Biological H2(g) Production and Modelling with Computational Fluid
Dynamics (CFD)
RUKİYE ÖZTEKİN*, DELİA TERESA SPONZA
Department of Environmental Engineering,
Dokuz Eylül University,
Tınaztepe Campus, 35160 Buca/Izmir,
TURKEY
*Corresponding Author
Abstract: - In this study, bio-hydrogen gas [bio-H2(g)] production and modeling with a three-phase
computational fluid dynamics (CFD) model, heat and mass transfer of bio-hydrogen production, reaction
kinetics, and fluid dynamics; It was investigated by dark fermentation process in an anaerobic continuous plug
flow reactor (ACPFR). The three-phase CFD model was used to determine the bio-H2(g) production in an
ACPFR. The effect of different operating parameters, increasing hydrolic retention times (HRTs) (1, 2, 4, 8,
and 12 days), different pH values (4.0, 5.0, 6.0, 7.0, and 8.0), and increasing feed rate as organic loading rates
(OLRs) (0.5, 1.0, 2.0, 4.0, 8.0 and 10.0 g COD/l.d) on the bio-H2(g) production rates were operated in
municipal sludge wastes (MSW) with Thermoanaerobacterium thermosaccharolyticum SP-H2 methane
bacteria during dark fermentation for bio-H2(g) production. The effect of HRT, pH, and feed rate on the bio-
H2(g) efficiencies and H2(g) production rates were examined in the simulation stage. Production of volatile
fatty acids (VFAs) namely, acetic acids, butyric acids, and propionic acids were important points influencing
the bio-H2(g) production yields. The artificial neural network (ANN) model substrate inhibition on bio-H2(g)
production to the methane (CH4) bacteria was also investigated. The reaction kinetics model used Thermotoga
neapolitana microorganisms with the Andrews model of substrate inhibition. Furthermore, the ANN model was
well-fitted to the experimental data to simulate the bio-H2(g) production from chemical oxygen demand (COD).
Key-Words: - Acetic acids; Anaerobic continuous plug flow reactor (ACPFR); Artificial neutral network
(ANN) model; Biological hydrogen gas [Bio-H(g)] production; Butyric acids; Computational
fluid dynamics (CFD) model; Propionic acids; Thermoanaerobacterium thermosaccharolyticum
SP-H2; Thermotoga neapolitana, Volatile fatty acids (VFAs).
Received: August 18, 2022. Revised: October 15, 2023. Accepted: November 17, 2023. Published: December 12, 2023.
1 Introduction
The technologies of energy production based on
burning fossil fuels constitute the world's main
energy production source and cause pollution and
degradation of the natural environment, [1], [2].
Fossil fuels and non-renewable energy; It causes
environmental damage and climate change by
causing destruction in soil, water, and air. The
partial and complete combustion of fossil fuels
emits greenhouse pollutants like COx, NOx, SOx,
CxHy, ash, and other organic compounds in the
environment, [3]. For energy production based on
renewable resources; New clean technologies need
to be developed, [4], [5]. The development of these
technologies is supported by social pressure, carbon
neutrality requirements, the political environment,
and appropriate legal regulations. Because it reduces
carbon dioxide [CO2(g)] and methane [CH4(g)]
emissions into the air and has the potential to
improve the quality of life of future generations; It
fits the main theme of the decarboxylation approach.
An increase in new renewable energy expenditures,
technologies that include the energy use of biomass
to bio-H2(g) may provide a solution to these
challenges, [6].
H2(g) is nontoxic, colorless, odorless, [7],
tasteless, and the third most abundant element on
Earth, [8]. H2(g) as an environmentally friendly gas
is an energy carrier that could play a significant role
in the reduction of greenhouse gases, [9], [10]. Due
to water (H2O) production during the combustion
process, H2(g) is considered a clean fuel. H2(g) is
regarded as an ideal energy with a high energy yield
of 122 kJ/g, which is 2.75 folds greater than that of
hydrocarbon fuels, [11]. For this reason, H2(g) is
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
321
Volume 20, 2023
one of the cleanest and most promising energy
sources.
In comparison with conventional anaerobic
process [fermentative CH4(g) production], due to
some inconsistency and drawbacks, the H2(g)
production processes by dark fermentation are less
well developed than the CH4(g) production. During
anaerobic digestion of organic wastes, such as solid
waste and wastewater, CH4(g) is produced and its
production processes have been well established
commercially. H2(g) is a more valuable energy
carrier and chemical feedstock compared with
CH4(g), [12], [13]. Therefore, dark fermentation can
treat agro-industrial effluents and also valorized
them through energy production, [12]. However, a
bottleneck for the widespread utilization of these
processes is the relatively low hydrogen production
rates (HPR), and thus several strategies to enhance it
have been proposed, [14]. One of them is the use of
feedback control, which is the case explored in this
communication. Different ways of H2(g) production
methods are summarized at Figure 1 (Appendix).
Bio-H2(g) is a feasible, promising, and clean
alternative fuel with no CO2(g) emissions and high
energy content per unit weight (141.9 J/kg);
Additionally, only H2O is produced as a result of
H2(g) combustion, [15]. Using the dark fermentation
process, organic wastes can be decomposed into
bio-H2(g), CO2(g), and VFAs, and converted into
metabolites, which can be utilized in other
fermentation processes based on the carboxylate
platform, [16]. Considering the conventional
treatment costs and the energy and chemical needs
in the processes used, by reducing the carbon
footprint; Dark fermentation, which produces clean
and renewable H2(g) from wastewater, is an
alternative to reduce fossil fuel consumption, [17]. It
compared different methods applied for H2(g)
production, such as photo-fermentation, dark
fermentation, electrolysis, electrodialysis and
photocatalysis, which include environmental,
economic, energy and energetic impacts, [18].
According to Bio-H2(g) production methods, the
most economical and efficient process occurs in
dark fermentation (Figure 2, Appendix).
Biological H2(g) production offers many
advantages, such as clean gas, simple technology,
and cheap high-intensity energy (122 kJ/g).
Additionally, its use does not produce any
greenhouse gases and has some significant
economic and environmental advantages. Biofuels
are considered solid, liquid, and gaseous fuels
produced predominantly using biomass. Various
fuels such as ethanol, methanol, H2(g), and CH4(g)
can be obtained from biomass, [19]. Plant biomass,
from which biofuels are produced, accumulates
solar energy. First-generation biofuels are produced
using traditional methods such as fermentation or
esterification, which do not require large amounts of
energy. In the case of traditionally so-called first-
generation biofuels, production is based on edible
plants such as sugar beet root, corn, sugarcane,
cereals, potatoes (starch), or vegetable oils (such as
rapeseed, palm, or jatropha). In contrast, bioethanol
is produced through alcoholic fermentation, while
biodiesel is produced through the esterification of
vegetable oils. Bio-H2(g) production can be carried
out in batch, fed-batch, and continuous modes.
Reactor types for bio-H2(g) production can be
grouped as open and closed systems. Closed
fermentation systems can be tubular reactors, bubble
columns, or airlift systems. A bioreactor used in
H2(g) production significantly affects the efficiency
and effectiveness of H2(g) production.
Biological methods allow us to cost-effectively
produce bio-H2(g) via dark fermentation;
Additionally, the photo-fermentation process can
also be applied to produce bio-H2(g) from various
sources, [20], [21]. However, production efficiency
largely depends on temperature, pH, and light
intensity. Another method worth mentioning is the
combined photo and dark fermentation method for
bio-H2(g) production, which can increase the
production efficiency by 20% to 189%, [22]. Two-
stage hybrid processes were applied to produce bio-
H2(g) from diluted solid waste, [23]. Typically, bio-
H2(g) production relies on a continuously stirred
tank reactor; Tube anaerobic packed bed reactor can
be considered a promising technology for bio-H2(g)
production because a high organic loading rate can
be achieved by using recirculation and a large
surface area to ensure better microorganism contact,
[23].
The last few years have seen an increase in the
use of more advanced techniques such as
computational fluid dynamics (CFD) for the design
and optimization of wastewater treatment systems.
[24], [25], [26]. Bio-H2(g) production CFD
simulation tool is used to simulate all kinds of
complex problems arising from the variability of
parameters of biological models. CFD methods
enable the determination of variables such as
volume fraction, shear strain rate, or turbulent
kinetic energy. They also facilitate the reliable
prediction of the relevant hydrodynamic variables,
the computational times, and fluid dynamics
coupling, as well as mass transfer and kinetic
variables. Additionally, the influence of inner
geometry and mass phase transfer can be used in
numerical simulations, [27]. Fluid particles pass
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
322
Volume 20, 2023
through the reactor with little or no longitudinal
mixing and exit from the reactor in the same
sequence in which they entered. Their identity
remains in the reactor for a time equal to the
theoretical detention time. This type of flow is
approximated in long open tanks with a high
length/width ratio where longitudinal dispersion is
minimal or absent, [28]. Operational factors to be
considered in selecting the type of reactor to be used
in the treatment process include the nature of the
wastewater to be treated, the nature of the reaction
(homogeneous/heterogeneous), the reaction kinetics
that governs the treatment process, process
performance requirements, and local environmental
conditions.
The integration of physical and biological
processes still poses great challenges. Although it is
necessary to develop and apply new methods to
improve reactor hydrodynamics, heat, and mass
transfer, [29], there are very few publications on
modeling in packed bed bio-H2(g) production, [30],
[31]. CFD methods can be used to optimize the
reactor configuration and therefore improve the
performance of a bio-H2(g) production reactor. CFD
software is able to predict the hydrodynamics of
fluid flow, heat and mass transfer, chemical
reactions in a reactor, and other related events by
solving a series of partial differential equations.
