Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Cleaner Production 372 (2022) 133744

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Charcoal briquette production from waste in the coffee production process


using hydrothermal and torrefaction techniques: A comparative study with
carbonization technique
Siriwan Srisang a, Kittisak Phetpan a, Nuttapong Ruttanadech a, Warunee Limmun a,
Pannipa Youryon b, Pornprapa Kongtragoul b, Naruebodee Srisang a, *,
Thatchapol Chungcharoen a
a
Department of Engineering, King Mongkut’s Institute of Technology Ladkrabang, Prince of Chumphon Campus, Chumphon, 86160, Thailand
b
Department of Agricultural Technology, King Mongkut’s Institute of Technology Ladkrabang, Prince of Chumphon Campus, Chumphon, 86160, Thailand

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

Handling Editor: Zhen Leng This research aimed to study charcoal briquette production from coffee production waste, i.e., coffee parchment
(CP) and coffee cherry pulp (CCP). The carbonization technique (CT) was studied in five mixtures (CP and CCP ≈
Keywords: 0–90%) and three pressures (1000–1600 psi) to determine the appropriate conditions. The torrefaction technique
Charcoal briquette (TT) and hydrothermal technique (HT) were subsequently performed under proper conditions from the CT to
Coffee production waste
investigate further the fit temperatures (200–260 ◦ C) and reaction times (40–120 min). The fuel characteristics
Carbonization technique
were examined regarding the calorific value (CV), proximate and ultimate analyses, mechanical properties, and
Torrefaction technique
Hydrothermal technique utilization properties. The results demonstrated the notable influence of interaction between the mixture and
Fuel characteristics pressure factors and individual mixture on the fuel properties, whereas personal pressure have an insignificant
effect. The ratio of CP ≈ 90% and binder ≈10% at a pressure of 1600 psi in the CT prepared appropriate fuel
properties (calorific value ≈ 27 MJ/kg, fixed carbon content ≈ 65%). Interestingly, almost all conditions of the
TT and HT provided greenhouse gas emissions lower than the CT. The TT and HT at 260 ◦ C for 120 min provided
high calorific values (25–26 MJ/kg) with other fuel characteristics in the acceptable standard, except for the
fixed carbon content.

1. Introduction technique (CT), torrefaction technique (TT), and hydrothermal tech­


nique (HT) are used to produce biofuels under appropriate conditions to
Coffee is popularly consumed as a beverage because of its unique obtain desirable fuel properties. The CT is simply a thermochemical
sensory flavor and health advantages (diabetes and cancer prevention). conversion of biomass waste into charcoal under an operating temper­
Coffee production vastly increases to respond the consumer demand, ature of 300–900 ◦ C. Aransiola et al. (2019) exhibited that the suitable
which causes large amounts of residues. These wastes are disposed in binder type, binder content, and compression pressure conditions in the
landfills, resulting in high greenhouse gas (GHG) emissions. Wet coffee CT provided the low moisture content (MC) and high compressive
bean manufacturing is widely used and causes enormous residues such strength in fuel briquette. Lubwama et al. (2020) showed that the
as coffee cherry pulp (CCP), coffee parchment (CP), and defective coffee carbonization process aided in reducing the MC and volatile matter
beans (undesirable size and ochratoxin A contamination) after produc­ content (VM), which resulted in an increase in CV and fixed carbon
tion. Biomass waste reduction via recycling is a sustainable method that content (FC). The feedstock type and quantity in the CT should be
can create various products, such as tiles (Srisang and Srisang, 2021) considered because they may cause high MC and ash content (AC) in the
and biomass fuel (Chungcharoen and Srisang, 2020). fuel. Haykiri-Acma et al. (2013) observed that the fuel qualities (FC and
Coffee residues produce biomass fuel with a high calorific value (CV) CV) were improved from the CT compared with raw material; none­
(CP pellet ≈ 20.21 MJ/kg) (de Souza et al., 2020). The carbonization theless, the seaweed charcoal still had a high AC, and briquette

* Corresponding author.
E-mail address: naruebodee.sr@kmitl.ac.th (N. Srisang).

https://doi.org/10.1016/j.jclepro.2022.133744
Received 11 March 2022; Received in revised form 12 August 2022; Accepted 19 August 2022
Available online 24 August 2022
0959-6526/© 2022 Elsevier Ltd. All rights reserved.
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

