You are on page 1of 11

Bioresource Technology 361 (2022) 127734

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Process modeling and evaluation of optimal operating conditions for


production of hydrogen-rich syngas from air gasification of rice husks using
aspen plus and response surface methodology
Emmanuel Yeri Kombe a, b, Nickson Lang’at c, Paul Njogu d, Reiner Malessa e,
Christian-Toralf Weber b, Francis Njoka a, *, Ulrich Krause f
a
Department of Energy Technology, Kenyatta University, P. O. Box 43844 – 00100, Nairobi, Kenya
b
Department of Engineering and Industrial Design, Hochschule Magdeburg-Stendal, Breitscheidstr. 2, 39114 Magdeburg, Germany
c
Department of Agriculture and Biosystems Engineering, Kenyatta University, P. O. Box 43844 – 00100, Nairobi, Kenya
d
Institute of Energy and Environmental Technology, Jomo Kenyatta University of Agriculture and Technology, P. O. Box 62000 – 00200, Nairobi, Kenya
e
Department of Engineering, Technische Hochschule Brandenburg Fachbereich Technik, University of Applied Sciences, Magdeburger Str. 50, 14770 Brandenburg an der
Havel, Germany
f
Department of Instrumental and Environmental Technology, Otto-von-Guericke-University, Universitätspl. 2, 39106 Magdeburg, Germany

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• A robust three-phase numerical simula­


tion model of a fixed-bed gasifier is
developed.
• The generated regression models have a
high degree of accuracy.
• High temperature favors CO and H2
production, H2 yield, LHVSyngas, CGE,
and PCG.
• Optimal performance is achieved at
temperatures of 820–1090 ◦ C and ER of
0.06–0.10.
• An optimum H2 yield of 46.39 g/kg and
PCG of 75.93% is observed.

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

Keywords: Biomass gasification is recognized as a viable avenue to accelerate the sustainable production of hydrogen. In this
Rice husk gasification work, a numerical simulation model of air gasification of rice husks is developed using the Aspen Plus to investigate
Hydrogen production the feasibility of producing hydrogen-rich syngas. The model is experimentally validated with rice husk gasification
Aspen Plus
results and other published studies. The influence of temperature and equivalence ratio on the syngas composition,
Response surface methodology
Optimization
H2 yield, LHVSyngas, H2/CO ratio, CGE, and PCG was studied. Furthermore, the synchronized effects of temperature
and ER are studied using RSM to determine the operational point of maximizing H2 yield and PCG. The RSM analysis
results show optimum performance at temperatures between 820 ◦ C and 1090 ◦ C and ER in the range of 0.06–0.10.
The findings show that optimal operating conditions of the gasification system can be achieved at a more refined
precision through simulations coupled with advanced optimization techniques.

* Corresponding author.
E-mail address: njoka.francis@ku.ac.ke (F. Njoka).

https://doi.org/10.1016/j.biortech.2022.127734
Received 23 June 2022; Received in revised form 29 July 2022; Accepted 30 July 2022
Available online 3 August 2022
0960-8524/© 2022 Elsevier Ltd. All rights reserved.
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