Processes that describe mass, momentum, energy,
and species balances are advanced methods that are
widely and successfully used, [32], [33]. CFD is an
effective tool for hydrodynamic-biokinetic analysis
of anaerobic digestion of partial differential
equations, [34]. It is complex because it requires a
complex biological kinetic process using the user-
defined function (UDF) written in C. It also allows
optimizing performance costs without increasing the
cost of prototyping, [35]. To see the flow pattern of
the liquid, CFD can be used, which focuses on
predicting hydrodynamic patterns through porous
media. Flow dispersity patterns through a packed
bed reactor resulted in arranged spherical particles
using the discrete element method (DEM) coupling
with a biokinetics model, which is an efficient
method to predict bio-H2(g) concentration in a
continuous tube reactor, [36]. Parameters such as
HRT, H2(g) production rate, substrate conversion,
pressure drop, and flow dispersity can be calculated
by CFD to see the flow pattern of liquid through
porous media, [37].
The artificial neural network (ANN) model is
used to examine relationships in complex nonlinear
data due to the data classification and learning
ability of the ANN model; It is a good tool that is
widely used and works according to human nervous
systems and brain, [38], [39], [40], [41]. In the last
decade, ANN models have been used in
environmental engineering fields such as biological
treatment of wastewater, membrane filtration,
pollution adsorption, and electrodialysis of salt
water, [38], [39], [40], [41].
Until now, there are few parameters for online
monitoring in bioreactors, the most frequent are
temperature, pH, oxidation-reduction potential
(ORP), dissolved oxygen (DO), and dissolved CO2.
A useful approach is the use of mathematical
models with these online determinations for the
estimation of the fermentative products. For this
purpose, the ANN has been successfully used, since
they are based on the connectivity of biological
neurons that have an incredible capability for
emulation, analysis, prediction, association, and
adaptation, [12], [42]. For instance, pH,
temperature, and NaCl concentration were used to
estimate maximum specific growth rate and
bacteriocin production in Streptococcus
macedonicus ACA-DC 198 cultures using
feedforward ANNs, [43]. By applying a recurrent
neural network, DO, feeding rate, and liquid volume
were used to determine biomass concentration in
Saccharomyces cerevisiae cultures, [42]. ORP, and
backpropagation neural network were used to
predict ethanol and biomass production in non-
axenic cultures, [44].
Biomass gasification, as an attractive technology
for the conversion of various types of biowastes to
energy, is known to be a sustainable procedure to
produce H2(g), [45], [46]. The gasification of
biowastes has been investigated in several research
works from the view point of performance analysis,
[47], [48], [49], [50], [51], [52], [53], [54], [55],
[56], [57]. Nevertheless, just a few works on
performance analysis of linked gasification–H2(g)
production have been reported, [58], [59], [60]. In
order to have a comprehensive analysis of a H2(g)
production system via water–gas shift reactors,
different modeling approaches based on
thermodynamic equilibrium, kinetics, CFD, and
ANNs can be developed. The models derived by
equilibrium approaches are independent of the
gasifier structure, so can be applied for ideal
systems and typical thermodynamic characteristics.
However, for a widely complex process, accurate
kinetic parameters are needed that are used in
kinetic modeling. In calculations relying on CFD, a
series of equations of energy, momentum, mass, and
species through a specific area of the gasifier are
solved simultaneously and can then predict the
distribution of temperature and concentration. The
methods based on ANN require a huge amount of
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
323
Volume 20, 2023
data and then use a set of mathematical regressions
for correlations among input and output data, [55],
[61], [62], [63], [64], [65], [66]. This method has
gained importance recently because it can estimate
nonlinear functions without the need for a
mathematical explanation of events in the system.
However, when critical interactions of complex
nonlinearities such as biomass conversion are
included in a data set, ANN models are attractive for
outcome prediction, [67], [68], [69], [70]. Therefore,
very little work has been reported on modeling
biomass gasification using the ANN model, and
nothing in the field of downdraft gasifiers coupled
with water-gas shift reactors for bio-H2(g)
production.
In this study, a new design project ACPFR
model to analyze the heat and mass transfer,
reaction kinetics, and fluid dynamics of bio-H2(g)
production through fermentation and full transient
three-phase CFD modeling of bio-H2(g) production
was investigated. Different operating parameters,
increasing HRTs (1, 2, 4, 8 and 12 days), different
pH values (4.0, 5.0, 6.0, 7.0 and 8.0) and feed rate
(0 .5, 1.0, 2.0, 4.0, 8.0) and 10.0 g COD/l.d) on bio-
H2(g) production rates, bio-H2(g) yields and H2(g)
for its production, it was run in MSW with
Thermoanaerobacterium thermosaccharolyticum
SP-H2 methane bacteria during anaerobic dark
fermentation in an ACPFR. Rates in the simulation
phase; The production of VFAs, namely acetic,
butyric, and propionic acids, are important points
affecting bio-H2(g) production efficiency.
Furthermore, the aim of this study is to develop an
ANN to predict H2(g) production in genetically
modified Thermoanaerobacterium
thermosaccharolyticum SP-H2 fermentations based
on online measurements of ORP, pH, and dissolved
CO2, respectively.
2 Materials and Methods
2.1 Microorganisms
The hydrogen-producing bacterium SP-H2 was
isolated from a thermophilic acidogenic reactor
inoculated with municipal sewage sludge and
processed a carbohydrate-rich simulated food waste.
Based on the 16S rRNA gene sequence, the
bacterium was identified as
Thermoanaerobacterium thermosaccharolyticum
SP-H2. The maximum growth rate was observed at
55–60°C and at optimum pH=7.5.
2.1.1 Inoculum, Substrates, Mineral Medium
The effluent from the thermophilic acidogenic
reactor, inoculated with municipal sewage sludge
and treating high-strength simulated wastewater,
[71], was used to isolate a new bacterial strain
known as Thermoanaerobacterium
thermosaccharolyticum SP-H2.
Pfennig`s medium, [72], containing the
following: 330 mg/l NH4Cl, 500 mg/l MgCl2.6H2O,
168 mg/l CaCl2, 330 mg/l KCl, 330 mg/l KH2PO4
was used to isolate for Thermoanaerobacterium
thermosaccharolyticum SP-H2 bacteria. The
medium was supplemented with 500 mg/l yeast
extract, 2500 mg/l NaHCO3 as well as with a trace
element solution, [73], and a vitamin solution, [74].
0.5 g/l sodium sulfide and 0.5 g/l cysteine were used
as reducing agents. The final pH was 6.8–7.0.
2 ml of effluent was inoculated into an 18 ml
medium containing 5 g/l potato starch as the carbon
source in a 60 ml serum bottle and cultured in
anaerobic condition at 55°C. The enrichment culture
was sequentially sub-cultured into a series of 5–10
serum bottles (dilution 10–6- 10–11), introducing 10%
inoculum into each of them. From the last serum
bottle, in which the growth and formation of bio-
H2(g) were recorded, re-inoculation was repeated on
a series of serum bottles.
Incubation of Thermoanaerobacterium
thermosaccharolyticum SP-H2 was carried out at
55°C in the dark on a thermostat shaker with 100
rpm.
2.1.2 Conducting Batch Test for Microorganisms
Batch experiments were carried out in 60 ml serum
bottles with a working volume of ≈20 ml containing
2 ml (10% v/v) of the strain SP-H2 suspension at
exponential phase (OD600=0.8-1.0), 18 ml of
Pfennig`s medium, [72]. Cheese whey was added as
the carbon source. The volume of cheese whey
added was 5.0 ml, respectively. The chemical
oxygen demand (COD) of the added substrate in all
bottles was 3840 mg O2/l. The initial pH of the
medium was adjusted to 7.0 with a 10% solution of
HCl or NaOH. The serum bottles were closed with
rubber stoppers and aluminum caps and purged with
nitrogen gas to create anaerobic conditions. During
the dark fermentation, biogas components and
soluble metabolite products (SMPs) were
monitored. Control serum bottles did not contain
any carbon source, except endogenous carbon from
cell suspension and some components (yeast extract,
cysteine) of the medium. The hydrogen production
in the control bottles were subtracted from the
hydrogen production in treatment bottles to obtain
the actual hydrogen production from wastewater by
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
324
Volume 20, 2023
SP-H2. Incubation was carried out at 55°C in the
dark on a thermostat shaker with 100 rpm. All
treatments were conducted in triplicates.
2.2 Experimental Set-up
The experiments were conducted by an up-flow
anaerobic continuous plug flow reactor (ACPFR)
packed with immobilization materials. ACPFR is a
tube reactor. The tube reactor diameter was 10 cm,
whereas the length of the reactor was 45 cm to
achieve a total working volume of 3.53 liter. The
outer shell, which is a cylinder with a central
rotation axis, was also considered. A packed bed
reactor should have 4 times more length than
diameter to achieve the best-packed biofilm
composition. The volume of the reactor was 3532.5
cm3 (3.53 l). The packed bed reactor density was
0.94 g/cm3, and its porosity was 85%. The geometry
was divided into a number of discrete cells, and the
governing equations were solved numerically until
the time-step horizon was converged for a transient
study.
The DEM coupled with CFD was used to create
a packed bed with a spherical particle tube reactor.
Bio-H2(g) production on the immobilized culture
was via a continuously operated biofilm as a 2.1 mm
thickness layer on the packed bed. The inlet was a
liquid, while gaseous effluent was collected at the
top side of the reactor. A mesh refinement study was
made during grid spacing by reducing the factor
from 0.5 to 0.0025 mm until the results were
unremarkably changed with the grid size reduction.
It was found that 0.015 mm is fine enough to obtain
grid-independent results. The total number of mesh
elements was 0.15984 ml. Overall, the mesh
provided the best accuracy and was adopted for the
CFD simulation. To estimate the grid convergence
uncertainty of the CFD solution, this study used the
grid convergence index (GCI) method based on the
Richardson extrapolation. The initial wall boundary
y + spacing remained the same for each grid
refinement level.
2.3 Analytical Methods
The biogas composition was analyzed by a gas
chromatography–mass spectrometry (GC-MS); a
gas chromatograph (GC) (Agilent Technology
model 6890N) equipped with a mass selective
detector (Agilent 5973 inert MSD) by injecting a
sample volume of 2 ml. Mass spectra were recorded
using a VGTS 250 spectrometer equipped with a
capillary SE 52 column (HP5-MS 30 m, 0.25 mm
ID, 0.25 μm) at 220°C with an isothermal program
for 10 min. The initial oven temperature was kept at
50oC for 1 min, then raised to 220oC at 25oC/min
and from 200 to 300oC at 8oC/min, and was then
maintained for 5.5 min. High purity Helium [He(g)]
was used as the carrier gas at constant flow mode
(1.5 ml/min, 45 cm/s linear velocity). The
calibration was carried out with a standard gas
composed of 25% CO2(g), 2% O2(g), 10% N2(g) and
63% CH4(g), respectively.
