Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Chemical Engineering Journal 466 (2023) 142760

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Modelling and analysis of hydrodynamics and flow phenomena of fluid


with formic acid as pollutant in the reactive area of a flat plate
photocatalytic reactor with top and bottom turbulence promote
M.G. Rasul a, *, S. Ahmed a, M.A. Sattar a, b, M.I. Jahirul a
a
School of Engineering and Technology, Central Queensland University, Rockhampton, Queensland 4702, Australia
b
Mechanical and product design engineering, Swinburne University of Technology Hawthorn 3122, Australia

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

Keywords: The inadequate mixing of pollutant and its low mass transfer in the reactive surface reduces the efficiency of
Photocatalytic reactor photocatalytic reactor. A computational model is developed, and numerical analysis is conducted to examine the
Pollutant effect of top and bottom baffles on the mass transfer and hydrodynamic performance of the reactor compared to
Hydrodynamics
reactor without baffle in this study. The hydrodynamic flow phenomena such as local streamlines, velocity,
Baffle
Turbulence
turbulence characteristic, pressure-drops, and wall shear stress were investigated. The experiment was done in a
lab-scale flat-plate photocatalytic reactor using Formic Acid as pollutant in feed water for model validation. The
FA concentration versus feed flow rate and velocity as a function of time showed good agreement between
simulation and experimental results (only about 5 % variation). From the simulation, it was found that the
pressure drops increase with increase in baffle height and the lowest was found to be 0.1 kPa and 0.4 kPa when
the baffle height was 5 mm at velocity of 0.15 m/s and 0.5 m/s respectively. The highest turbulent kinetic energy
and the dissipation was found 0.215 m2/s2 and 46 m2/s3 respectively when the baffle height was 12 mm in the
present simulation. Baffles significantly changes the hydrodynamic performance such as turbulence, velocity,
mass of pollutant and the pressure drops. The velocity magnitude also increases and the highest was found to be
9.2 times of the free stream velocity when 12 mm top baffle was used. Similar trends were found for bottom
baffles. The highest magnitude of turbulent kinetic energy was found to be four times higher than the plain
reactor without baffle. Therefore, baffle will significantly promote turbulence and improve the hydrodynamic
performance thus enhance the photocatalytic rector performance.

1. Introduction reactor to degrade phenol [8]. They observed that the operating pa­
rameters including initial phenol concentration, coating disc diameter,
Photocatalysis is the most promising advance oxidation process due liquid flow rate, nozzle-to-disc distance, etc, affects the photocatalytic
to its effectiveness of degradation of pollutant and handle of wide range activity. They found that the new reactor is cheaper to degrade the
of pollutants with negligible post processing of stream [1]. Heteroge­ pollutant and the reactor achieve the removal of TOC and COD of 79 %
neous catalysis is the most promising technology to degrade organic and and 84 %, respectively. Bai et al designed a novel Photocatalytic Fuel
inorganic pollutant for water purification because of its high efficiency Cell system using BiVO4/TiO2 nano tubes and ZnO/CuO nanowires and
using TiO2 and UV light [2–6]. Riaz and Park has presented the review of applied it to decompose methylene blue, methyl orange, and Congo red
the photocatalytic membrane reactors based on TiO2 by considering the and observed 90 %, 76 %, and 83 % degradation respectively after 80
different parameter for wastewater and polluted water treatment [7]. mins reaction [9]. They concluded that the suggested photoelectrodes
Their findings suggest that the performance of membrane is influenced provide great possibility for high-efficiency and cost-effective
by the irradiation time, the source of irradiation, and the TiO2 and they degradation.
identified that the sun light can be the better replacement of the UV Tahir et al investigated the methane (CH4) and carbon dioxide (CO2)
light. Jafarikojour et al investigated the impinging jet stream for a photo conversion to produce fuels using photocatalytic conversion technique

* Corresponding author.
E-mail address: m.rasul@cqu.edu.au (M.G. Rasul).

https://doi.org/10.1016/j.cej.2023.142760
Received 29 December 2022; Received in revised form 7 March 2023; Accepted 1 April 2023
Available online 5 April 2023
1385-8947/Crown Copyright © 2023 Published by Elsevier B.V. This is an open access article under the CC BY-NC license (http://creativecommons.org/licenses/by-
nc/4.0/).
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 1. Schematic diagram of the reactor (a) Top and bottom flange; (b) Reactive area cross section area and picture of (c) Reactor (d) Illumination system.

Fig. 2. Schematic diagram of the reactive module with top (upper figure) and bottom (lower figure) baffle, h = baffle height, H = reactor depth, L = reactive
module length.

[10]. They found that the CH4 and CO2 can effectively be converted to effectiveness of novel composite nanofibers of TiO2/ZnO/Bi2O3 as
fuels using light irradiations and suggest that the process will be bene­ photocatalytic reactor for NO and o-xylene degradation using solar ra­
ficial for the environment as it uses light to convert. Rathna et al diation [12]. They observed that the novel TiO2/ZnO/Bi2O3 composite
investigated the use of TiO2–WO3 nanoparticles into the Polyaniline nanofibers provides better photocatalytic activity than commercial TiO2
polymer matrix for improving its property in photocatalytic membrane reactor. Hurtado et al investigated the effect of silver on the Li1-xAgxV­
reactor [11]. They found that incorporation of nanoparticles improves MoO6 on the performance of photocatalytic reactor and compared with
the degradation of pollutant in the reactor. Pei et al investigated the TiO2. They found that the Li1-xAgxVMoO6 shows better performance of

