You are on page 1of 15

Energy Reports 9 (2023) 5977–5991

Contents lists available at ScienceDirect

Energy Reports
journal homepage: www.elsevier.com/locate/egyr

Research paper

Development of carbonized rice husks briquettes: Synergy between


emissions, combustion, kinetics and thermodynamic characteristics

Vianney Andrew Yiga a,b , , Andrew Nuwamanya a , Agatha Birungi a , Michael Lubwama a,b,c ,
Harriet Nalubega Lubwama b
a
Department of Mechanical Engineering, Makerere University, Kampala, Uganda
b
Department of Research and Development, Yosevi Engineering Services Limited, Kampala, Uganda
c
Africa Center of Excellence in Materials, Product Development and Nanotechnology, Makerere University, Kampala, Uganda

article info a b s t r a c t

Article history: Briquettes uptake continues to be low, despite their ability to offset the negative impacts caused
Received 27 March 2023 by firewood/charcoal. Possible reason for this is lack of literature towards key parameters of the
Received in revised form 2 May 2023 developed briquettes. In the current study, emissions, combustion, kinetics and thermodynamic
Accepted 18 May 2023
characteristics of rice husks agricultural residue carbonized briquettes were obtained. The water
Available online 30 May 2023
boiling test and emissions monitoring system for CO, CO2 and PM2.5 were used to determine fuel
Keywords: and energy consumption, thermal efficiency and emissions while thermogravimetric analysis was
Briquettes utilized in obtaining combustion, kinetic and thermodynamic parameters. The activation energy was
Combustion determined by employing four different model-free methods. Drop strength and particle density of the
Emissions briquettes were 67.8% and 548.15kg/m3 , respectively. Briquettes‘ moisture, volatile matter, ash content
Kinetics and fixed carbon were 7.2%, 16.6%, 47.4% and 28.8% respectively. Average activation energy attained by
Pyrolysis the developed briquettes by Kissinger–Akahira–Sunose, Ozawa–Flynn–Wall, Starink and Tang methods
Rice husks
were 78.3kJ/mol, 83.3kJ/mol, 78.7kJ/mol and 75.0kJ/mol, respectively. Low energy barrier (5.2kJ/mol)
TGA
between activation energy and enthalpy changes indicated that initiation occurs easily. The obtained
thermodynamic parameters indicated that the thermal degradation process was endothermic in nature.
High energetic densities (10,688.9MJ/m3 ) and Fuel Value Indices (255.5MJ/m3 ·%) were attributable to
the higher heating value of the developed briquettes (19.5MJ/kg).
© 2023 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/).

1. Introduction from 48 million tonnes in 2015 to 105 million tonnes per annum
by 2040, which corresponds to an annual wood supply deficit of
Africa’s population at 1.21 billion (or 17% of the world total) 72,600 tonnes by 2030 (UBOS, 2020). When firewood/charcoal is
is expected to further rise, indicating a corresponding increase in burnt, its combustion reaction produces heat and emissions such
demand sources of energy (Oluwafemi‘Femi’Mimiko, 2022). Over as vapors, gases, and particulates (Yadav and Devi, 2018). The
70% of the population in sub-Saharan Africa does not have access smoke produced from using firewood/charcoal for domestic cook-
to electricity (Szabó et al., 2016). In Uganda, Uganda Bureau of ing applications not only has negative consequences to the envi-
Statistics (UBOS) reports that only 20% Ugandans have access ronment, but also has negative health implications on women and
to electricity (Cartland et al., 2022). For homes with access to children especially (Raymer, 2006; Rehfuess and World Health
electricity, it is majorly used for lighting purposes and an unno- Organization, 2006). This is mainly because these fuels are often
ticeable composition for cooking. Over 90% of Ugandans depend burnt in poorly ventilated kitchens or traditional cook-stoves
on charcoal and firewood as a source of energy for domestic characterized by very low efficiencies or any exhaust-cleaning
cooking purposes (Gianvenuti et al., 2022). This has significantly mechanisms, thus resulting to high emission of pollutants. Ac-
contributed to high rates of deforestation which is linked to tually, the smoke emitted during firewood/charcoal combustion
climate change. With increasing population growth and urban- contains toxic components that include a high concentration of
ization in Uganda, the demand for firewood/charcoal is expected particulate matter (PM), carbon monoxide (CO), nitrogen oxides
to further increase. The demand is projected to more than double (NOX ), sulphur oxides (SOX ), formaldehyde, and polycyclic or-
ganic matter (Sana et al., 2019). Sana et al. (2019) shows many
∗ Corresponding author at: Department of Mechanical Engineering, Makerere studies which have highlighted the link between indoor air pol-
University, Kampala, Uganda. lution and the occurrence of various health problems, including
E-mail address: vianney.yiga@mak.ac.ug (V.A. Yiga). cardiovascular and respiratory diseases in the short and medium

https://doi.org/10.1016/j.egyr.2023.05.066
2352-4847/© 2023 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/).
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

term, as well as chronic exposures. Emissions from the com-


bustion of firewood/charcoal are estimated to be responsible for
about 4.6% of the global burden of disease (Lim et al., 2012).
The World Health Organization (WHO) has held emissions from
burning of firewood/charcoal responsible for about 4.3 million of
global premature deaths worldwide (WHO, 2014). The emissions
to the environment from combustion of firewood/charcoal have
led to an upsurge in global warming, causing long-lasting changes
to the climate system, which threatens irreversible consequences
(Gioda, 2019). To curb these challenges, government policy has
been geared towards production of greener fuels for domestic
cooking applications (Mudoko, 2013). One such alternatives is
carbonized agricultural residue briquettes.
Agricultural residue briquettes initially attracted success as
alternatives to firewood/charcoal because of the large of amounts
of agricultural residues generated from agricultural activities. The
theoretical potential production of agriculture residues in Uganda
is approximated to be around 4.4 million metric tons annually
(Miito and Banadda, 2016). Rice husks are one such residues,
whose current production capacity is 136,000 metric tons annu- Fig. 1. Cross-section of step-down kiln for rice husks carbonization.
ally (Buyondo et al., 2020). Their current disposal means include
open burning, a practice which further exacerbates the effects of
climate change. Rice husks agricultural residue briquettes have
been developed but their uptake remains generally very low
(Nakibuuka, 2022; Nakiyinji, 2022; Lubwama et al., 2020a; Yiga
and Lubwama, 2020; Lubwama and Yiga, 2018). One of the major
reasons for this trend could be possible hazardous emissions
produced during their combustion. Additionally, their combus-
tion, kinetics as well as thermodynamic characteristics have not
been reported in the literature, further constraining customers’
perceptions towards utilizing rice husks carbonized briquettes
for domestic cooking applications. Combustion characteristics as
well as kinetic and thermodynamic parameters can be obtained
by thermogravimetric analysis (TGA) (Yiga et al., 2023; Liu et al.,
2021; Li et al., 2019).
In this study therefore, carbonized rice husks briquettes were Fig. 2. Rice husks (a); rice husks biochar (b).
developed using low pressure compaction (≤7MPa) and their
emissions, combustion, kinetics and thermodynamic character-
istics were presented. Additionally, characterizations based on
2.3. Briquette development
SEM/EDX, FTIR, drop strength, particle density, higher heating
value and energetic parameters were also presented.
Briquettes were produced by mixing rice husks biochar
2. Materials and methods (≤600 µm size) with cassava flour peelings binder before low
pressure (≤5 MPa) compaction. To form binder, 100 g of cassava
2.1. Materials flour peelings were mixed in 750 cm3 boiling of water, followed
by stirring to produce a homogeneous starchy paste. 1000 g of
Rice husks (K98 variety) were obtained from Zhong’s Indus- biochar were then mixed with the developed binder to form
tries Limited, Kalungu District, Uganda. Cassava peelings were ob- a homogenous composite material that was compressed into
tained Wandegeya market. The brown outer layer was removed cylindrical briquettes using a mold. The produced briquettes were
so as to obtain a high starch-containing cortex. The peelings were then oven-dried for 16 h at 65 ◦ C.
washed in water and dried for 8 h per day for one week to reach
a < 13% moisture content, before milling into flour.
2.4. Microstructural/EDX studies and FTIR spectrometry
2.2. Carbonization
SEM was used to observe developed briquettes’ networks.
Rice husks were carbonized using a step down kiln (see Fig. 1). Samples were observed under a Tescan Vega 3 scanning electron
To form biochar, rice husks (see Fig. 2a) were placed in saggars microscope, Pleasanton, USA, coupled with energy dispersive X-
of 80L capacity and 0.77 m height. They were then covered with ray (EDX) spectroscopy. They were fixed on double sided adhesive
flat ceramic covers to allow uniform heat flow to the saggars so as
carbon tape, vacuum dried and scanned at an acceleration volt-
to regulate heat transfer. The covers were always smeared using
age of 10 kV. EDX spectra were obtained at an acceleration
mud for pyrolysis to occur gradually from end to another. The
voltage of 20.0 kV and collected for 19 s. The FTIR spectra of
main entrance of the kiln was closed using bricks and mud, such
that the heat generated from burning firewood was maintained the developed briquette samples were collected in the range of
in the combustion area, to allow smooth heat exchange from 4000 to 400 cm−1 using a Jasco FT/IR-6600 type A machine,
the fuel to rice husks. Firing of the fuel was done to provide Cremella, Italy. The resolution was 4 cm−1 , the scanning speed
sufficient heat for carbonization until a temperature range of was 2 mm/sec. The samples were ground to obtain very fine pow-
400–500 ◦ C for 5 h (Guo et al., 2020). The kiln was closed for ders using a mortar and pestle. The presence of free functional
24 h exchanging enough heat to these saggars, thereby forming groups in the samples was determined based on the FTIR spectra
biochar (see Fig. 2b). Obtained biochar was then water-cooled. (Adeleke et al., 2021a).
5978
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

2.5. Mechanical characterization mine proximate analysis of raw materials as well as developed
briquettes (Lubwama et al., 2020b). Additionally, TGA was used
To obtain the compaction integrity of the briquettes, they to obtain briquettes’ mass loss with increase in temperature
were elevated up to 2 m and then dropped onto a thick steel (Lubwama and Yiga, 2017). TGA was carried out according to the
plate. Drop strength was then obtained as the ratio of the weight ASTM E1131 standard (ASTM, 2014). The samples were heated
after dropping to the weight before dropping (Lubwama and from 40 ◦ C to 600 ◦ C at three different heating rates of 20, 40
Yiga, 2017). Drop strength is an extremely important factor in and 50 ◦ C/min. High-purity compressed air (Oxygen: Nitrogen =
briquette packaging, storage and transportation (Anggraeni et al., 21:79, > 99.99%) was used for cleaning the crucibles and chamber
2021; Fengmin and Mingquan, 2011). The particle density of prior to TGA experimentation. Nitrogen gas was used as the purge
the developed briquettes was obtained by standard procedures gas for pyrolysis experimentation. The flow rate was maintained
involving ratios of mass and volume (Dao et al., 2022; Lubwama at 20 mL/min and the sample masses averaged 20 mg. TGA
et al., 2020b). It is an important parameter in understanding if also provided combustion explanations in terms of differential
they will have enough heat for boiling as well as transportation thermogravimetry (DTG), peak temperatures and mean reactivity
abilities without crumbling. of the rice husks.