Bio-H2(g) was measured with GC-MS (Agilent
6890N GC - Agilent 5973 inert MSD) according to
the above operating parameters. In addition to VFAs
(acetic, butyric, and propionic acids) were analyzed
at the end of the experimental set-up. After
centrifugation (13000 rpm, 30 min), VFA
concentrations were measured by same GC-MS
(Agilent Technology model 6890N GC - Agilent
5973 inert MSD). The gas carrier of the flow was
nitrogen [N2(g)].
All experimental parameters were measured
according to the Standard Methods (2022), [28].
2.4 Kinetics Model for Bio-H2(g) Production
The reaction kinetic model was created using the
Andrews substrate inhibition model with
Thermotoga neapolitana microorganisms, [75].
Thermotoga neapolitana is a rod-shaped, gram-
negative bacterium, [76], distinguishable by a thick
periplasmic cell wall, [77]. They are generally 0.2-5
μm in size, but can also reach sizes up to 100 μm. It
is sporeless, with its rod shape and gram-negative
features, and is characteristic of the order
Thermotogales, [77]. Thermotoga neapolitana is
considered thermophilic, with a habitable
temperature range of 50–95oC. The optimum
temperature is 77oC, making it almost
hyperthermophilic, [77]. There is also evidence that
it can be found in saline environments due to its
ability to grow in moderately halophilic
environments, [78].
The kinetic model of bio-H2(g) production was
used Thermotoga neapolitana bacteria with the
Andrews model of substrate inhibition in Eq. (1),
[75]:



(1)
where;
: is the maximum H2(g) specific
production rate (mmol H2/g.h), S: is the substrate
concentration (g/l) and Ks: is the inhibition constant
(g/l), respectively.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
325
Volume 20, 2023
2.4.1 Computational Fluid Dynamic (CFD)
Model
The kinetics conservation equation was
implemented in the CFD software by UDF function
written in C code. The model contains protein,
which makes up 27-30% of the dry weight of
Thermotoga neapolitana bacteria. A numerical CFD
study of hydrodynamics-biokinetics aspects for
interactions between multiphase is an important part
of how hydrodynamics influence biokinetics for bio-
H2(g) production. This numerical model contains
kinetics information, which is applicable to a set of
process parameters, thereby complicating process
analysis and the design of the fermentation system.
Adoption of kinetics models is also complicated
across different studies making application of
process analysis and design of fermentation system.
CFD modeling involves the use of numerical
methods and algorithms to solve the fundamental
governing equations of fluid dynamics (i.e.
continuity, momentum, and energy equations). In
traditional CFD software, the solutions to these
equations are found by solving a set of partial
differential equations called the Navier-Stokes
equations. The Navier-Stokes equations describe the
motion of a fluid and how the pressure, velocity,
temperature, and density of a moving fluid are
related. For a three-dimensional (3D) system, they
consist of one continuity equation for the
conservation of mass, three equations for the
conservation of momentum, and one equation for
the conservation of energy, [79].
The continuity equation is presented in Eq. (2),
and it states that the mass in the control volume
cannot be created, destroyed, or transformed:
 (2)
where, ρ: is the density, t: is the time, and (.V): is
the divergence of the velocity vector field,
respectively.
Traditional CFD packages are used to solve
partial differential equations related to fluid flow; It
uses finite volume, finite difference, or finite
element methods. The use of advanced methods for
simulating fluid flows, such as the mesh Boltzmann
method, [80], and the computational fluid
dynamics/discrete element method, [81], or
meshless methods, such as smoothed particle
hydrodynamics, [82], has increased over the last few
years.
The use of CFD to study and optimize slurry
anaerobic digesters has already been undertaken by
several authors, [83], [84], [85]. In general, the
hydrodynamics of common sludge anaerobic
digesters can be adequately simulated assuming
single-phase (liquid) or two-phase (gas/liquid) flow,
as shown in most studies reviewed, [86]. A common
approach to model the rheological behaviour of
slurry digesters is to use a non-Newtonian model for
the liquid phase. This allows the model to account
for the effects of the total solids content in the
viscosity of the wastewater without having to
include a solid phase in the model, hence reducing
the number of phases to be simulated, [84], [87],
[88]. In general, this is a reasonably good
assumption as the size of the solids dispersed on the
flow is very small compared to the size of the
reactors.
For high-rate anaerobic granular sludge reactors
(AGSRs), such as up-flow anaerobic sludge blanket
(UASB) reactors, expanded granular sludge bed
(EGSB) reactors, and internal circulation (IC)
reactors; The role of biogas bubbles, the influence
of the granules and the influence of the mixture
should not be underestimated. UASB reactors are
thought to be self-mixing by the upstream
movement of biogas bubbles and liquid flow
through the reactor, [89]. Additionally, high-rate
systems can retain biomass granules (high solids
residence time). This is ensured by a combination of
reactor design, settling characteristics of the
granules, and liquid up-flow velocity. An accurate
CFD model would include the effects of biogas
bubbles on the overall flow characteristics. It would
also capture the effects of increased flow rates of
wastewater and biogas on the loss of biomass
(sludge wash-out). In this context, CFD simulations
stand out as a tool capable of aiding in the design
and study of AGSRs by allowing for design
iterations for optimizations without the need to
construct and build reactors.
The use of CFD for the simulation of AGSRs has
begun in the last two decades; however, multiphase
simulations are generally computationally
demanding, especially when a granular (solid) phase
is included. Furthermore, a CFD model should only
be used to guide real-life design decisions if it has
been carefully verified and validated, [90].
Therefore, knowing the state of the art in terms of
previous studies and validated models will enable
the development of further research.
Previously, published studies in the field focused
on CFD applied to anaerobic digesters in a more
general approach, without focusing on granular
reactors, [26], [27], [86], [91]. Other reviews
focused on aspects such as protocols for the
simulations and validation of the models, [24], [92].
The use of CFD applied to specific tasks such as
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
326
Volume 20, 2023
modeling of mixing in anaerobic digestion reactors
has been the core of some reviews, [93], [94], [95].
Modeling of anaerobic granular reactors has been
studied from the perspective of hydrodynamics and
general modeling, [96], [97]. Modeling of granular
sludge reactors (aerobic, anaerobic, and nitriding-
anammox) is reviewed from a mechanistic modeling
perspective (i.e., mass balance-based models with
transport and reaction terms), [25]. Some studies
have reported using CFD to study the transport of
solid, liquid and gas phases, [25]. While we discuss
the forces involved in momentum transfer between
phases, no details are given about CFD modeling.
2.4.2 Artificial Neural Network (ANN) Model
An ANN model coming from the simulation results
for the considered gasification system, relying on
features and output matrixes, is established. The
research aim is to develop an ANN model linked
with an equilibrium for the estimation of the specific
mass flow rate of H2(g) production (smH2) from
different operating conditions (HRTs, pH, and feed
rate). Then, an attempt is made to investigate the
relative impact of biomass properties and operating
parameters on smH2. At the end, to have a
comprehensive analysis, variations of the inputs on
smH2 regarding H2(g) content are compared and
analyzed together.
The ANNs always consist of three layers
including (i) input, (ii) hidden, and (iii) output
layers. The outputs of a neuron are calculated using
Eq. (3):

 (3)
where, n: is the input number, xj: is the jth input to
the neutron, ωj: is the jth synaptic weight, and f: is a
non-linear function, respectively.
For converting output data between 1 and + 1,
the hyperbolic tangent formula was applied as Eq.
(4):
󰇛󰇜
 (4)
During the training process of input and output
data set, the network weights are adjusted to achieve
the similar outputs as seen in the training data set.
For this purpose, the data were divided into two
subsets for training model and validation purposes.
The Pearson correlation coefficient (r2) and mean
standard error (MSE) were computed to evaluate the
performance of the developed models according to
the following formulas, [98], as Eq. (5) and Eq. (6):




(5)

 
 (6)
In order to avoid numerical overflows related to
very large or small weights, all of data were
converted to normalized values using as Eq. (7):
 󰇡 
󰇢  (7)
Due to lack of related studies, the performance
and efficiency of ACPFR for biological H2(g)
production and wastewater treatment in different
experimental conditions for bio-H2(g) production
were evaluated in the present study. In addition, the
ANN model was developed to predict the
performance of the ACPFR for wastewater
treatment and bio-H2(g) production.
ANN predicts the output of a process given the
values of process input and process control
variables, [99]. It is often used to model
relationships between large sets of varying data.
ANN can either feed the results to an operator to
make process control adjustments or implement
appropriate control adjustments automatically, [99].
Many researchers have used this type of approach
with great success and recommended the use of
neural network models especially when the exact
relationship between inputs and outputs is not
known and where strong non-linear relationships
exist. It was reported that the real-life anaerobic
process for biogas yield is very complex, [100], and
non-linear, [101], as well as highly dependent on
different substrate characteristics, [102], and various
operating conditions such as organic loading rate,
pH, retention time, carbon/nitrogen ratio,
temperature, pressure, agitation rate, etc., [103].
However, one way to understand the relationships
between the substrates’ characteristics and the
optimum biogas yield is through machine learning
facilitated by models and equations, [104].
However, determining an exact mathematical model
is rigorous because the relationships are very
complex and highly non-linear, [99].
2.5 Other Artificial Intelligence Techniques
Artificial intelligence has found a wide range of
applications in many fields such as environmental
sciences, agricultural sciences, basic and applied
sciences, anthropological studies, medical fields,
and general engineering family. To achieve
reasonable and logical results; They are used to
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
327
Volume 20, 2023
model, predict, and simulate processes. Modeling a
multivariate system; It is quite difficult due to the
complexity of processes that exhibit non-linear
behavior, which are difficult to describe with linear
mathematical models, [104]. Therefore, artificial
intelligence technology can predict nonlinear
relationships extremely quickly and reliably.