impact of various gasification conditions on its performance (Gómez-


Nomenclatures Barea and Leckner, 2010). Many commercially available computer-
based modeling and simulation software, such as Chemkin, Fluent,
Adj SS Adjusted sum of squares Aspen Hysys, and Aspen Plus, have been used in biomass gasification
BBD Box-Behnken design simulation (Singh and Tirkey, 2021). Among those programs, Aspen
CCD Central composite design Plus is mostly preferred for biomass gasification simulations (Singh and
CCRD Central composite rotatable design Tirkey, 2021) because it contains a pre-installed library model for
CGE Cold gas efficiency calculating solid characteristics. It is, therefore, able to determine solid
CS50:OPF50 coconut shells and oil palm fronds blending ratio of components better than the other software. Besides this, Aspen Plus is
50:50 paired with an imperative programming language referred to as
DF Degree of freedom FORTRAN code for numerical calculations, making it easier to create
ER Equivalence ratio customizations and numerical computations.
HHV Higher heating value of syngas Several extensive studies have been conducted on modeling the
LHVbm Lower heating value of biomass feedstock (MJ/kg) biomass gasification process for hydrogen production using Aspen Plus
LHVSyngas Lower heating value of syngas (MJ/Nm3) simulator. For instance, Shen et al. (2008) simulated steam gasification
PCG Percentage of combustible gas of straw in an interconnected fluidized bed system for hydrogen pro­
Qyield Syngas yield (Nm3/kg) duction. They studied how temperature and steam to biomass ratio (S/B)
RMSE Root mean square error affected the composition of syngas, hydrogen yield, and carbon con­
RSM Response surface methodology version efficiency. Tan and Zhong (2010) too simulated hydrogen pro­
S/B Steam to biomass ratio duction from steam gasification of sawdust. The authors investigated the
SSR Sum of squares of residual impact of reactor temperature, S/B, and pressure on CO/CO2 ratio and
SST Total sum of squares syngas composition. Hussain et al. (2021) simulated steam gasification
of palm kernel shells in a fluidized bed gasification plant for hydrogen
production. They investigated how temperature and S/B affected the
syngas composition. Abuadala and Dincer (2010) conducted a compre­
1. Introduction hensive simulation study of steam gasification of wood and sawdust to
predict H2-rich syngas production. The authors studied how tempera­
A sustainable energy system in the future projects hydrogen as the ture and S/B influence energy and exergy efficiencies of hydrogen
most crucial energy carrier (Shen et al., 2008). Biomass, the world’s production.
primary source of energy after fossil fuels (Tavares et al., 2020), has the Faraji and Saidi (2021) developed a comprehensive simulation
potential to accelerate the acceptance of hydrogen. Its major charac­ model that integrates pyrolysis and air gasification process of various
teristics are apparent zero net CO2 (Roy et al., 2019) and worldwide algal biomass materials for hydrogen production. The authors studied
availability (Zaman et al., 2020). the influence of rector temperature, gasifier pressure, and airflow rate
A substantial amount of agro-industrial, municipal, and forestry on syngas composition, hydrogen yield, and H2/CO ratio. Cohce et al.
waste is discarded each year; instead, it can be recovered and converted (2010) simulated air gasification of oil palm shells coupled with steam-
into energy via thermochemical or biological conversion methods (Li methane reforming and shift reaction to investigate its performance in
et al., 2019). Annually, the world generates over 120 million tons (Yoon hydrogen production. They further evaluated the energy efficiency,
et al., 2012) of rice husks. Generally, the consistent yearly production of exergy efficiency, and cold gas efficiency (CGE), which are essential in
rice husks makes it possible to secure a sustainable raw material supply designing and optimizing hydrogen production.
(Yoon et al., 2012). Rice husks do not require a pre-treatment process, Favas et al. (2017) simulated air and steam plasma gasification of
and this reduces the need for pre-treatment costs (Yoon et al., 2012). forest residue, coffee husks, and vine pruning. The authors investigated
Furthermore, rice husks have a uniform particle size and chemical how temperature, equivalence ratio (ER), and S/B affected the compo­
composition, consequently enabling efficient combustion (Yoon et al., sition of syngas and the syngas lower heating value (LHVSyngas). They
2012). A couple of years back, rice husk was regarded to be of no eco­ further studied the combined effect of S/B and temperature on hydrogen
nomic value to millers (Kate and Chaurasia, 2018) and was often production using a 3D diagram. Tavares et al. (2020) simulated air and
dumped in the ground or burned in open fields, thus creating land and steam gasification of Portuguese forest residue in a downdraft gasifier.
air pollution (Bakar et al., 2016). Due to the increase in demand for The authors investigated the influence of temperature and S/B on syngas
waste-to-energy utilization, researchers are actively searching for composition and LHVSyngas. Furthermore, they used a 3D diagram to
alternative avenues of using rice husks as fuel. Gasification of solid fuels study the combined effect of temperature and S/B, thereby optimizing it
is widely acknowledged as the most efficient method of generating en­ to magnify hydrogen production. Safarian et al. (2021) simulated air
ergy from a variety of wastes (Faraji and Saidi, 2021). It is incredibly and a mixture of air–steam gasification of timber and wood waste for
adaptable when it comes to using different types of fresh biomass hydrogen production. They considered three zones of the gasification
feedstocks (Heidenreich and Foscolo, 2015). This technique improves process: drying, pyrolysis, and gasification coupled with a water gas
the efficiency of waste management (Kombe et al., 2022) and offers reactor and product recovery unit. The authors studied how temperature
excellent potential for producing renewable and low-carbon energy and S/B affected the composition of syngas, LHVSyngas, and hydrogen
(Kumar and Samadder, 2017). production efficiency. Doranehgard et al. (2017) described the catalytic
From a theoretical point of view, gasification is a thermochemical behavior of calcium oxide in air and steam gasification of rice husk
process that converts raw biomass materials to a gaseous fuel by based on a developed semi-kinetic model. The authors investigated how
partially oxidizing them with gasifying agents such as oxygen, steam, temperature, S/B, and ER influenced the yield of hydrogen and CO2
air, steam, or their mixtures (Bridgwater, 2003). It is characterized by absorption ratio.
drying (100 ◦ C – 200 ◦ C), pyrolysis (200 ◦ C – 500 ◦ C), oxidation, and Response surface methodology (RSM) is a mathematical and statis­
reduction stages (Faraji and Saidi, 2021). The key reactions that take tical tool with a wide range of applications in engineering where
place in a gasifier are reported in a recent publication (Kombe et al., simultaneous optimization of different response variables is required
2022) and elsewhere (Doherty et al., 2009; Han et al., 2017). (Roy et al., 2020). It is a popular method for optimizing new and current
Due to the gasification process’s complexity, gasification models are modeled systems or processes (Zaman and Ghosh, 2021). However, very
required to understand and predict its behavior as well as to study the few studies have reported biomass gasification process optimization

2
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

using RSM for hydrogen production. Seçer and Hasanoğlu (2020) used Table 1
RSM with a Box-Behnken design (BBD) to optimize an experimental Proximate and ultimate analysis of rice husk, rubber wood, and rape straw.
study of co-gasification of Çan lignite and sorghum biomass for Rice husk Rubber wood Rape straw
hydrogen production. Okolie et al. (2020) developed a model and (Jayah et al., 2003) (Striūgas et al., 2014)
optimized hydrothermal gasification of synthetic biomass for the pro­ Ultimate analysis (wt%, dry)
duction of hydrogen based on BBD in RSM. Kang et al. (2015) optimized C 37.42 50.60 39.60
hydrogen production from noncatalytic supercritical water gasification H 5.17 6.50 5.60
of lignin central composite design (CCD) in RSM. Yusup et al. (2014) N 0.13 0.20 0.78
S 0.64 0.0 0.08
optimized hydrogen production from the steam gasification process of O 46.28 42.0 48.54
palm kernel shell with in-situ catalytic adsorption using RSM based on a
central composite rotatable design (CCRD).
Proximate analysis (wt%, dry)
Although there are several studies on the production of hydrogen Moisture content 9.75 14.0–18.5 14.9
from biomass gasification through modeling using Aspen Plus, many of Volatile matter 72.82 80.1 62.50
the simulation studies have focused on the fundamental process of Fixed carbon 16.82 19.2 17.20
hydrogen production and the inclusion of sub-models considering the Ash 10.36 0.7 5.40