2.6. Higher Heating Value (HHV) and Water boiling test (WBT) 2.8.1. Combustion characteristics parameters
For each heating rate, ignition temperature, Ti is a measure
Higher heating value of the developed briquettes was obtained of difficulty to ignite as low ignition temperature suggests easy
according to Eq. (1) through relationships with ash content, fixed ignition, and vice versa. Burnout temperature, Tf refers to the
carbon and moisture content (Lubwama et al., 2020b). temperature of the combustible substances in the completely
burnt state. The ignition and burnout temperatures of the de-
( )
MJ
HHV = 17300 − 117.51AC + 165.551FC − 232.69MC (1) veloped briquettes were determined according to the tangent
kg
method proposed by Liu et al. (2016). The flammability index,
The Water Boiling Test (WBT) protocol version 4.2.3 was used F, reflects the reactivity of the early stages of combustion and is
to determine pollutant emissions during briquettes combustion expressed by Eq. (2).
(CCA, 2014). An improved cook-stove was used and performance ( dα )
analysis was based on the boiling of 5 liters of water in a flat- dt max
F = (2)
bottomed cylindrical cross-section aluminum cooking pan of 7 Ti 2
liters maximum capacity. Tests conducted on the samples simu-
where, ddtα max is the maximum combustion rate. The larger the
( )
lated normal regular household cooking practices and specifically
determined thermal performance of the fuels and monitored value of flammability index, the more combustible a material is.
gaseous emissions namely: carbon monoxide (CO), carbon diox- The combustion characteristic index, S, reflects how quick the
ide (CO2 ) and aerosol emissions (Particulate matter, PM2.5 ). Spe- combustion rate is and is expressed by Eq. (3) (Li et al., 2017).
cific operational parameters measured during a WBT included ( dα ) ( dα )
thermal efficiency, specific fuel consumption, time to boil, burn- dt dt
max mean
ing rate, and fire power for all three WBT phases namely: High- S= (3)
Ti 2 Tf
power (cold-start); High power (hot-start) and Low-power (sim-
where, ddtα mean is the average combustion rate. The larger the
( )
mering) (Arora et al., 2014; Obeng et al., 2017). During high-
power (cold start), 5 litres of water were brought to a boil in a combustion characteristic index, the better the combustion char-
7-litre pot using a stove at ambient temperature. For high-power acteristics, and the more intense combustion is (Song et al., 2019).
(hot-start), 5 litres of water were brought to a boil in a 7-litre pot
using a pre-heated stove; and during low-power (simmering), the 2.8.2. Kinetics modeling
water temperature is kept at about 3 ◦ C below boiling point for
( )
−E
From the Arrhenius equation (k = Ae RT ), the rate constant in
62 min (Arora et al., 2014; Obeng et al., 2017).
the kinetic equation is closely related to temperature. A is a pre-
2.7. Emissions testing exponential factor and T is the absolute temperature. According
to Xiao et al. (2020), a non-isothermal process can be regarded as
A Laboratory Emissions Measurement System (LEMS) that an isothermal process within an infinitesimal short time interval.
measures gaseous emissions carbon monoxide (CO) and carbon As such, the kinetic equation for an isothermal homogeneous
dioxide (CO2 ) was used. In the particulate matter (PM2.5 ) system, phase reaction can be adopted for briquette pyrolysis under a
the exhaust is sampled from the duct through an ultra-sharp programmed heating condition by the means of thermal analysis
cut in the PM2.5 cyclone to remove particles larger than 2.5 µm. (see Eq. (4)).
The flow through the cyclone is controlled by a critical precision dα
orifice. Particles less than 2.5 µm are deposited on a fibre glass = k × f (α) (4)
dt
filter using a gravimetric filter system that is humidity condi-
where k is the rate constant of the reaction, α is the immediate
tioned (Kivumbi et al., 2021a; Arora et al., 2014; Obeng et al.,
weight loss ratio during pyrolysis and is given by Eq. (5), f (α)
2017). Weights before and after each experiment are taken on a
is a function that reflects the apparent kinetics of pyrolysis of
calibrated microbalance with the difference in mass of the filter
substances, which is expressed as shown in Eq. (6).
before and after testing yielding the emission of PM2.5 (Kivumbi
et al., 2021a). Emission factors for CO, CO2 and PM2.5 based on (m0 − m)
carbon and on energy were determined (Arora et al., 2014; Obeng
α= ( ) (5)
m0 − mf
et al., 2017).
Where m0 , m, mf are the initial, instantaneous and final weights
2.8. Thermogravimetric analysis of the sample, respectively.
f (α) = (1 − α)n (6)
Thermogravimetric analyzer (Eltra Thermostep non isothermal
Thermogravimetric analyzer, Haan, Germany) was used to deter- Where n is the order of the reaction.
5979
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991
( )
βi 1
Combining Eqs. (4)–(6), we obtain: A plot of ln against for a given value of conversion,
T1.894661 Tαi
dα α (yields )a straight line with slope −1.052Eα
( )
−E
and an intercept
= Ae RT
(1 − α)n (7) R
dt ln R.g (α) + 3.633504 − 1.894661(ln Eα ).
Aα Eα

For a given heating rate, β = dT


dt
, Eq. (8) can be obtained as the
non-isothermal reaction rate. 2.8.3. Thermodynamic analysis
dα dα dt The KAS, OFW, Starink and Tang methods were used to ob-
= × (8) tain thermodynamic characteristics of the developed briquettes
dT dt dT at a given heating rate, including change in Gibbs free energy
Therefore, Substituting Eq. (7) into Eq. (8) gives: (∆G), change in enthalpy (∆H) and change in entropy (∆S) using
dα Eqs. (14)–(16) (Yiga et al., 2023; Saha et al., 2022).
( )
A −E
= e RT
(1 − α)n (9)
β
( )
dT KB T m
∆G = Eα + (RTm ) ln (14)
In the current study, the data obtained from TG/DTG were used hA
to determine the kinetic parameters (activation energy (E) and Eα
( )
β Eα e RTm
pre-exponential factor (A)), based on Arrhenius equation. Fur- Where A =
RTm 2
ther, these kinetic parameters can be estimated graphically by
integrating Eq. (9) and then applying mathematical approxima- ∆H = Eα − (RTm ) (15)
tion for exponential term. Based on previous literatures, five ∆H − ∆G
( )
different approximations were carried out and correspondingly, ∆S = (16)
Tm
four isoconversional (model free) models and a model fitting
method were adopted to calculate the apparent activation energy where KB is the Boltzmann constant, h is the Plank’s constant and
(E) at specific conversion time (α ) (Chandrasekaran et al., 2017; Tm is the peak temperature during combustion.
Das and Tiwari, 2017). Among the above five models, Kissinger–
Akahira–Sunose (KAS), Ozawa–Flynn–Wall (OFW), Starink and 2.9. Energetic characterization
Tang models were used in this study.
The energy per unit volume (energetic density) of developed
(a) Kissinger–Akahira–Sunose (KAS) method
briquettes was obtained by multiplying their particle density with
In the KAS method, mathematical approximation for exponential
their higher heating value (see Eq. (17)). The Fuel Value Index
term is assumed and after approximation and rearrangement, the
(FVI) of the developed briquettes was calculated according to
solution is given by Eq. (10) (Akahira, 1971). Eq. (18) (Yiga et al., 2021; Massuque et al.).
βi
( ) ( )
A α Eα Eα
ln = ln − (10) ED = PD × HHV (17)
Tαi 2 R.g (α) RTαi 3
where ED is Energetic density (MJ/m ), PD is particle density
(kg/m3 ) and HHV is Higher Heating Value (MJ/kg).
( )
βi
A plot of ln against −1
for a given value of conversion, α
Tαi 2 Tαi
( ) ED
−Eα Aα E α FVI = (18)
yields a straight line with slope and an intercept ln R.g (α)
,
R AC
from which Eα can be calculated.
where FVI is Fuel Value Index (MJ/m3 · %), ED is Energetic density
(b) Ozawa–Flynn–Wall (OFW) method
(MJ/m3 ) and AC is Ash content (%).
The OFW method uses the Doyle’s approximation (Doyle, 1965)
and OFW equation is expressed by Eq. (11) (Ozawa, 1965). 3. Results and discussions
( )
A α Eα Eα
ln βi = ln − 5.332 − 1.052 (11) 3.1. Surface morphologies and elemental analysis
R.g (α) RTαi
Where g (α) is constant at a given value of conversion. A plot of The morphology of the developed briquettes is presented in
ln βi against T1 for a given value of conversion, α yields a straight Fig. 3. There were observed structures of granular surface, ir-
αi ( ) regular arrangements and charging effects, possibly due to the
−1.052Eα
line with slope R
and an intercept ln Aα Eα
R.g (α)
− 5.332. high volatility of the developed briquettes (Zhong et al., 2017). A
(c) Starink method porous structure was vivid, which allows more paths for airflow
The Starink method is a non-isothermal model-free method ex- during ignition of briquettes, hence enhancing burning efficiency
pressed by Eq. (12) (Starink, 2003). of the developed briquettes (Kivumbi et al., 2021b; Carnaje et al.,
2018). Moreover, a web of cellulosic fibers that are contained
βi
( )
Eα by an enveloping membrane was observed, which is similar to
ln = Constant − 1.0008 (12)
T 1.92 RTαi the structure of inherent rice husks material (Yiga et al., 2021).
( ) The reason for this observance is because rice husks contain high
βi 1
A plot of ln T1.92
against Tαi
for a given value of conversion, amounts of cellulose (≈30%) (Yiga et al., 2022).
−1.0008Eα Elemental analysis by EDX revealed that the developed bri-
α yields a straight line with slope R
and an intercept
quettes had carbon (64.8%) and oxygen (27.5%) as the main
Constant.
constituents with other elements such as potassium and sili-
(d) Tang method
con having very small percentages (see Table 1). High carbon
The Tang method is a non-isothermal model-free method ex-
composition signals that the developed briquettes will have high
pressed by Eq. (13) (Tang et al., 2003).
fixed carbon compositions, and hence high calorific values (Yiga
βi
( )
and Lubwama, 2020). As expected there was a high carbon–
ln oxygen ratio (2.4:1), which confirms presence of highly thermal
T 1.894661
( ) constituents of rice husks like lignin because it has a higher
Aα Eα Eα
= ln + 3.633504 − 1.894661(ln Eα ) − 1.052 carbon–oxygen ratio compared to cellulose. Similar results were
R.g (α) RTαi reported by Adeleke et al. (2021a). From EDX also, only trace
(13) amounts of Ca, Al and Fe were detected.
5980
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Fig. 3. Surface morphologies of developed briquettes at magnifications of 50 µm (a); 100 µm (b); 200 µm (c).