Advances in computing power are minimizing the
time required to develop models as well as the time
required to retrain models to incorporate new data
and reflect process changes. The four main types of
artificial intelligence approaches based on major
branches are summarized and shown in Figure 3
(Appendix), [105]:
Artificial intelligence can be developed without
quantifying the micro-scale interactions that occur.
In the anaerobic process, such interactions are often
poorly understood, thus making it impossible to
develop useful mechanistic process models. Instead
of this, several researchers have applied artificial
intelligence techniques to optimize the dark
fermentation process. While it is understood that the
application of artificial intelligence in biogas
production is still a growing phenomenon, its
application in bio-H2(g) production is still a very
new application.
Nature-inspired computing (NIC) is a recently
developed branch of artificial intelligence
techniques. Natural systems (living and non-living)
have an innate ability to evolve, often in parallel and
against each other, in a dialectic way. The harmony,
beauty, and vigor of life underlie this complexity of
evolution. Even without a central control, the
processing of information happens in a distributed,
self-organized, and optimal way. Equilibrium is
maintained in nature through optimal searching, and
this forms the basis of algorithm development for
optimization problems in process engineering.
Algorithms are iterative procedures for providing
calculations or guidelines in a step-wise manner
tailored for specific goals. Computational
optimization aims to create algorithms to design,
implement, and test for solving optimization
problems, [106].
Optimization works on various levels, including
maximization of performance, efficiency, and profit,
or minimization of energy and economics. If infinite
time were available, any problem could be solved,
but that is not the case with real situations. When
time and resources are constrained, intelligent
techniques are required. To address non-linear
systems like the dark fermentation process in an
ACPFR, computer simulation becomes an
indispensable tool.
3 Results and Discussions
3.1 CFD Model for Bio-H2(g) Production
from ACPFR
Three-dimensional, unstable, incompressible,
multiphase CFD tube reactor model to simulate bio-
H2(g) production from ACPFR; The Lagrangian-
Eulerian approximation was implemented with a
two-stage model using the appropriate Reynolds
stress closure solved by boundary conditions. The
fluid flow was laminar according to the general
criteria of Reynolds number (Re) at the inlet. The
fluid properties were constant except for the
formulation of the buoyancy term. The governing
equations of continuity are Eq. (8) and momentum
Eq. (9) can be written as follows:

 󰇛󰇜 (8)
 󰇛󰇜 󰇛󰇜  
(9)
where, ρ: is the volume average density, : is the
flow velocity, p: is the static pressure, : is the
stress tensor, ρ represents the gravitational body
force and
represents the external force.
The continuity equation for the gas phase is
shown as Eq. (10):

 
󰇍
󰇍
󰇍
󰇍
(10)
where, : is the gas density,
󰇍
󰇍
󰇍
󰇍
: is the gas velocity,
: is the interphase mass transfer terms for the gas-
solid interface reactions, and : is the volume
fraction of the gas phase.
The energy equation for a fluid region is given
by Eq. (11):
 󰇛󰇜 󰇛󰇜 󰇟󰇛 󰇜󰇠 (11)
where, ρ, k, T and Sh: are the density, molecular
conductivity, temperature and the volumetric heat
source, respectively; : is the flow velocity, while
: is the heat conductivity due to turbulent
transport.
The momentum equation for the gas phase is Eq.
(12):

󰇍
󰇍
󰇍
󰇍

󰇍
󰇍
󰇍
󰇍

󰇍
󰇍
󰇍
󰇍
󰇍
󰇍
󰇍
󰇍


(12)
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
328
Volume 20, 2023
where,
󰇍
󰇍
󰇍
󰇍
: is the second order stress tensor of the
gas,
: is the interaction force representing
momentum transfer between the gas and solid
phase, : is the gravitational force, and sm: is the
mass source.
The gas species is Eq. (13):


󰇍
󰇍
󰇍
󰇍
 
(13)
where, Deff: is the effective mass diffusion
coefficient, Yi: is the mass fraction of gas species, I,
: is the species source term from the particle,
and : is the species source term from reactions,
respectively.
The liquid phase of continuity and momentum
equations are Eq. (14) and Eq. (15):
 󰇛󰇜 󰇛󰇜 (14)
 󰇛󰇜 󰇛󰇜 
󰇛 󰇜 (15)
where, : is the void fraction of the fluid phase, :
is the density of the fluid phase, : is the velocity of
the fluid phase, Tv: is the momentum viscous tensor,
TR: is the Reynolds tensor and SM: is the momentum
of the source term.
3.2 Simulations of Numerical
The Lagrangian–Eulerian approach is adopted to
describe the biomass slurry flow behavior of the
liquid–gas phase in laminar flow. Both phases were
treated as a continuous process that intertwined and
interacted with each other in the computational
domain. The recirculation flow rate coefficient was
0.32. Initial and boundary conditions are shown in
Table 1. The operating temperature was between 30
and 45oC. This long-term simulation lasted 40 days.
The concentration of gaseous products (H2, CH4,
etc.) and soluble metabolites such as VFAs (acetic
acids, butyric acids, and propionic acids) were
evaluated at specified time intervals throughout all
operational phases.
All terms of the governing equations are discrete
using the second-order upwind scheme. The
PRESTO (pressure staggering option) algorithm
was used for the pressure-velocity coupling. The
Green-Gauss cell-based method was used for the
discretization of the gradient. Each case was
simulated in ANSYS Fluent 2016 software
(Cannonsburg, PA, USA) with the initialization
procedure for simulations with second-order
schemes. Convergence was evaluated based on the
low mass flow rate imbalance below 1.0x105 kg/s.
The second step was to generate an ANSYS Fluent
(Cannonsburg, PA, USA) journal file to
automatically run the flow case with the prescribed
boundary conditions from the algorithm for the
pressure, flow rate, and temperature. The numerical
simulations were performed on an i7-3770 CPU
3.70 GHz processer Intel computer with 16 GB
RAM and a 64-bit operating system.
3.3 The Effect of Increasing HRTs for Bio-
H2(g) Production at ACPFR
Increasing HRTs values (1, 2, 4, 8, and 12 days)
were examined in MSW with
Thermoanaerobacterium thermosaccharolyticum
SP-H2 methane bacteria during dark fermentation
for bio-H2(g) production in ACPFR, at 37oC (Fig.
4). 0.070, 0.109, 0.121 and 0.078 mg/l.h bio-H2(g)
production rates were measured for 1, 2, 4, and 12
days HRTs, respectively, in ACPFR after dark
fermentation process, at 37oC (Figure 4, Appendix).
The maximum 0.134 mg/l.h bio-H2(g) production
rate was observed for 8 days HRTs in ACPFR after
dark fermentation process at 37oC (Figure 4,
Appendix).
The bio-H2(g) yield in different fractions of
ACPFR was 0.201-0.567 mg/h/l. ACPFR reactor
temperature was 25oC and HRT was 2 hours. Bio-
H2(g) concentrations filled the packed bed
differently. The fluid velocity profile is laminar,
with maximum velocity occurring at the center of
the ACPFR. It was observed that the temperature
gradient occurred only at the ACPFR inlet due to
heat transfer from the ACPFR wall.
0.952, 1.067, 1.183 and 1.075 mol H2/mol
substrate bio-H2(g) yields were obtained for 1, 2, 4
and 12 days HRTs, respectively, in ACPFR after
dark fermentation process, at 37oC (Fig. 5). The
maximum 0.202 mol H2/mol substrate Bio-H2(g)
yield was found for 8 days HRTs in ACPFR after
dark fermentation process, at 37oC (Figure 5,
Appendix).
3.4 The Effect of Different pH Values for
Bio-H2(g) Production at ACPFR
Different pH values (4.0, 5.0, 6.0, 7.0 and 8.0) were
examined in MSW with Thermoanaerobacterium
thermosaccharolyticum SP-H2 methane bacteria
during dark fermentation for bio-H2(g) production in
ACPFR, at 37oC (Fig. 6). 0.94, 1.03, 0.85 and 0.69
kmol/m3 bio-H2(g) production values were
measured for pH=4.0, pH=5.0, pH=7.0, pH=8.0
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
329
Volume 20, 2023
respectively, in ACPFR after dark fermentation
process, at 37oC (Figure 6, Appendix). The
maximum 1.22 kmol/m3 bio-H2(g) production value
was found for pH=6.0 in ACPFR after dark
fermentation process at 37oC (Figure 6, Appendix).
The model of dark fermentation shows that the
most optimal pH for the bio-H2(g) production
process is pH=5.0-6.0. Lowering pH effectively
reduces the process of methanogenesis,
simultaneously increasing bio-H2(g) production.
3.5 The Effect of Increasing Feed Rate as
OLR for Bio-H2(g) Production at
ACPFR
Increasing feed rates as OLR values (0.5, 1, 2, 4, 8,
and 10 g COD/l.d) were operated in MSW with
Thermoanaerobacterium thermosaccharolyticum
SP-H2 methane bacteria during dark fermentation
for bio-H2(g) production in ACPFR, at 37oC (Figure
7, Appendix). 1.36, 1.22, 1.14, 0.81 mol H2/mol
substrate bio-H2(g) yields were observed for 1, 2, 4,
8 and 10 g COD/l.d OLR, respectively, in ACPFR
after dark fermentation process, at 37oC (Figure 7,
Appendix). The maximum 1.61 mol H2/mol
substrate bio-H2(g) yield was obtained for 0.5 g
COD/l.d OLR in ACPFR after dark fermentation
process at 37oC (Figure 7, Appendix).
The increase of OLR up to 8 g COD/l.d has a
negative effect on the biological H2(g) yield. This
situation may be due to VFAs accumulation at
higher OLR, in addition, the supersaturation of
H2(g) in liquid phase may be related to the lower
biogas production and relieving inhibition due to
bio-H2(g) production, [107].