gasification process’s specific aspects such as syngas condition through


water–gas-shift reaction to improve its predicting capabilities. To our the simulation were estimated by employing the Peng-Robinson equa­
knowledge, very limited work has comprehensively considered the in­ tion of state with the Boston-Mathias alpha function (PR-MB).
clusion of syngas purification coupled with RSM optimization for
hydrogen production. Moreover, it can be observed from the brief
2.3. Aspen plus model flowsheet
literature review on RSM that no work has been done on a compre­
hensive simulation of air-gasification of rice husk integrated with syngas
The generated flowsheet for the fixed-bed gasifier simulation process
purification and RSM. Furthermore, 400 runs of Aspen Plus simulated
of air gasification of rice husks is shown in Fig. 1. The red dotted line
results were used to construct an RSM design matrix, offering another
depicts the energy flow. The Aspen Plus reactor blocks used in devel­
mileage of the present study. This work, therefore, sought to develop a
oping the process simulation model are described in Table 2. The default
robust numerical simulation model of a fixed bed gasifier integrated
ID refers to the software developer-defined block name, and the user
with syngas purification and RSM with the ability to perform a rigorous
creates the assigned ID.
multi-objective optimization (using both contour, overlay contour, 3D
surface, and optimization plots). This was then used to predict the
feasibility of hydrogen production from air-gasification of low-cost 2.4. Simulation procedures
biomass materials such as rice husks, thereby improving the efficiency
of its utilization. Biomass drying was simulated in the block ’DRIER’ to reduce its
moisture content. This was achieved through a FORTRAN statement
2. Materials and methods written in a calculator block. Then, the partially evaporated moisture is
separated by a separator block MOISTSEP as EXIT. Thermal decompo­
2.1. Biomass characteristic sition of the pre-dried biomass (DRYBIOM) was subsequently carried out
in the reactor block DECOMP. This reactor simulates biomass decom­
Rice husk was sourced locally from a rice mill factory in Mwea irri­ position from a non-conventional component to its conventional com­
gation scheme region, Kirinyaga County, Kenya. The sample’s proximate ponents by defining the yield distribution in a calculator block using a
analysis was determined according to ASTM standards. The biomass FORTRAN command according to its proximate and ultimate analysis,
moisture content was determined as per ASTM E 871 using a binder shown in Table 1. Subsequently, air was used as the gasifying agent in
drying and heating chamber (Model: E 28), ASTM E 872, and ASTM E the RGibbs reactor block OXID to oxidize the disintegrated components
1755 for volatile matter and ash content, respectively, using a furnace partially. The thermal needs of the reactions were met by supplying the
(Model: WiseTherm). The sample’s ultimate analysis was then per­ heat of reaction (Q-COMP). To simulate the gasifier’s reduction zone, a
formed using a Leco C–H–N 628 with add-on 628 S modules following second RGibbs reactor block REDUC was used. RGibbs reactor is useful
ASTM D 5373-02 for C–H–N and ASTM D 4239-05 for sulfur content. because of its capacity to handle multiphase chemical equilibrium. The
The proximate and ultimate analysis results of rice husk is shown in OXID and REDUC reactors calculate the final syngas composition by
Table 1. Gibbs free energy minimization. Syngas purification and dehumidifica­
tion were simulated using a series of units block consisting of a cyclone,
2.2. Model development a cooler, a scrubber, and a separator block. The block ’CYCLONE’ was
used to separate ash and unreacted carbon from the product gas. The hot
Aspen Plus Version 10 was used to model the air gasification of rice ’SYNGAS’ was cooled to 25 ◦ C by ’COOLER’ and brought in contact with
husks. A three-phase numerical simulation model of air gasification of water ’FRESH-SO’ in the ’SCRUBBER’ to remove contaminants such as
rice husk was developed. The phases comprised biomass drying, biomass particulate matter, nitrogen compounds (NH3, HCN), and tars as
gasification, and syngas purification. In the biomass drying phase, fresh described in detail in reference (Woolcock and Brown, 2013). The
biomass moisture content is lowered before it is loaded into the biomass moisture in the ’CL-GAS’ stream was then condensed and separated from
gasification phase, which comprises decomposition, oxidation, and dry syngas by a water separator, ’WATERSEP’. The reactor unit oper­
reduction reactors. The syngas produced is then purified by passing ating and input conditions of the feed stream in the simulator is shown in
through a cyclone, cooler, scrubber, and a separator block in the third Table 3.
phase.
The Aspen Plus software’s standard component database excludes 2.5. Model assumptions
biomass feedstock and ash. As a result, the global stream class was
specified as MIXCINC, which contains MIXED, CISOLID, and NC sub- A number of assumptions were made in developing the numerical
streams. This global stream class option is advocated for when simulation model. The gasification process was considered a steady-
modeling non-conventional feed like biomass material and ash (Adeniyi state condition with constant pressure (atmospheric pressure). Char is
et al., 2021). The physical properties of non-conventional components of assumed to contain only carbon. The time reactants take in the reactor is

3
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

Fig. 1. Process simulation flowsheet for air gasification of rice husk in Aspen Plus.

the composition of syngas, H2 yield, LHVSyngas, H2/CO ratio, and PCG


Table 2
were then studied.
Aspen Plus reactor block description used in the simulation.
Block ID Default Description
ID
2.7. Response surface methodology (RSM)
DRIER RStoic Simulate wet biomass drying by specifying the
stoichiometric reaction This study aimed to develop a regression model that would define
MOISTSEP Sep2 Separation of excess moisture from dry biomass
system performance by considering the interactions between the key
DECOMP RYield Decompose biomass material into conventional
components by defined yield distribution based on the factors. The simulated data from Aspen Plus were used to construct a
biomass proximate and ultimate analysis design matrix in Minitab. The objective response regression models were
OXID RGibbs Model oxidation of the disintegrated biomass created by performing an Analysis of variance (ANOVA) at a 95% con­
components
fidence level and thus evaluating the model’s quality. The response-
REDUC RGibbs Simulate biomass gasification
CYCLONE SSplit Simulate solid separation from the hot syngas
variable interaction was established as per the 2nd order quadratic
COOLER Heater Cool hot syngas to 25 ◦ C equation (Mojaver et al., 2019) given in Eq. (1).
SCRUBER RadFrac Remove contaminants such as particulate matter,
nitrogen compounds (NH3, HCN), and tars from the ∑
n ∑
n n− 1 ∑
∑ n

product gas
y = β0 + β i xi + βii x2i + βij xi xj + ∈ (1)
i=1 i=1 i=1 j=1
WATERSEP Sep2 Separate moisture from the purified syngas.
Where y represents the output response; x stands for the decision
parameter; βi are coefficients; n denotes the total sum of factors, and ∈
Table 3 stand for the statistical error.
Reactor unit operating and feed stream input conditions for simulation. The regression coefficient (R2) and the adjusted regression coeffi­
Block information Operating conditions cient (R2adj ) values were used to quantify the accuracy of the developed
Type Name Temperature Pressure regression model. These parameters were determined as follows
(◦ C) (atm) (Mojaver et al., 2019):
RStoic DRIER 109.85 1
⎛ ⎞
( )
RYield DECOMP 499.5 1 ⎜ SSR 2
(n− p) ⎟
⎟ = 1 − 1 − R (n − 1)
RGibbs OXID 800 1 R2adj = 1 − ⎜
⎝ SST ⎠ (2)
RGibbs REDUC 350–1100 1 (n− 1)
1− p