formation required a high pressure for compression (2.71 × 104 psi). Table 1
The TT or dry torrefaction technique is a mild pyrolysis method Advantages, disadvantages, and research gaps from the CT, TT, and HT litera­
(200–300 ◦ C) with an appropriate reaction time, and it is more stable ture review.
than conventional pyrolysis (Lin et al., 2021). The TT prepares the Thermochemical Advantages Disadvantages Research gaps
abated hygroscopicity and additional grindability (Barskov et al., 2019). technique
Kai et al. (2019) stated that the temperature in the TT had a greater CT • The CT is a • The CT requires • Most of the
effect on the fuel quality than the reaction time. Lin et al. (2021) simple a high coffee
observed that the FC and AC in the biofuel raised with increasing tem­ technique and temperature for production
can be used combustion waste is
perature and could reduce GHG emissions when burned together with
with many (900 ◦ C) transformed via
coal. Tu et al. (2018) reported that the TT had a higher pellet strength biomass types (Haykiri-Acma the
than the HT, and the ameliorated pellet’s CV and FC when the enhanced such as et al., 2013) and conventional
temperature in the TT and HT (220–300 ◦ C). Wu et al. (2018) demon­ corncobs a high pressure CT into
strated that the biomass briquette from the TT had poor density and (Aransiola for briquetting alternative fuel
et al., 2019), (3.34 × 104 psi) for household
strength owing to diminished hydrogen bond construction between the seaweed (Lubwama and cooking
polymer chains. (Haykiri-Acma Yiga, 2018), (Lubwama and
The HT or wet torrefaction combines dry torrefaction and water et al., 2013), which depend Yiga, 2018;
leaching. The HT is performed at a modest temperature (120–300 ◦ C) rice husks, on biomass Lubwama et al.,
coffee husks, type. 2020);
and pressure (2–10 MPa) with a short reaction time (5–60 min) (Zhuang
and groundnut • Most of the fuel therefore, the
et al., 2020). Notably, wet biomass waste can be treated using the HT to shells utilization from TT and HT
obtain hydrochar with low AC and sulfur content (Lee et al., 2018). Wu (Lubwama CT is limitedly should be
et al. (2018) provided that charcoal briquette via the HT received better et al., 2020). used in applied to
density and strength than the TT and corresponded with the commercial • The CT households investigate the
provides owing to a high feasibility of the
barbecue charcoal. Song et al. (2020) demonstrated that different
satisfactory fuel AC (barbecue fuel quality
feedstocks caused physical and chemical distinctions after the HT pro­ characteristics charcoal ≈ upgrade and
cess, resulting in fuel quality improvements (CV, FC, and combustion (low MC, high 12.8%) (Wu result in
characteristics). Yu et al. (2018) produced hydrochar from sewage FC, and low et al., 2018). worldwide
VM) for • The GHG usage.
sludge using the HT at 160–240 ◦ C for 60 min, which increased the
domestic emission of • Factors of
heating value and AC with increasing temperature. cooking wood charcoal feedstock type
Table 1 summarizes and compares the strengths and weaknesses of (Lubwama and (460 gCO2e) and quantity,
the CT, TT, and HT, which lead to the research gap. Yiga, 2018). from traditional compression
The above research indicated that three techniques have the poten­ CT is high pressure,
(Chungcharoen reaction time,
tial to produce alternative fuel briquettes from coffee production waste,
and Srisang, and
which lead to clean and sustained fuel production. Therefore, this 2020). temperature
research investigated the influence of mixture and compression pressure TT • The TT can • The TT gives a significantly
factors in the CT on charcoal briquette characteristics (proximate anal­ reduce poor strength of affect the fuel
hygroscopicity fuel owing to characteristics;
ysis, CV, and hardness) to determine suitable conditions. These condi­
and add a the hydrogen hence, these
tions were used in the TT and HT to develop the fuel qualities with the grindability bond abatement factors should
proximate and ultimate analyses, density, porosity, thermal efficiency (Barskov et al., (Wu et al., be determined
(ηth ), flame temperature (FT), and GHG emissions. The briquette char­ 2019) and a CV 2018). to provide the
acteristics of the three techniques were compared. to the fuel (Lin • The TT is proper
et al., 2021) complex conditions for
• The TT because the the desired fuel
2. Materials and methods improves the temperature production.
coalification and reaction • Limited
2.1. Feedstocks preparation degree of fuel time must be information is
to approach intensively available for
that of coal, controlled for each
The CP and CCP are residues from Robusta coffee production via wet which causes the required technique’s
processing obtained from Small and Medium Enterprises in Chumphon the GHG fuel investigation
Province. The CP and CCP had an initial moisture content of approxi­ emissions characteristics and comparison
mately 10% and 13% (wet basis, w.b.), respectively. Each feedstock was reduction for preparation (Kai of fuel
use in power et al., 2019). utilization and
transformed into charcoal using a conventional carbonization process at
plants (Lin GHG emission.
500 ◦ C, which had the preliminary properties as shown in Table 2. et al., 2021).
• The TT uses a
2.2. Charcoal briquettes preparation low
temperature
(200–300 ◦ C)
The CP and CCP were transformed into charcoal using a thermo­ (Wu et al.,
chemical process with three different techniques to study the fuel 2018; Lin et al.,
properties. The briquette formation used a binder, which was tapioca 2021) for the
flour (TF). The TF was homogeneously dissolved in water at 10% of the reaction, which
aids in
total mass because it was the appropriate amount, as Chungcharoen and economizing
Srisang (2020) reported. energy.
HT • The wet • The HT is more
2.2.1. Carbonization technique (CT) feedstock can complex than
be executed via the TT because
The CP and CCP had a size range of 0.5–1.5 cm and were placed on a
the HT to it combines TT
screen (60 × 60 cm2) in a heating furnace (Cerawan Supply Company), obtain and water
setting the heating rate at 10 ◦ C/min to obtain a carbonization tem­ (continued on next page)
perature of 500 ◦ C. The CP and CCP were carbonized for 2 h into

2
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Table 1 (continued )
Thermochemical Advantages Disadvantages Research gaps
technique

hydrochar with leaching


low AC and (Zhuang et al.,
sulfur (Lee 2020).
et al., 2018). • Several
• The HT offers parameters such
more fuel as feedstock
strength under type (Song
the same et al., 2020),
temperature temperature,
Fig. 2. Charcoals after the grinding process: A) CP and B) CCP.
and reaction and reaction
time as the TT time (Yu et al.,
(Wu et al., 2018) in HT
2018). strongly relate
to the required
fuel qualities,
which results in
an arduous
operation.

Table 2
Ultimate analysis of CP and CCP charcoals.
Charcoal Carbon Hydrogen Nitrogen Sulfur Oxygen
(%) (%) (%) (%) (%)

