 180 min is mainly pure C. Interestingly, for
intermediate torrefaction times, e.g., 60 min, the amount of B increases
for mild torrefaction temperatures (250 ◦ C) and decreases again at se­
vere torrefaction conditions (300 ◦ C), also well in line with data in the
literature [49]. All in all, for all torrefaction temperatures, the effect of
Fig. 7. Arrhenius plot for the kinetic parameters presented in Table 3. Notice
the oxygen concentration on the solid composition is slight, as shown in
that the oxidative reactions would have the concentration of oxygen XO2 as
Fig. 9.
multiplying factor of the rate constant ki.
The contribution of each gas released considered in the extended 2-
step mechanism, namely, V1, V2, V1o, V2o, and V3o, to the global mass
multiplied by the oxygen concentration available in the atmosphere,
loss during isothermal torrefaction depends on the temperature and
XO2, as shown in Eqs. (14)-(21). The rate of the pyrolytic reactions is
oxygen concentration. The contribution of the reaction of each pseudo-
higher than that of the oxidative reactions, however, the rate of
component to the global mass loss is depicted in Fig. 10 for all oxygen
extraction of gases V1o is only slightly lower than that of volatiles V1.
concentrations and torrefaction temperatures of 200, 250 and 300 ◦ C.
Thus, the release of gases from solid A by pyrolysis, V1, and by oxidation,
The mass loss during inert torrefaction is caused exclusively by the
V1o, will compete when torrefaction occurs in presence of oxygen. The
release of volatiles V1 and V2. The release of volatiles in the first step is
Arrhenius plot also shows that the release of gases V2o has a rate con­
faster than in the second. It is well known that the mass loss during inert
stant several orders of magnitude lower than that of the rest of reactions.
torrefaction at 200 ◦ C is caused mainly by pyrolysis of hemicellulose,
Therefore, this reaction will be negligible for all oxygen concentrations,
with a minor contribution of cellulose pyrolysis [49,11,16]. For low
since it will be always much slower than the release of volatiles V2. In the
temperatures, mainly the degradation of hemicellulose occurs [16],
case of gases V3o, the rate constant is also very low, nonetheless, this
which explains the low presence of V2 in Fig. 10 [16], derived princi­
reaction might be relevant once solids A and B are consumed and the
pally from the decomposition of cellulose. For higher torrefaction tem­
composition of the solid remaining is pure C. In this case, even though
peratures, A is consumed faster, and the amount of released volatiles V1
kV3o is very low, this would be the only possible reaction occurring for
stabilizes in a short period of time, after around 10 min for temperature
the solid C.
of 300 ◦ C. At severe torrefaction conditions, volatiles V2 are generated as
Using the kinetics parameters in Table 3, an estimation of the
a product of the thermal degradation of B, in good agreement with Chen
remaining percentage of mass, X, can be obtained from the extended 2-
et al. [16]. These results are consistent with results present in the
step model, by solving the differential equation system Eqs. (14)-(21)
literature for inert torrefaction [49,11], in which the devolatilization of
and Eq. (23). The results of the predictions are compared to the exper­
B yielding V2 is barely present for light torrefaction conditions, and
imental measurements in Fig. 8 for the five torrefaction temperatures
largely increases with the torrefaction severity.
between 200 and 300 ◦ C.
The effect of oxygen on the evolution of the volatiles V1 and V2 is
The accuracy of the extended 2-step model was quantified by
weak. However, the presence of oxygen accelerates the consumption of
calculating the Root Mean Square Error (RMSE) between the model
A, enhancing the global mass loss. Up to 200 ◦ C, V1o increases with
predictions and the TGA measurements for each torrefaction
oxygen concentration. Higher temperature contributes to releasing more

8
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

Fig. 8. Comparison of the estimations of the extended 2-step mechanism for the remaining percentage of mass and the experimental TGA measurements for each
torrefaction temperature and oxygen concentration.