2
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

overall effectiveness. A suitable strategy should be adopted to minimize


the thickness of the deposition on the surface to ensure good operational
stability. Proper mixing of chemical is an essential and conventional
technique in an effective chemical reactor. Additionally, insufficient
mixing of reactant can lead to low pollutant mass transfer in the reactive
surface, which reduces the effectiveness of the reactor [16,17]. The
optimum mixing predominantly depends on comparative distribution
and turbulent kinetic energy. Addition of baffles enhances the turbu­
lence and pollutant mass transfer in photocatalytic reactor [18].
Literature review reveals that there is very limited data on this part of
the photocatalytic system. The oxidation process of the reactor depends
on four crucial stages: mixing of the reactant, mass transfer of pollutant,
reaction, and separation product. As mixing is important in a photo­
catalytic reactor and in subsequent processing steps, this area has
received substantial attention [19,20]. Batch and continuous reactors
Fig. 3. Velocity at 2 mm from the catalyst coated layer for different mesh size. are usually employed in photo-chemical reactions, but the use of a
stirred tank has several disadvantages such as excessive stagnation of
fluid and non-homogeneous mixing, huge operating energy, and large
space [19]. Gordanshekan et al investigated the effect of UV irradiation,
CFX concentration and photocatalyst load ratio, pH, light intensity, and
anions on the Photocatalytic adsorption/degradation. They found that
91 % and 94 % of CFX removal for Bi2WO6/TiO2 and Bi2WO6/g-C3N4
respectively [21]. Matiazo et al investigated the influence of light
number and arrangement on the oxidation of n-decane using NETmix
milli-photocatalytic reactor numerically [22]. They found that the sys­
tem can degrade the pollutant 95 % when the ideal distance of LEDs is
12 mm.
The application of baffles can provide continuous mixing in a pho­
tocatalytic reactor without moving parts which offers various benefits
such as less space and energy consumption, improved process control,
less residence time and enhanced reaction selectivity. Improvement in
mixing using turbulence in low mass transfer region in an annular
photoreactor and static mixer was discussed by [23–26]. They reported
Fig. 4. Comparison of experimental results with CFD prediction [38]. that resistance of an immobilized photocatalytic reactor greatly in­
fluences the pollutant degradation and mass transfer. These results
indicate the importance of further investigation on which will consid­
erably increase mixing by creating vortices. The photocatalytic reactor
should be designed for uniform flow and shorter residence time. These
expectations, however, are not accurate and the reactor hydrodynamic
trends may cause the design ineffective. Numerous approaches are uti­
lized to enhance hydraulic performance of reactor and reduce the
magnitude of stagnant areas. Nonetheless the use of incorrect baffles
would decrease the reactor performance compared to reactor without a
baffle. Therefore, it is necessary to examine height of the baffles and
correct position of the baffles.
The literatures indicate that there are only a few studies available on
the CFD modelling of water disinfection in photoreactor, however the
modelling of the hydrodynamics of a photocatalytic flat plate reactor for
Fig. 5. Comparison of outlet formic acid between experiment and CFD. wastewater purification under different arrangements of baffle is
limited. Furthermore, research on developing appropriate reactor
design for degrading pollutants at higher rate is challenging [27] which
photocatalytic conversion than that of TiO2 [13]. Zhuang et al studied
has been addressed in this study. The CFD simulation offer more options
the effect of carbon nanotube in TiO2 on the photocatalytic performance
for parametric investigation for the optimal design of reactor that helps
of TiO2 [14]. They noticed an excellent adsorption ability and photo­
evaluate energy dissipation rate due to turbulence, kinetic energy, and
catalytic oxidation capability for degrading HgO when carbon doped
other parameters that are complicated to observe and vary in complex
nanotube were used.
geometries experimentally. As a result, CFD modelling has been used to
The use of photocatalytic reactor with a catalyst in the presence of
optimize, design, and scale-up of photocatalytic systems [28–31]. In this
light source has gained considerable interest among researchers because
study, the influence of top and bottom baffles of various heights as
of its efficacy in mineralizing and degrading the organic pollutants in
mixing and turbulence promoters on the hydrodynamics performance
wastewater. It is impractical to employ suspended catalyst therefore,
are investigated, which has got significant attention recently due to its
catalyst is coated in the reactor with the supporting materials. The
capability in producing large stream-wise vortices, improving turbulent
photocatalytic efficiency decreases because the degraded by-products
energy dissipation and mass and momentum-transfer in the flow over
deposits on the catalyst surface which is fixed on the surface [8,15].
those without baffles. Several runs were completed using CFD code
Furthermore, high level of pollutant in water fouled the catalyst surface.
FLUENT to calculate the distribution of significant hydrodynamic pa­
Therefore, the light can’t penetrate properly to the catalyst surface
rameters in the reactive section for various baffle arrangements and the
through the bulk liquid which may decrease the photocatalytic reactor’s
results obtained are discussed and compared in this study.

3
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 6. Effect of top baffle height on turbulent kinetic energy (m2/s2) distribution for (a) Re = 7500 and (b) Re = 2150.

2. Modelling of turbulence promoter the RNG k–ε model is reported to be more effective than other k–ε
models because it solve eddy activities at relatively low Reynolds
2.1. Model development numbers accurately [34]. RNG k-ε model needs less computational time
using a rigorous statistical technique than algebraic stress the Reynolds
The mass and momentum of fluid was dealt with Reynolds averaged, stress models. In this study, depending on the inlet velocity, the RNG k-ε
Navier-Stokes (RANS) equation by considering Newtonian, incom­ model was chosen from the available models for CFD simulation. The
pressible, isothermal, and non-reactive fluid with constant properties. transport equations for k and ε can be written as,
The reactor flow field was assumed to be turbulent steady state condi­ [( ) ]
∂ ∂ ∂ μ ∂k
tions. Using RANS turbulence modelling, [32] the CFD modelling solves (ρk) + (ρkui ) = μ+ t + Pk − ρ ∊ (4)
∂t ∂xi ∂xj σ∊ ∂xj
the numerical solution of the continuity equation Eqn. (1), RANS
equation Eqn. (2) and conservation of species equation Eqn. (3) by a [( ) ]
∂ ∂ ∂ μt ∂∊ ∊ ∊2
finite volume technique which are expressed as: (ρ∊) + (ρ∊ui ) = μ+ *
+ C1∊ Pk − C2∊ ρ (5)
∂t ∂xi ∂xj σ∊ ∂xj k k
∂(ρ)
+ ∇.(ρu) = 0 (1)
∂t where,
( )
η
∂ Cμ η3 1 −
(ρ∇) + ∇.(ρuu) = − ∇P − ∇.τ (2) *
C2∊ = C2∊ +
η0
(6)
∂t 1 + β η3
(ρUmk + ρuṁk ) = − ∇Jk (3) In Equations (4) and (5), Pk is the production of turbulent kinetic
energy due to mean velocity gradient. Pk can be modelled in RNG k-ε
The bar above the symbol is time average value. Here, ρ is the density
model similar to the standard k-ε which can be defined as:
(kg/m3), U is the fluid velocity (m/s), u is the fluctuating flow velocity. P
and τ are pressure (Pa) and the viscous stress tensor respectively. mk and ∂ui
(7)
′ ′
Pk = − ρui uj
Jk are the mass fraction and the diffusive flux of species k respectively. ∂xi
To consider turbulence correctly, Reynolds stresses are modelled to
Reynolds stresses was related with mean velocity gradient and tur­
achieve closure of Eqn. (2). In generally, the extensively applied tur­
bulent viscosity (μt) using Boussinesque approximation.
bulence models in many engineering applications are the realizable k–ε,
Pk can be calculated using the Boussinesque hypothesis as in the
mixing length, RNG k–ε, and standard k–ε [33]. To simulate more
following equation:
complex flows especially in a flow channel containing different barriers,