CP 50.63 ± 5.97 ± 0.93 ± 0.06 ± 40.40 ±


0.03b 0.01b 0.06a 0.00a 0.12b
CCP 49.96 ± 5.46 ± 1.52 ± 0.13 ± 38.90 ±
0.02a 0.02a 0.04b 0.01b 0.30a
Fig. 3. Charcoal briquette preparation: A) mold and B) charcoal after
a,b
Different superscripts in the same column indicate significant differences at p compression into briquettes.
≤ 0.05.
mL/min and a heating rate of 10 ◦ C/min. The TT was performed at
charcoal (Fig. 1A and B). These charcoals were ground using a pulver­ different temperatures (200, 230, and 260 ◦ C) and reaction times (40,
izing machine (NANO TECH®, model NT-500D) and were screened 80, and 120 min) because these temperatures and times were sufficient
through a sieve opening of 0.71 mm (mesh No. 25) and 0.85 mm (mesh for the variation in compositions inside the biomass (volatile matter,
No. 20) until their size was in the range of 0.71–0.85 mm (Fig. 2A and B). carbon, hydrogen, and oxygen contents) and structural changes that
Both feedstocks were mixed with the TF (10% of the total briquette affected the fuel attributes (Barskov et al., 2019). The charcoal samples
mass) in five ratios of the percentage by weight of CP:CCP:TF as were then compressed into briquettes.
following: 0:90:10 (mixture A), 30:60:10 (mixture B), 45:45:10 (mixture
C), 60:30:10 (mixture D), and 90:0:10 (mixture E). Whole mixtures were 2.2.3. Hydrothermal technique (HT)
placed in a mold (Fig. 3A) to form cylindrical briquettes (inside diam­ The HT used the same conditions as the TT to compare fuel charac­
eter ≈ 1 cm, outside diameter ≈ 5 cm, and length ≈ 10 cm) (Fig. 3B). A teristics. The HT applied the method of Wu et al. (2018) by placing the
hydraulic press (WINNER, model Workshop press) was used to compress feedstock in a reactor (500 mL) and mixing it with deionized water at 1 g
at pressures of 1000, 1300, and 1600 psi for 40 s. The weight and MC of biomass per 5 mL. Nitrogen gas flowed into the reactor at a rate of 200
charcoal briquette were approximately 200 g and below 10% (w.b.), mL/min for 20 min to exhaust oxygen. The HT was operated at the
respectively. temperature (200–260 ◦ C) and reaction time (40–120 min), which suf­
ficiently provided for the dehydration, polymerization, and carboniza­
2.2.2. Torrefaction technique (TT) tion reactions to transform organic wastes into a rich carbon fuel (Shen,
The proper condition (mixture and pressure) from the CT were 2020). The charcoal sample was washed with deionized water, dried at
selected to study and compare the fuel characteristics via the TT, which room temperature, and formed into a charcoal briquette.
was applied using the method of Lin et al. (2021). The feedstock was
placed in a cylindrical reactor (500 mL) (TEFIC BIOTECH Company,
2.3. Charcoal briquette characteristics
CHINA) under an inert atmosphere with a nitrogen flow rate of 200
2.3.1. Calorific value and proximate analysis
The CV was determined using a bomb calorimeter (PARR, model
PARR-6300, USA) according to the American Society for Testing and
Materials (ASTM) standard method (D 5865-11a). The proximate anal­
ysis consisted of the MC, VM, AC, and FC, which were inspected ac­
cording to the ASTM standard of D 3173–87, D 3175–11, D 3174–89,
and E 711-87, respectively.

2.3.2. Ultimate analysis


The chemical components of carbon (C), hydrogen (H), nitrogen (N),
oxygen (O), and sulfur (S) in the charcoal samples were examined using
an elemental analyzer (LECO, model TruSpec, GERMANY), which was
Fig. 1. Charcoals after the carbonizing process: A) CP and B) CCP. conducted according to ASTM standards D 5373 and D 4239.