Table 1
Elemental compositions for EDX of the developed briquettes.
Element C O K Si Ca Al Fe
wt.% 64.8 27.5 3.4 2.3 0.8 0.7 0.6

3.2. Surface functional groups

Fig. 4 shows the FTIR spectra of developed briquettes. Peaks


around 3000 cm−1 and 1100 cm−1 are typical characteristic of
cellulose backbone in rice husks that ascribed to –OH and C–O–C
stretching bands, respectively (Yakaboylu et al., 2021). The broad-
band at around 3300 cm−1 is where moisture content could take
part in the formation of hydrogen bonds is associated with O–H
vibrations in hydroxyl groups due to the presence of pheno-
lic hydroxyl groups (Adeleke et al., 2021a,b). Hydroxyl groups
signal presence of an alcohol, which can contribute to higher
flammability index and hence enhanced combustion character-
istics (Carnaje et al., 2018). Deformation vibration of C–H in
aliphatic carbon (–CH3 ) and –C(CH3 )2 appeared in at around
2800–2975 cm−1 and around 1200 cm−1 . At around 1384 cm−1 , Fig. 4. FTIR spectra of developed briquettes.
the observed band is due to aromatic C==C stretching (Montiano
et al., 2015). The aromatic C==C bonds and characteristic phe-
nolic OH groups were established along with oxygen bridges obtained in the current study, in comparison to those obtained
as strengtheners for briquettes when exposed to curing in inert by Anggraeni et al. (2021) (≥82%), Lubwama et al. (2020b) (>
environment (Adeleke et al., 2021b; Zhong et al., 2017). According 85%), and Obi and Okongwu (2016) (76%–100%), respectively.
to Carnaje et al. (2018), the presence of aliphatic and aromatic This could have been low pressure of ≤5 MPa utilized in the
hydrocarbons in the fuel means that it contains fats and oils that current study. Particle density is extremely important as shows
enhance ignition properties of the developed briquettes. the compactness of a briquette and is directly proportional to
its compressive strength (Granado et al., 2021). Particle densities
3.3. Mechanical characterization reported in this study indicate the briquettes’ ability to burn for
long time thus emitting large quantities of heat during burning.
Drop strength and particle density of the developed briquettes
were 67.8% and 548.15 kg/m3 , respectively. Similar density values 3.4. Proximate analysis
were reported for rice husks briquettes developed by Deshan-
navar et al. (2018) and Arewa et al. (2016) but lower densities Proximate analysis for the raw materials used in making bri-
were obtained by Obi and Okongwu (2016) (333–499 kg/m3 ). quettes as well as accruing briquettes are shown in Fig. 5. Rice
The density obtained by Kipngetich et al. (2023), Suryaningsih husks biochar had the highest moisture content (22.5%), because
et al. (2019a, 2017) for carbonized rice husks briquettes was it was characterized in its moist form as obtained after water
much higher (838–1058 kg/m3 , 873–1158 kg/m3 and 680 kg/m3 cooling from the kiln. The moisture content of cassava peeling
respectively). Drop strength and particle density provide an in- flour (9.1%) is lower than of cassava peelings (10.9%) because
dication of handle-ability, transportation and storage integrity when the peelings are milled its bulk density increases and thus
of the developed carbonized briquettes (Anggraeni et al., 2021; lose moisture to the atmosphere. Cassava peelings flour had the
Deshannavar et al., 2018). It also shows the extent to which highest volatile matter compositions (72.9%) among the sam-
briquettes can be stacked on top of each other during domestic ples. Rice husks biochar obtained the highest ash content (47.1%)
or kiln cooking (Lubwama et al., 2020b). Lower drop strength was due to silica formation during temperature degradation of rice
5981
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

firewood and charcoal, respectively. Clearly, the briquettes devel-


oped in the current study are superior to firewood. Their HHV
are within the ranges of 16.3–22.7 MJ/kg, 16.4–23.7 MJ/kg, 19.9–
23.7 MJ/kg and 13.5–21.5 MJ/kg for carbonized rice husks bri-
quettes developed by Kipngetich et al. (2023, 2022), Suryaningsih
et al. (2019a) and Obi and Okongwu (2016), respectively. Higher
calorific value ranges were obtained by Deshannavar et al. (2018)
(18.6–20.1 MJ/kg) and Elinge et al. (2019) (17.3–34.6 MJ/kg).
HHV of briquettes significantly influences their energy density,
which is an invaluable factor to consider for grouping one or
more residues at the time of briquetting as it allows for stan-
dardization of the briquettes density (de Oliveira et al., 2022).
Fig. 6a shows a stove loaded with briquettes before ignition.
A yellowish bioethanol gel (see Fig. 6b) was used for igniting
briquettes, because it is much thicker than kerosene and produces
smaller blue flames which be sustained for longtime producing a
lot of heat content (Kivumbi et al., 2021a). As heat is exchanged
from the bioethanol gel, the briquettes adjacent to the ash tray
gets heated which later exchanges heat to the other parts of the
briquettes, through Fourier law of conduction (Lubwama et al.,
Fig. 5. Proximate analysis. 2020b). It took 5 min to ignite 230 g of rice husks briquettes
and 7 min to ignite 375 g or rice husks briquettes (see Fig. 6c).
Clearly, ignition time depends on the quantity of briquettes as less
husks to form biochar (Javed et al., 2022; Chang, 2020). The amount of fuel is easier to ignite, compared to higher amounts
least ash contents were obtained in cassava peelings (3.2%) and of fuel. After briquettes have completely burnt, ash remains (see
cassava peelings flour (3.7%). The highest fixed carbon content Fig. 6d) are noticeable because of the high amounts of ash already
was obtained in rice husks biochar (20.8%) because carbonization noticed from proximate analysis (see Fig. 5).
increases carbon formation (Amarasekara et al., 2017). Briquettes‘
moisture, volatile matter, ash content and fixed carbon were 7.2%, 3.5.2. Emissions in the three phases
16.6%, 47.4% and 28.8% respectively. These results agree with The concentrations of CO2 and CO increased during ignition
previous findings on carbonized rice husks briquettes (Kipngetich phase with time, however during this experiment in this experi-
et al., 2023, 2022; Suryaningsih et al., 2019a; Lubwama and ment three phases have been studied. Maximum emissions of CO2
Yiga, 2018; Suryaningsih et al., 2017). Briquettes with such low and CO were obtained during the hot-start phase and less emis-
moisture contents are feasible for combustion because excessive sions of were obtained during the other phases of cold start, hot
energy is needed for drying and higher moisture content is indeed start and simmer phase (see Fig. 7a). CO concentrations remained
a challenge while burning (Afsal et al., 2020). Moreover, low non recognizable at the start of the cold-start phase until about
moisture content is synonymous with high calorific values and the 20th minute, before exponentially increasing by 55th minute.
easy fuel ignition (Suryaningsih et al., 2019a). The ash contents of The significant increase in CO concentration shows incomplete
the developed briquettes in the present study were in range with combustion, related to evaporation of some moisture from the
previous works of Kipngetich et al. (2022) (47.33–51.24%) and briquettes surface (Obeng et al., 2017). Additionally, it is possibly
Kipngetich et al. (2023) (38.4–48.7%). The fixed carbon composi- due to low burning temperature, insufficient oxygen, poor mixing
tions obtained in the briquettes are coherent with the high carbon of fuel with the combustion air and too short residence time of
the combustion gases in the combustion zone (Xiu et al., 2018).
content already observed from the EDX results (see Table 1).
CO2 rate in the first phase is high because significant moisture
It should be noted that the fixed carbon compositions obtained
content is expelled from the developed briquettes in this phase
are much higher than those for firewood (14.6%) but less than
(Kpalo et al., 2021). During hot start phase, developed briquettes
charcoal (56.2%) from the study of Lutaaya et al. (2023).
lost their volatile matter, fixed carbon, thus easily thus emitting
high emissions. In this phase, CO emissions increased gradually
3.5. Emissions performance results up. This finding is contrary to findings of Kivumbi et al. (2021a)
who reported that the hot start phase produced less emissions
3.5.1. Ignition properties compared to other cold start and simmer. During the simmer
Ignition is a rapid transition process by which an exothermic phase, the CO and CO2 emissions gradually decrease, indicating
oxidation reaction and self-supported combustion are initiated a constant dissemination of emissions to the atmosphere. The
(Chavan et al., 2022). When the briquettes are ignited, their tem- emissions of PM2.5 generated ranged between 314.7 ppm and
peratures increase, which makes them decompose into another 317.0 ppm (see Fig. 7b). These emissions are related to the inher-
solid phase that is less compact. Some of their constituents such ent ash contents in the developed briquettes (Deng et al., 2019).
as volatile matter escape during the burning process that occur. Although the trendlines indicated high emission levels for CO
Other constituents such as moisture content turn into vapor during the test, personal exposure to CO emissions is relatively
phase gas phase. Corresponding higher heating value (HHV) of low, since the laboratory experiment was conducted under con-
the developed briquettes was 19.5 MJ/kg, which is much higher trolled conditions that required placing the stove directly under
than other carbonized rice husks previously developed by Lub- the hood which is ideal (Obeng et al., 2017).
wama and Yiga (2018) (≈16.5 MJ/kg), Suryaningsih et al. (2018)
(≤13.1 MJ/kg) and Suryaningsih et al. (2017) (≈14 MJ/kg). In 3.6. Water Boiling Test (WBT) performance metrics
fact, it is also in range than charcoal’s calorific value reported
by Tumutegyereize et al. (2016) (18.3–27.3 MJ/kg). Lutaaya et al. Table 2 shows the IWA performance metrics as well as the
(2023) reported calorific values of 17.6 MJ/g and 29.4 MJ/kg for standard performance measures of the developed briquettes. It
5982
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Fig. 6. Burn stove (a); bioethanol gel (b); burning briquettes (c); and burnt briquettes covered with ash (d).. (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)

Fig. 7. (CO/CO2 ) emissions (a); and PM2.5 emissions (b) produced during the water boiling test.