Feed Operating conditions SSR and SST are determined as per Eqs. (3) and (4), respectively.
streams
Temperature Pressure Flow rate (kg/hr) ∑
n
( )2
(◦ C) (atm)
SSR = yi − yj (3)
i=1
WETBIOM 24.85 1 20
( ∑n )
AIR-D 132 1 100 ∑
n 2
i=1 yi
AIR 24.85 1 10 SST = y2i − (4)
FRESH-SO 25 1 150 i=1
n

Where yi and yj represent the observations and fitted observations,


considered sufficiently long for the reactions to attain both chemical and respectively.
thermodynamic equilibrium. Finally, the product gas stream consists The P-value, which is a crucial parameter in the model, is often
mainly of CH4, CO2, CO, H2, N2, and H2O, and all the gases follow the regarded as insignificant if its value exceeds 0.05 (Roy et al., 2020). R2 is
Peng-Robinson equation of state with the Boston-Mathias alpha function a measure of the degree of fit, it was computed as per Eq. (5) (Mojaver
(PR-MB). et al., 2019).
SSR
2.6. Sensitivity analysis procedures R2 = 1 − (5)
SST

The developed simulation model was used to perform sensitivity Both R2 and R2adj values ranged from 0 to 100%. A value greater than
analysis by varying temperature from 350 ◦ C to 1100 ◦ C and ER in the 90% indicated that the model was precise (Mojaver et al., 2019; Zaman
ranges of 0.06 to 0.42. The influence of temperature and ER variation on and Ghosh, 2021). The difference between R2 and R2adj of less than 0.2,

4
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

further indicated a suitable model (Zaman and Ghosh, 2021).


H2 mass flow in gasifier (g/hr)
Hydrogen yield (g/kg of biomass) = (9)
Biomass mass flow rate (kg/hr)
2.8. Model validation
percentage volume of hydrogen (dry basis)
A full-scale locally made pilot updraft gasifier at Jomo Kenyatta H2 /CO ratio = (10)
percentage volume of carbon monoxide (dry basis)
University of Agriculture and Technology (JKUAT) was used to validate
the simulation results. The unit comprises a 400 mm diameter and 2000 The percentage of combustible gas (PCG) in the syngas was calcu­
mm height reactor coupled with an electric blower (Black and Decker lated as per Eq. (11) (Duan et al., 2015).
KTX5000 Blower 220–240 model) to provide the air needed for gasifi­ nCO + nH2 + nCH4
cation. The gasifier was loaded with 19.6 kg of rice husks at the PCG = × 100 (11)
ntotal
beginning of the experiment and manually ignited by dropping a
burning piece of paper into the gasifier through the opening. The gasifier Where ntotal denotes the number of moles of syngas, and nCH4 , nH2 , and
feeding door was closed after the ignition of biomass. Gasification nCO represents the number of moles of CH4, H2, and CO in the syngas,
temperature was measured and monitored using K-type thermocouples. respectively.
The reactor product gas purification was carried out in a series of gas-
cleaning devices, consisting of a wet scrubber, a cyclone, and two fil­ 3. Results and discussion
ters arranged in series. A gas chromatography with a thermal conduc­
tivity detector (GC – TCD) model (GC-8A, Shimadzu Corp., Kyoto, 3.1. Model validation
Japan) was used to analyze the collected product gas samples.
The simulated results were further validated using a set of two The comparison between the simulated and experimental results is
published experimental results obtained at similar conditions as the shown in Fig. 2. The average RMSE is found to be 7.73. From the figure,
operation limits of the simulations. The first validation was carried out the RMSE reduces with a rise in temperature. The rates of reactions are
with the test results of Jayah et al. (2003). In this validation, nine runs of slower at low temperatures, causing the reactions to take longer to attain
test results obtained from air gasification of rubber wood in a downdraft a state of equilibrium (Aydin et al., 2017), resulting in a decrease in the
gasifier at 900 ◦ C and 1 atm were used. Secondly, test results by Striūgas equilibrium models’ performance. Aydin et al. (2017) also observed a
et al. (2014) were used in the comparison. In this comparison, a single similar trend in their semi-empirical model study of various wood-based
run of test results obtained from air gasification of rape straw in a biomass feedstock in a downdraft gasifier. Although the model predicted
downdraft gasifier operated at 1000 ◦ C was used. The proximate and CH4, H2, and N2 concentration to a good degree, it underestimated CO
ultimate analysis of rubber wood and rape straw is shown in Table 1. and overestimated CO2 concentration by a large margin. It is reasoned
Using the root mean square error (RMSE), the deviation between the that the bed temperature was insufficient for achieving the equilibrium
experimental and the simulation results were quantified. This was state, leading to underestimating CO and overestimating CO2
determined as per Eq. (6) (Upadhyay et al., 2019). concentration.
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ The second comparison between the simulated results and the test