3
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

2.4. Mechanical properties 2.6. Statistical analysis

2.4.1. Hardness The experimental results were obtained from at least three repeti­
The hardness value indicated the charcoal briquette strength, tions, and statistical differences were analyzed using SPSS software
measured with a shore D hardness tester (model EQUQTIP), and re­ (version 14). One-way analysis of variance (ANOVA) with Tukey’s
ported in the Brinell hardness (HB) unit. honest significant difference (HSD) test (p ≤ 0.05) indicated a statisti­
cally significant difference in each condition. The influence of mixture
2.4.2. Porosity and pressure factors and their interactions on charcoal briquette quali­
Porosity was calculated from the apparent and specific densities ties in the CT were analyzed using a two-way analysis of variance with
(Equation (1)). The apparent density was evaluated using the mass Tukey’s HSD test at three significance levels (p ≤ 0.05, p ≤ 0.01, and p ≤
divided by the volume of the briquette. The briquette mass was deter­ 0.001).
mined using a digital scale with a precision of 0.01 g (Jadever Corp.,
model SNUG III-300). The cylindrical briquette volume was calculated 3. Results and discussion
using the diameter and length, which were measured using a vernier
caliper with a precision of 0.05 mm (Mitutoyo Corp., model 530–104). 3.1. CV and proximate analysis of charcoal briquettes from the CT
The specific density was determined using a gas pycnometer (Micro­
meritics Instrument Corp., model Accupyc 1340, Norcross, USA). Fig. 4 shows that all calorific values of the charcoal briquettes from
( ) the different mixtures and pressures in the CT (>17.5 MJ/kg) satisfied
Specific density – Apparent density
Porosity ​ (%) ​ = × 100 (1) the German Institute for Standardization (DIN) 51731. Most of the CV
Specific density
significantly increased when the CP quantity was increased in each
mixture, except for mixtures B, C, and D at 1000–1300 psi. In particular,
2.5. Utilization properties
the CV essentially increased at every pressure when the CP quantity
increased to 90% (without CCP) in mixture E with the high CV (26.7–27
2.5.1. Thermal efficiency (ηth )
MJ/kg), whereas the CCP quantity of 90% (without CP) in mixture A had
The water heating capability of charcoal briquette was estimated
the low CV (18.9–19.4 MJ/kg). These results may come from the higher
using Equation (2) (Sawadogo et al., 2018). The 1 L of water quantity
carbon content in the CP than in the CCP (as reported in Table 2),
was boiled using the heat from a briquette (200 g).
leading to the higher CV when the CP content vastly increases. de Souza
( )
Mw × Cw × Twf – Twi + (Mwe × Le ) et al. (2020) presented a higher CV and carbon content of CP (20.21
ηth ​ (%) = × 100 (2) MJ/kg and C ≈ 47.66%) than that of the coffee husk (18.89 MJ/kg and
(Mb × LCV)
C ≈ 46.16%).
where ηth - Thermal efficiency (%) The increase in pressure from 1000 to 1600 psi did not significantly
affect the CV, except for mixture D at 1600 psi. We expected that an
Mw - Initial water mass (kg) increase in compression pressure would enhance the density and raise
Cw - Specific heat capacity of water (4.187 kJ/kg◦ C) the energy value. However, the pressure range in this study insignificant
Twf - Final water temperature (◦ C) affected the CV as collated with the effect of mixtures, which intensively
Twi - Initial water temperature (◦ C) connected with the biomass compositions within briquette (Lela et al.,
Mwe - Evaporated water mass (kg) 2016; de Souza et al., 2020). Furthermore, the CV of briquettes may be
Le - Latent heat for water evaporation (2257 kJ/kg) influenced by the inorganic content within biomass (Thabuot et al.,
Mb - Consumed briquette mass (kg) 2015).
LCV - Low calorific value of briquette (MJ/kg) Fig. 5A shows that almost all conditions produced charcoal bri­
quettes with sufficient MC (<10% w.b.), except for mixture A. Most of
2.5.2. Flame temperature (FT) the MC significantly decreased with the CCP quantity in each mixture
The conventional stove with water of approximately 1 L was boiled because it contained a high MC (13% w.b.) similar to the high MC
from briquette combustion. The FT was the maximum blaze temperature (13.19% w.b.) of Annona squamosa peel, which was used to produce
measured every 5 min using a thermoscan camera (FLIR model E series, biofuel (Lin et al., 2021). Most of the MC in mixtures B, D, and E
SINCERE NETWORK, Thailand). insignificantly altered with the expanded pressure, but mixtures A and C