3.6 The Measurements of VFAs after Bio-
H2(g) Production at ACPFR
Different VFAs (acetic acid, butyric acid, and
propionic acid) were observed in MSW with
Thermoanaerobacterium thermosaccharolyticum
SP-H2 methane bacteria during dark fermentation in
ACPFR, at 37oC (Figure 8, Appendix). 52%, 55%,
58%, and 61% acetic acid proportions were
measured for 1, 2, 4, and 8 days HRTs, respectively,
in ACPFR after dark fermentation, at 37oC (Figure
8, Appendix). 73% maximum acetic acid value was
found for 12 days HRTs in ACPFR after dark
fermentation process, at 37oC (Figure 8, Appendix).
18%, 22%, 19%, and 4% butyric acid
proportions were observed for 2, 4, 8, and 12 days
HRTs, respectively, in ACPFR after dark
fermentation process, at 37oC (Fig. 8). 25%
maximum butyric acid value was obtained for 1 day
HRTs in ACPFR after dark fermentation process, at
37oC (Figure 8, Appendix).
23%, 20%, 20%, and 22% propionic acid
proportions were measured for 1, 4, 8, and 12 days
HRTs in ACPFR after dark fermentation process, at
37oC (Figure 8, Appendix). The maximum 27%
propionic acid value was observed for 2 days HRTs
in ACPFR after dark fermentation process, at 37oC
(Figure 8, Appendix).
3.7 The Comparison between Experimental
Bio-H2(g) Production Yield and ANN
Values in ACPFR
A comparison between experimental values of bio-
H2(g) production yields and the predicted ANN
values (Figure 9, Appendix). The value of R2 for the
ANN model was found to be up to 0.98 (data not
shown). 0.952, 1.067, 1.183, 1.202, and 1.075 mol
H2/mol substrate bio-H2(g) production yields were
measured for 1, 2, 4, 8, and 12 days HRTs,
respectively, in ACPFR after dark fermentation
process, at 37oC (Figure 9, Appendix).
0.947, 1.056, 1.175, 1.199 and 1.071 mol H2/mol
substrate ANN values were observed for 1, 2, 4, 8,
and 12 days HRTs, respectively, from ANN
simulation (Figure 9, Appendix). The ANN was an
excellent model because of the lowest error and the
highest coefficient values. The obtained results
indicated that the simulation model based on the
ANN is practical.
4 Conclusions
The maximum is 0.134 mg/l.h bio-H2(g) production
rate was observed for 8 days HRTs in ACPFR after
dark fermentation process at 37oC.
The maximum 0.202 mol H2/mol substrate Bio-
H2(g) yield was found for 8 days HRTs in ACPFR
after dark fermentation process, at 37oC
The maximum 1.22 kmol/m3 bio-H2(g)
production value was found for pH=6.0 in ACPFR
after dark fermentation process at 37oC. The model
of dark fermentation shows that the most optimal
pH for the bio-H2(g) production process is
pH=5.0-6.0. Lowering pH effectively reduces the
process of methanogenesis, simultaneously
increasing bio-H2(g) production.
The maximum 1.61 mol H2/mol substrate bio-
H2(g) yield was obtained for 0.5 g COD/l.d OLR in
ACPFR after the dark fermentation process at 37oC.
73% maximum acetic acid value was found for
12 days HRTs in ACPFR after dark fermentation
process, at 37oC.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
330
Volume 20, 2023
25% maximum butyric acid value was obtained
for 1 day HRTs in ACPFR after dark fermentation
process, at 37oC.
The maximum 27% propionic acid value was
observed for 2 days HRTs in ACPFR after dark
fermentation process, at 37oC.
Reduced H2(g) content is caused by the
production of CO2(g) by bacteria species that do not
produce bio-H2(g). Moreover, short HRT increases
the rate of substrate conversion and generates higher
substrate flow. For HRT equal to 1 h, acetic acid
dominates because the productivity of metabolites
decreases with decreasing HRT with decreasing
substrate conversion. The acetic pathway is the most
effective pathway in the dark fermentation process.
The acetic fermentative pathway was the main route
for bio-H2(g) production. The fermentative pathway
implied hydrogen in reduced metabolites. The low
biomass retention may contribute to the changing
metabolic pathways of acetogenic bacteria.
The use of validated CFD models to predict the
multiphase flow, accounting for the interfacial
momentum transfer between the three phases
present in ACPFR is still a challenge. It is generally
accepted that the ultimate goal of having validated
multiphase CFD models is the coupling with
biokinetics models for accurate modeling of the
biogas generation within the reactors, once
hydrodynamics and biochemical effects are
interdependent in ACPFR. A functional coupled
CFD-biokinetics model would allow for the scale-up
of processes while correctly predicting the
generation of biogas, and also taking into
consideration realistic mixing conditions within the
ACPFR. Some of the main challenges and hence
opportunities for future work towards this end goal
are listed hereafter.
The first and more complex challenge is the
inclusion to predict biomass degradation as well as
biogas generation. For that to be possible, a three-
phase CFD model must be used to account for the
biomass as well as the biogas inside the ACPFR.
The granules’ apparent density is not fixed, as it
is linked to the biogas generated. Including the
effects of the granular apparent density changes due
to bubble entrapment/attachment would lead to a
better prediction of the movement of the granules in
the sludge bed as well as the effects of granular
wash-out. Therefore, leading to a more accurate
prediction of loss of biomass.
The established ANN-based model in this work
indicates satisfactory and sound results with an R2
value of more than 0.980 and an RMSE value lower
than 0.25 for smhydrogen as a product from a
gasification system connected with a H2(g) plant.
Almost all of the inputs show a significant impact
on the smhydrogen output. Significantly, gasifier
temperature, SBR, moisture content, and H2(g) have
the highest impacts on the smhydrogen with
contributions of 19.85%, 17.29%, 15.41%, and
11.53%, respectively. In addition, other variables of
feed properties like carbon (C), oxygen (O), sulphur
(S) and nitrogen (N) contribute in the range of
1.31%–9.4% and proximate components like VM,
FC and A contribute in the range of 3.21%–7.71%
to the impact on smhydrogen.
It also examined the adaptability, governing
equations, processing time, flexibility, and
applicability of artificial intelligence in dark
fermentation process in an ACPFR for bio-H2(g)
production. In addition to, this study established that
artificial intelligence modeling has the potential to
drastically reduce the process development time for
dark fermentation process in an ACPFR at
anaerobic conditions of substrates, although at
varying degrees.
The accurate results obtained for bio-H2(g)
production through the gasification system
connected to ACPFR reactors confirm the strong
predictive ability of the developed ANN-based
model by applying a backpropagation algorithm
with a hidden layer of 13 neurons. The developed
model has the ability to be used with a wide range
of biomass. The results show the relative influence
of various biomass properties and operating
parameters on the bio-H2(g) output from the system.
Finally, the developed ANN model can be
practically used to screen suitable biomasses for
H2(g) extraction based on a gasification system
connected to W-G shifting and H2(g) recovery unit.
Developing standardized and practical
procedures for selecting algorithm and determining
dataset size. Developing such procedures will
require a comprehensive understandings of the
impacts/effectiveness of different algorithms and
training data samples to solve various bioenergy
problems. More case studies for bioenergy systems
with different biomass feedstock, conversion
technologies, and products will be needed.
CFD has proven useful in evaluating reactor
performance; It allows local and instantaneous
analysis of reaction rate; interphase hydrogen flows
and other processes taking place within the tank.
Numerical simulations were used to study bio-H2(g)
production and removal at different times to
evaluate the local behavior of the equipment and
suggest geometric changes. Experimental results
show that bio-H2(g) production significantly reduces
production costs; It clearly shows the opportunity to
be used as an environmentally friendly biofuel.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
331
Volume 20, 2023
Process design, modeling, and simulation of bio-
H2(g) production in an ACPFR; It represents an
essential tool for process optimization and scale-up.
Bio-H2(g) process efficiency production depends
significantly on the type of substrate from various
sources. Furthermore, in the process of scaling up
experimental parts to real application and
commercialization; Different bioreactor
configurations combined with kinetic models will be
realized.
Acknowledgement:
This research study was undertaken in the
Environmental Microbiology Laboratories at Dokuz
Eylül University Engineering Faculty
Environmental Engineering Department, İzmir,
Turkey. The authors would like to thank this body
for providing financial support.
References:
[1] G.Peixoto, N.K. Saavedra, M.B.A.
Varesche, M. Zaiat, Hydrogen production
from soft-drink wastewater in an upflow
anaerobic packed-bed reactor, International
Journal of Hydrogen Energy, Vol.36, 2011,
pp. 8953–8966.
[2] A.D.N. Ferraz Junior, C. Etchebehere, M.
Zaiat, High organic loading rate on
thermophilic hydrogen production and
metagenomic study at an anaerobic packed-
bed reactor treating a residual liquid stream
of a Brazilian biorefinery, Bioresource
Technology, Vol.186, 2015, pp. 81–88.
[3] M.Y. Azwar, M.A. Hussain, A.K. Abdul-
Wahab, Development of biohydrogen
production by photobiological, fermentation
and electrochemical processes: A review,
Renewable and Sustainable Energy
Reviews, Vol.31, 2014, pp. 158–173.
[4] A. Smolinski, K. Stanczyk, N. Howaniec,
Steam gasification of selected energy crops
in a fixed bed reactor, Renewable Energy,
Vol.35, 2010, pp. 397–404.
[5] N. Howaniec, A. Smolinski, Biowaste
utilization in the process of co-gasification
with bituminous coal and lignite, Energy,
Vol.118, 2017, pp. 18–23.
[6] G. Kumar, S. Shobana, D. Nagarajan, D.J.
Lee, K.S. Lee, C.Y. Lin, C.Y. Chen, J.S.
Chang, Biomass based hydrogen production
by dark fermentation: Recent trends and
opportunities for greener processes, Current
Opinion in Biotechnology, Vol.50, 2018, pp.
136–145.
[7] B. Lumbers, D.W. Agar, J. Gebel, F. Platte,
Mathematical modeling and simulation of
the thermo-catalytic decomposition of
methane for economically improved
hydrogen production, International Journal
of Hydrogen Energy, Vol.47, 2022, pp.