volatiles V2, similar to that in the absence of oxygen. Similarly, the 4.3. Validation of the extended 2-step model for non-isothermal
amount of V1o also increases with temperature. In fact, the relative torrefaction
contribution of the pyrolysis reactions, V1 and V2, to the global mass loss
diminished with oxygen concentration due to the competition with the Once the kinetic parameters were determined and the accuracy of
oxidation reactions releasing gases V1o. The weak contribution of the extended 2-step mechanism was corroborated, the validity of the
oxidation of C, V3o, to the global mass loss can be seen at high tem­ model was further evaluated for experimental conditions different to
peratures and after long resident time when the solid composition is those used to derive the kinetic parameters, such as non-isothermal
mainly C, as shown in Fig. 9. Nonetheless, the relative contribution of C torrefaction processes. For that purpose, the non-isothermal experi­
oxidation to the global mass loss is slight. In contrast to the rest of mental results obtained from the torrefaction of olive stones with oxygen
gaseous species, a negligible influence of the oxidation of B, V2o, to the concentrations, XO2 = 0, 0.05, 0.1, and 0.21, and heating rates, β = 1.00,
global mass loss can be inferred in Fig. 10 from the null contribution of 1.78, 3.16, 5.62, and 10.0 K/min, were used. The predictions were ob­
V2o for all oxygen concentrations and temperatures, as concluded by the tained by solving the differential equation system Eqs. (14)-(21) with the
analysis of the kinetics parameters presented in Section 4.1. initial condition of Eq. (22). The kinetic parameters of each reaction
The analysis of the contribution of each gaseous species considered reported in Table 3 were employed, considering the variable tempera­
in the extended 2-step mechanism to the global mass loss concluded that ture during the non-isothermal process, calculated as a function of the
2 reactions accurately describe the inert torrefaction of olive stones, heating rate. Results are depicted in Fig. 11 together with the experi­
whereas a third reaction should be considered in the case of oxidative mental measurements obtained from the TGA. Both the experimental
torrefaction. However, for high torrefaction temperatures, an extra re­ values and the estimations of the model were obtained for a temperature
action should be considered in cases with long residence times of the range from 150 ◦ C, temperature for which the samples have already
biomass in the torrefaction reactor. These results obtained from the been subjected to a drying process, to 300 ◦ C, maximum temperature
study of the isothermal torrefaction process are consistent with the used for the isothermal torrefaction tests. Therefore, the time required
fundamental knowledge of the torrefaction process gained from the for this temperature increase of 150 ◦ C differs for each heating rate used,
evaluation of the non-isothermal process. from 150 min for a heating rate of β = 1.00 K/min to 15 min for β = 10.0
K/min. A fairly good agreement between the estimations of the extended

9
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

Fig. 9. Evolution of the solid composition as a function of the torrefaction operating conditions.

10
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

Fig. 10. Contribution of the reaction of each pseudo-component to the global mass loss.

11
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

Fig. 11. Comparison of the experimental TGA results for the percentage of mass remaining during non-isothermal torrefaction of olive stones and the estimations of
the extended 2-step model.