presented significant MC differences. These unclear variation trends of
2.5.3. Greenhouse gas (GHG) emission the MC may have resulted from moisture reabsorption from the sur­
The GHG emission was the CO2 release quantity during briquette roundings after heat treatment (Wu et al., 2018). The MC was expected
combustion, which was presented with the carbon dioxide equivalent to diminish owing to the raised compression force. Lela et al. (2016)
value (CO2e) using Equation (3) (Chungcharoen and Srisang, 2020). The observed a slight MC reduction when the compression force was
emission factor value (EF) was used to convert the heating value and increased.
consumed fuel amount into the GHG emission. The EF in this study was Fig. 5B demonstrates that all charcoal briquettes after the CT at a
approximately 100 gCO2/MJ (Eggleston et al., 2006), used for the CO2 high temperature of 500 ◦ C provided a lower VM (23–26%) than the
release evaluation from the solid biomass combustion in residence. Ananas comosus peel briquette (66–89%) (Lin et al., 2021) because the
VM was removed mainly at high temperatures. Nevertheless, the low
GHG ​ emission ​ (gCO2 e) ​ = ​ Mb × LCV × EF (3) VM of charcoal briquettes in this study was sufficient (<40%) for igni­
tion amelioration with the soft release of VM products (CO2 and CO)
where GHG emission - Release quantity of carbon dioxide equivalent
(Lubwama et al., 2020). Most of the VM insignificantly diminished from
(gCO2e)
the change in CP and CCP amounts, except for mixtures D and E at some
pressures. Mixtures A, B, and C provided high VM because the carbo­
Mb - Consumed briquette mass (kg)
hydrate within the CCP (≈45–89%) (Oliveira and Franca, 2015) was
LCV - Low calorific value of briquette (MJ/kg)
transformed into volatiles during combustion (Haykiri-Acma et al.,
EF - Emission factor (gCO2/MJ)
2013). An elevated pressure from 1000 to 1600 psi did not significantly
influence most of the VM in charcoal briquette.
Fig. 5C shows that FC significantly increased with the increase in CP

4
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Fig. 4. CV of charcoal briquettes obtained using the CT.

Fig. 5. Proximate analysis result of A) MC, B) VM, C) FC, and D) AC in charcoal briquette from the CT.

content in every mixture compared to the same pressure level. Mixture E A. This low AC can mitigate slagging and agglomeration problems (Lin
had a high FC (>60%) and agreed with the European standard (EN) (Wu et al., 2021). The AC from mixtures A, B, C, and E insignificantly
et al., 2018), whereas mixture A provided a lower FC (<45%). The vastly changed with pressure in this study, implying that pressure had less
different FC results in mixtures A and E may have resulted from the influence on the AC than the mixture. Lela et al. (2016) demonstrated
apparent difference in carbon contents between CP and CCP (Table 2). the change in the AC of fuel briquettes due to the process’s mixture and
The higher FC of the briquette was a desirable attribute because it drying temperature factors.
induced an expanded heating value and diminished the VM (Wu et al., The CV and proximate analysis results of charcoal briquette
2018). The compression pressure mainly influenced the FC at 1300 psi demonstrated the CT could provide the ideal fuel attributes with high
for mixture E and 1600 psi for mixtures B and D. Both pressures may be CV, high FC, low MC, and the high AC.
sufficient to cause intensive biomass compaction and result in a higher
FC.
3.2. Mechanical property of charcoal briquette from the CT
Fig. 5D shows that the AC amount significantly decreased with
increased CP quantity, except for some mixtures at 1000 and 1300 psi.
Fig. 6 shows that the hardness significantly increased when the CP
The high AC in mixture A (19.69–20.10%) caused by a high carbohy­
and CCP amounts apparently contrasted, as observed in mixtures A and
drate amount within the CCP was transformed into inorganic compo­
E. The highest CP content in mixture E provided the maximum hardness
nents during combustion, resulting in a high AC (Haykiri-Acma et al.,
(143 HB). The residual lignin after carbonization (500–900 ◦ C) (Lub­
2013). In contrast, the low AC in mixture E (4.27–4.41%) took place
wama et al., 2020) in CP exhibited a more solid bridge bonding of lignin
from almost all the CP combustion into heat. However, the low AC in
during compression (Lubwama and Yiga, 2018), which improved the
almost all the mixtures satisfied EN1860-2 (<18%), except for mixture
mechanical durability and hardness (de Souza et al., 2020). Most of the