4
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 7. Effect of top baffle height on turbulent dissipation rate (m2/s3) distribution for (a) Re = 7500 and (b) Re = 2150.

P k = μt S 2 (8) Where Ji and Ri are the diffusion flux and net rate production of
species i due to chemical reaction. Yi is the mass fraction of reactants and
1/2
Here S = (2Sij Sij ) is the modulus of mean rate-of-strain tensor. The final products in each cell of the domain. The reaction kinetic was
turbulent viscosity (μt) is computed using the relationship below, evaluated from experimental work and used in this CFD modelling. The
Reynolds number for the present research was estimated without having
K2
μt = ρ C μ (9) calculated the transfer coefficients in the reactor. Re values could be
ε
determined using Reynolds, Schmidt, and Sherwood dimensionless
Here model coefficents Cμ = 0.0845 and C1ε = 1.42, C2ε = 1.68, ηo = number correlation [35]. However, for our small set-up Re values was
4.38 β = 0.012, and dimensionless expansion parameter η = Sk/∊. The calculated using the dimensions of the reactive section of the photo­
model constants were put in the GUI (Graphic user interface) in ANSYS catalytic reactor.
Fluent.
To model the reaction of the formic acid the following equation was
used. 2.2. Model geometry of reactive module and methodology

∂(ρYi)
+ ∇.(ρ→

u Yi) = − ∇. Ji + RSi (10) The numerical model was developed for the reactive section of the
∂t experimental model shown in Fig. 1(c). Fig. 1 shows both photo and the

5
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 8. Effect of top baffle height on wall shear stress for (a) Re = 7500 (b) 2150.

schematic of the experimental set-up of a flat plate photo reactor. The electronic microscope (SEM) was used to measure the thickness of the
reactor consists of top and bottom flanges, TiO2 (anatase) films and coated surface and five coated thicknesses of approximately 0.52, 1.21,
reaction area. TiO2 has high efficiency of degrading and mineralizing 2.24, 4.77 and 7.01 μm were prepared. TiO2 films were deposited on the
different types of organic pollutant to purify water using UV light as glass (220 × 100 × 3.5 mm) and placed on the bottom flange and the
suggested in the literatures [36,37], that is why TiO2 (anatase) was used upper part has quartz windows which allow the UV light with wave­
as catalyst in this study. The glass plate was degreased using H2O2 so­ length of 254 nm.
lution and cleaned with deionized water. The surface was etched with Flat-plate photocatalytic reactors were used which are generally
diluted nitric acid for better adhesion and the surface was cleaned and regarded as efficient as they provide high surface area to volume ratio
dried before applying the TiO2 coating. TiO2 powder was added into the for water treatment. Authors have selected flat reactor because TiO2 can
reagent ethanol to prepare the solution then the solution was sonicated be deposited uniformly and the desired thickness of TiO2 can be ach­
for an hour for homogenization. A magnetic stirrer was used to agitate ieved. Degradation of many organic compound produces formic acid
the suspension during the coating process. Spray coating deposition which needs to be further degraded therefore it is considered as a model
technique was employed to deposit the TiO2 solution on the surface and pollutant. Although formic acid was chosen as model pollutant, other
after each spray the surface was placed inside an incubator for two hours types of pollutants can be used in this reactor. The dimension of the
to dry the surface. The process was repeated for several times to get the reactor was selected by investigating the hydrodynamics of the rescore
desired thickness and the coated surface was placed inside an oven at using CFD which has been published in authors’ another paper [38].
400◦ C for several hours to stabilize the TiO2 on the surface. Scanning The position of the baffle was related with the length of the reactor L.

6
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Inlet velocity=0.5m/s

Inlet velocity=0.15 m/s

Fig. 9. Effect of top baffle height on pressure drop for (a) Re = 7500 and (b) Re = 2150.