Table 2
IWA performance metrics and standard performance measures for developed briquettes.
High Power Thermal Efficiency (%) 47.4%
Low Power Specific Consumption (MJ/min/L) 0.003
High Power CO (g/MJd) 10.31
Low Power CO (g/min/L) 0.05
IWA performance metrics
High Power PM (mg/MJd) 60.1
Low Power PM (mg/min/L) 0.36
Indoor Emissions CO (g/min) 0.65
Indoor Emissions PM (mg/min) 3.8
Fuel to Cook 5L (g) 259.0
CO to Cook 5L (g) 29.5
PM to Cook 5L (mg) 184.2
Standard performance measures
Energy to Cook 5L (kJ) 4,749
Time to Boil (min) 56.4
CO2 to Cook 5L (g) 546.5

took 56.4 minutes and 259 g of briquettes at an energy consump- and simmer phases, the soot emission had reduced considerably
tion of 4,748.8kJ to boil 5 liters of water, hence emitting 546.5 g, (Kivumbi et al., 2021a). The respective total emissions of PM2.5
29.5 g and 184.2mg of CO2 , CO and PM2.5 , respectively. The cor- were 112 mg, 93 mg and 68 mg (see Fig. 8a). Similarly, emissions
responding low power specific consumption rate and high-power and emissions per MJ reduced with change in phase from cold-
thermal efficiency were 0.003 MJ/min/L and 47.4%, respectively. start to hot-start and simmer. The total emissions and emissions
The obtained results showed much lower emission dissipation per MJ for CO and CO2 during the simmer phase were lower
compared to a study by Obeng et al. (2017) who obtained the than those obtained during the cold-start and hot-start phases,
emissions performance of different wood species using different because simmering took less time (Mitchell et al., 2016). These
cook-stoves. findings were in contradiction to those obtained by Kivumbi et al.
Fig. 8 shows the total emissions and emissions per MJ for PM2.5 (2021a) who reported higher emissions and emissions per MJ in
(see Fig. 8a), CO (see Fig. 8b) and CO2 (see Fig. 8c) of the developed the simmer phase compared to both cold and hot-start phases.
briquettes. The highest amount of PM2.5 emissions was gener- As expected, the emissions per MJ were in line with the total
ated during the cold-start phase, since the fuel burned with a emissions of CO, CO2 and PM2.5 generated. As such, more CO and
yellow flame accompanied with emission of soot. In the hot-start PM2.5 emissions per MJ were generated in the cold-start phases,
5983
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Fig. 8. Total emissions and emissions per MJ for PM2.5 (a); CO (b); and CO2 (c) for developed briquettes.

compared to the other phases. This is because during the cold- will be emitted as compared to PM2.5 . However, since rice husks
start phase, a lot of power is focused to evaporating the existing briquettes are carbon-neutral, the generated emissions during
volatiles in the rice husks briquettes (Lubwama and Yiga, 2018). their combustion can be sustainably absorbed by plants during
Moreover, the incomplete combustion due to inherent moisture photosynthesis (Suryaningsih et al., 2019b; Thao et al., 2012).
in rice husks briquettes causes formation of high CO emissions.
On the other hand, CO2 emissions per MJ are highest in the 3.7. TG and DTG results
simmer phase (160 g/MJ) possibly due to sufficient time for air to
mix, causing higher CO2 generation (see Fig. 8c). Similar results Fig. 9 presents the TG, DTG and conversion curves for the
were obtained by Kivumbi et al. (2021a). developed briquettes at three different heating rates of 20, 40
The basic operational and energy consumption parameters as and 50 ◦ C/min. The TG curves (see Fig. 9a) show the typical
well as emission factors for cold-start, hot-start and simmering appearance of pyrolysis of carbonized rice husks briquettes and
operations for the developed briquettes are presented in Ta- from them, the thermal phases for each of the heating rates
ble 3. It took 55 min to boil water in the cold-start phase as can be located (Lubwama and Yiga, 2018; Yiga and Lubwama,
opposed to the 50 min taken in the hot-start phase. This trend 2020). It is evident that the complete pyrolysis process for all the
is due to the fact that less energy is required for the reaction samples can be divided into three major stages namely; dehydra-
in the latter. The burning rate was 1.89g/min, 1.93g/min and tion, devolatilization and carbonization (Setter and Oliveira, 2022;
0.39g/min for cold-start, hot-start and simmer phases, respec- Setter et al., 2020). The first stage from 40 to 150 ◦ C causes degra-
tively. Lower rates were reported by Anggraeni et al. (2021) dation and hence weight loss (≤8%) of some light components,
(≤0.81g/min) while works by Arewa et al. (2016) reported val- including inbound water and light volatile components (Ogwang
ues in range of the current study (0.7–2.0g/min). Kivumbi et al. et al., 2021; Ojha et al., 2021; Hu et al., 2015)). Corresponding
(2021a) obtained higher rates of 3.7–8.2g/min, 2.1–4.2g/min, and peaks for water evaporation can be observed in the DTG curves
1.1–1.7g/min for cold-start, hot-start and simmer phases, respec- at this temperature range (see Fig. 9b). Water evaporation formed
tively for charcoal fines briquettes while Nwabue et al. (2017) a number of slit-shaped cracks on the briquettes’ surface and
obtained average burning rates of 1300–3800g/min for bio-coal these cracks acted as the gateway for pyrolysis gases spilling out
briquettes. This implies that the briquettes used in this study burn in the subsequent pyrolysis process (Hu et al., 2021). From the
slower; thus, the cook-stove does not have to be loaded with fuel findings of Kuthe et al. (2021), lignocellulosic materials with less
frequently (Kivumbi et al., 2021a). Lutaaya et al. (2023) reported than 15% water content are feasible for making briquettes for
burning rates of 8g/min and 4g/min for firewood and charcoal combustion purposes. The reason behind this is that excessive
respectively. Thermal efficiencies were 46%, 49% and 65% for moisture content poses major issues in grinding and immoderate
the cold-start, hot-start and simmer phases, respectively. These strength is required for drying.
were much higher than those obtained for briquettes developed The second stage, between 150 and about 400 ◦ C causes the
by Kivumbi et al. (2021a) (21.8%–54.6%) and Lask et al. (2015) maximum weight loss during thermal degradation. From Fig. 9a,
(≈40%) as well as firewood (31%) and charcoal (36%) from Lutaaya it is clear that an increase in heating rate leads to a clear shift
et al. (2023). Specific fuel consumption was 42.8 g/L, 39.7 g/L of this phase to the right-hand side, which confirms that higher
and 7.5 g/L for cold-start, hot-start and simmer phases, respec- temperatures are required to cause decomposition of hemicellu-
tively. Higher specific fuel consumption values were reported by lose and cellulose (de Souza et al., 2022; Setter and Oliveira, 2022;
Anggraeni et al. (2021) (≤560 g/L) in rice husk briquettes as Hu et al., 2021; Kumar et al., 2020; Cao et al., 2014). The shift
well as Lutaaya et al. (2023) (80 g/L) in firewood. Corresponding is more vivid in the DTG curves as peak temperatures increased
specific energy consumption rates from the current study were from 354.9 ◦ C to 377.5 ◦ C to 380.7 ◦ C as heating rates increased
0.014 MJ/min/L, 0.015 MJ/min/L and 0.003 MJ/min/L, respectively. from 20 ◦ C/min to 40 ◦ C/min and to 50 ◦ C/min respectively (see
The reason for this behavior was the high net calorific value of Fig. 9b). The shift is due to the variation in the heat flux within the
the developed briquettes (18.3 MJ/kg) which ensured constant rice husks particles caused by the different heating rates (Setter
heat delivery through the briquettes during cooking. Emission and Oliveira, 2022; Setter et al., 2020). As such, at the lower
factors are a critical parameter in determining the greenhouse heating rate, sufficient time is available for heating because of the
gas emissions from the combustion of the developed briquettes. linear temperature profile between the outer surface and inner
They quantify the magnitude of emission normalized by fuel core of the developed briquettes, while at the higher heating rate,
or energy consumption (Still et al., 2011). It is clear that CO2 the temperature gradient is sufficient between outer and inner
emissions are by far the most dominant emission category in core of the briquettes (Kumar et al., 2020). The peak temperatures
the determination of the total greenhouse gas emissions while obtained in the current study are similar to those reported by Hu
PM2.5 were least significant. The significance is that more CO2 et al. (2015) on carbonized rice husks briquettes (358–388 ◦ C).
5984
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Table 3
Basic operational and energy consumption parameters as well as emission factors for the developed
briquettes.
Time to boil (min) 55
Burning rate (g/min) 1.89
Basic operation Thermal efficiency (%) 46
Specific fuel consumption (g/L) 42.8
Firepower (W) 577
Net Calorific Value (MJ/kg) 18.3
Energy consumption rate (kJ/min) 35
Specific energy consumption (MJ/min/L) 0.014
Cold-start Energy consumption parameters
Total energy consumed (kJ) 1,923
Energy delivered (MJ) 0.878
Average cooking power (kW) 0.247
CO (g/kg) 115.81
CO2 (g/kg) 1651
PM2.5 (g/kg) 0.68
Emission factors
CO (g/MJ) 4.7
CO2 (g/MJ) 66.4
PM2.5 (mg/MJ) 27.6
Time to boil (min) 50
Burning rate (g/min) 1.93
Basic operation Thermal efficiency (%) 49
Specific fuel consumption (g/L) 39.7
Firepower (W) 589
Net Calorific Value (MJ/kg) 18.3
Energy consumption rate (kJ/min) 35
Specific energy consumption (MJ/min/L) 0.015
Hot-start Energy consumption parameters
Total energy consumed (kJ) 1,787
Energy delivered (MJ) 0.861
Average cooking power (kW) 0.267
CO (g/kg) 109.34
CO2 (g/kg) 1662
PM2.5 (g/kg) 0.63
Emission factors
CO (g/MJ) 5.1
CO2 (g/MJ) 76.8
PM2.5 (mg/MJ) 28.9
Burning rate (g/min) 0.39
Thermal efficiency (%) 65
Basic operation
Specific fuel consumption (g/L) 7.5
Firepower (W) 118
Net Calorific Value (MJ/kg) 18.3
Energy consumption rate (kJ/min) 7
Specific energy consumption (MJ/min/L) 0.003
Energy consumption parameters
Simmer Total energy consumed (kJ) 323
Energy delivered (MJ) 0.209
Average cooking power (kW) 0.078
CO (g/kg) 70.55
CO2 (g/kg) 1722
PM2.5 (g/kg) 0.49
Emission factors
CO (g/MJ) 4.7
CO2 (g/MJ) 114.7
PM2.5 (mg/MJ) 32.3

Fig. 9. TG curves (a); DTG curves (b); and conversion curves (c) for developed briquettes.