(Xe − Xp )2 results of Jayah et al. (2003) and Striūgas et al. (2014) is shown in
RMSE = (6) Table 4. The observed RMSEs of 2.498 and 2.381 are within the
N
acceptable range since the simulation model neglects the gasification
Where Xe and Xp denote the test and the simulated result, respec­ systems’ kinetics and fluid dynamics. The numerical simulation model
tively, while N represents the number of data sets. yielded nearly 100% of CH4, which is an underestimated value
compared to the experimental value. Due to short residence time, actual
2.9. Performance parameters gasifiers cannot achieve thermodynamic equilibrium resulting in under-
forecasting CH4 (Zaman et al., 2020). This error analysis indicates a
The LHVSyngas, CGE, H2 yield, H2/CO ratio, and the percentage of reliable numerical simulation model in its predictive capacity.
combustible gas (PCG) are among the key parameters for assessing the
gasifier’s performance in response to the variation in the gasification 3.2. Sensitivity analysis
process’s operating conditions.
The lower heating value (LHVSyngas) refers to the quantity of heat 3.2.1. Effects of temperature
released by completely combusted syngas, excluding the heat of The effect of reactor temperature on syngas composition, H2 yield,
vaporization of water present in the products of combustion. The LHVSyngas, H2/CO ratio, CGE, and PCG is illustrated in Fig. 3. Temper­
LHVSyngas was determined by the composition of the combustible gases ature enhancement favors the production of CO-rich syngas while
in the product gas, such as methane, carbon monoxide, and hydrogen reducing the concentration of CO2 and CH4 (Fig. 3 (a)). For temperatures
and their respective lower heating values, as expressed in Eq. (7) between 350 ◦ C and 750 ◦ C, the concentration of H2 in the syngas and H2
(Kaewluan and Pipatmanomai, 2011). yield increased with a temperature rise. It then decreased with a further
temperature rise (Fig. 3 (a) and (b)). Similar behavior in H2, CH4, CO2
LHVSyngas (MJ/Nm3 ) = 35.81xCH4 + 12.63xCO + 10.79xH2 (7)
and CO concentrations was observed through a simulation study by Tan
Where Xi denotes the fractional volume on a dry basis of each syngas and Zhong (2010).
constituent. As observed from Fig. 3 (c) and (d), LHVSyngas, CGE, and PCG
CGE is the ratio of the product gas’ to the feedstock’s lower heating increased while H2/CO ratio decreased with a temperature rise. Han
value, as illustrated in Eq. (8) (Atnaw et al., 2013; Guangul et al., 2014). et al. (2017) reported similar behavior in LHVSyngas and H2/CO ratio in
their simulation study.
Qyield × LHVSyngas
CGE = × 100% (8) A temperature rise shifts the equilibrium of steam-methane reform­
LHVbm
ing reaction (R5), Boudouard reaction (R1), and water gas reaction (R2)
Where Qyield represents the syngas yield in (Nm3/kg), and LHVbm towards the formation of products resulting in increasing CO and H2
stands for the biomass’s lower heating value in (MJ/kg). while reducing CO2 and CH4 concentration in the syngas (Shen et al.,
Hydrogen yield and H2/CO ratio were calculated as per Eq. (9) 2008). The fluctuation in the concentration of H2 might result from the
(Hussain et al., 2017; Shen et al., 2008) and Eq. (10) (Han et al., 2017), mutual effect of the reactions occurring in the gasifier. Water-gas shift
respectively. reaction (R4) prevails at lower temperatures, producing H2-rich syngas

5
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

Fig. 2. Comparison between Aspen Plus simulation model and experimental results at, (a) temperature = 398 ◦ C, ER = 0.370; (b) temperature = 449 ◦ C, ER = 0.361;
and (c) temperature = 467 ◦ C, ER = 0.357.

Table 4
Comparison between the experimental and simulated results.
Run No Parameters Experiment (Jayah et al., 2003) Aspen Plus RMSE

MC (%) A/F CO (%) H2 (%) CO2 (%) CH4 (%) N2 (%) CO (%) H2 (%) CO2 (%) CH4 (%) N2 (%)

1 18.5 2.03 19.600 17.200 9.900 1.400 51.900 20.744 18.897 10.199 0.004 50.155 1.362
2 16.0 2.20 20.200 18.300 9.700 1.100 50.700 19.650 17.900 10.709 0.003 51.738 0.867
3 14.7 2.37 19.400 17.200 9.700 1.100 52.600 18.346 16.712 11.317 0.002 53.623 1.097
4 16.0 1.96 18.400 17.000 10.600 1.300 52.700 18.760 17.010 11.130 0.002 53.030 0.664
5 15.2 2.12 19.700 13.200 10.800 1.300 55.000 20.665 18.824 10.236 0.004 50.271 3.374
6 14.0 2.29 18.900 12.500 8.500 1.200 59.100 19.286 17.568 10.879 0.003 52.264 3.991
7 14.7 1.86 19.100 15.500 11.400 1.100 52.900 23.574 21.474 8.881 0.008 46.064 4.690
8 13.8 2.04 22.100 12.700 10.500 1.300 53.400 21.845 19.900 9.686 0.005 48.564 3.940
9 12.5 2.36 19.100 13.000 10.700 1.200 56.00 19.003 17.311 11.011 0.003 52.673 2.498
Average 2.498

MC ER Experiment (Striūgas et al., 2014) Aspen Plus


1 14.9 0.29 14.590 14.430 11.800 4.040 54.960 16.874 13.448 14.186 0.003 55.492 2.381

but is hindered at temperatures above 750 ◦ C. At temperatures beyond similar trend in their simulation study.
750 ◦ C, a reversed R4 is promoted, resulting in a decrease in H2 pro­
duction (Duan et al., 2015). Temperature enhancement favors H2 and 3.2.2. Effects of equivalence ratio
CO yields, resulting in increased LHVSyngas (Lahijani and Zainal, 2011) The effects of ER on the composition of syngas, H2 yield, LHVSyngas,
and PCG. CGE depends on the LHVSyngas; an increase in LHVSyngas results H2/CO ratio, CGE, and PCG are shown in Fig. 4. With ER enhancement,
in a corresponding increase in the CGE. Duan et al. (2015) reported a the concentration of H2, CO, CH4, and the H2 yield reduced while the

6
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

Fig. 3. The effect of gasification temperature on (a) syngas composition, (b) H2 yield, (c) LHVSyngas and H2/CO ratio, and (d) PCG and CGE.