The distance between them was 0.125L to make it eight baffles in were employed to discretise pressure. The residual was set 10− 4 for the
equally spaced. The model was run with irradiance of 70.6, 57.9, 37.1 continuity, momentum variables, and κ-ε, and 10− 6 for concentration for
and 20.4 Wm− 2. Seven 15 W ultra-violet lamps were used in this convergence criteria. One point in the domain where there is high
experiment whose position can be controlled vertically to change fluctuation of velocity was selected as observing point for convergence.
irradiance. The model solves the continuity and momentum equation first then
A 2D CFD model of the reactor was used in modelling as demon­ solve the conservation of species to minimize the computation time and
strated in Fig. 2 [38]. The reactor comprises of reactive area, inlet, and stabilize the solution [41,42].
outlet. The same model with added baffle was applied in this research to
assess the impact of baffle on the hydrodynamic performance of the 2.3. Boundary conditions and grid dependency
reactor. The 2D model was used for reducing the computational time
and occurrence of recirculating and dead regions in the reactor [39]. The In the present simulation, all walls and baffles surface were consid­
reactor outlet design was simplified for reducing the complexity and the ered a no-slip boundary condition. The inlet was assigned as VELOCITY-
amount of construction material required. The height (H) and length (L) INLET with the inlet velocity of 0.15 m/s and 0.5 m/s normal to inlet
of 15 mm and 200 mm, respectively between the inlet and outlet of the surface and outlet was considered as PRESSURE-OUTLET boundary
reactive part was considered in this study (Fig. 2). conditions. The pressure at the outlet of reactor was assigned at 101,325
A set of four top baffles having heights of 12 mm, 10 mm, 7.5 mm, Pa (Atmospheric). The hydraulic diameter of outlet and inlet boundary
and 5 mm was fitted at 0.125L, 0.25L, 0.375L, 0.5L, 0.675L, 0.75L and was specified at 15 mm with turbulence intensity (TI) of 5 %. A fully
0.875L respectively. The CFD model was developed using GAMBIT developed flow state was assumed at the outlet. Formic acid (FA) was
2.3.16 [40] and the domain was meshed of 347,809 cells and grid chosen as a model pollutant to assess the effect of top and bottom baffle
independency test was carried out. The cells near the baffle wall were position and height on the FA mass transport. The concentration of FA
made finer to cope with the fluctuation of velocity and shear stress using was fixed at 0.0034 (mass fraction) at the inlet with zero-diffusive flux at
the fixed size cells option. The reactive section with a bottom baffle the top wall. To make sure that the solutions is mesh independent, a
(lower figure) is illustrated in Fig. 2. This arrangement was chosen to series of tests using different mesh sizes were conducted prior to the
decrease the time for computational and to reduce the storage require­ actual simulations. The wall velocity profiles at 2 mm above the bottom
ment [41]. wall of the reactive section for three mesh sizes of 0.1, 0.2 and 0.3 mm
A 2D model was developed for reducing computational time and are compared in Fig. 3. The mesh sizes that were tested included a total
power. Numerical diffusion was tackled by employing a second order of 34189, 105637, 54,183 quadrilateral cells respectively. A negligible
upwind differencing scheme. SIMPLE algorithm was used for pressur­ difference was observed among the velocity profiles obtained by the
e–velocity-coupling scheme for better convergence. PRESTO schemes grids used. Additionally, the simulated hydrodynamic parameters were

7
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 10. Effect of top baffle height on local stream functions with an inlet velocity of 0.15 m/s (top figure) and 0.5 m/s (bottom figure) for A) 12 mm, B) 10 mm, C)
7.5 mm, D) 5 mm baffle and E) plain reactor respectively.

found to be same for all three mesh sizes examined. Thus, the subsequent 3.2. Simulated results of photocatalytic reactor with top baffle
modellings were performed with a mesh size of 0.1 mm.
3.2.1. Effect of top baffle height on Turbulence, wall shear and pressure
3. Results and discussion characteristics
The turbulent energy dissipation rate and kinetic energy (TKE) for
3.1. Experimental results and model validation different baffle heights (12 mm, 10 mm, 7.5 mm, 5 mm) and plain
reactor, respectively, are displayed in Fig. 6 and Fig. 7 for Re = 7500 and
To validate the numerical prediction of the model some experimental 2150 respectively. The model was simulated with variable baffles
measurements without baffles were done at different inlet velocities. heights at 0.125L, 0.25L, 0.375L, 0.5L, 0.625L, 0.75L and 0.875L from
The result from the experimental investigation and simulation for four the inlet with the highest TKE when the baffle height is 12 mm as shown
inlet velocities of 0.663 m/s, 1.32 m/s, 1.98 m/s and 2.65 m/s are in Fig. 6. A high turbulent kinetic energy region was observed for 10 mm
compared in Fig. 4. Only about 5 % variation between experimental and baffle height near the main flow which produces a powerful effect on the
simulation results were found, as shown in Fig. 4, which implies that the reactor performance.
model can simulate the performance of reactor correctly in other However, very low turbulence intensity was found downstream and
scenarios. close to each baffle in the reactor. A lowest turbulent kinetic energy
The optimum cell was selected by conducting a grid-independency region is predicted in the plain reactor without baffles as shown in Fig. 6.
test to alleviate the error from improper meshing (Fig. 3). As shown in At the same time, higher turbulence dissipation rate was found for 12
Fig. 5, the FA concentration at the reactor outlet between experiment mm top baffle as shown in Fig. 7. As expected, the baffles frequently
and simulation at different feed flow rate was found to be very close. The change the direction and intensity of the flow. This leads to high tur­
variation was within 5 %. bulence in local flow and improves the fluid mixing. Thus, the pollutants
mass transfer from the fluid to the reactive surface are enhanced.
As shown in Fig. 8, the wall shear stress fluctuates more for wall

8
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 11. Effect of top baffle height on velocity contours (ms− 1) with an inlet velocity of 0.15 m/s (top figure) and 0.5 m/s (bottom figure) for A) 12 mm, B) 10 mm, C)
7.5 mm, D) 5 mm baffle and E) plain reactor respectively.