5
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Fig. 6. Mechanical property of hardness of charcoal briquettes from the CT.

hardness was not significantly altered when the CP and CCP quantities fuel attributes. Kpalo et al. (2019) demonstrated that a sufficient pres­
were 30–60% (mixtures B, C, and D) because the CP and CCP contents sure (1480 psi) causes an explicit change in density and MC in some
did not inadequately differ. The increased pressure should enhance the mixtures. The low pressure of approximately 609–1015 psi was suffi­
density and strength of the briquette. However, the hardness change in cient for low-cost briquette production with good quality (Lubwama and
this study was insignificant and uneven with increased pressure because Yiga, 2017).
of the deficient pressure level and micro-pore occurrence during bri­ The charcoal briquette preparation under this research’s production
quetting after carbonization. These results contrasted with the conditions had the following satisfactory fuel attributes: the CV (>17.5
high-pressure use (5800–11600 psi), which could increase the density MJ/kg) satisfied the DIN51731 standard, and most of the MCs were in
and vigor of the briquettes, as reported by Song et al. (2020). the range of 7.9–9.8% w.b. (European standard) (de Souza et al., 2020),
in particular, mixture E had the maximum CV (27 MJ/kg) and the
minimum MC (6–7% w.b.); the VM was considerably low (<40%),
3.3. Influence of production parameters on briquette attributes from the
which assisted in reducing the smoke during combustion, and there was
CT
sufficient amount to promote fuel ignitability. An FC of approximately
40–65% supported the incremental CV at mixture E (FC ≈ 62–65%), and
Fig. 7A–E presents the altered trend of briquette properties from each
the AC at mixture E (AC < 4%) was satisfactory (<18%), following the
influence of mixtures and compression pressures. Most of the CV, MC,
European standard EN 1860–2.
FC, and AC properties changed from the mixture factor, as shown by the
Fig. 8 depicts the hardness property. The elevated CP amount of
steep slope of the graph line, particularly mixtures A and E, which had
about 90% in mixture E caused the apparent hardness upgrade
significantly different CP and CCP amounts. The mixture factor slightly
compared with the other mixtures. The pressure factor did not cause an
affected the VM compared with the other properties. The flat graph line
apparent variation in the hardness. However, the hardness (100–132
indicated that the pressure factor did not cause a noticeable variation in

Fig. 7. Influence of production parameters on A) CV, B) MC, C) VM, D) FC, and E) AC.

6
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Lin et al. (2021) demonstrated the augmented CV of biofuel with further


temperatures and times in the TT.
Fig. 9B shows that the MC in the HT decreased significantly with
increasing temperature. The MC in the TT significantly reduced when
the increased temperature was over 200 ◦ C because of the decrease in
hydrogen bond formation, improving the hydrophobic properties (Bar­
skov et al., 2019). Conversely, the MC notably increased when the
expanded time was over 80 min in the HT due to the longer water ab­
sorption time during the hydrothermal process. The HT had a signifi­
cantly lower MC than the TT at the same temperature (200 and 260 ◦ C),
as agreed by Tu et al. (2018). However, the MC from the TT and HT was
significantly less than the CT, which implied the mitigated GHG
emissions.
Fig. 9C indicates that the VM from the CT was significantly lower
than that from the TT and HT because the higher temperature (500 ◦ C)
caused more VM removal. The high VM induced a smoke problem,
whereas it contained a large amount of methane and other hydrocarbons
Fig. 8. Influence of production parameters on hardness. (Pandey and Dhakal, 2013), which could be easily ignited and be
conducive to fuel combustion (Sawadogo et al., 2018). The VM is
HB) were satisfactory, corresponding with those of Chungcharoen and essentially mitigated with the temperature raised in the TT and HT
Srisang (2020), who presented a median hardness value (103–123 HB). (260 ◦ C). Wu et al. (2018) observed a decrease in the VM when the
Table 3 shows the influence level from the overall parameters, i.e., temperature in the TT and HT increased. The increase in time caused the
the mixture factor, pressure factor, and interaction of both factors on the significant VM reduction in the HT and TT (200 ◦ C), which may arise
briquette characteristics. The greater value of the sum of squares value from the longer decomposition time and led to more VM releases. Lin
points to a more significant influence on the fuel properties. The inter­ et al. (2021) demonstrated that the VM in the TT decreased with the
action between the mixture and pressure had the greatest impact on temperature and time increment in some conditions. Most of the VM in
hardness (p ≤ 0.001), whereas the mixture had the second greatest in­ the TT was significantly lower than that in the HT due to the more VM
fluence on the FC. Under some conditions, the mixture slightly impacted displacement from tremendous pore quantities (Reza et al., 2014).
the VM and hardness properties, as confirmed by the low sum of squares Fig. 9D shows that the FC in the CT was significantly higher than that
(p ≤ 0.05). The compression pressure did not significantly affect any of in the TT and HT because the higher temperature stimulated the poly­
the fuel properties. merization reaction and caused a greater FC increment (Putra et al.,
The overall briquette properties indicated that mixture E provided 2018). Most of the FC in the TT and HT significantly increased with
the supreme values of the CV, FC, and hardness, with the lowest content increasing temperature and time (Wu et al., 2018; Lin et al., 2021). Most
of MC and AC. A pressure of 1600 psi was selected to investigate the of the FC in the HT was lower than that of the TT because of the higher
effects of torrefaction and hydrothermal techniques on fuel attributes discharge of components in liquid or gaseous forms during the hydro­
because this maximal pressure level assured the briquette strength after thermal process (Wu et al., 2018).
forming. Therefore, the mixture E and 1600 psi were used to produce Fig. 9E shows that the AC in the TT and HT significantly diminished
briquettes via the TT and HT at different temperatures (200, 230, and compared with the CT. The least AC (1.4%) in the HT was caused by
260 ◦ C) and reaction times (40, 80, and 120 min). substantial inorganics displacement during the water wash process
(Chen et al., 2007). Most of the AC in the TT was essentially elevated
with an increasing temperature that may have resulted from the incre­
3.4. Proximate and ultimate analyses in the TT and HT mental decomposition of hemicellulose and cellulose (Tu et al., 2018).
The elevated temperature in the HT at 200–230 ◦ C did not significantly
Fig. 9A shows that the CV in the CT (27 MJ/kg) was significantly affect the elevated AC that may arise from the water wash process. The
higher than that in the TT (20–26 MJ/kg) and HT (20–25 MJ/kg) additional time in TT and HT hardly influenced the AC, except for some
because higher temperatures decompose low-energy chemical compo­ conditions. The AC alteration also depended on the influence of the
nents and create high-energy chemical elements (Fan et al., 2018). The feedstock type and operating conditions (temperature and time) (Wu
increase in temperature and time in TT explicitly ameliorated the CV, et al., 2018; Lin et al., 2021).
except for the CV at 260 ◦ C did not significantly rise with the expanded Fig. 10A and B shows that most of the HT provided more carbon
time. The temperature and time effects on the CV in the HT exhibited a content than the TT and the CT owing to the more profound decompo­
similar tendency, except for some conditions. The increase in time sition (Tu et al., 2018). Most of the carbon content in the TT and HT
elevated the reaction severity and improved CV (Barskov et al., 2019). exhibited an increasing trend when the temperature enhanced over
230 ◦ C. The increased time in the HT caused improved carbon content
Table 3 because of structural disruption within the material (physical and
Analysis of variance (ANOVA) of charcoal briquette properties as influenced by chemical) and related to the temperature, pressure, and catalyst condi­
production parameters. tions (Putra et al., 2018). Except for some conditions, the nitrogen
Charcoal briquette properties Sum of squares (P-value) amounts in the TT and HT were less than in the CT. The sulfur content
was entirely removed in the TT and HT. The low nitrogen and sulfur
Mixture (M) Pressure (P) MxP
amount in briquette caused the low NOx and SO2 emissions during
CV (MJ/kg) 106.35*** 1.55(NS) 3.10*** combustion.
MC (% w.b.) 177.38*** 0.03(NS) 6.83***
VM (% dry mass) 59.16* 1.33(NS) 15.04***
The oxygen and hydrogen contents in the TT and HT were lower than
FC (% dry mass) 2370.74*** 23.65(NS) 22.29*** in the CT. The low amount of oxygen and hydrogen promoted the
AC (% dry mass) 1115.93*** 12.80(NS) 37.30*** heating value (Tu et al., 2018) and reduced CO2, CO, and H2O during
Hardness (HB) 71.60* 98.84(NS) 3208.71*** combustion (Fan et al., 2018). The oxygen quantities declined with
The levels of statistical significance of the mean values are * ≤ 0.05, *** ≤ 0.001, expanded time in the HT and TT (260 ◦ C). The higher temperature
and NS = not significant. caused a lower tendency of oxygen quantity in the TT and HT, except for