4265–4283.
[8] M.R. Talaghat, N. Naamaki, Mathematical
modeling of hydrogen production using
methanol steam reforming in the coupled
membrane reactor when the output materials
of the reformer section are used as feed for
the combustion section, International
Journal of Hydrogen Energy, Vol.46, 2021,
pp. 2282–2295.
[9] K. Christopher, R. Dimitrios, A review on
exergy comparison of hydrogen production
methods from renewable energy sources,
Energy and Environmental Sciences, Vol.5,
No.5, 2012, pp. 6640-6651.
[10] M.M. Amin, B. Bina, E. Taheri, M.R. Zare,
M. Ghasemian, S.W. Van Ginkel, A.
Fatehizadeh, Metabolism and kinetic study
of bioH2 production by anaerobic sludge
under different acid pretreatments, Process
Biochemistry, Vol.61, 2017, pp. 24-29.
[11] T. Lipman, An Overview of Hydrogen
Production and Storage Systems with
Renewable Hydrogen Case Studies,
Montpelier: Clean Energy States Alliance;
2011.
[12] P.C. Hallenbeck, D. Ghosh, Advances in
fermentative biohydrogen production: the
way forward? Trends in Biotechnology,
Vol.27, No.5, 2009, pp. 287-297.
[13] L.B. Brentner, J. Peccia, J.B. Zimmerman,
Challenges in developing biohydrogen as a
sustainable energy source: implications for
a research agenda, Environmental Science
& Technology, Vol.44, No.7, 2010, pp.
2243-54.
[14] P. Sivagurunathan, G. Kumar, P. Bakonyi,
S.H. Kim, T. Kobayashi, K.Q. Xu, G.
Lakner, G. Toth, N. Nemestothy, K. Belafi-
Bako, A critical review on issues and
overcoming strategies for the enhancement
of dark fermentative hydrogen production in
continuous systems, International Journal
of Hydrogen Energy, Vol.41, 2016,
3820e36.
[15] Y. Wong, T. Wu, J. Juan, A review of
sustainable hydrogen production using seed
sludge via dark fermentation, Renewable
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
332
Volume 20, 2023
and Sustainable Energy Reviews, Vol.34,
2014, pp. 471-482.
[16] M. Agler, B. Wrenn, S. Zinder, L.
Angenent, Waste to bioproduct conversion
with undefined mixed cultures: the
carboxylate platform, Trends in
Biotechnology, Vol.29, 2011, pp. 70-78.
[17] Y. Chen, Y. Yin, J. Wang, Recent advance
in inhibition of dark fermentative hydrogen
production, International Journal of
Hydrogen Energy, Vol.46, 2021, pp. 5053-
5073.
[18] M. Aydin, A. Karaca, A.M. Qureshy, I.
Dincer, A comparative review on clean
hydrogen production from wastewaters,
Journal of Environmental Management,
Vol.279, 2021, 111793.
[19] M.P. Sudhakar, B.R. Kumar, T. Mathimani,
K. Arunkumar, A review on bioenergy and
bioactive compounds from microalgae and
macroalgae-sustainable energy perspective,
Journal of Cleaner Production, Vol.228,
2019, pp. 1320–1333.
[20] H. Zhu, W. Parker, R. Basnar, A. Proracki,
P. Falletta, M. Béland, P. Seto, Buffer
requirements for enhanced hydrogen
production in acidogenic digestion of food
wastes, Bioresource Technology, Vol.100,
2009, pp. 5097–5102.
[21] C.L. Alvarez-Guzmán, S. Cisneros-de la
Cueva, V.E. Balderas-Hernández, A.
Smolinski, A. De Leon-Rodriguez,
Biohydrogen production from cheese whey
powder by Enterobacter asburiae: Effect of
operating conditions on hydrogen yield and
chemometric study of the fermentative
metabolites, Energy Reports, Vol.6, 2020,
pp. 1170–1180.
[22] Z. Trad, J.P. Fontaine, C. Larroche, C. Vial,
Multiscale mixing analysis and modeling of
bio-hydrogen production by dark
fermentation, Renewable Energy, Vol.98,
2016, pp. 264–282.
[23] X. Wang, J. Ding, W.Q. Guo, N.Q. Ren, A
hydrodynamics reaction kinetics coupled
model for evaluating bioreactors derived
from CFD simulation, Bioresource
Technology, Vol.101, 2010, pp. 9749–9757.
[24] I. Nopens, D. Sudrawska, W. Audenaert, D.
Fernandes del Pozo, U. Rehman, Water and
wastewater CFD and validation: are we
losing the balance? Water Science and
Technology, Vol.81, 2020, pp. 1636–1645.
[25] J.E. Baeten, D.J. Batstone, O.J. Schraa,
M.C.M. van Loosdrecht, E.I.P. Volcke,
Modelling anaerobic, aerobic and partial
nitritation-anammox granular sludge
reactors - A review, Water Research,
Vol.149, 2019, pp. 322–341.
[26] R.W. Samstag, J.J. Ducoste, A. Griborio, I.
Nopens, D.J. Batstone, J.D. Wicks, S.
Saunders, E.A. Wicklein, G. Kenny, J.
Laurent, CFD for wastewater treatment: an
overview, Water Science and Technology,
Vol.74, 2016, pp. 549–563.
[27] B. Wu, Advances in the use of CFD to
characterize, design and optimize bioenergy
systems, Computers and Electronics in
Agriculture, Vol. 2013, pp. 195–208.
[28] W.C. Lipps, E.B. Braun-Howland, T.E.
Baxter, Standard Methods for the
Examination of Water and Wastewater,
(24th. Edition), W.C. Lipps, E.B. Braun-
Howland, T.E. Baxter, (editors), American
Public Health Association (APHA),
American Water Works Association
(AWWA), Water Environment Federation
(WEF), Elevate Your Standards. American
Public Health Association 800 I Street, NW
Washington DC: 20001-3770, USA,
December 1, 2022; ISBN:9780875532998.
[29] J.S. Chang, K.S. Lee, P.J. Lin, Bio-
hydrogen production with fixed-bed
bioreactors, International Journal of
Hydrogen Energy, Vol.27, 2002, pp. 1167
1174.
[30] G. Kumar, A. Mudhoo, P. Sivagurunathan,
D. Nagarajan, A. Ghimire, C.H. Lay, C.Y.
Lin, D.J. Lee, J.S. Chang, Recent insights
into the cell immobilization technology
applied for dark fermentative hydrogen
production, Bioresource Technology,
Vol.219, 2016, pp. 725–737.
[31] G. Kumar, G. Buitrón, Fermentative bio-
hydrogen production in fixed bed reactors
using ceramic and polyethylene carriers as
supporting material, Energy Procedia,
Vol.142, 2017, pp. 743–748.
[32] F. Hawkes, I. Hussy, G. Kyazze, R.
Dinsdale, D. Hawkes, Continuous dark
fermentative hydrogen production by
mesophilic microflora: Principles and
progress. Int. J. Hydrogen Energy 2007, 32,
172–184.
[33] Saady, N.M.C. Homoacetogenesis during
hydrogen production by mixed cultures dark
fermentation: Unresolved challenge,
International Journal of Hydrogen Energy,
Vol.38, 2013, pp. 13172–13191.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
333
Volume 20, 2023
[34] A. Wodołazski, Co-simulation of CFD-
multiphase population balance coupled
model aeration of sludge flocs in stirrer tank
bioreactor, International Journal of
Multiphase Flow, Vol.123, 2020, pp. 103–
162.
[35] T. Weide, E. Brugging, C. Wetter, A.
Ierardi, M. Wichern, Use of organic waste
for biohydrogen production and volatile
fatty acids via dark fermentation and further
processing to methane, International
Journal of Hydrogen Energy, Vol.44, 2019,
pp. 24110–24125.
[36] R.Y. Kannah, S. Kavitha, P.
Sivashanmugham, G. Kumar, D.D. Nguyen,
S.W. Chang, J.R. Banu, Biohydrogen
production from rice straw: Effect of
combinative pretreatment, modeling
assessment and energy balance
consideration, International Journal of
Hydrogen Energy, Vol.44, 2019, pp. 2203–
2215.
[37] P.C. Ri, J.S. Kim, T.R. Kim, C.H. Pang,
H.G. Mun, G.C. Pak, N.Q. Ren, Effect of
hydraulic retention time on the hydrogen
production in a horizontal and vertical
continuous stirred-tank reactor,
International Journal of Hydrogen Energy,
Vol.44, 2019, pp. 17742–17749.
[38] H. Chen, A.S. Kim, Prediction of permeate
flux decline in crossflow membrane
filtration of colloidal suspension: a radial
basis function neural network approach,
Desalination, Vol.192, No.1-3, 2006, pp.
415-428.
[39] W.-Q. Guo, N.-Q. Ren, Z.-B. Chen, B.-F.
Liu, X.-J. Wang, W.-S. Xiang, J. Ding,
Simultaneous biohydrogen production and
starch wastewater treatment in an
acidogenic expanded granular sludge bed
reactor by mixed culture for long-term
operation, International Journal of
Hydrogen Energy, Vol.33, No.24, 2008, pp.
7397-7404.
[40] M. Sadrzadeh, T. Mohammadi, J. Ivakpour,
N. Kasiri, Neural network modeling of Pb2+
removal from wastewater using
electrodialysis, Chemical Engineering and
Processing-Process Intensification, Vol.48,
No.8, 2009, pp. 1371-1381.
[41] A.K. Giri, R.K. Patel, S.S. Mahapatra,
Artificial neural network (ANN) approach
for modeling of arsenic (III) biosorption
from aqueous solution by living cells of
Bacillus cereus biomass, Chemical
Engineering Journal, Vol.178, 2011, pp.
15-25.
[42] L. Chen, S. Nguang, X. Li, X. Chen, Soft
sensors for on-line biomass measurements,
Bioprocess and Biosystems Engineering,
Vol.26, 2004, pp. 191-195.