concentration of CO2 increased (Fig. 4(a) and (b)). Fig. 4(c) and (d) show
a decrease in the LHVSyngas, CGE, and PCG and an increase in H2/CO PCG (%) = − 33.44 + 0.21673 Temp − 78.99 ER − 0.00009 Temp
ratio. Han et al. (2017) reported a similar trend in the behavior of × Temp + 189 ER × ER − 0.19745 Temp × ER (12)
LHVSyngas and the H2/CO ratio in their simulation study.
Increasing ER shifts the process towards complete combustion, H2 yield (g/kg) = − 69.58 + 0.24644 Temp + 86.33 ER − 0.000125 Temp
resulting in an increase in CO2 and a decrease in H2 and CO. Higher × Temp − 44.6 ER × ER − 0.16994 Temp × ER
temperatures hinder methanation reaction (R3), lowering the concen­
(13)
tration of CH4 in the syngas. Furthermore, R3 requires H2 as a reactant,
and its decrease leads to the production of low CH4-concentrated syngas. The ANOVA results for H2 yield and PCG are shown in Table 5.
Favas et al. (2017) observed a similar behavior in the concentration of The p-values for the overall model are zero with corresponding high
CO2, CO, and H2, while Lan et al. (2018) and Rupesh et al. (2016) F-values of 1067.28 and 3197.44 for the H2 yield and PCG regression
observed similar trends in CO2, CO, CH4, and H2 concentrations through model, respectively, showing the significance of the models. From the
their simulation studies. Syngas dilution by nitrogen attributes to the ANOVA results, the linear, square, and 2-way interaction terms are
lowering of LHVSyngas (Lahijani and Zainal, 2011). Similar behavior in significant. R2 values of 93.12% for H2 yield and 97.59% for PCG
LHVSyngas was observed through a simulation study by Favas et al. regression models were observed. The high R2 values show that the H2
(2017). yield and PCG regression models fit with acceptable precision into the
experimental results. Besides, the Adj-R2 values of 93.04% and 97.56%
3.3. Statistical analysis for H2 yield and PCG, respectively, are satisfactorily close to their cor­
responding R2 values. Those results show a minimal chance of incor­
3.3.1. Regression models and analysis of variance porating an insignificant term in the model. Therefore, the regression
The final model regression equations for PCG and H2 yield in terms of models can accurately determine the response variables.
the coded factors with significant variables are depicted in Eqs. (12) and
(13), respectively. 3.3.2. Multi-objective optimization
A synchronized effect of the critical parameters is studied to deter­
mine the operational point at which the maximum H2 yield and PCG can

7
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

Fig. 4. The effect of ER on (a) syngas composition, (b) H2 yield, (c) LHVSyngas and H2/CO ratio, and (d) PCG and CGE.

Table 5
ANOVA results for H2 yield and PGE.
H2 Yield PCG

Constant DF Adj SS F-Value P-Value Adj SS F-Value P-Value

Model 5 51014.4 1067.28 0.000 131,513 3197.44 0.000


Linear 2 29349.5 1535.07 0.000 112,086 6812.82 0.000
Temp 1 11784.3 1232.70 0.000 29,566 3594.14 0.000
ER 1 17565.2 1837.43 0.000 82,521 10031.50 0.000
Square 2 13946.3 729.43 0.000 9007 547.44 0.000
Temp*Temp 1 13846.0 1448.38 0.000 7207 876.11 0.000
ER*ER 1 100.3 10.49 0.001 1800 218.77 0.000
2-Way Interaction 1 7718.6 807.41 0.000 10,420 1266.67 0.000
Temp*ER 1 7718.6 807.41 0.000 10,420 1266.67 0.000
Error 394 3766.5 3241
Total 399 54780.9 134,754

Coefficient of determination
R2 93.12% 97.59%
Adj -R2 93.04% 97.56%

simultaneously be attained. The combined effect of reactor temperature temperature and ER is also observed in Fig. 5 (b). It is evident from Fig. 5
and ER on the H2 yield and PCG is illustrated in Fig. 5. A high value of H2 (c) that a PCG of greater than 70% can be attained at high gasification
yield (more than 40 g/kg of biomass) can be observed at high temper­ temperatures and low ER, as also observed in the 3-D surface plot (Fig. 5
atures and low ER, as illustrated in Fig. 5 (a). A similar trend in the (d)).
behavior of H2 yield with the synchronized effect of gasification Fig. 5 (e) shows that the white region represents the optimum zone.

8
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

Fig. 5. The synchronized effects of gasification temperature and ER on the H2 yield, PCG, and the graphical representation of the multi-objectively determined
optimal zone.

The optimum working point of the system can be selected from any­ biomass and PCG value of 75.93%, as obtained from Minitab. Consid­
where in the optimum zone. For instance, a point selected at random ering the same conditions, the Aspen Plus run yields 46.90 g/kg of
within the optimum zone and representing operation conditions of biomass and 75.60%% for H2 yield and PCG, respectively. These results
950 ◦ C and 0.06 gives a predicted H2 yield value of 46.39 g/kg of show that Aspen Plus model results agree with the values predicted by