7
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Fig. 9. Charcoal briquette attributes of A) CV, B) MC, C) VM, D) FC, and E) AC via the TT and HT compared with the CT.

Fig. 10. Ultimate analysis of charcoal briquette via the A) TT and B) HT compared with the CT.

the TT at 40 min. The hydrogen amounts in the TT and HT diminished TT (260 ◦ C). The deoxygenation, dehydration, and decarboxylation re­
with increasing temperature (Zhuang et al., 2020; Lin et al., 2021) when actions in the TT decrease the hydrogen and oxygen contents (Lin et al.,
the temperature raised over 230 ◦ C, except for the TT for 40 min. The 2021). The HT’s dehydration, demethylation, and decarboxylation re­
additional time caused the hydrogen quantity reduction in the HT and actions lead to diminished hydrogen and oxygen amounts (Putra et al.,

8
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

2018). than the CT, which helped to economize the energy consumption in the
Therefore, the CT, TT, and HT provided satisfactory amounts of process. The ηth in the TT significantly increased with increasing tem­
carbon (30–60%), hydrogen (5–6%), oxygen (30–40%), nitrogen perature at the reaction time for 40 min, while the increase in time
(<1%), and sulfur (<1%) (Chaney, 2010), except for the excess oxygen hardly affected the ηth . The ηth in the HT crucially improved when the
content in the CT. temperature enhanced over 230 ◦ C and the growth in time at 260 ◦ C.
Fig. 11 shows the Van Krevelen diagram to indicate the coalification The uneven temperature and time effects on the ηth may have resulted
degree of briquettes. The low H/C and O/C ratios indicated good fuel from the AC content and stove efficiency (Lubwama and Yiga, 2018).
qualities like coal. The TT and HT provided lower H/C and O/C ratios Fig. 13B shows most of the FT in the TT and HT are not significantly
than the CT, and these ratios were classified in the biomass, peat, and changed with the improved temperatures and reaction times. The
lignite groups. The H/C and O/C ratios linearly diminished with the maximum FT from the three techniques (643–650 ◦ C) insignificantly
escalated time in the HT, while the TT showed a similar trend only at differed and was higher than those of the bio-composite briquette
260 ◦ C. Interestingly, the TT and HT at 260 ◦ C for 80–120 min upgraded (570 ◦ C) (Chungcharoen and Srisang, 2020). Remarkably, the FT was
briquette attributes resembled that of lignite (Tu et al., 2018). correlated with the ηth namely the high FT caused low fuel consumption
and resulted in the high. ηth .
3.5. Mechanical properties in the TT and HT Fig. 14 shows that almost all briquettes in the TT and HT had a lower
GHG emission than that of the CT and wood charcoal (460 gCO2e) and
Fig. 12A and B shows that most of the specific and apparent densities kerosene (478 gCO2e) (Chungcharoen and Srisang, 2020). The GHG
in the TT and HT insignificantly differed, which resulted in the close emissions in TT increased significantly with the elevated temperature,
porosity in the TT and HT (56–60%). Both densities in CT were signif­ whereas the increase in time caused the substantial augmentation of
icantly higher than those in the TT and HT. The porosity in CT did not GHG emissions only at 200 ◦ C. The GHG emissions in HT greatly
diverge substantially from that of other techniques, which may have enhanced with reaction time but insignificantly changed with the
resulted from the slight difference between the specific and apparent enhanced temperature (below 260 ◦ C). These escalated GHG emissions
densities and led to a nearby porosity. The temperature and time in the may have resulted from the augmented heating value, which stimulated
TT and HT did not affect the density or porosity. The CT had higher the component decomposition. Lin et al. (2021) demonstrated that the
hardness than the TT and HT, although there were the same porosities. different temperatures and reaction times in the TT did not significantly
The higher hardness in the CT is caused by the remaining lignin after affect the GHG emission. Furthermore, the different feedstock types and
carbonization at 500 ◦ C, resulting in more bonding during compression. operating conditions may also affect GHG emissions. Most of the GHG
The hardness in the HT was primarily lower than in the CT because the emissions in the TT were higher than in the HT. The low GHG emission
biomass components were less stable under hydrothermal conditions, indicates fuel-friendly to the environment, which mitigated the air
resulting in the lignin decomposition at low temperatures (Wu et al., pollution emissions. Lin et al. (2021) reported that the fuel combustion
2018). The HT provided a higher hardness than the TT because the of biochar mixed with bituminous coal could reduce GHG emissions by
organic compounds accumulated on a surface in the HT (Liu and Zhang, approximately 83–94%.
2008) and caused the liquid bridge formation with the nearby particle,
which improved the briquette strength (Simons, 2007). However, most 4. Conclusion
of the apparent densities in TT, HT, and CT were at a pleasant level
(>0.6 g/cm3) (Kpalo et al., 2020). The manufacturing factors of fuel briquettes from coffee production
waste produced via the CT, TT, and HT were investigated in terms of fuel
quality. The essential outcomes are summarized as follows:
3.6. Utilized properties in the TT and HT
1. The interaction between the mixture and pressure factors in
Fig. 13A shows that the ηth in the CT, TT, and HT are sufficient for
briquette production via the CT had the greatest influence on hard­
domestic cooking (ηth ≈ 33%) (Sawadogo et al., 2018). The ηth in the CT
ness. In contrast, the mixture had the second-highest influence on the
were significantly higher than most of the TT and HT, indicating supe­
FC. The proper conditions of mixture E and 1600 psi provided the
rior fuel conversion for heat utilization and implied saving of energy,
cost, and resources. However, the TT and HT used a lower temperature

Fig. 11. Van Krevelen diagram of charcoal briquette via the A) TT and B) HT compared with the CT.

9
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Fig. 12. Charcoal briquette properties of A) specific density, B) apparent density, C) porosity, and D) hardness via the TT and HT compared with the CT.

Fig. 13. Utilization properties of A) ηth and B) FT via the TT and HT compared with the CT.

fuel briquette with prominent characteristics (high CV and FC values with the conventional CT. Understanding the fuel characteristics of each
and low MC and AC values). technique can provide a guideline for modifying the process and
2. The TT and HT can produce charcoal briquettes with acceptable fuel creating additional pleasant briquette attributes with the required uti­
attributes (CV, proximate and ultimate results, mechanical and uti­ lization and environmental friendliness.
lized properties) under lower temperature and GHG emission
compared with the CT, except that the FC is low (<60%). The CRediT authorship contribution statement
appropriate conditions for TT and HT should be 260 ◦ C for 120 min
with better overall properties and the upgraded briquette attributes Siriwan Srisang: Writing – review & editing, Methodology, Inves­
equal in lignite. tigation. Kittisak Phetpan: Visualization, Validation. Nuttapong Rut­
tanadech: Resources, Validation. Warunee Limmun: Data curation,
Thus, torrefaction and hydrothermal techniques have the potential Validation. Pannipa Youryon: Resources. Pornprapa Kongtragoul:
to produce charcoal briquettes from coffee production waste compared Data curation. Naruebodee Srisang: Writing – original draft,

10
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Fig. 14. GHG emissions of charcoal briquette via the TT and HT compared with the CT.