velocity corresponds to both Re = 7500 and 2150. The lowest and transfer to the bulk stream. Subsequently, the deposition of by-product
highest wall shear stress repeatedly occurs in that wall velocity. The will be less for reactor with baffle than plain reactor and thus the
cyclic highest and lowest occurrence of wall shear and velocity enhances reactor with baffle will be more efficient.
the mass transfer on the wall [34]. For 12 mm baffle height with 0.5 m/s The pressure-drop in the reactive part for different baffle heights is
inlet velocity, the wall shear stress of about 12 and 9 Pa were found, displayed in Fig. 9 which was calculated from the static pressures be­
respectively, average of which is 10.5 Pa that are also much greater than tween the outlet (x = 0.2 m) and the inlet (x = 0 m). The calculated
without baffles (about 0.25 Pa). For 10 mm baffle, 6 Pa and 3 Pa was pressure drops at 0.5 m/s inlet velocity were found to be 0.5 kPa, 1 kPa,
found to be the highest and lowest wall shear stress, respectively. 3 kPa, and 10 kPa, for baffles with height of 5 mm, 7.5 mm, 10 mm, and
For 7.5 mm baffle, the average wall shear stress was found to be 2 Pa, 12 mm respectively. At inlet velocity of 0.15 m/s, the pressure-drops
which is also higher than the wall shear stress of a reactor without were found to be 1.75 kPa, 0.5 kPa and 0.2 kPa and 0.12 kPa for baf­
baffles. The wall shear stress was small in front of the first baffle for each fles with 12 mm, 10 mm, 7.5 mm, and 5 mm heights respectively. The
case. This could be due to change in cross-section of fluid-flow. In higher pressure-drop can raise in extra energy costs because of changes
contrast, the highest wall shear stress was 0.7 Pa for an inlet velocity of direction of flow and fluctuation of velocity. This will significantly in­
0.15 m/s. No variation was observed for 10 mm, 7.5 mm and 5 mm crease the friction loss of fluid flow. The eddy created at the back of the
baffle height for 0.15 m/s inlet velocity. Two peaks of wall shear stress baffles will also causes an extra energy due to simultaneous turbulent
that spans from 25 mm to 50 mm in the reactive section for both flow energy dissipation.
conditions were observed. This observation may be attributed to the
higher wall velocities as can be seen in Fig. 8. Shear induced diffusion 3.2.2. Effect of top baffle height on flow field and mass transfer
caused by wall shear stress considerably stimulate the pollutant mass The baffles in a reactor for water treatment is more effective and

9
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 12. Mass fraction of formic acid at the photocatalyst surface (a) Re = 2150 and (b) 7500.

Table 1
A summary of hydrodynamic parameters for top baffles with different heights.
Parameters Inlet velocity = 0.15 m/s Inlet velocity = 0.5 m/s
(Re = 2150) (Re = 7500)

12 mm 10 mm 7.5 mm 5 mm 12 mm 10 mm 7.5 mm 5 mm

Pressure drop (kPa) 1.9 0.52 0.2 0.1 15.1 4 1.05 0.4
2 1
Turbulent kinetic energy Max. 6.8 × 10− 2.72 × 0.019 7.47 × 6.14 × 10− 0.23 0.121 7.87 ×
(m2/S2) 10− 2 10− 3 10− 2
10 10
Min. 3.01 × 10− 3.32 × 1.87 × 2.19 × 8.24 × 10− 1.27 × 3.75 × 3.6 ×
10− 10 10− 11 10− 10 10− 2 10− 9 10− 9
Wall shear (Pa) Max. 0.7 2 1.5 1 22.5 6 3 1.5
Min. 0.25 0.25 0.1 0.05 13 2 1.5 0.5
Avg. 0.3 1 0.5 0.45 10 3.5 2 0.72
Turbulent dissipation rate Max. 3.0 2.77 0.783 0.21 7.0 3.5 5 2
(m2/S3)
Min. 1.75 0.461 0.157 2.98 × 1.75 0.75 0.45 0.25
10− 2
% Contact loss 0 0 0 0 0 0 0 0

economical which needs less maintenance due to no mechanical parts. the dead regions fades out as the baffle height decreases. Clearly, the
Within the reactor, the flow pattern is significantly affected by the recirculation region and its size depend on baffle height and location,
baffles position and height, therefore, determining the correct position therefore the correct design (height and location) of baffle should be
and optimal height of baffle is essential for enhancing reactor’s hydro­ done to reduce the creation of circulation regions to ensure uniform
dynamic performance. In this study, the heights of baffle were set to 67 flow. Therefore, selecting the correct number of baffles is essential to
%, 50 %, 33 %, and 20 % of the reactor height (H) (Fig. 10(A-D)) and its prevent the development of dead regions inside the reactor. It can be
impact on the flow was analysed by placing them at 0.125L, 0.25L, seen from the figure that 12 mm baffle produces higher velocity at the
0.375L, 0.5L, 0.625L, 0.75L and 0.875L respectively. reaction surface than that of other baffles. At the same time the dead
Fig. 10 presents the streamlines of the flow. From Fig. 10, a recir­ zone span between the baffles is more in 12 mm baffle than thar of other
culation region is developed at the inter-side of each baffle. The extent of baffles for both velocities considered.

10
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 13. Effect of bottom baffle aspect ratios on turbulent kinetic energy (m2/s2) distribution for Reynolds number 7500 (top figure) and 2150 (bottom figure)
respectively.

Table 2
Summary of hydrodynamic parameters for various bottom baffles located at 100 mm from inlet.
Parameters Inlet velocity = 0.15 m/s Inlet velocity = 0.5 m/s
(Re = 2150) (Re = 7500)

1.5 mm 2.5 mm 5 mm 7.5 mm 1.5 mm 2.5 mm 5 mm 7.5 mm

Pressure-drop (kPa) 0 0 0.1 0.2 0.05 0.1 0.5 0.7


3 2
Turbulent kinetic energy Max. 4.25 × 10− 5.25 × 5.96 × 1.7 × 1.99 × 10− 3.18 × 4.51 × 7.55 ×
(m2/S2) 10− 3 10− 3 10− 2 10− 2 10− 2 10− 2
10 10
Min. 3.61 × 10− 8.5 × 3.61 × 8.51 × 4.35 × 10− 6.94 × 2.5 × 6.94 ×
10− 4 10− 10 10− 4 10− 8 10− 3 10− 8
Wall shear, τw (Pa) Max. 0.25 0.5 0.8 1.0 0.5 0.7 0.8 1.5
Min. 0 0 0 0 0.1 0.1 0.15 0.25
Avg. 0.15 0.25 0.4 0.55 0.3 0.4 0.6 0.75
2 2 2 2 2
Turbulent dissipation rate Max. 3.47 × 10− 9.1 × 12 × 10− 1.1 1 × 10− 2 × 10− 3.0 × 4 × 10−
(m2/S3) 10− 2 10− 2
5 2
Min. 1.96 × 10− 6.51 × 1.75 × 6.1 × 0 0 0 1 × 10−
10− 3 10− 2 10− 2
% Contact loss 10 10 12.5 25 15 25 40 50