2-step mechanism and the experimental measurements can be observed solid residue.
in Fig. 11, with average RMSE values of 1.9, 2.2, 2.4 and 1.7 %, for XO2 The extended 2-step model was validated for non-isothermal torre­
= 0, 0.05, 0.1 and 0.21, respectively. Similar values of the average RMSE faction conditions by comparison with an independent experimental
were found for all the different heating rates tested, confirming the data set conducted in the TGA for various heating rates and oxygen
robustness of the extended 2-step mechanism for a wide range of tor­ concentrations, obtaining a good match with the experimental results.
refaction heating rates. Similar values of the average RMSE were found The simple, yet robust, extended 2-step mechanism proposed is shown to
for all the different heating rates tested, confirming the robustness of the be consistent with the fundamentals of the inert and oxidative torre­
extended 2-step mechanism for a wide range of torrefaction heating faction processes, and is recommended when torrefaction of biomass
rates. occurs under oxygen-lean conditions. Future work should apply the
same model to other types of biomass.
5. Conclusions
CRediT authorship contribution statement
A novel extended 2-step first-order kinetics mechanism was proposed
to predict the evolution of biomass for both inert and oxidative Antonio Soria-Verdugo: Conceptualization, Methodology, Soft­
isothermal torrefaction. Non-isothermal torrefaction measurements ware, Validation, Formal analysis, Investigation, Resources, Data cura­
were used to apply a model-fitting kinetics method, concluding that 2 tion, Writing – original draft, Visualization, Funding acquisition.
reactions are required to model inert torrefaction, whereas for oxidative Eduardo Cano-Pleite: Software, Validation, Formal analysis, Investi­
torrefaction a third reaction should be considered. The extended 2-step gation, Data curation, Writing – original draft, Writing – review &
mechanism was validated by comparing the time evolution of the editing, Visualization. Aidin Panahi: Software, Validation, Formal
remaining mass predicted by the model and measurements in a TGA for analysis, Investigation, Data curation, Writing – original draft, Writing –
isothermal torrefaction of olive stones under inert and oxidative atmo­ review & editing, Visualization. Ahmed F. Ghoniem: Conceptualiza­
spheres. The results of extended 2-step mechanism were in good tion, Methodology, Formal analysis, Investigation, Resources, Writing –
agreement with the measurements, with deviations below 2.0 % for all review & editing, Visualization, Supervision, Project administration,
temperatures and oxygen concentration tested. The extended 2-step Funding acquisition.
mechanism considers 2 reactions for pyrolytic torrefaction, whereas a
third reaction is also relevant for in the case of oxidative torrefaction. In
addition, an extra reaction should be considered for oxidative torre­ Declaration of Competing Interest
faction at high temperature when the residence time of biomass in the
reactor is long enough to take into account the slow oxidation of the The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence

12
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

the work reported in this paper. [22] Gilbert P, Ryu C, Sharifi V, Swithenbank J. Effect of process parameters on
pelletisation of herbaceous crops. Fuel 2009;88(8):1491–7.
[23] González JF, González-García CM, Ramiro A, González J, Sabio E, Gañán J, et al.
Data availability Combustion optimisation of biomass residue pellets for domestic heating with a
mural boiler. Biomass Bioenerg 2004;27:145–54.
Data will be made available on request. [24] Gul S, Ramzan n., Hanif M.A., Bano S.. Kinetic, volatile release modeling and
optimization of torrefaction. J Anal Appl Pyrol 2017;128:44–53.
[25] Holstein A, Bassilakis R, Wojtowicz MA, Serio MA. Kinetics of methane and tar
Acknowledgments evolution during coal pyrolysis. Proc Combust Inst 2005;30:2177–85.
[26] Kanwal S, Chaudhry N, Munir S, Sana H. Effect of torrefaction conditions on the
physicochemical characterization of agricultural waste (sugarcane bagasse). Waste
The research that led to this publication was conducted with the Manag 2019;88:280–90.
support of a US-Spain Fulbright grant co-sponsored by the Spanish [27] Kung KS, Ghoniem AF. A decentralized biomass torrefaction reactor concept. Part
Ministry of Universities (“Ministerio de Educación, Cultura y Deporte en II: Mathematical model and scaling law. Biomass Bioener 2019;125:204–11.
[28] Kung KS, Shanbhogue S, Slocum AH, Ghoniem AF. A decentralized biomass
el marco del Programa Estatal de Promoción del Talento y su Emplea­ torrefaction reactor concept. Part I: Multi-scale analysis and initial experimental
bilidad en I + D + i, Subprograma Estatal de Movilidad, del Plan Estatal validation. Biomass Bioener 2019;125:196–203.
de I + D + I”). The authors acknowledge the financial support by the [29] Kung KS, Thengane SK, Shanbhogue S, Ghoniem AF. Parametric analysis of
torrefaction reactor operating under oxygen-lean conditions. Energy 2019;181:
Foundation Seed Fund MIT - Spain “la Caixa”. ECP also acknowledges 603–14.
support from the CONEX-Plus program funded by Universidad Carlos III [30] Kung KS, Thengane SK, Ghoniem AF. Functional mapping of torrefied product
de Madrid and the European Union’s Horizon 2020 program under the characteristics with index of torrefaction. Fuel Process Technol 2020;202:106362.
[31] Lu KM, Lee WJ, Chen WH, Liu SH, Lin TC. Torrefaction and low temperature
Marie Sklodowska-Curie grant agreement No. 801538.
carbonization of oil palm fiber and eucalyptus in nitrogen and air atmospheres.
Bioresour Technol 2012;123:98–105.
References [32] Martín-Lara MA, Ronda A, Blázquez G, Pérez A, Calero M. Pyrolysis kinetics of the
lead-impregnated olive stone by non-isothermal thermogravimetry. Process Saf
[1] Anca-Couce A, Zobel N, Berger A, Behrendt F. Smouldering of pine wood: Kinetics Environ 2018;113:448–58.
and reaction heats. Combust Flame 2012;159(4):1708–19. [33] Mediavilla I, Barro R, Borjabad E, Peña D, Fernández MJ. Quality of olive stone as a
[2] Arias B, Pevida C, Fermoso J, Plaza MG, Rubiera F, Pis JJ. Influence of torrefaction fuel: Influence of oil content on combustion process. Renew Energ 2020;160:
on the grindability and reactivity of woody biomass. Fuel Process Technol 2008;89: 374–84.
169–75. [34] Mi BB, Liu ZJ, Hu WH, Wei PL, Jiang ZH, Fei BH. Investigating pyrolysis and
[3] Arrhenius S. Über die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker combustion characteristics of torrefied bamboo, torrefied wood and their blends.
durch Säuren (On the reaction velocity of the inversion cane sugar by acids). Z Phys Bioresour Technol 2016;209:50–5.
Chem 1889;4:226–48. [35] Miranda T, Esteban A, Rojas S, Montero I, Ruiz A. Combustion analysis of different
[4] Barr M, Kung KS, Thengane SK, Mohan V, Sweeney D, Ghoniem AF. olive residues. Int J Mol Sci 2008;9:512–25.
Characterization of aggregate behaviors of torrefied biomass as a function of [37] Nguyen Q, Nguyen DD, Vothi H, He C, Goodarzi M, Bach QV. Isothermal
reaction severity. Fuel 2020;266:117152. torrefaction kinetics for sewage sludge pretreatment. Fuel 2020;277:118103.
[5] Bates RB, Ghoniem AF. Biomass torrefaction: Modeling of volatile and solid [38] Niu Y, Lv Y, Lei Y, Liu S, Liang Y, Wang D, et al. Biomass torrefaction: properties,
product evolution kinetics. Bioresour Technol 2012;124:460–9. applications, challenges, and economy. Renew Sust Energ Rev 2019;115:109395.
[6] Bates RB, Ghoniem AF. Biomass torrefaction: Modeling of reaction [39] Panahi A, Tarakcioglu M, Schiemann M, Delichatsios M, Levendis YA. On the
thermochemistry. Bioresour Technol 2013;134:331–40. particle sizing of torrefied biomass for co-firing with pulverized coal. Combust
[7] Bates RB, Ghoniem AF. Modeling kinetics-transport interactions during biomass Flame 2018;194:72–84.
torrefaction: The effects of temperature, particle size, and moisture content. Fuel [40] Panahi A, Toole N, Xinyu W, Levendis YA. On the minimum oxygen requirements
2014;137:216–29. for oxy-combustion of single particles of torrefied biomass. Combust Flame 2020;
[8] Cahyanti MN, Doddapaneni TRKC, Kikas T. Biomass torrefaction: An overview on 213:426–40.