5985
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Table 4
Combustion characteristics parameters of developed briquettes.
Heating Temperature Combustion rate Flammability Combustion Mean
rate (◦ C) (%/min) index (10−6 ) characteristic reactivity
index (10−10 ) (10−3 )
(◦ C/min) Ignition Peak Burnout Maximum Average (%/ min ·◦ C2 ) (%2 //min2 · ◦ C3 ) (%/min/◦ C)
20 196.7 354.9 439.7 0.07 0.03 1.78 1.02 0.19
40 219.3 377.5 457.2 0.06 0.02 1.31 0.70 0.17
50 251.1 380.7 452.8 0.06 0.02 1.02 0.47 0.17

In third stage, from about 400 to 600 ◦ C, lignin in the devel- termined using KAS, OFW, Starink and Tang methods based on the
oped briquettes is decomposed off, and the weight loss continues thermogravimetric analysis data obtained under different heating
until char residues are left at the maximum decomposition tem- rates. In order to compute the kinetic parameters, the same values
perature (Setter and Oliveira, 2022; de Souza et al., 2022; Ogwang of conversion rate (α ) in the range of 0.2–0.8 were considered for
et al., 2021; Setter et al., 2020; Kumar et al., 2020). High amount the three heating rates (Pinzi et al., 2020; Buratti et al., 2016).
of char residues is proportionally linked to high HHV (de Souza This range was considered because during the fitting of data using
et al., 2022). Increasing char residues were also noted with in- the KAS, OFW, Starink and Tang methods, the lowest conversion
creasing heating rate due to heat transfer constraints as higher value (0.1) and the highest conversion value (0.9) did not fit well
heating rates reduce the time required to reach the pyrolysis because of poor correlation factors (Kaur et al., 2018). In fact, they
temperature, and the pyrolysis time thus becomes shorter (Quan may contain errors because at these values thermal degradation
et al., 2009). Similar results were obtained by Liu et al. (2021) for does not reflect the overall pyrolytic behavior of the developed
corn straw briquettes. briquettes (Gan et al., 2018). Figs. 10a, 10b, 10c and 10d show
The variation in degree of conversion of the developed bri- the KAS, OFW, Starink and Tang linear plots, respectively, for the
quettes as a function of temperature at different heating rates of conversion values within 0.2–0.8 for the developed briquettes
20, 40 and 50 ◦ C/min are shown in Fig. 9c. From Eq. (5), when (Yiga et al., 2023). From their slopes and intercepts, calculated
temperature rises, conversion rates increased due to reduction activation energies and pre-exponential factors from the slopes
in original weight of the briquettes. The trend followed by the and intercepts are listed in Table 5. The average correlation val-
conversion curves is very similar to that presented in TG and ues of 0.8851, 0.9117, 0.8864 and 0.8868 for KAS, OFW, Starink
DTG curves, which present a decomposition in original weight and tang methods, respectively, indicated the accuracy of the
(see Figs. 9a and 9b). Therefore, increase in heating rate gener- individual isoconversional methods and the calculated activation
ally shifted the conversion curve to the right, signaling higher energies were reliable (Kumar and Ramesh, 2022; Ni et al., 2022).
temperatures are required to decompose constituents of devel- Corresponding average activation energy values were 78.3 kJ/mol,
oped briquettes at higher heating rates. Similar observations were 83.3 kJ/mol, 78.7 kJ/mol and 75.0 kJ/mol, respectively. The ob-
reported by Kumar and Ramesh (2022). tained activation energy values for the developed briquettes are
much lower than those for rice husks materials characterized
3.8. Combustion characteristics parameters
by Jia et al. (2020) (170.4–326.4 kJ/mol) and Loy et al. (2019)
(153.1–279.9 kJ/mol). This confirms the successful conversion of
Table 4 shows the briquettes’ combustion characteristic pa-
rice husks into briquettes with suitable HHV.
rameters. Increase in heating rate generally led to an increase
The relationship between activation energy and conversion
in ignition, peak and burnout temperatures, which confirmed
rate (α ) is shown in Fig. 11, depicting that the activation energy
the results for thermal stability (see Fig. 9). Similar observations
of the developed briquettes deeply depended on the conversion
were noted by Gao et al. (2022) for bamboo briquettes and
rate. It is clear that increase in conversion rate led to increasing
Liu et al. (2021) for corn straw briquettes. Increase in heat-
activation energy, due to thermal degradation of different com-
ing rate from 20◦ C/min to 40◦ C/min led to increase in ignition
ponents of biomass with increasing temperature. This trend is
temperature from 196.7◦ C to 219.3◦ C and further increase in
synonymous with that obtained by Chen et al. (2021), Jia et al.
heating rate to 50◦ C/min increased ignition temperature further
(2020) and Kumar et al. (2020) for rice husks materials. The
to 251.1 ◦ C. In a similar trend, increasing heating rate from 20
◦ increase in activation energy with increasing conversion rate is
C/min to 40 ◦ C/min led to increase in burnout temperature
from 439.7◦ C to 457.2◦ C while further increase in heating rate to caused by the decomposition of chars in the developed briquettes
50◦ C/min decreased burnout temperature by about 1%. Increas- (Liu et al., 2021). It should be noted that OFW method presented
ing ignition and burnout temperatures with increasing heating higher activation energy values compared to the KAS, Starink and
rate is because less time is available for heating combustible OFW methods, but the variations among these methods at each
elements in the developed briquettes. Average maximum and av- conversion rate were minimal.
erage combustion rates of the developed briquettes were around
0.07 %/min and 0.02 %/min. Average flammability index and com- 3.10. Thermodynamic parameters
bustion characteristic index values were 1.37×10−6 %/min· ◦ C2
and 0.73×10−10 %2 /min2 · ◦ C3 , respectively. Average mean reac- Thermodynamic parameters of the developed briquettes at
tivity of the developed briquettes was 0.18×10−3 %/min/◦ C. In- the three heating rates of 20, 40 and 50 ◦ C/min were calculated
creasing heating rate led to reducing mean reactivity as well using the OFW method, as shown in Eqs. (14)–(16) and are
as flammability and combustion characteristic indices because presented in Table 6. Enthalpy of reaction (∆H) and Gibbs free
less time was available for heating combustible elements in the energy (∆G) increased with increase in conversion rate. The low
developed briquettes. energy barrier (5.2 kJ/mol) between the activation energies of the
developed briquettes (see Fig. 11) and their ∆H values for the
3.9. Kinetics analysis process indicated that reaction initiation can occur easily, due
to the fact that the lower difference between ∆H and activation
Kinetic parameters for pyrolysis of the developed briquettes energy favors the complex formation (Kumar and Ramesh, 2022;
including activation energy and pre-exponential factor were de- Simões et al., 2022). This is not surprising, since the inherent
5986
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Fig. 10. Kinetic plots for developed briquettes using KAS method (a); OFW method (b); Starink method (c); and Tang method (d).

Table 5
Kinetic parameters of developed briquettes.
Conversion rate Activation energy (kJ/mol) Pre-exponential factor (min−1 ) R2
(α ) KAS OFW Starink Tang KAS OFW Starink Tang KAS OFW Starink Tang
0.2 27.7 32.5 28.0 26.7 6.9E−05 1.4E+04 5.9E−01 2.6E−05 0.9604 0.9734 0.9611 0.9613
0.3 18.5 24.3 18.8 18.0 4.6E−03 1.2E+03 2.8E+00 1.8E−04 0.9325 0.9652 0.9345 0.9351
0.4 31.6 37.9 31.9 30.5 3.1E+01 1.2E+04 7.0E−01 9.8E−06 0.9386 0.9616 0.9398 0.9402
0.5 41.4 49.0 41.8 39.9 3.7E−04 2.5E+04 3.8E−01 9.7E−07 0.4337 0.5424 0.4384 0.4400
0.6 98.8 103.9 99.2 94.5 9.7E−01 5.6E+08 1.0E+01 3.2E−02 0.9844 0.9874 0.9846 0.9846
0.7 114.4 119.0 114.7 109.3 9.5E+00 5.9E+09 1.2E+01 1.7E−01 0.9975 0.998 0.9975 0.9975
0.8 216.0 216.2 216.3 205.9 9.8E+07 7.1E+16 2.9E+01 7.6E+05 0.9487 0.9536 0.9489 0.949
Average 78.3 83.3 78.7 75.0 1.4E+07 1.0E+16 7.9E+00 1.1E+05 0.8851 0.9117 0.8864 0.8868

rice husks also have low energy barriers, which indicate that Gibbs free energy (∆G), which reveals the overall energy
rice husk are suitable for pyrolysis (Kumar et al., 2020; Loy change of the system, was found to decrease with increase in
et al., 2019). As observed with previous studies, heating rate conversion rate. Lower ∆G values indicates favorable decomposi-
did not significantly affect the briquettes’ enthalpy (Kumar and tion which signals that as conversion rate increased, more com-
Ramesh, 2022). Average ∆H value ranges (depending on heating bustible elements in the developed briquettes were converted
rate) were in the ranges of 69.8–78.0 kJ/mol, 69.6–77.9 kJ/mol to volatiles. Average ∆G value ranges were 162.7–163.5 kJ/mol,
and 69.5–77.8 kJ/mol for 20 ◦ C/min, 40 ◦ C/min and 50 ◦ C/min
165.3–166.2 kJ/mol and 165.0–165.9 kJ/mol, for 20 ◦ C/min,
heating rates, respectively. The obtained positive values for ∆H
40 ◦ C/min and 50 ◦ C/min heating rates, respectively. Clearly, the
values indicated that endothermic reactions were dominant for
heating rate had a mild effect on ∆G, as is the case in the original
the thermal devolatilization of the developed briquettes (Kumar
and Ramesh, 2022; Kali et al., 2022). Additionally, it was clear rice husks (Kumar et al., 2020). Previous findings on ∆G of rice
that the ∆H values are much lower than those of rice husks husks showed that they were in the range of 191.8–319.9 kJ/mol.
from which the briquettes are made, signaling that less energy is The corresponding significant drop in ∆G when briquettes are
required to dissociate the complex bonds of briquettes compared made implies that less energy was needed to form activated
to rice husks for formation of new chemical bonds (Kumar et al., complex (Loy et al., 2019). These positive values suggested that
2020). the reactions were not spontaneous (Simões et al., 2022). The
5987
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Table 6
Thermodynamic parameters of developed briquettes.
∆G ∆H ∆S
Method α
(kJ/mol) (kJ/mol) (kJ/mol.K)
20 ◦ C/min 40 ◦ C/min 50 ◦ C/min 20 ◦ C/min 40 ◦ C/min 50 ◦ C/min 20 ◦ C/min 40 ◦ C/min 50 ◦ C/min
0.2 166.95 169.78 169.48 22.48 22.29 22.27 −0.23 −0.23 −0.23
0.3 169.06 171.96 171.68 13.28 13.09 13.07 −0.25 −0.24 −0.24
0.4 166.26 169.07 168.77 26.38 26.19 26.17 −0.22 −0.22 −0.22
0.5 164.85 167.61 167.30 36.18 35.99 35.97 −0.20 −0.20 −0.20
KAS
0.6 160.31 162.90 162.57 93.58 93.39 93.37 −0.11 −0.11 −0.11
0.7 159.54 162.11 161.78 109.18 108.99 108.97 −0.08 −0.08 −0.08
0.8 156.23 158.67 158.32 210.78 210.59 210.57 0.09 0.08 0.08
Average 163.31 166.02 165.70 73.12 72.93 72.91 −0.14 −0.14 −0.14
0.2 166.11 168.92 168.62 27.28 27.09 27.07 −0.22 −0.22 −0.22
0.3 167.63 170.49 170.20 19.08 18.89 18.87 −0.24 −0.23 −0.23
0.4 165.31 168.09 167.78 32.68 32.49 32.47 −0.21 −0.21 −0.21
0.5 163.97 166.70 166.38 43.78 43.59 43.57 −0.19 −0.19 −0.19
OFW
0.6 160.05 162.63 162.30 98.68 98.49 98.47 −0.10 −0.10 −0.10
0.7 159.34 161.90 161.56 113.78 113.59 113.57 −0.07 −0.07 −0.07
0.8 156.22 158.67 158.32 210.98 210.79 210.77 0.09 0.08 0.08
Average 162.66 165.34 165.02 78.04 77.85 77.82 −0.13 −0.13 −0.13
0.2 166.89 169.72 169.43 22.78 22.59 22.57 −0.23 −0.23 −0.22
0.3 168.97 171.88 171.59 13.58 13.39 13.37 −0.25 −0.24 −0.24
0.4 166.21 169.02 168.72 26.68 26.49 26.47 −0.22 −0.22 −0.22
0.5 164.80 167.56 167.25 36.58 36.39 36.37 −0.20 −0.20 −0.20
Starink
0.6 160.29 162.88 162.55 93.98 93.79 93.77 −0.11 −0.11 −0.11
0.7 159.53 162.10 161.76 109.48 109.29 109.27 −0.08 −0.08 −0.08
0.8 156.22 158.67 158.31 211.08 210.89 210.87 0.09 0.08 0.08
Average 163.27 165.97 165.66 73.45 73.26 73.24 −0.14 −0.14 −0.14
0.2 167.14 169.98 169.68 21.48 21.29 21.27 −0.23 −0.23 −0.23
0.3 169.20 172.11 171.83 12.78 12.59 12.57 −0.25 −0.25 −0.24
0.4 166.45 169.26 168.96 25.28 25.09 25.07 −0.22 −0.22 −0.22
0.5 165.04 167.81 167.50 34.68 34.49 34.47 −0.21 −0.20 −0.20
Tang
0.6 160.54 163.15 162.81 89.28 89.09 89.07 −0.11 −0.11 −0.11
0.7 159.78 162.36 162.02 104.08 103.89 103.87 −0.09 −0.09 −0.09
0.8 156.48 158.93 158.58 200.68 200.49 200.47 0.07 0.06 0.06
Average 163.52 166.23 165.91 69.75 69.56 69.54 −0.15 −0.15 −0.15