Fig. 11 displays the velocity contours inside the reactor for various mm, 7.5 mm, and 5 mm. From the result it can be deduced that better
baffles located at 0.125L, 0.25L, 0.375L, 0.5L, 0.625L, 0.75L and 0.875L. fluid mixing can be attained with 12 mm baffles height.
It shows the flow zones and the recirculation zones. When fluid flows Fig. 12 demonstrates the contours of mass fractions of the formic acid
under the baffle from the upstream to the narrow section, a powerful at the reactive surface when the Reynolds number is 2150 and 7500. For
shear layer at the tip of baffle is created. For all baffles, the wall ve­ Re = 2150, the FA mass fraction decreases as the fluid flows toward the
locities differ significantly depending on the baffle height. exit of reactor for all baffles. The figure also shows that the FA mass
For the 12 mm baffle height, the maximum velocity was around 4.61 fraction at the reactive surface is less for the plain and 5 mm baffles than
m/s between 0.02 m and 0.04 m from the inlet. This can be attributed to other baffles. Two low concentration regions were created at the end of
the decreases in cross-sectional area and increase in kinetic energy. A 7.5 mm and 12 mm baffles. High FA concentration was found at the
low velocity area can be found at the space between baffles and before upper half area of 0.5L baffles. In these zones, higher concentration was
and after of the first and last baffle. This phenomenon continues until the observed because of high velocity gradients. This situation was
reattachment point which creates flow reversal. A bigger interfacial confirmed by examining the velocity field in the reactive section.
zone between fluids can cause significant increase in fluid mixing. of the Nevertheless, at Re = 7500, two low concentration areas were observed
maximum velocity of 0.465 m/s was found in the above-mentioned at the back of 12 mm and 5 mm baffles. This finding relates to the cre­
zones. The liquid velocity also increases for the baffles of height of 10 ation of low velocity region.

11
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 14. Effect of bottom baffle spacing on streamline for baffle height 0.17H (upper three) and 0.1H (lower two) respectively.

Fig. 15. Effect of bottom baffle spacing on velocity (ms− 1) for baffle height a) 0.17H and b) 0.1H respectively.

Low FA concentration was seen at the reactive surface which can be


described by the low TKE. The FA concentration was less at turbulent
flow (Re = 7500) than in laminar flow (Re = 2150). Several studies
suggested that the introduction of baffles and other mixing mechanism
will enhance the mass transfer performance if catalyst is used in the
reactor [18,23,43–45].
A summary of hydrodynamic parameters obtained with top baffles
for various heights computed at Re = 2150 and 7500 is presented in
Table 1. The values of the parameters computed at Re = 7500 was
significantly higher than at Re = 2150. No % loss of contact can be seen
for both flow conditions.

3.3. Simulated result of photocatalytic reactor with bottom baffle


Fig. 16. Effect of bottom baffle positions on local streamline contours
computed at (i) Re = 7500 and (ii) Re = 2150 for A) 0.125L B) 0.25L C) 0.5L 3.3.1. Effect of bottom baffle height on hydrodynamic of photocatalytic
respectively. reactor
The baffles were placed at the middle of the reactive section with an
aspect ratio of 1.5, 2.5, 5, and 7.5 and the results of turbulent kinetic
energy are presented in Fig. 13. The peak turbulent intensity predicted

12
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 17. Velocity contours (ms− 1) for bottom baffle position at A) 0.125L B) 0.25L C) 0.5L for (i) Re = 7500 and (ii) Re = 2150 respectively.

are noticed on the front section of baffles top and is stretched over an and fluid mixing through the creation of wall vortices. For a given baffle
area near to the main flow which causes strong turbulence intensity. height, recirculation size was minimal when the baffle interval was fixed
However, the turbulent intensity is very weak in the reactive module at 25 mm in the reactive section. In contrast, the range of the recircu­
when plain reactor was used as shown Fig. 13. The plain reactor exhibits lation zone was observed to be smaller when the baffle height was set at
a lower turbulent kinetic energy as compared to the reactor with baffles 50 mm. The extent of recirculation enhances the interfacial deformation
because plain reactor does not produce turbulence and the scouring with a larger interfacial space between bulk fluid and pollutants in the
effects. Therefore, the baffles will enhance the fluid mixing thus will feed stream which causes significant improvement in mixing. For the
increase the mass transfer of the pollutant from the liquid to the catalyst simulated entire length (L = 200 mm), when the baffles were placed at
surface. The highest magnitude of turbulent kinetic energy was observed intervals of 0.1L, 0.125L, 0.15L and 0.25L, the required numbers of
to be four times than that of the plain reactor without baffle. baffle were 9, 3, 3 and 2 respectively. The results reveal that the wall
In contrast, the highest magnitude of turbulent kinetic energy was shear-stress decreases when spacing between baffles is increases and
observed to be four times lower than that obtained with smaller fluid hence the pressure-drop [38]. To optimize the design of baffles in a
flow velocity. A summary of significant hydrodynamic parameters reactive module, the height and spacing of baffles should be considered
achieved with an inlet velocity of 0.15 m/s and 0.5 m/s for various as the two important parameters, because they can have a significant
bottom baffles is shown in Table 2. effect on wall shear stress and pressure-drop.

3.3.2. Effect of bottom baffle position on the hydrodynamics of 3.3.3. Effect of bottom baffle position and height on flow field and mass
photocatalytic reactor transfer
The baffle was placed at intervals of 0.1L, 0.125L, 0.15L and 0.25L To find the ideal location, several numerical modellings have been
and the effect on the flow was investigated numerically and presented in conducted by positioning a baffle with 5 mm height (30 % of the reactor
Fig. 14. The baffle heights were chosen to be 10 % and 17 % of the height depth (H)) at three different locations at 0.125L, 0.25L, and 0.5L,
of reactor for minimal recirculation region close to the baffle. respectively and its influence on the local streamlines was examined.
The distance of 0.1H was chosen between two successive baffles so Fig. 16 presents the streamlines obtained with simulation for inlet ve­
that they correspond to the locations in the wake of the first baffle. locities of 0.15 m/s and 0.5 m/s respectively, which shows creation of a
Remarkably, the second baffle did not affect the velocity considerably recirculation region downstream of each of the baffle. The magnitude of
particularly when two successive baffles were located near to each other the recirculation regions is affected by the baffle position and the inlet
as indicated in Fig. 14. velocities of the reactor. For a given baffle location, the extent of the re-
It is evident from Fig. 15 that the velocity is significantly influenced circulation zone is larger at high Reynolds number flow compared with
by the spacing between the baffles which will improve the mass transfer one that formed at low Reynolds number flow.