Conceptualization, Formal analysis, Supervision. Thatchapol Chung­ Kpalo, S., Zainuddin, M., Halim, H.A., Ahmad, A., Abbas, Z., 2019. Physical
characterization of briquettes produced from paper pulp and Mesua ferrea mixtures.
charoen: Validation, Visualization.
Biofuels 1–8.
Kpalo, S.Y., Zainuddin, M.F., Manaf, L.A., Roslan, A.M., 2020. A review of technical and
Declaration of competing interest economic aspects of biomass briquetting. Sustainability 12, 4609.
Lee, J., Lee, K., Sohn, D., Kim, Y.M., Park, K.Y., 2018. Hydrothermal carbonization of
lipid extracted algae for hydrochar production and feasibility of using hydrochar as a
The authors declare that they have no known competing financial solid fuel. Energy 153, 913–920.
interests or personal relationships that could have appeared to influence Lela, B., Barišić, M., Nižetić, S., 2016. Cardboard/sawdust briquettes as biomass fuel:
the work reported in this paper. physical–mechanical and thermal characteristics. Waste Manag. 47, 236–245.
Lin, Y.-L., Zheng, N.-Y., Hsu, C.-H., 2021. Torrefaction of fruit peel waste to produce
environmentally friendly biofuel. J. Clean. Prod. 284, 124676.
Data availability Liu, Z., Zhang, F.-S., 2008. Effects of various solvents on the liquefaction of biomass to
produce fuels and chemical feedstocks. Energy Convers. Manag. 49, 3498–3504.
Lubwama, M., Yiga, V.A., 2017. Development of groundnut shells and bagasse briquettes
No data was used for the research described in the article. as sustainable fuel sources for domestic cooking applications in Uganda. Renew.
Energy 111, 532–542.
Acknowledgments Lubwama, M., Yiga, V.A., 2018. Characteristics of briquettes developed from rice and
coffee husks for domestic cooking applications in Uganda. Renew. Energy 118,
43–55.
This research funded by King Mongkut’s Institute of Technology Lubwama, M., Yiga, V.A., Muhairwe, F., Kihedu, J., 2020. Physical and combustion
Ladkrabang (KREF026401). properties of agricultural residue bio-char bio-composite briquettes as sustainable
domestic energy sources. Renew. Energy 148, 1002–1016.
Oliveira, L.S., Franca, A.S., 2015. An overview of the potential uses for coffee husks.
References Coffee in health and dis. prevent. 283–291. Elsevier.
Pandey, S., Dhakal, R.P., 2013. Pine needle briquettes: a renewable source of energy. Int.
Aransiola, E., Oyewusi, T., Osunbitan, J., Ogunjimi, L., 2019. Effect of binder type, J. Energy Sci 3, 254–260.
binder concentration and compacting pressure on some physical properties of Putra, H.E., Damanhuri, E., Dewi, K., Pasek, A.D., 2018. Hydrothermal carbonization of
carbonized corncob briquette. Energy Rep. 5, 909–918. biomass waste under low temperature condition. In: MATEC Web of Conferences.
Barskov, S., Zappi, M., Buchireddy, P., Dufreche, S., Guillory, J., Gang, D., Hernandez, R., EDP Sciences, 01025.
Bajpai, R., Baudier, J., Cooper, R., 2019. Torrefaction of biomass: a review of Reza, M.T., Uddin, M.H., Lynam, J.G., Coronella, C.J., 2014. Engineered pellets from dry
production methods for biocoal from cultured and waste lignocellulosic feedstocks. torrefied and HTC biochar blends. Biomass Bioenergy 63, 229–238.
Renew. Energy 142, 624–642. Sawadogo, M., Tanoh, S.T., Sidibé, S., Kpai, N., Tankoano, I., 2018. Cleaner production
Chaney, J., 2010. Combustion Characteristics of Biomass Briquettes. University of in Burkina Faso: case study of fuel briquettes made from cashew industry waste.
Nottingham Nottingham, UK. J. Clean. Prod.
Chen, S.-F., Mowery, R.A., Scarlata, C.J., Chambliss, C.K., 2007. Compositional analysis Shen, Y., 2020. A review on hydrothermal carbonization of biomass and plastic wastes to
of water-soluble materials in corn stover. J. Agric. Food Chem. 55, 5912–5918. energy products. Biomass Bioenergy 134, 105479.
Chungcharoen, T., Srisang, N., 2020. Preparation and characterization of fuel briquettes Simons, S.J., 2007. Liquid bridges in granules. In: Handbook of Powder Technology.
made from dual agricultural waste: cashew nut shells and areca nuts. J. Clean. Prod. Elsevier, pp. 1257–1316.
256, 120434. Song, X., Zhang, S., Wu, Y., Cao, Z., 2020. Investigation on the properties of the bio-
de Souza, H.J.P.L., Arantes, M.D.C., Vidaurre, G.B., Andrade, C.R., Carneiro, A.d.C.O., de briquette fuel prepared from hydrothermal pretreated cotton stalk and wood
Souza, D.P.L., de Paula Protásio, T., 2020. Pelletization of eucalyptus wood and sawdust. Renew. Energy 151, 184–191.
coffee growing wastes: strategies for biomass valorization and sustainable bioenergy Srisang, S., Srisang, N., 2021. Recycling spent bleaching earth and oil palm ash to tile
production. Renew. Energy 149, 128–140. production: impact on properties, utilization, and microstructure. J. Clean. Prod.
Eggleston, H., Buendia, L., Miwa, K., Ngara, T., Tanabe, K., 2006. 2006 IPCC Guidelines 294, 126336.
for National Greenhouse Gas Inventories. Thabuot, M., Pagketanang, T., Panyacharoen, K., Mongkut, P., Wongwicha, P., 2015.
Fan, F., Yang, Z., Li, H., Shi, Z., Kan, H., 2018. Preparation and Properties of Hydrochars Effect of applied pressure and binder proportion on the fuel properties of holey bio-
from Macadamia Nut Shell via Hydrothermal Carbonization, 5. Royal Society open briquettes. Energy Proc. 79, 890–895.
science, 181126. Tu, R., Jiang, E., Yan, S., Xu, X., Rao, S., 2018. The pelletization and combustion
Haykiri-Acma, H., Yaman, S., Kucukbayrak, S., 2013. Production of biobriquettes from properties of torrefied Camellia shell via dry and hydrothermal torrefaction: a
carbonized brown seaweed. Fuel Process. Technol. 106, 33–40. comparative evaluation. Bioresour. Technol. 264, 78–89.
Kai, X., Meng, Y., Yang, T., Li, B., Xing, W., 2019. Effect of torrefaction on rice straw
physicochemical characteristics and particulate matter emission behavior during
combustion. Bioresour. Technol. 278, 1–8.

11
S. Srisang et al. Journal of Cleaner Production 372 (2022) 133744

Wu, S., Zhang, S., Wang, C., Mu, C., Huang, X., 2018. High-strength charcoal briquette nutrients availability and energy gain from produced hydrochar. Appl. Energy 229,
preparation from hydrothermal pretreated biomass wastes. Fuel Process. Technol. 88–95.
171, 293–300. Zhuang, X., Song, Y., Zhan, H., Bi, X.T., Yin, X., Wu, C., 2020. Pyrolytic conversion of
Yu, Y., Lei, Z., Yang, X., Yang, X., Huang, W., Shimizu, K., Zhang, Z., 2018. Hydrothermal biowaste-derived hydrochar: decomposition mechanism of specific components.
carbonization of anaerobic granular sludge: effect of process temperature on Fuel 266, 117106.

12

You might also like