process parameters, economic and environmental aspects and recent [41] Peng JH, Bi XT, Lim J, Sokhansanj S. Development of torrefaction kinetics for
advancements. Bioreour Technol 2020;301:122737. British Columbia softwoods. Int J Chem React Eng 2012;10(1):1542–80.
[9] Cano-Pleite E, Rubio-Rubio M, Garcia-Hernando N, Soria-Verdugo A. Microalgae [42] Pentananunt R, Rahman ANMM, Bhattacharya SC. Upgrading of biomass by means
pyrolysis under isothermal and non-isothermal conditions. Algal Res 2020;51: of torrefaction. Energy 1990;15(12):1175–9.
102031. [43] Pimchuai A, Dutta A, Basu P. Torrefaction of agriculture residue to enhance
[10] Cano-Pleite E, Rubio-Rubio M, García-Hernando N, Soria-Verdugo A. Evaluation of combustible properties. Energ Fuel 2010;24(9):4638–45.
the number of first-order reactions required to accurately model biomass pyrolysis. [44] Prins MJ, Ptasinski KJ, Janssen FJJG. More efficient biomass gasification via
Chem Eng J 2021;418:127291. torrefaction. Energy 2006;31(15):3458–70.
[11] Chai M, Xie L, Yu X, Zhang X, Yang Y, Rahman MM, et al. Poplar wood torrefaction: [45] Prins MJ, Ptasinski KJ, Janssen FJJG. Torrefaction of wood Part 1. Weight loss
Kinetics, thermochemistry and implications. Renew Sustain Energy Rev 2021;143: kinetics. J Anal Appl Pyrol 2006;77:28–34.
110962. [46] Senneca O, Chirone R, Salatino P. A thermogravimetric study of nonfossil solid
[12] Chen Q, Zhou JS, Liu BJ, Mei QF, Luo ZY. Influence of torrefaction pretreatment on fuels. 2. Oxidative pyrolysis and char combustion. Energ Fuel 2002;16(3):661–8.
biomass gasification technology. Chin Sci Bull 2011;56:1449–56. [47] Senneca O, Chirone R, Salatino P. Oxidative pyrolysis of solid fuels. J Anal Appl
[13] Chen WH, Lu KM, Lee WJ, Liu SH, Lin TC. Non-oxidative and oxidative torrefaction Pyrol 2004;71:959–70.
characterization and SEM observations of fibrous and ligneous biomass. Appl Energ [48] Shang L, Ahrenfeldt J, Holm JK, Barsberg S, Zhang RZ, Luo YH, et al. Intrinsic
2014;114:104–13. kinetics and devolatilization of wheat straw during torrefaction. J Anal Appl Pyrol
[14] Chen WH, Peng J, Bi XT. A state-of-the-art review of biomass torrefaction, 2013;100:145–52.
densification and applications. Renew Sust Energ Rev 2015;44:847–66. [49] Silveira EA, Luz SM, Leao RM, Rousset P, Caldeira-Pires A. Numerical modeling
[15] Chen D, Chen F, Cen K, Cao X, Zhang J, Zhou J. Upgrading rice husk via oxidative and experimental assessment of sustainable woody biomass torrefaction via
torrefaction: Characterization of solid, liquid, gaseous products and a comparison coupled TG-FTIR. Biomass and Bioener 2021;146:105981.
with non-oxidative torrefaction. Fuel 2020;275:117936. [50] Singh RN. Equilibrium moisture content of biomass briquettes. Biomass Bioenerg
[16] Chen W-H, Eng CF, Lin Y-Y, Bach Q-V, Ashokkumar V, Show P-L. Two-step 2004;26(3):251–3.
themodegradation kinetics of cellulose, hemicellulose, and lignin under isothermal [51] Soria-Verdugo A, Goos E, García-Hernando N. Effect of the number of TGA curves
torrefaction analyzed by particle swarm optimization. Energy Convers Manag employed on the biomass pyrolysis kinetics results obtained using the Distributed
2021;237:114116. Activation Energy Model. Fuel Process Technol 2015;134:360–71.
[17] Coats AW, Redfern JP. Kinetic parameters from thermogravimetric data. Nature [52] Soria-Verdugo A, Goos E, Arrieta-Sanagustín J, García-Hernando N. Modeling of
1964;201(4914):68–9. the pyrolysis of biomass under parabolic and exponential temperature increases
[18] Czajka K, Kisiela A, Moron W, Ferens W, Rybak W. Pyrolysis of solid fuels: using the distributed activation energy model. Energy Convers Manag 2016;118:
Thermochemical behaviour, kinetics and compensation effect. Fuel Process 223–30.
Technol 2016;142:42–53. [53] Soria-Verdugo A, Goos E, Morato-Godino A, García-Hernando N, Riedel U.
[19] de Palma KR, García-Hernando N, Silva MA, Tomaz E, Soria-Verdugo A. Pyrolysis Pyrolysis of biofuels of the future: Sewage sludge and microalgae -
and combustion kinetics study and complementary ash fusibility behavior of Thermogravimetric analysis and modelling of the pyrolysis under different
sugarcane bagasse, sugarcane straw, and their pellets – Case of study of agro- temperature conditions. Energy Convers Manag 2017;138:261–72.
industrial residues. Energy Fuels 2019;33:3227–38. [54] Soria-Verdugo A, Rubio-Rubio M, Goos E, Riedel U. Combining the lumped
[20] Di Blasi C, Lanzetta M. Intrinsic kinetics of isothermal xylan degradation in inert capacitance method and the simplified distributed activation energy model to
atmosphere. J Anal Appl Pyrol 1997;40–41:287–303. describe the pyrolysis of thermally small biomass particles. Energy Convers
[21] Duan H, Zhang Z, Rahman MM, Guo X, Zhang X, Cai J. Insight into torrefaction of Manage 2018;175:164–72.
woody biomass: Kinetic modeling using pattern search method. Energy 2020;201: [55] Soria-Verdugo A, Rubio-Rubio M, Goos E, Riedel U. On the characteristic heating
117648. and pyrolysis time of thermally small biomass particles in a bubbling fluidized bed
reactor. Renew Energ 2020;110:312–22.