9
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

Minitab software, thus validating the developed regression models with Cohce, M.K., Dincer, I., Rosen, M.A., 2010. Thermodynamic analysis of hydrogen
production from biomass gasification. Int. J. Hydrogen Energy 35, 4970–4980.
adequate precision.
https://doi.org/10.1016/j.ijhydene.2009.08.066.
Doherty, W., Reynolds, A., Kennedy, D., 2009. The effect of air preheating in a biomass
4. Conclusion CFB gasifier using ASPEN Plus simulation. Biomass Bioenergy 33, 1158–1167.
https://doi.org/10.1016/j.biombioe.2009.05.004.
Doranehgard, M.H., Samadyar, H., Mesbah, M., Haratipour, P., Samiezade, S., 2017.
This work developed a robust numerical model of air-gasification of High-purity hydrogen production with in situ CO 2 capture based on biomass
rice husks integrated with syngas purification and RSM; and successfully gasification. Fuel 202, 29–35. https://doi.org/10.1016/j.fuel.2017.04.014.
Duan, W., Yu, Q., Wang, K., Qin, Q., Hou, L., Yao, X., Wu, T., 2015. ASPEN Plus
validated with experimental results of rice husks, rubber wood, and rape
simulation of coal integrated gasification combined blast furnace slag waste heat
straw. ER and temperature were investigated. High temperature and low recovery system. Energy Convers. Manage. 100, 30–36. https://doi.org/10.1016/j.
ER favored H2 production and PCG. The generated H2 yield and PCG enconman.2015.04.066.
Faraji, M., Saidi, M., 2021. Hydrogen-rich syngas production via integrated configuration
regression models are found to accurately determine the response var­
of pyrolysis and air gasification processes of various algal biomass: Process
iables. Based on RSM analysis results, the optimum performance was simulation and evaluation using Aspen Plus software. Int. J. Hydrogen Energy 46,
observed at temperatures 820 ◦ C − 1090 ◦ C and ER of 0.06–0.10. In the 18844–18856. https://doi.org/10.1016/j.ijhydene.2021.03.047.
future, reaction kinetics and hydrodynamics could be incorporated to Favas, J., Monteiro, E., Rouboa, A., 2017. Hydrogen production using plasma gasification
with steam injection. Int. J. Hydrogen Energy 42, 10997–11005. https://doi.org/
enhance the proposed model’s accuracy. 10.1016/j.ijhydene.2017.03.109.
Gómez-Barea, A., Leckner, B., 2010. Modeling of biomass gasification in fluidized bed.
Prog. Energy Combust. Sci. 36, 444–509. https://doi.org/10.1016/j.
CRediT authorship contribution statement
pecs.2009.12.002.
Guangul, F.M., Sulaiman, S.A., Ramli, A., 2014. Study of the effects of operating factors
Emmanuel Yeri Kombe: Conceptualization, Investigation, Meth­ on the resulting producer gas of oil palm fronds gasification with a single throat
odology, Formal analysis, Validation, Data curation, Writing – original downdraft gasifier. Renewable Energy 72, 271–283. https://doi.org/10.1016/j.
renene.2014.07.022.
draft. Nickson Lang’at: Supervision, Methodology, Writing – review & Han, J., Liang, Y., Hu, J., Qin, L., Street, J., Lu, Y., Yu, F., 2017. Modeling downdraft
editing. Paul Njogu: Supervision, Writing – review & editing. Reiner biomass gasification process by restricting chemical reaction equilibrium with Aspen
Malessa: Methodology, Data curation, Writing – review & editing, Plus. Energy Convers. Manage. 153, 641–648. https://doi.org/10.1016/j.
enconman.2017.10.030.
Project administration. Christian-Toralf Weber: Resources, Supervi­ Heidenreich, S., Foscolo, P.U., 2015. New concepts in biomass gasification. Prog. Energy
sion. Francis Njoka: Visualization, Writing – review & editing. Ulrich Combust. Sci. 46, 72–95. https://doi.org/10.1016/j.pecs.2014.06.002.
Krause: Supervision, Software. Hussain, M., Tufa, L.D., Yusup, S., Zabiri, H., Taqvi, S.A., 2017. Aspen Plus® Simulation
Studies of Steam Gasification in Fluidized Bed Reactor for Hydrogen Production
Using Palm Kernel Shell. In: Mohamed Ali, M.S., Wahid, H., Mohd Subha, N.A.,
Sahlan, S., Md Yunus, M.A., Wahap, A.R. (Eds.), Modeling, Design and Simulation of
Declaration of Competing Interest Systems, Communications in Computer and Information Science. Springer
Singapore, Singapore, pp. 628–641. https://doi.org/10.1007/978-981-10-6463-0_
The authors declare that they have no known competing financial 54.
Hussain, M., Zabiri, H., Uddin, F., Yusup, S., Tufa, L.D., 2021. Pilot-scale biomass
interests or personal relationships that could have appeared to influence
gasification system for hydrogen production from palm kernel shell (part A): steady-
the work reported in this paper. state simulation. Biomass Conv. Bioref. doi: 10.1007/s13399-021-01474-1.
Jayah, T.H., Aye, L., Fuller, R.J., Stewart, D.F., 2003. Computer simulation of a
downdraft wood gasifier for tea drying. Biomass Bioenergy 25, 459–469. https://doi.
Data availability
org/10.1016/S0961-9534(03)00037-0.
Kaewluan, S., Pipatmanomai, S., 2011. Potential of synthesis gas production from rubber
No data was used for the research described in the article. wood chip gasification in a bubbling fluidised bed gasifier. Energy Convers. Manage.
52, 75–84. https://doi.org/10.1016/j.enconman.2010.06.044.
Kang, K., Azargohar, R., Dalai, A.K., Wang, H., 2015. Noncatalytic Gasification of Lignin
Acknowledgments in Supercritical Water Using a Batch Reactor for Hydrogen Production: An
Experimental and Modeling Study. Energy Fuels 29, 1776–1784. https://doi.org/
10.1021/ef5027345.
This study was funded by the Federal Ministry for Economic Coop­
Kate, G.U., Chaurasia, A.S., 2018. Gasification of rice husk in two-stage gasifier to
eration and Development (BMZ) and accorded to EYK for his doctoral produce syngas, silica and activated carbon. Energy Sources Part A 40, 466–471.
study (Grant Ref. 91672311). The authors also express gratitude for the https://doi.org/10.1080/15567036.2017.1423418.
Kombe, E.Y., Lang’at, N., Njogu, P., Malessa, R., Weber, C.-T., Njoka, F., Krause, U.,
access to computing resources provided by the Otto-von-Guericke-
2022. Numerical investigation of sugarcane bagasse gasification using Aspen Plus
University, Instrumental and Environmental Technology Department. and response surface methodology. Energy Convers. Manage. 254, 115198 https://
doi.org/10.1016/j.enconman.2021.115198.
Kumar, A., Samadder, S.R., 2017. A review on technological options of waste to energy
Appendix A. Supplementary data for effective management of municipal solid waste. Waste Manage. 69, 407–422.
https://doi.org/10.1016/j.wasman.2017.08.046.
Supplementary data to this article can be found online at https://doi. Lahijani, P., Zainal, Z.A., 2011. Gasification of palm empty fruit bunch in a bubbling
fluidized bed: A performance and agglomeration study. Bioresour. Technol. 102,
org/10.1016/j.biortech.2022.127734.
2068–2076. https://doi.org/10.1016/j.biortech.2010.09.101.
Lan, W., Chen, G., Zhu, X., Wang, X., Liu, C., Xu, B., 2018. Biomass gasification-gas
References turbine combustion for power generation system model based on ASPEN PLUS. Sci.
Total Environ. 628–629, 1278–1286. https://doi.org/10.1016/j.
scitotenv.2018.02.159.
Abuadala, A., Dincer, I., 2010. Efficiency evaluation of dry hydrogen production from
Li, S., Zheng, H., Zheng, Y., Tian, J., Jing, T., Chang, J.-S., Ho, S.-H., 2019. Recent
biomass gasification. Thermochim Acta 507–508, 127–134. https://doi.org/
advances in hydrogen production by thermo-catalytic conversion of biomass. Int. J.
10.1016/j.tca.2010.05.013.
Hydrogen Energy 44, 14266–14278. https://doi.org/10.1016/j.
Adeniyi, A.G., Ighalo, J.O., Amosa, M.K., 2021. Modelling and simulation of banana
ijhydene.2019.03.018.
(Musa spp.) waste pyrolysis for bio-oil production. Biofuels 12, 879–883. https://doi.
Mojaver, P., Khalilarya, S., Chitsaz, A., 2019. Multi-objective optimization using
org/10.1080/17597269.2018.1554949.
response surface methodology and exergy analysis of a novel integrated biomass
Atnaw, S.M., Sulaiman, S.A., Yusup, S., 2013. Syngas production from downdraft
gasification, solid oxide fuel cell and high-temperature sodium heat pipe system.
gasification of oil palm fronds. Energy 61, 491–501. https://doi.org/10.1016/j.
Appl. Therm. Eng. 156, 627–639. https://doi.org/10.1016/j.
energy.2013.09.039.
applthermaleng.2019.04.104.
Aydin, E.S., Yucel, O., Sadikoglu, H., 2017. Development of a semi-empirical equilibrium
Okolie, J.A., Nanda, S., Dalai, A.K., Kozinski, J.A., 2020. Optimization and modeling of
model for downdraft gasification systems. Energy 130, 86–98. https://doi.org/
process parameters during hydrothermal gasification of biomass model compounds
10.1016/j.energy.2017.04.132.
to generate hydrogen-rich gas products. Int. J. Hydrogen Energy 45, 18275–18288.
Bakar, R.A., Yahya, R., Gan, S.N., 2016. Production of High Purity Amorphous Silica
https://doi.org/10.1016/j.ijhydene.2019.05.132.
from Rice Husk. Procedia Chem. 19, 189–195.
Roy, D., Samanta, S., Ghosh, S., 2019. Techno-economic and environmental analyses of a
Bridgwater, A.V., 2003. Renewable fuels and chemicals by thermal processing of
biomass based system employing solid oxide fuel cell, externally fired gas turbine
biomass. Chem. Eng. J. 91, 87–102. https://doi.org/10.1016/S1385-8947(02)
00142-0.