13
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 18. Effect of bottom baffle height A)7.5 mm, B) 5 mm, C) 2.5 mm, D) 1.5 mm, E) Plain reactor on local stream functions for Reynolds number 7500 (top figure)
and 2150 (bottom figure) respectively (baffle positioned at 0.5L from inlet).

The velocity profiles of baffle located at 0.125L, 0.25L and 0.5L in the to the baffle in both sides. However, the size of the low concentration
reactive zone are shown in Fig. 17. Fig. 17 indicates that the normal area slowly vanishes as the height of baffle reduces from 7.5 mm to 1.5
features of flow structure lie in the recirculation zones and main flow mm. This finding can be described by the changes in velocity gradient
regions. When the fluid particles pass over the baffle, the flow from the and recirculation regions created close to baffle tips. A close examina­
upstream into the narrow area creates a powerful shear layer at the tip of tion of the velocity field revealed that a strong recirculation zone with
baffle. A velocity magnitude of 1.19 m/s was observed at the tip of the different sizes appears to persist downstream from the baffles which
baffle when it is placed at 0.125L. This could be attributed combinedly creates a higher velocity gradient.
to the reduced cross-sectional area and increased kinetic energy. At the The impact of baffle position on the FA mass fraction is presented in
downstream of the flow, a low-velocity area is observed until a reat­ Fig. 20 (b) at Re = 7500. The height of the baffle was considered as
tachment area which is due to reverse flow. A velocity of 0.566 m/s was 0.33H and located at 0.125L, 0.25L and 0.5L respectively. FA concen­
noticed between the regions mentioned above. Similar trends were tration is displayed in Fig. 20 (b) which shows that the region of low
observed when the baffles were placed at 0.25L and 0.5L. concentration of FA extends about 0.25L of the reactive section
The impact of height of baffle on the local streamlines in the reactor remained at and adjacent to each of the baffle positions. These phe­
are presented in Fig. 18 at Reynolds number of 7500. In modelling, the nomena could be associated to a recirculation region close to the baffle
location of the baffle is fixed at 100 mm from the inlet and the height of positions which lead to a region with a high velocity gradient.
baffle was varied for a range from 1.5 mm to 7.5 mm. Fig. 17 indicates Numerous studies have suggested that the application of baffles
that the area of the recirculation zone increases from 50 mm to 100 mm would enhance the performance and mass transfer of pollutant by
when height of baffle increases from 1.5 mm to 7.5 mm. improving mixing in a reactor [17,23,46,47]. By analysing the hydro­
Fig. 19 shows the effect of baffle height on the velocity. The highest dynamic parameters obtained with the top and bottom baffles, it is
velocity for the 7.5 mm baffle height was determined to be 1.59 m/s and suggested that the top baffle with 12 mm height can be a promising
0.458 m/s at Re = 7500 and Re = 2150 respectively at the top of the approach as turbulence promoters in the flat plate reactor. Because there
baffle. The friction factor increases if the ratio of baffle to reactor height is no contact loss of streamlines from the catalyst surface for the 12 mm
increases for a given flow rate and baffle location. High velocity exists at top baffle, using this baffle, the highest level of turbulence was observed
the tip of baffle for all baffle heights as there is a small cross-sectional near the catalyst coated surface which is likely to benefit by minimizing
area at the top of baffle. Results implies that for a specific location, the pollutant mass transport resistance and removing the degraded by-
the baffle height is crucial to ensure a shorter-circuiting free region. products from the reactive surface. In contrast, the use of a bottom
The effect of baffle height, 1.5 mm to 7.5 mm, on the FA mass baffle in the reactive section causes a significant loss of streamlines
fraction at the inlet at velocity of 0.5 m/s (Re = 7500) is presented in contact which may limit the effective exposure of catalyst to the
Fig. 20(a). The baffle is located at 0.5L from the inlet and the results contaminated water stream.
obtained shows that the FA concentration in the reactive surface reduces The influence of baffle spacing on the hydrodynamic parameters was
as it passes the baffle, and a low concentration area appears to stay close studied by varying the baffle spacing from 12 mm to 20 mm in the

14
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 19. Effect of bottom baffle height on velocity (ms− 1) for Reynolds number 7500 (top) and 2150 (bottom) respectively (Baffle positioned at 0.5L from inlet).

reactive section. The height of the baffle was set at 1.5 mm which was 4. Conclusions
fixed at the bottom wall. The simulation results obtained using various
spacings is presented in Table 3. The calculated average wall velocities The effect of position and height of baffle on the fluid flow phe­
at 2.5 mm inside the bottom wall are found to be 0.2 m/s, 0.45 m/s, and nomena and FA concentration in a flat plate photocatalytic reactor was
0.5 m/s for 20 mm, 15 mm and 12 mm spacing respectively. The in­ investigated in this study using CFD simulation. The reactive section of
crease in TKE and wall shear stress is linked with the decrease of baffle the reactor was considered for simulation. The influence of hights of top
spacing, which is advantageous to improve the mass transfer of pollutant and bottom baffles on the flow phenomena was analysed. The hydro­
in the reactor. On the contrary, the increase in pressure drop can also be dynamically important flow characteristics like local streamlines dis­
seen which is related to the energy consumption. However, the amount tribution, velocity contours, turbulence characteristics, wall shear stress,
of pressure drop was found to vary slightly over the spacing of baffle vortex formation, pressure drops, boundary layers separation, etc, were
under the simulation conditions. investigated. The simulation results were validated with the experi­
As presented above in sections 3.2 and 3.3, the reaction of the mental measurement of FA concentration as a function of flow rate and
pollutant occurs at the TiO2 coated surface, so the pollutant comes inlet velocity with respect to time which shows very good agreement
across the TiO2 coated surface. After degradation, the degraded with less than 5 % variation. The effect of baffle heights on pollutant
pollutant goes away from the surface and the new water layer with mass fraction are also investigated for low and high Reynolds numbers.
pollutant comes across TiO2 coated surface. This mechanism improves The experimentally validated simulated results provided useful infor­
the turbulence (Hydrodynamics improvement), so the reaction rate in­ mation to realize the influence of top and bottom baffles on fluid flow
creases. This was the main idea of using promoters in this research that features and characteristics and their role on improving fluid mixing in
the turbulent kinetic energy and dissipation can be greatly enhanced by the reactor. The detailed simulated result showed that both flow phe­
using the top and bottom promoter. However, it should be mentioned nomena and pollutant mass transfer on the reactive surface of the
that addition of baffles increases turbulent energy dissipation but causes photocatalytic reactor are significantly influenced by the baffles. The
extra energy. This extra energy consumption can be compensated with pressure drops increase with the increase of the baffle height. A set of
the enhancement of the efficiency. four top baffles having 12 mm, 10 mm, 7.5 mm, and 5 mm heights were
used and among them the lowest pressure drops of 0.1 kPa and 0.4 kPa
were found when the baffle height was 5 mm at velocity of 0.15 m/s and
0.5 m/s, respectively. The velocity magnitude also increases with baffle