13
A. Soria-Verdugo et al. Energy Conversion and Management 267 (2022) 115892

[56] Svanberg M, Olofsson I, Flodén J, Nordin A. Analysing biomass torrefaction supply [64] Xin S, Huang F, Liu X, Mi T, Xu Q. Torrefaction of herbal medicine wastes:
chain costs. Bioresour Technol 2013;142:287–96. characterization of the physicochemical properties and combustion behaviors.
[57] Thengane SK, Burek J, Kung KS, Ghoniem AF, Sanchez DL. Life cycle assessment of Bioresour Technol 2019;287:121408.
rice husk torrefaction and prospects for decentralized facilities at rice mills. J Clean [65] Yang W, Shimanouchi T, Iwamura M, Takahashi Y, Mano R, Takashima K, et al.
Prod 2020;275:123177. Elevating the fuel properties of Humulus lupulus, Plumeria alba and Calophyllum
[58] Thengane SK, Kung KS, Gupta A, Ateia M, Sanchez DL, Mahajani SM, et al. inophyllum L. through wet torrefaction. Fuel 2015;146:88–94.
Oxidative torrefaction for cleaner utilization of biomass for soil amendment. [66] Yang Y, Sun MM, Zhang M, Zhang K, Wang DH, Lei C. A fundamental research on
Cleaner Engineering and Technology 2020;1:100033. synchronized torrefaction and pelleting of biomass. Renew Energ 2019;142:
[60] Uemura Y, Omar W, Othman NA, Yusup S, Tsutsui T. Torrefaction of oil palm EFB 668–76.
in the presence of oxygen. Fuel 2013;103:156–60. [67] Zhang C, Ho SH, Chen WH, Fu Y, Chang JS, Bi X. Oxidative torrefaction of biomass
[61] van der Stelt MJC, Gerhauser H, Kiel JHA, Ptasinski KJ. Biomass upgrading by nutshells: Evaluations of energy efficiency as well as biochar transportation and
torrefaction for the production of biofuels: A review. Biomass Bioenerg 2011;35: storage. Appl Energ 2019;235:428–41.
3748–62. [68] Zhang Z, Duan H, Zhang Y, Guo X, Yu X, Zhang X, Rahman M, Cai J. The following
[62] Wang C, Peng J, Li H, Bi XT, Legros R, Lim CJ, et al. Oxidative torrefaction of references were also included in the revised manuscript: Investigation of kinetic
biomass residues and densification of torrefied sawdust to pellets. Bioresour compensation effect in lignocellulosic biomass torrefaction: Kinetic and
Technol 2013;127:318–25. thermodynamic analyses. Energy 2020;207.
[63] Wang Z, Li H, Lim CJ, Grace JR. Oxidative torrefaction of spruce-pine-fir sawdust
in a slot-rectangular spouted bed reactor. Energ Convers Manage 2018;174:
276–87.

14

You might also like