(Simões et al., 2022). The low and negative value obtained con-
firmed that extended time is required for thermal decomposition
of active material in the developed briquettes, hence their burn-
ing for long durations. Moreover, these are lower, compared to
those of rice husks, which confirms that there was transforma-
tion of rice husks structure into more complex structures with
briquette formation (Kumar et al., 2020; Naqvi et al., 2019). It
should be noted that increase in heating rate showed no effect
on the ∆S values. This observation is similar to the results of
Kumar and Ramesh (2022). Moreover, variations in entropy con-
firmed existence of complex reactions occurring during thermal
conversion of the developed briquettes.

3.11. Energetic properties

The obtained high values for energetic densities


(10,688.9 MJ/m3 ) and Fuel Value Indices (255.5 MJ/m3 · %) are
attributable to the high higher heating value of the developed
briquettes (19.5 MJ/kg) (Yiga et al., 2021). Increasing energetic
properties have been reported with densification (Kipngetich
et al., 2023; de Souza et al., 2022). The high energetic densities
are also possibly due to the high carbon contents of the developed
Fig. 11. Activation energy versus conversion rates for developed briquettes. briquettes (Kipngetich et al., 2023). These are noted to facilitate
storage and transport issues (de Oliveira et al., 2022; Hu et al.,
2015)). The high energetic densities provide a greater amount
relatively high ∆G values indicated unfavorable decomposition of heat per unit volume, which is desired for domestic cooking
of the developed briquettes. purposes (de Oliveira et al., 2022; Massuque et al.). Moreover, the
Change in entropy (∆S), a measure of disorders (Simões et al., high FVI is due to the high particle density of the developed bri-
2022), was recorded as −0.14 kJ/mol.K, −0.13 kJ/mol.K, quettes (Mierzwa-Hersztek et al., 2019; Marquez-Reynoso et al.,
−0.14 kJ/mol.K and −0.15 kJ/mol.K for KAS, OFW, Starink and 2017). It should be noted that the high energetic parameters are
Tang methods respectively. A negative ∆S value depicts de- also possibly because rice husks themselves inherently have high
creased disorders of the system, i.e., ordered product formation ED and FVI (Yiga et al., 2021).
5988
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

4. Conclusions References

Adeleke, A.A., Odusote, J.K., Ikubanni, P.P., Agboola, O.O., Balogun, A.O., La-
In this study, a water boiling test and emissions monitoring
sode, O.A., 2021b. Tumbling strength and reactivity characteristics of hybrid
system for CO, CO2 and PM2.5 were used to determine emis- fuel briquette of coal and biomass wastes blends. Alex. Eng. J. 60 (5),
sions released over time while thermogravimetric analysis was 4619–4625.
utilized in obtaining combustion, kinetic and thermodynamic Adeleke, A., Odusote, J., Ikubanni, P., Lasode, O., Malathi, M., Pasawan, D., 2021a.
Physical and mechanical characteristics of composite briquette from coal and
parameters. The high drop strength (67.8%) and particle density pretreated wood fines. Int. J. Coal Sci. Technol. 8 (5), 1088–1098.
(548.15 kg/m3 ) imply that it is easy to store and transport the Afsal, A., David, R., Baiju, V., Suhail, N.M., Parvathy, U., Rakhi, R.B., 2020.
developed briquettes. Moisture, volatile matter, ash content and Experimental investigations on combustion characteristics of fuel briquettes
fixed carbon were 7.2%, 16.6%, 47.4% and 28.8% respectively. Igni- made from vegetable market waste and saw dust. Mater. Today: Proc. 33,
3826–3831.
tion, peak and burnout temperatures increased with increase in Akahira, T., 1971. Trans. Joint convention of four electrical institutes. Res. Rep.
heating rates. Average maximum and average combustion rates of Chiba Inst. Technol. 16, 22–31.
the developed briquettes were around 0.07%/min and 0.02%/min. Amarasekara, A., Tanzim, F.S., Asmatulu, E., 2017. Briquetting and carbonization
Average flammability index, combustion characteristic index val- of naturally grown algae biomass for low-cost fuel and activated carbon
production. Fuel 208, 612–617.
ues and mean reactivities were 1.37 × 10−6 %/min· ◦ C2 and Anggraeni, S., Hofifah, S.N., Nandiyanto, A.B.D., Bilad, M.R., 2021. Effects of
0.73 × 10−10 %2 /min 2 · ◦ C3 and 0.18 × 10−3 %/min/◦ C respectively. particle size and composition of cassava peels and rice husk on the briquette
Average activation energy attained by the developed briquettes performance. J. Eng. Sci. Technol. 16 (1), 527–542.
by Kissinger–Akahira–Sunose, Ozawa–Flynn–Wall, Starink and Arewa, M.E., Daniel, I.C., Kuye, A., 2016. Characterisation and comparison of rice
husk briquettes with cassava peels and cassava starch as binders. Biofuels 7
Tang methods were 78.3 kJ/mol, 83.3 kJ/mol, 78.7 kJ/mol and (6), 671–675.
75.0 kJ/mol, respectively. The low energy barrier (5.2 kJ/mol) Arora, P., Das, P., Jain, S., Kishore, V.V.N., 2014. A laboratory based comparative
between activation energy and enthalpy changes indicated that study of Indian biomass cookstove testing protocol and water boiling test.
initiation occurs easily. The obtained thermodynamic parameters Energy Sustain. Dev. 21, 81–88.
ASTM, 2014. Standard test method for compositional analysis by thermogravime-
indicated that the thermal degradation process was endothermic try. In: ASTM Standards. ASTM International, West Conshohocken, http:
in nature. High energetic densities (10,688.9 MJ/m3 ) and Fuel //dx.doi.org/10.1520/E1131-08R14.
Value Indices (255.5 MJ/m3 · %) were attributable to the high Buratti, C., Mousavi, S., Barbanera, M., Lascaro, E., Cotana, F., Bufacchi, M.,
2016. Thermal behaviour and kinetic study of the olive oil production chain
higher heating value of the developed briquettes (19.5 MJ/kg).
residues and their mixtures during co-combustion. Bioresour. Technol. 214,
Further works are recommended optimization of the production 266–275.
process of carbonized rice husks briquettes as well as their life Buyondo, K.A., Olupot, P.W., Kirabira, J.B., Yusuf, A.A., 2020. Optimization of
cycle assessment (LCA) in order to assess their total impact on production parameters for rice husk ash-based geopolymer cement using
response surface methodology. Case Stud. Constr. Mater. 13, e00461.
the environment.
Cao, J.P., Shi, P., Zhao, X.Y., Wei, X.Y., Takarada, T., 2014. Catalytic reforming
of volatiles and nitrogen compounds from sewage sludge pyrolysis to clean
CRediT authorship contribution statement hydrogen and synthetic gas over a nickel catalyst. Fuel Process. Technol. 123,
34–40.
Carnaje, N.P., Talagon, R.B., Peralta, J.P., Shah, K., Paz-Ferreiro, J., 2018. Devel-
Vianney Andrew Yiga: Conceptualization, Methodology, Ex- opment and characterisation of charcoal briquettes from water hyacinth
perimental design, Formal analysis, Investigation, Writing– origi- (Eichhornia crassipes)-molasses blend. PLoS One 13 (11), e0207135.
Cartland, R., Sendegeya, A.M., Hakizimana, J.D.D.K., 2022. Socio-economic analysis
nal draft, Writing – review & editing, Supervision. Andrew Nuwa- of solar photovoltaic-based mini-grids in rural communities: A Ugandan case
manya: Methodology, Experimental design, Writing– original study. J. Energy South. Afr. 33 (3), 36–50.
draft. Agatha Birungi: Methodology, Experimental design, Writing– Chandrasekaran, A., Ramachandran, S., Subbiah, S., 2017. Determination of kinetic
original draft. Michael Lubwama: Conceptualization, Methodol- parameters in the pyrolysis operation and thermal behavior of Prosopis
juliflora using thermogravimetric analysis. Bioresour. Technol. 233, 413–422.
ogy, Experimental design, Formal analysis, Investigation, Writing– Chang, S.H., 2020. Rice husk and its pretreatments for bio-oil production via fast
original draft, Writing – review & editing, Resources, Supervision. pyrolysis: A review. BioEnergy Res. 13 (1), 23–42.
Harriet Nalubega Lubwama: Methodology, Writing – original Chavan, D., Manjunatha, G.S., Singh, D., Periyaswami, L., Kumar, S., Kumar, R.,
draft, Writing – review & editing. 2022. Estimation of spontaneous waste ignition time for prevention and
control of landfill fire. Waste Manage. 139, 258–268.
Chen, C., Qu, B., Wang, W., Wang, W., Ji, G., Li, A., 2021. Rice husk and rice straw
Declaration of competing interest torrefaction: Properties and pyrolysis kinetics of raw and torrefied biomass.
Environ. Technol. Innov. 24, 101872.
Clean Cooking Alliance (CCA), 2014. The Water Boiling Test Version 4.2.3. 1750
The authors declare the following financial interests/personal Pennsylvania Ave NW, Suite 300 Washington, D.C, 20006.
relationships which may be considered as potential competing Dao, C.N., Salam, A., Oanh, N.T.K., Tabil, L.G., 2022. Effects of length-to-diameter
ratio, pinewood sawdust, and sodium lignosulfonate on quality of rice straw
interests: Agatha Birungi reports fnancial support was provided
pellets produced via a flat die pellet mill. Renew. Energy 181, 1140–1154.
by Yosevi Engineering Services. Das, P., Tiwari, P., 2017. Thermal degradation kinetics of plastics and model
selection. Thermochim. Acta 654, 191–202.
Deng, M., Li, P., Shan, M., Yang, X., 2019. Characterizing dynamic relationships
Data availability between burning rate and pollutant emission rates in a forced-draft gasifier
stove consuming biomass pellet fuels. Environ. Pollut. 255, 113338.
Data will be made available on request Deshannavar, U.B., Hegde, P.G., Dhalayat, Z., Patil, V., Gavas, S., 2018. Production
and characterization of agro-based briquettes and estimation of calorific
value by regression analysis: An energy application. Mater. Sci. Energy
Acknowledgments Technol. 1 (2), 175–181.
Doyle, C.D., 1965. Series approximations to the equation of thermogravimetric
data. Nature 207 (4994), 290–291.
This work was financially supported by Yosevi Engineering Elinge, C.M., Birnin-Yauri, A.U., Senchi, D.S., Ige, A.R., Ajakaye, J., Yusuf, A.,
Services Limited, Uganda, www.yosevi.com. Technical support Abubakar, R.K., 2019. Studies on the combustion profile of briquettes
produced from carbonized rice husk using different binders at moderate
from Mr. Andrew Wabwire from the Department of Mechanical
temperature and die pressure. Studies 5 (3).
Engineering, Makerere University, Uganda is gratefully acknowl- Fengmin, L., Mingquan, Z., 2011. Technological parameters of biomass briquetting
edged. of macrophytes in Nansi Lake. Energy Procedia 5, 2449–2454.