[43] P. Poirazi, F. Leroy, M.D. Georgalaki, A.
Aktypis, L. De Vuyst, E. Tsakalidou, Use of
artificial neural networks and a gamma-
concept-based approach to model growth of
and bacteriocin production by Streptococcus
macedonicus ACA-DC 198 under simulated
conditions of kasseri cheese production,
Applied and Environmental Microbiology,
Vol.73, 2007, pp. 768-776.
[44] P. Escalante-Minakata, V. Ibarra-Junquera,
H. Rosu, A. De León-Rodríguez, R.
González García, On-line monitoring of
Mezcal fermentation based on redox
potential measurements, Bioprocess and
Biosystems Engineering, Vol.32, 2009, pp.
47-52.
[45] S. Frigo, G. Spazzafumo, Cogeneration of
power and substitute of natural gas using
biomass and electrolytic hydrogen,
International Journal of Hydrogen Energy,
Vol.43, 2018, pp. 11696–11705.
[46] Q. Li, G. Song, J. Xiao, T. Sun, K. Yang,
Exergy analysis of biomass staged-
gasification for hydrogen-rich syngas,
International Journal of Hydrogen Energy,
Vol.44, 2019, pp. 2569–2579.
[47] N. Kobayashi, M. Tanaka, G. Piao, J.
Kobayashi, S. Hatano, Y. Itaya, S. Mori,
High temperature air-blown woody biomass
gasification model for the estimation of an
entrained down-flow gasifier, Journal of
Waste Management, Vol.29, 2009, pp. 245–
251.
[48] T. Damartzis, S. Michailos, A. Zabaniotou,
Energetic assessment of a combined heat
and power integrated biomass gasification–
internal combustion engine system by using
aspen plus®, Fuel Processing Technology,
Vol.95, 2012, pp. 37–44.
[49] T.L.T. Nguyen, J.E. Hermansen, R.G.
Nielsen, Environmental assessment of
gasification technology for biomass
conversion to energy in comparison with
other alternatives: The case of wheat straw,
Journal of Cleaner Production, Vol.53,
2013, pp. 138–148.
[50] A. Porcu, S. Sollai, D. Marotto, M.
Mureddu, F. Ferrara, A. Pettinau, Techno-
economic analysis of a small-scale biomass-
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
334
Volume 20, 2023
to energy bfb gasification-based system,
Energies, Vol.12, 2019, 494.
[51] D. Roy, S. Samanta, S. Ghosh, Thermo-
economic assessment of biomass
gasification-based power generation system
consists of solid oxide fuel cell,
supercritical carbon dioxide cycle and
indirectly heated air turbine, Clean
Technologies and Environmental Policy,
Vol.21, 2019, pp. 827–845.
[52] S. Safarian, C. Richter, R. Unnthorsson,
Waste biomass gasification simulation using
aspen plus: Performance evaluation of wood
chips, sawdust and mixed paper wastes,
International Journal of Power and Energy
Engineering, Vol.7, 2019, pp. 12–30.
[53] S. Safarian, R. Unnþórsson, C. Richter, A
review of biomass gasification modeling,
Renewable and Sustainable Energy
Reviews, Vol.110, 2019, pp. 378–391.
[54] S. Safarian, R. Unnthorsson, C. Richter,
Techno-economic analysis of power
production by using waste biomass
gasification, International Journal of Power
and Energy Engineering, Vol.8, 2020, pp.
1–8.
[55] S. Safarian, R. Unnthorsson, C. Richter,
Performance analysis and environmental
assessment of small-scale waste biomass
gasification integrated chp in Iceland,
Fermentation, Vol.197, 2020, 117268.
[56] S. Safarian, R. Unnthorsson, C. Richter,
Simulation of small-scale waste biomass
gasification integrated power production: A
comparative performance analysis for
timber and wood waste, International
Journal of Applied Power Engineering
(IJAPE), Vol.9, 2020, pp. 147–152.
[57] S. Safarian, R. Unnthorsson, C. Richter,
Techno-economic and environmental
assessment of power supply chain by using
waste biomass gasification in Iceland,
Biophysical Economics and Sustainability,
Vol.5, No.7, 2020, pp. 1-13.
[58] J. Gil, J. Corella, M.A.P. Aznar, M.A.
Caballero, Biomass gasification in
atmospheric and bubbling fluidized bed:
Effect of the type of gasifying agent on the
product distribution, Biomass and
Bioenergy, Vol.17, 1999, pp. 389–403.
[59] E. Shayan, V. Zare, I. Mirzaee, Hydrogen
production from biomass gasification; a
theoretical comparison of using different
gasification agents, Energy Conversion and
Management, Vol.159, 2018, pp. 30–41.
[60] V. Marcantonio, M. De Falco, M. Capocelli,
E. Bocci, A. Colantoni, M. Villarini,
Process analysis of hydrogen production
from biomass gasification in fluidized bed
reactor with different separation systems,
International Journal of Hydrogen Energy,
Vol.44, 2019, pp. 10350–10360.
[61] V. Rostampour, A.M. Motlagh, M.H.
Komarizadeh, M. Sadeghi, I. Bernousi, T.
Ghanbari, Using artificial neural network
(ann) technique for prediction of apple
bruise damage, Australian Journal of Crop
Science, Vol.7, 2013, pp. 1442–1448.
[62] J. George, P. Arun, C. Muraleedharan,
Assessment of producer gas composition in
air gasification of biomass using artificial
neural network model, International
Journal of Hydrogen Energy, Vol.43, 2018,
pp. 9558–9568.
[63] V. Nasir, S. Nourian, S. Avramidis, J. Cool,
Classification of thermally treated wood
using machine learning techniques, Wood
Science and Technology, Vol.53, 2019, pp.
275–288.
[64] V. Nasir, S. Nourian, S. Avramidis, J. Cool,
Prediction of physical and mechanical
properties of thermally modified wood
based on color change evaluated by means
of “group method of data handling” (gmdh)
neural network, Holzforschung, Vol.73,
2019, pp. 381–392.
[65] G. Capizzi, G.L. Sciuto, C. Napoli, M.
Wozniak, G. Susi, A spiking neural
network-based long-term prediction system
for biogas production, Neural Networks,
Vol.129, 2020, pp. 271–279.
[66] S. Safarian, S.M.E. Saryazdi, R.
Unnthorsson, C. Richter, Artificial neural
network integrated with thermodynamic
equilibrium modeling of downdraft biomass
gasification-power production plant,
Energy, Vol.213, 2020, 118800.
[67] M. Puig-Arnavat, J.A. Hernández, J.C.
Bruno, A. Coronas, Artificial neural
network models for biomass gasification in
fluidized bed gasifiers, Biomass and
Bioenergy, Vol.49, 2013, pp. 279–289.
[68] D. Baruah, D. Baruah, M. Hazarika,
Artificial neural network-based modelling
of biomass gasification in fixed bed
downdraft gasifiers, Biomass and
Bioenergy, Vol.98, 2017, pp. 264–271.
[69] A. Schmidt, W. Creason, B.E. Law,
Estimating regional effects of climate
change and altered land use on biosphere
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
335
Volume 20, 2023
carbon fluxes using distributed time delay
neural networks with bayesian regularized
learning, Neural Networks, Vol.108, 2018,
pp. 97–113.
[70] S. Safarian, S.M.E. Saryazdi, R.
Unnthorsson, C. Richter, Artificial neural
network modeling of bioethanol production
via syngas fermentation, Biophysical
Economics and Sustainability, Vol.6, 2021,
pp. 1–13.
[71] A.A. Kovalev, D.A. Kovalev, Y.V. Litti,
I.V. Katraeva, Biohydrogen production in
the two-stage process of anaerobic
bioconversion of organic matter of liquid
organic waste with recirculation of digester
effluent, International Journal of Hydrogen
Energy, Vol.45, No.51, 2020, pp. 26831–
26839.
[72] N. Pfennig, Anreicherungskulturen fürote
und grüne Schwefelbakterien, Zentralblatt
Fur Bakteriologie Mikrobiologie Und
Hygiene I- Abteilung Originale C-
Allgemeine Angewandte Und Okologische
Mikrobiologie, Vol.1, 1965, pp. 179189.
[73] N. Pfenning, K.D. Lippert, Über das
Vitamin B12 Bedürfnis phototropher
Schwefelbakterien, Archives of
Microbiology, Vol.55, 1966, pp. 245246.
[74] E.A. Wolin, M.J. Wolin, R.S. Wolfe,
Formation of methane by bacterial extracts,
Journal of Biological Chemistry, Vol.238,
1963, pp. 28822886.
[75] D. Frascari, M. Cappelletti, J.D.S. Mendes,
A. Alberini, F. Scimonelli, C. Manfreda, L.
Longanesi, D. Zannoni, D. Pinelli, S. Fedi,
A kinetic study of bio-hydrogen production
from glucose, molasses and cheese whey by
suspended and attached cells of Thermotoga
neapolitana, Bioresource Technology,
Vol.147, 2013, pp. 553–561.
[76] S.A. Van Ooteghem, S.K. Beer, P.C. Yue,
Hydrogen production by the thermophilic
bacterium Thermotoga Neapolitana, Applied
Biochemistry and Biotechnology, Vol.98–
100, No.1–9, 2002, pp. 177–189.
[77] S. Belkin, C.O. Wirsen, H.W. Jannasch, A
new sulfur-reducing, extremely
Thermophilic Eubacterium from a
submarine thermal vent, Applied and
Environmental Microbiology, Vol.51, No.6,
1986, pp. 1180–1185.
[78] E. Windberger, R. Huber, A. Trincone, H.
Fricke, K.O. Stetter, Thermotoga
Thermarum Sp. Nov. and Thermotoga
Neapolitana occurring in African
continental solfataric springs, Archives of
Microbiology, Vol.151, No.6, 1989, pp.
506–512.
[79] J. Anderson, Computational Fluid
Dynamics: The Basics with Applications,
1st ed. In: J.P. Holman, J.R. Lloyd, (Eds.).
McGraw-Hill Education, New York, NY,
USA, 1995.