10
E.Y. Kombe et al. Bioresource Technology 361 (2022) 127734

and organic Rankine cycle. J. Cleaner Prod. 225, 36–57. https://doi.org/10.1016/j. Tan, W., Zhong, Q., 2010. Simulation of Hydrogen Production in Biomass Gasifier by
jclepro.2019.03.261. ASPEN PLUS. In: 2010 Asia-Pacific Power and Energy Engineering Conference. IEEE,
Roy, D., Samanta, S., Ghosh, S., 2020. Performance optimization through response Chengdu, China, pp. 1–4. https://doi.org/10.1109/APPEEC.2010.5449270.
surface methodology of an integrated biomass gasification based combined heat and Tavares, R., Monteiro, E., Tabet, F., Rouboa, A., 2020. Numerical investigation of
power plant employing solid oxide fuel cell and externally fired gas turbine. Energy optimum operating conditions for syngas and hydrogen production from biomass
Convers. Manage. 222, 113182 https://doi.org/10.1016/j.enconman.2020.113182. gasification using Aspen Plus. Renewable Energy 146, 1309–1314. https://doi.org/
Rupesh, S., Muraleedharan, C., Arun, P., 2016. ASPEN plus modelling of air–steam 10.1016/j.renene.2019.07.051.
gasification of biomass with sorbent enabled CO2 capture. Resour.-Effic. Technol. 2, Upadhyay, D.S., Sakhiya, A.K., Panchal, K., Patel, A.H., Patel, R.N., 2019. Effect of
94–103. https://doi.org/10.1016/j.reffit.2016.07.002. equivalence ratio on the performance of the downdraft gasifier – An experimental
Safarian, S., Unnthorsson, R., Richter, C., 2022. Hydrogen production via biomass and modelling approach. Energy 168, 833–846. https://doi.org/10.1016/j.
gasification: simulation and performance analysis under different gasifying agents. energy.2018.11.133.
Biofuels 13 (6), 717–726. Woolcock, P.J., Brown, R.C., 2013. A review of cleaning technologies for biomass-
Seçer, A., Hasanoğlu, A., 2020. Evaluation of the effects of process parameters on derived syngas. Biomass Bioenergy 52, 54–84. https://doi.org/10.1016/j.
co–gasification of Çan lignite and sorghum biomass with response surface biombioe.2013.02.036.
methodology: An optimization study for high yield hydrogen production. Fuel 259, Yoon, S.J., Son, Y.-I., Kim, Y.-K., Lee, J.-G., 2012. Gasification and power generation
116230. https://doi.org/10.1016/j.fuel.2019.116230. characteristics of rice husk and rice husk pellet using a downdraft fixed-bed gasifier.
Shen, L., Gao, Y., Xiao, J., 2008. Simulation of hydrogen production from biomass Renewable Energy 42, 163–167. https://doi.org/10.1016/j.renene.2011.08.028.
gasification in interconnected fluidized beds. Biomass Bioenergy 32, 120–127. Yusup, S., Khan, Z., Ahmad, M.M., Rashidi, N.A., 2014. Optimization of hydrogen
https://doi.org/10.1016/j.biombioe.2007.08.002. production in in-situ catalytic adsorption (ICA) steam gasification based on Response
Singh, D.K., Tirkey, J.V., 2021. Modeling and multi-objective optimization of variable air Surface Methodology. Biomass Bioenergy 60, 98–107. https://doi.org/10.1016/j.
gasification performance parameters using Syzygium cumini biomass by integrating biombioe.2013.11.007.
ASPEN Plus with Response surface methodology (RSM). Int. J. Hydrogen Energy 46, Zaman, S.A., Ghosh, S., 2021. A generic input–output approach in developing and
18816–18831. https://doi.org/10.1016/j.ijhydene.2021.03.054. optimizing an Aspen plus steam-gasification model for biomass. Bioresour. Technol.
Striūgas, N., Zakarauskas, K., Džiugys, A., Navakas, R., Paulauskas, R., 2014. An 337, 125412 https://doi.org/10.1016/j.biortech.2021.125412.
evaluation of performance of automatically operated multi-fuel downdraft gasifier Zaman, S.A., Roy, D., Ghosh, S., 2020. Process modeling and optimization for biomass
for energy production. Appl. Therm. Eng. 73, 1151–1159. https://doi.org/10.1016/ steam-gasification employing response surface methodology. Biomass Bioenergy
j.applthermaleng.2014.09.007. 143, 105847. https://doi.org/10.1016/j.biombioe.2020.105847.

11

You might also like