5989
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Gan, D.K.W., Loy, A.C.M., Chin, B.L.F., Yusup, S., Unrean, P., Rianawati, E., Liu, H., J, E., Ma, X., Xie, C., 2016. Influence of microwave drying on
Acda, M.N., 2018. Kinetics and thermodynamic analysis in one-pot pyrolysis the combustion characteristics of food waste. Drying Technol. 34 (12),
of rice hull using renewable calcium oxide based catalysts. Bioresour. 1397–1405.
Technol. 265, 180–190. Liu, J., Jiang, X., Cai, H., Gao, F., 2021. Study of combustion characteristics
Gao, Q., Zhang, T., Feng, Z., Yang, J., Ni, L., Hu, W., Liu, Z., 2022. Energy and kinetics of agriculture briquette using thermogravimetric analysis. ACS
performances of molded charcoals from bamboo and Chinese fir blends: Omega 6, 15827–15833.
Influence of pyrolysis temperatures and residence times. Ind. Crop. Prod. Loy, A.C.M., Yusup, S., How, B.S., Yiin, C.L., Chin, B.L.F., Muhammad, M., Gwee, Y.L.,
177, 114500. 2019. Uncertainty estimation approach in catalytic fast pyrolysis of rice
Gianvenuti, A., Bedijo, N.G., Jalal, R., Hitimana, L., Walter, S., Linhares-Juvenal, T., husk: Thermal degradation, kinetic and thermodynamic parameters study.
Xia, Z., 2022. Woodfuel consumption in Refugee Hosting Areas and its impact Bioresour. Technol. 294, 122089.
on the surrounding forests—The case of Uganda. Forests 13 (10), 1676. Lubwama, M., Yiga, V.A., 2017. Development of groundnut shells and bagasse
Gioda, A., 2019. Residential fuelwood consumption in Brazil: Environmental and briquettes as sustainable fuel sources for domestic cooking applications in
social implications. Biomass Bioenergy 120, 367–375. Uganda. Renew. Energy 111, 532–542.
Granado, M.P.P., Suhogusoff, Y.V.M., Santos, L.R.O., Yamaji, F.M., De Conti, A.C., Lubwama, M., Yiga, V.A., 2018. Characteristics of briquettes developed from rice
2021. Effects of pressure densification on strength and properties of cassava
and coffee husks for domestic cooking applications in Uganda. Renew. Energy
waste briquettes. Renew. Energy 167, 306–312.
118, 43–55.
Guo, Z., Wu, J., Zhang, Y., Wang, F., Guo, Y., Chen, K., Liu, H., 2020. Characteristics
Lubwama, M., Yiga, V.A., Lubwama, H.N., 2020a. Effects and interactions of
of biomass charcoal briquettes and pollutant emission reduction for sulfur
the agricultural waste residues and binder type on physical properties and
and nitrogen during combustion. Fuel 272, 117632.
calorific values of carbonized briquettes. Biomass Convers. Biorefinery 1–21.
Hu, M., Deng, W., Hu, M., Chen, G., Zhou, P., Zhou, Y., Su, Y., 2021. Preparation
Lubwama, M., Yiga, V.A., Muhairwe, F., Kihedu, J., 2020b. Physical and combus-
of binder-less activated char briquettes from pyrolysis of sewage sludge for
tion properties of agricultural residue bio-char bio-composite briquettes as
liquid-phase adsorption of methylene blue. J. Environ. Manag. 299, 113601.
sustainable domestic energy sources. Renew. Energy 148, 1002–1016.
Hu, Q., Shao, J., Yang, H., Yao, D., Wang, X., Chen, H., 2015. Effects of binders on
Lutaaya, S.M., Olupot, P.W., Wakatuntu, J., Kasedde, H., 2023. Effects of waste pa-
the properties of bio-char pellets. Appl. Energy 157, 508–516.
per on fuel and mechanical properties of biogas digestate-derived briquettes.
Javed, M.H., Sikandar, M.A., Ahmad, W., Bashir, M.T., Alrowais, R., Wadud, M.B.,
2022. Effect of various biochars on physical, mechanical, and microstructural Biomass Convers. Biorefinery 1–17.
characteristics of cement pastes and mortars. J. Build. Eng. 57, 104850. Marquez-Reynoso, M.I., Ramírez-Marcial, N., Cortina-Villar, S., Ochoa-Gaona, S.,
Jia, C., Chen, J., Liang, J., Song, S., Liu, K., Jiang, A., Wang, Q., 2020. Pyrolysis 2017. Purpose, preferences and fuel value index of trees used for firewood
characteristics and kinetic analysis of rice husk. J. Therm. Anal. Calorim. 139, in El Ocote Biosphere Reserve, Chiapas, Mexico. Biomass Bioenergy 100, 1–9.
577–587. Massuque, J., De Assis, M.R., Trugilho, P.F., Characterization of miombo species
Kali, A., Amar, A., Loulidi, I., Hadey, C., Jabri, M., Alrashdi, A.A., Boukhlifi, F., et al., used by rural communities as fuelwood in northern mozambique. Energy
2022. Efficient adsorption removal of an anionic azo dye by lignocellulosic Sources A 1–10.
waste material and sludge recycling into combustible briquettes. Colloid. Mierzwa-Hersztek, M., Gondek, K., Jewiarz, M., Dziedzic, K., 2019. Assessment of
Interface. 6 (2), 22. energy parameters of biomass and biochars, leachability of heavy metals and
Kaur, R., Gera, P., Jha, M.K., Bhaskar, T., 2018. Pyrolysis kinetics and phytotoxicity of their ashes. J. Mater. Cycle. Waste Manag. 21 (4), 786–800.
thermodynamic parameters of castor (Ricinus communis) residue using Miito, G.J., Banadda, N., 2016. Waste to energy technologies for solid waste
thermogravimetric analysis. Bioresour. Technol. 250, 422–428. management a case study of Uganda. Agric. Eng. Int. CIGR J. 18 (3), 136–146.
Kipngetich, P., Kiplimo, R., Tanui, J.K., Chisale, P.C., 2022. Optimization of Mitchell, E.J.S., Lea-Langton, A.R., Jones, J.M., Williams, A., Layden, P., Johnson, R.,
combustion parameters of carbonized rice husk briquettes in a fixed bed 2016. The impact of fuel properties on the emissions from the combustion
using RSM technique. Renew. Energy 198, 61–74. of biomass and other solid fuels in a fixed bed domestic stove. Fuel Process.
Kipngetich, P., Kiplimo, R., Tanui, J.K., Chisale, P., 2023. Effects of carbonization on Technol. 142, 115–123.
the combustion of rice husks briquettes in a fixed bed. Clean. Eng. Technol. Montiano, M.G., Fernández, A., Díaz-Faes, E., Barriocanal, C., 2015. Tar from
13, 100608. biomass/coal-containing briquettes. Evaluation of PAHs. Fuel 154, 261–267.
Kivumbi, B., Jande, Y.A., Kirabira, J.B., Kivevele, T.T., 2021a. Water boiling test Mudoko, S.N., 2013. Uganda’s Policy on Energy and Power. ICA Training on
of carbonized briquettes produced from charcoal fines using African Elemi Energy Policy, Tokyo, Japan, p. 2.
(Canarium schweinfurthii) resin as an organic binder. Biomass Convers. Nakibuuka, A.K., 2022. Life Cycle Assessment of Rice Husk Briquette Production
Biorefinery 1–16. in Uganda (Doctoral dissertation). Makerere University.
Kivumbi, B., Jande, Y.A., Kirabira, J.B., Kivevele, T.T., 2021b. Production of Nakiyinji, J., 2022. Development and Characterization of Composite Briquettes
carbonized briquettes from charcoal fines using African elemi (Canarium for Domestic Cooking Application (Doctoral dissertation).
schweinfurthii) resin as an organic binder. Energy Sources A 1–17. Naqvi, S.R., Hameed, Z., Tariq, R., Taqvi, S.A., Ali, I., Niazi, M.B.K., Shahbaz, M., et
Kpalo, S.Y., Zainuddin, M.F., Abd Manaf, L., Roslan, A.M., 2021. Evaluation of al., 2019. Synergistic effect on co-pyrolysis of rice husk and sewage sludge by
hybrid briquettes from corncob and oil palm trunk bark in a domestic
thermal behavior, kinetics, thermodynamic parameters and artificial neural
cooking application for rural communities in Nigeria. J. Clean. Prod. 284,
network. Waste Manage. 85, 131–140.
124745.
Ni, L., Feng, Z., Zhang, T., Gao, Q., Hou, Y., He, Y., Liu, Z., et al., 2022. Effect
Kumar, M., Mishra, P.K., Upadhyay, S.N., 2020. Thermal degradation of rice husk:
of pyrolysis heating rates on fuel properties of molded charcoal: Imitating
Effect of pre-treatment on kinetic and thermodynamic parameters. Fuel 268,
industrial pyrolysis process. Renew. Energy 197, 257–267.
117164.
Nwabue, F.I., Unah, U., Itumoh, E.J., 2017. Production and characterization of
Kumar, T.T.A., Ramesh, S.K.T., 2022. Thermal decomposition kinetics of Prosopis
smokeless bio-coal briquettes incorporating plastic waste materials. Environ.
juliflora charcoal briquette using thermogravimetric analysis. Environ. Sci.
Technol. Innov. 8, 233–245.
Pollut. Res. 1–16.
Obeng, G.Y., Mensah, E., Ashiagbor, G., Boahen, O., Sweeney, D.J., 2017. Watching
Kuthe, N.V., Ingle, P.B., Gore, V.G., 2021. Biomass briquettes as an alternative
energy source compare to wood charcoal in boilers. Int. J. Sci. Res. Mech. the smoke rise up: Thermal efficiency, pollutant emissions and global
Mater. Eng. 5 (4), 16–40. warming impact of three biomass cookstoves in Ghana. Energies 10 (5), 641.
Lask, K., Booker, K., Han, T., Granderson, J., Yang, N., Ceballos, C., Gadgil, A., 2015. Obi, O.F., Okongwu, K.C., 2016. Characterization of fuel briquettes made from a
Performance comparison of charcoal cookstoves for Haiti: Laboratory testing blend of rice husk and palm oil mill sludge. Biomass Convers. Biorefinery 6,
with water boiling and controlled cooking tests. Energy Sustain. Dev. 26, 449–456.
79–86. Ogwang, I., Kasedde, H., Nabuuma, B., Kirabira, J.B., Lwanyaga, J.D., 2021. Kased
Li, J., Qiao, Y., Zong, P., Wang, C., Tian, Y., Qin, S., 2019. Thermogravimetric characterization of biogas digestate for solid biofuel production in Uganda.
analysis and isoconversional kinetic study of biomass pyrolysis derived from Sci. Afr. 12, e00735.
land, coastal zone, and marine. Energy Fuels 33 (4), 3299–3310. Ojha, D.K., Kumar, V.S.P., Vinu, R., 2021. Analytical pyrolysis of bagasse and
Li, Y., Xing, X., Xu, B., Xing, Y., Zhang, X., Yang, J., Xing, J., 2017. Effect of groundnut shell briquettes: Kinetics and pyrolysate composition studies.
the particle size on co-combustion of municipal solid waste and biomass Bioresour. Technol. Rep. 15, 100784.
briquette under N2/O2 and CO2/O2 atmospheres. Energy Fuels 31 (1), de Oliveira, P.R.S., Trugilho, P.F., de Oliveira, T.J.P., 2022. Briquettes of acai
932–940. seeds: Characterization of the biomass and influence of the parameters of
Lim, S.S., Vos, T., Flaxman, A.D., Danaei, G., Shibuya, K., Adair-Rohani, H., production temperature and pressure in the physical-mechanical and energy
Aryee, M., et al., 2012. A comparative risk assessment of burden of disease quality. Environ. Sci. Pollut. Res. 29 (6), 8549–8558.
and injury attributable to 67 risk factors and risk factor clusters in 21 regions, Oluwafemi‘Femi’Mimiko, N., 2022. Sino-African relations and trends for the post-
1990–2010: A systematic analysis for the global burden of disease study Covid-19 global order. In: The Palgrave HandBook of Africa and the Changing
2010. Lancet 380 (9859), 2224–2260. Global Order. Palgrave Macmillan, Cham, pp. 649–671.