[80] Z.G. Feng, E.E. Michaelides, E.E., The
immersed boundary-lattice Boltzmann
method for solving fluid–particles
interaction problems, Journal of
Computational Physics, Vol.195, No.2,
2004, pp. 602–628.
[81] H.R. Norouzi, R. Zarghami, N. Mostoufi,
New hybrid CPU-GPU solver for CFD-
DEM simulation of fluidized beds, Powder
Technology, Vol.316, 2017, pp. 233–244.
[82] M. Meister, W. Rauch, Wastewater
treatment modeling with smoothed particle
hydrodynamics, Environmental Modelling,
and Software, Vol.75, 2016, pp. 206–211.
[83] B. Wu, E.L. Bibeau, Development of 3-D
anaerobic digester heat transfer model for
cold weather applications, Transactions of
the ASABE, Vol.49, 2006, pp. 749–757.
[84] B. Wu, S. Chen, CFD simulation of non-
Newtonian fluid flow in anaerobic digesters,
Biotechnology and Bioengineering, Vol.99,
2008, pp. 700–711.
[85] S. Dabiri, P. Kumar, C. Ebner, W. Rauch,
On the effect of biogas bubbles in anaerobic
digester mixing, Biochemical Engineering
Journal, Vol.173, 2021, 108088.
[86] C. Sadino-Riquelme, R.E. Hayes, D. Jeison,
A. Donoso-Bravo, Computational fluid
dynamic (CFD) modelling in anaerobic
digestion: general application and recent
advances, Critical Reviews in
Environmental Science and Technology,
Vol.48, 2018, pp. 39–76.
[87] P.A. Lopez-Jimenez, J. Escudero-Gonzalez,
T. Montoya Martínez, V. Fajardo
Montanana, C. Gualtieri, Application of
CFD methods to an anaerobic digester: the
case of Ontinyent WWTP, Valencia, Spain,
Journal of Water Process Engineering,
Vol.7, 2015, pp. 131–140.
[88] S. Shrestha, S.P. Lohani, CFD analysis for
mixing performance of different types of
household biodigesters, Clean Energy,
Vol.6, 2022, pp. 325–334.
[89] C.A. de Lemos Chernicharo, 2007.
Anaerobic Reactors, Biological Treatment
Wastewater Series, Vol.4, IWA Publishing,
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
336
Volume 20, 2023
London, United Kingdom, 2007, ISBN 1: 1-
84339 164 3, ISBN 13: 9781843391647.
[90] S. Safarianbana, R. Unnthorsson, C.
Richter, Development of a New
Stoichiometric Equilibrium-Based Model
for Wood Chips and Mixed Paper Wastes
Gasification by Aspen Plus, In ASME
International Mechanical Engineering
Congress and Exposition; American Society
of Mechanical Engineers: Salt Lake City,
UT, USA, 2019, V006T006A00.
[91] D.J. Batstone, D. Puyol, X. Flores-Alsina, J.
Rodríguez, Mathematical modelling of
anaerobic digestion processes: applications
and future needs, Reviews in Environmental
Science and Biotechnology, Vol.14, 2015,
pp. 595–613.
[92] J. Laurent, R.W. Samstag, J.M. Ducoste, A.
Griborio, I. Nopens, D.J. Batstone, J.D.
Wicks, S. Saunders, O. Potier, A protocol
for the use of computational fluid dynamics
as a supportive tool for wastewater
treatment plant modelling, Water Science
and Technology, Vol.70, 2014, pp. 1575–
1584.
[93] J. Lindmark, E. Thorin, R. Bel Fdhila, E.
Dahlquist, Effects of mixing on the result of
anaerobic digestion: review, Renewable and
Sustainable Energy Reviews, Vol.40, 2014,
pp. 1030–1047.
[94] G. Leonzio, Studies of mixing systems in
anaerobic digesters using CFD and the
future applications of nanotechnologies,
Waste and Biomass Valorization, Vol.11,
2020, pp. 5925–5955.
[95] J. Li, M. Suvarna, L. Li, L. Pan, J. Perez-
Ramírez, Y.S. Ok, X. Wang, A review of
computational modeling techniques for wet
waste valorization: research trends and
future perspectives, Journal of Cleaner
Production, Vol.367, 2022, 133025.
[96] F. Liotta, P. Chatellier, G. Esposito, M.
Fabbricino, E.D. van Hullebusch, P.N.L.
Lens, F. Pirozzi, Current views on
hydrodynamic models of nonideal flow
anaerobic reactors, Critical Reviews in
Environmental Science and Technology,
Vol.45, 2015, pp. 2175–2207.
[97] J. Yang, Approaches for Modelling
Anaerobic Granule-Based Reactors,
Bacterial Biofilms, S. Dincer, M. Sümengen
Özdenefe, A. Arkut (Eds.), IntechOpen,
London, United Kingdom, 2020, number of
pages: 360. https://doi.org/10.5772/
intechopen.90201, ISBN: 978-1-78985-900-
3.
[98] U. Özdemir, B. Özbay, S. Veli, S. Zor,
Modeling adsorption of sodium dodecyl
benzene sulfonate (SDBS) onto polyaniline
(PANI) by using multi linear regression and
artificial neural networks, Chemical
Engineering Journal, Vol.178, 2011, pp.
183-190.
[99] V.V. Nair, H. Dhar, S. Kumar, A.K. Thalla,
S. Mukherjee, J.W.C. Wong, Artificial
neural network-based modeling to evaluate
methane yield from biogas in a laboratory-
scale anaerobic bioreactor, Bioresource
Technology, Vol.217, 2016, pp. 90–99.
[100] W. Uddin, K. Ayeshab, Z. Kamran, H. Aun,
K. Bilal, I. Saiful, M. Ishfaqa, K. Imran, M.
Adild, K. Hee Je. 2019. Current and future
prospects of small hydropower in Pakistan:
A survey, Energy Strategy Reviews, Vol.24,
2019, pp. 166–177.
[101] C. Rashama, G. Ijoma, T. Matambo, Biogas
generation from by-products of edible oil
processing: a review of opportunities,
challenges and strategies, Biomass
Conversion and Biorefinery Journal, Vol.9,
No.4, 2019, pp. 803–826.
[102] K.C. Surendra, D. Takara, A.G. Hashimoto,
S.K. Khanal. 2014. Biogas as a sustainable
energy source for developing countries,
Opportunities and Challenges Renewable
and Sustainable Energy Reviews, Vol.31,
2014, pp. 846–859.
[103] R. Arthur, M.F. Baidoo, E. Antwi. Biogas
as a potential renewable energy source: A
Ghanaian case study, Renewable Energy,
13, 2010, pp. 1–7
[104] M.M. Ghiasi, M. Arabloo, A.H.
Mohammadi, T. Barghi, Application of
ANFIS soft computing technique in
modeling the CO2 capture with MEA, DEA,
and TEA aqueous solutions, International
Journal of Greenhouse and Gas Control,
Vol.49, 2016, pp. 47–54.
[105] S.A. Kalogirou, Artificial intelligence for
the modeling and control of combustion
processes: A review, Progress in Energy
and Combustion Science, Vol.29, No.6,
2003, pp. 515–566.
[106] F. Manzano-agugliaro, F.G. Montoya, C.
Gil, A. Alcayde, J. Gómez, R. Banose,
Optimization methods applied to renewable
and sustainable energy: A review,
Renewable and Sustainable Energy
Reviews, Vol.15, 2011, pp. 1753–1766.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
337
Volume 20, 2023
[107] S.W. Van Ginkel, B. Logan, Increased
biological hydrogen production with
reduced organic loading, Water Research,
Vol.39, No.16, 2005, pp. 3819-3826.
APPENDIX
Fig. 1: Different ways of H2(g) production methods
Fig. 2: The methods of bio-H2(g) production.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
338
Volume 20, 2023
Fig. 3: Artificial intelligence categories and major techniques adopted from, [105].
Table 1. The parameters of initial and boundary conditions
Parameters
Values
Inlet flow rate (l/h)
0.2 - 1.0
HRTs (h)
1 - 12
Biomass / substrate ratio
0.154 – 0.352
Temperature (oC)
75 - 250
The size of the solid biomass particles (m)
0.00224 – 0.01167
Fig. 4: The effect of increasing HRTs for bio-H2(g) production rate at ACPFR, at 37oC.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
339
Volume 20, 2023
Fig. 5: The effect of increasing HRTs for Bio-H2(g) production yields at ACPFR, at 37oC.
Fig. 6: The Effect of different pH values for bio-H2(g) production at ACPFR, at 37oC.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
340
Volume 20, 2023
Fig. 7: The effect of increasing feed rate as OLR for bio-H2(g) production at ACPFR, at 37oC.
Fig. 8: The measurements of VFAs after bio-H2(g) production at ACPFR, at 37oC.
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
341
Volume 20, 2023
Fig. 9: The comparison between experimental bio-H2(g) production yields and ANN values in ACPFR, at 37oC.
Contribution of Individual Authors to the
Creation of a Scientific Article (Ghostwriting
Policy)
Prof. Dr. Delia Teresa Sponza and Post-Dr. Rukiye
Öztekin took an active role in every stage of the
preparation of this article.
The authors equally contributed in the present
research, at all stages from the formulation of the
problem to the final findings and solution.
Sources of Funding for Research Presented in a
Scientific Article or Scientific Article Itself
This research study was undertaken in the
Environmental Microbiology Laboratories at Dokuz
Eylül University Engineering Faculty
Environmental Engineering Department, İzmir,
Turkey. The authors would like to thank this body
for providing financial support.
Conflict of Interest
The authors have no conflicts of interest to declare.
Creative Commons Attribution License 4.0
(Attribution 4.0 International, CC BY 4.0)
This article is published under the terms of the
Creative Commons Attribution License 4.0
https://creativecommons.org/licenses/by/4.0/deed.en
_US
WSEAS TRANSACTIONS on BIOLOGY and BIOMEDICINE
DOI: 10.37394/23208.2023.20.33
Ruki
ye Özteki
n, Deli
a Teresa Sponza
E-ISSN: 2224-2902
342
Volume 20, 2023