5990
V.A. Yiga, A. Nuwamanya, A. Birungi et al. Energy Reports 9 (2023) 5977–5991

Ozawa, T., 1965. A new method of analyzing thermogravimetric data. Bull. Chem. Suryaningsih, S., Nurhilal, O., Yuliah, Y., Salsabila, E., 2018. Fabrication and
Soc. Japan 38 (11), 1881–1886. characterization of rice husk charcoal bio briquettes. AIP Conf. Proc. 1927
Pinzi, S., Buratti, C., Bartocci, P., Marseglia, G., Fantozzi, F., Barbanera, M., 2020. (1), 030044.
A simplified method for kinetic modeling of coffee silver skin pyrolysis Suryaningsih, S., Resitasari, R., Nurhilal, O., 2019a. Analysis of biomass briquettes
by coupling pseudo-components peaks deconvolution analysis and model based on carbonized rice husk and jatropha seed waste by using newspaper
free-isoconversional methods. Fuel 278, 118260. waste pulp as an adhesive material. J. Phys. Conf. Ser. 1280 (2), 022072.
Quan, C., Li, A., Gao, N., 2009. Thermogravimetric analysis and kinetic study Szabó, S., Moner-Girona, M., Kougias, I., Bailis, R., Bódis, K., 2016. Identification of
on large particles of printed circuit board wastes. Waste Manage. 29 (8), advantageous electricity generation options in sub-Saharan Africa integrating
2353–2360. existing resources. Nature Energy 1 (10), 1–8.
Raymer, A.K.P., 2006. A comparison of avoided greenhouse gas emissions when Tang, W., Liu, Y., Zhang, H., Wang, C., 2003. New approximate formula for
using different kinds of wood energy. Biomass Bioenergy 30 (7), 605–617. arrhenius temperature integral. Thermochim. Acta 408 (1–2), 39–43.
Rehfuess, E., World Health Organization, 2006. Fuel for Life: Household Energy Thao, P.T.M., Kurisu, K.H., Hanaki, K., 2012. Evaluation of strategies for utilizing
and Health. rice husk based on life cycle cost analysis in relation to greenhouse gas
Saha, D., Sinha, A., Pattanayak, S., Roy, B., 2022. Pyrolysis kinetics and ther- emissions in an giang province, Vietnam. Biomass Bioenergy 37, 122–131.
modynamic parameters of plastic grocery bag based on thermogravimetric Tumutegyereize, P., Mugenyi, R., Ketlogetswe, C., Gandure, J., 2016. A compar-
data using iso-conversional methods. Int. J. Environ. Sci. Technol. 19 (1), ative performance analysis of carbonized briquettes and charcoal fuels in
391–406. Kampala-urban, Uganda. Energy Sustain. Dev. 31, 91–96.
Sana, A., Meda, N., Badoum, G., Kafando, B., Bouland, C., 2019. Primary cooking Uganda Bureau of Statistics (UBOS), 2020. Uganda Wood & Forest Resources
fuel choice and respiratory health outcomes among women in charge of Accounts. Kampala, Uganda.
household cooking in ouagadougou, Burkina faso: Cross-sectional study. Int. World Health Organization WHO, 2014. WHO Guidelines for Indoor Air Quality:
J. Environ. Res. Public Health 16 (6), 1040. Household Fuel Combustion.
Setter, C., Oliveira, T.J.P., 2022. Evaluation of the physical-mechanical and energy Xiao, R., Yang, W., Cong, X., Dong, K., Xu, J., Wang, D., Yang, X., 2020.
properties of coffee husk briquettes with kraft lignin during slow pyrolysis. Thermogravimetric analysis and reaction kinetics of lignocellulosic biomass
Renew. Energy 189, 1007–1019. pyrolysis. Energy 201, 117537.
Setter, C., Silva, F.T.M., Assis, M.R., Ataíde, C.H., Trugilho, P.F., Oliveira, T.J.P., 2020. Xiu, M., Stevanovic, S., Rahman, M.M., Pourkhesalian, A.M., Morawska, L.,
Slow pyrolysis of coffee husk briquettes: Characterization of the solid and Thai, P.K., 2018. Emissions of particulate matter, carbon monoxide and
liquid fractions. Fuel 261, 116420. nitrogen oxides from the residential burning of waste paper briquettes and
Simões, L.M.S.A., Setter, C., Sousa, N.G., Cardoso, C.R., de Oliveira, T. J.P., other fuels. Environ. Res. 167, 536–543.
2022. Biomass to biofuel densification of coconut fibers: Kinetic triplet and Yadav, I.C., Devi, N.L., 2018. Biomass burning, regional air quality, and climate
thermodynamic evaluation. Biomass Convers. Biorefinery 1–18. change. In: Earth Systems and Environmental Sciences. In: Encyclope-
Song, A., Zha, F., Tang, X., Chang, Y., 2019. Effect of the additives on combustion dia of Environmental Health, Elsevier, http://dx.doi.org/10.1016/B978-0-12-
characteristics and desulfurization performance of cow dung briquette. 409548-9.11022-X.
Chem. Eng. Process. Process Intensif. 143, 107585. Yakaboylu, G.A., Jiang, C., Yumak, T., Zondlo, J.W., Wang, J., Sabolsky, E.M.,
de Souza, E.C., Gomes, J.P.S., Pimenta, A.S., de Azevedo, T.K.B., Pereira, A.K.S., 2021. Engineered hierarchical porous carbons for supercapacitor applications
Gomes, R.M., Dias Júnior, A.F., et al., 2022. Briquette production as a sus- through chemical pretreatment and activation of biomass precursors. Renew.
tainable alternative for waste management in the tannin extraction industry. Energy 163, 276–287.
Environ. Sci. Pollut. Res. 1–13. Yiga, V.A., Lubwama, M., 2020. Thermogravimetric analysis of agricultural residue
Starink, M.J., 2003. The determination of activation energy from linear heating carbonized briquettes for domestic and industrial applications. MRS Adv. 5
rate experiments: A comparison of the accuracy of isoconversion methods. (20), 1039–1048.
Thermochim. Acta 404 (1–2), 163–176. Yiga, V.A., Lubwama, M., Olupot, P.W., 2021. Characterization of rice husks as
Still, D., MacCarty, N., Ogle, D., Bond, T., Bryden, M., 2011. Test Results of potential reinforcement for polymer composites. Mater. Circ. Econ. 3 (1), 1–8.
Cook Stove Performance. Aprovecho Resarch Center, Shell Foundation, United Yiga, V.A., Lubwama, M., Olupot, P.W., 2022. Thermal stability of unmodified
States Environmental Protection Agency, p. 126. and alkali-modified rice husks for flame retardant fiber-reinforced PLA
Suryaningsih, S., Nurhilal, O., Widyarini, R.A., Suhendi, N., 2019b. The analysis composites. J. Therm. Anal. Calorim. 1–27.
of ignition and combustion properties of the burning briquettes made from Yiga, V.A., Lubwama, M., Olupot, P.W., 2023. Pyrolysis, kinetics and thermody-
mixed biomass of rice husk and corn cob. IOP Conf. Ser. Mater. Sci. Eng. 550 namic analyses of rice husks/clay fiber-reinforced polylactic acid composites
(1), 012006. using thermogravimetric analysis. J. Therm. Anal. Calorim. 1–20.
Suryaningsih, S., Nurhilal, O., Yuliah, Y., Mulyana, C., 2017. Combustion quality Zhong, Q., Yang, Y., Li, Q., Xu, B., Jiang, T., 2017. Coal tar pitch and molasses
analysis of briquettes from variety of agricultural waste as source of blended binder for production of formed coal briquettes from high volatile
alternative fuels. IOP Conf. Ser. Earth Environ. Sci. 65 (1), 012012. coal. Fuel Process. Technol. 157, 12–19.

5991

You might also like