15
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

Fig. 20. Mass fraction of FA for (a) different bottom baffle heights at 0.5L from inlet and (b) at different location with height 0.33H.

[2] A. Adesina, Industrial exploitation of photocatalysis: progress, perspectives and


Table 3 prospects, Catal. Surv. Asia 8 (4) (2004) 265–273.
Average values at distance 2.5 mm from bottom walls for different spacing of [3] Herrmann, J.-M., Heterogeneous photocatalysis: state of the art and present applications
In honor of Pr. RL Burwell Jr.(1912–2003), Former Head of Ipatieff Laboratories,
baffles inserted in the reactive section.
Northwestern University, Evanston (Ill). Topics in Catalysis, 2005. 34(1): p. 49-65.
Spacing Pressure drops Turbulent kinetic energy (m2/S2) Wall shear (Pa) [4] D. Wu, et al., Hierarchical TiO2 structures derived from F− mediated oriented
(Pa) assembly as triple-functional photoanode material for improved performances in
(Pa) CdS/CdSe sensitized solar cells, Electrochim. Acta 248 (2017) 79–89.
[5] H. Wang, et al., Preparation of TiO2 microspheres with tunable pore and chamber
2
12 mm 300 3 × 10− 7 size for fast gaseous diffusion in photoreduction of CO2 under simulated sunlight,
2
15 mm 200 2 × 10− 5 Journal of Colloid Interface Science 539 (2019) 194–202.
20 mm 100 7 × 10− 3
1 [6] P.S. Saud, et al., Effective photocatalytic efficacy of hydrothermally synthesized
silver phosphate decorated titanium dioxide nanocomposite fibers, J. Colloid
Interface Science 465 (2016) 225–232.
[7] S. Riaz, S.-J. Park, An overview of TiO2-based photocatalytic membrane reactors
heights and the highest was found to be 9.2 times higher of the free
for water and wastewater treatments, J. Indust. Eng. Chem. 84 (2020) 23–41.
stream velocity at the reactive surface when 12 mm top baffle was used, [8] M. Jafarikojour, et al., Application of a new immobilized impinging jet stream
and similar trends were found for bottom baffle. reactor for photocatalytic degradation of phenol: Reactor evaluation and kinetic
modelling, J. Photochem. Photobiol. A: Chem. 364 (2018) 613–624.
[9] J. Bai, et al., A solar light driven dual photoelectrode photocatalytic fuel cell (PFC)
Declaration of Competing Interest for simultaneous wastewater treatment and electricity generation, J. Hazard.
Mater. 311 (2016) 51–62.
[10] M. Tahir, et al., Enhanced photocatalytic carbon dioxide reforming of methane to
The authors declare that they have no known competing financial fuels over nickel and montmorillonite supported TiO2 nanocomposite under UV-
interests or personal relationships that could have appeared to influence light using monolith photoreactor, J. Clean. Prod. 213 (2019) 451–461.
the work reported in this paper. [11] R. Thara, J.B.P. Ettiyappan, D.R. Sudhakar, Fabrication of visible-light assisted
TiO2-WO3-PANI membrane for effective reduction of chromium (VI) in
photocatalytic membrane reactor, Environ. Technol. Innov. 24 (2021), 102023.
Data availability [12] C.C. Pei, W.-W.-F. Leung, Photocatalytic oxidation of nitrogen monoxide and o-
xylene by TiO2/ZnO/Bi2O3 nanofibers: Optimization, kinetic modeling and
mechanisms, Appl Catal B 174 (2015) 515–525.
Data will be made available on request.
[13] L. Hurtado, et al., Photocatalytic performance of Li1− xAgxVMoO6 (0⩽ x⩽ 1)
compounds, Chem. Eng. J. 234 (2013) 327–337.
References [14] H. Zhuang, et al., Advanced treatment of biologically pretreated coal gasification
wastewater by a novel integration of heterogeneous catalytic ozonation and
biological process, Bioresour. Technol. 166 (2014) 592–595.
[1] R. Binjhade, R. Mondal, S. Mondal, Continuous photocatalytic reactor: Critical
review on the design and performance, J. Environ. Chem. Eng. (2022), 107746.

16
M.G. Rasul et al. Chemical Engineering Journal 466 (2023) 142760

[15] K.V.S. Rao, M. Subrahmanyam, P. Boule, Immobilized TiO2 photocatalyst during [32] Ranade, V., Computational flow modelling for chemical reactor engineering. Vol. 5.
long-term use: decrease of its activity, Appl Catal B 49 (4) (2004) 239–249. 2002: Academic press.
[16] S. Gupta, H. Gomaa, M.B. Ray, Fouling control in a submerged membrane reactor: [33] Launder, B. and D. Spalding, The numerical computational of turbulent flows, Vol. 3
Aeration vs membrane oscillations, Chem. Eng. J. 432 (2022), 134399. Comp. Meth. Appl. Mech. Eng, 1974. 269.
[17] P.K. Dutta, A.K. Ray, Experimental investigation of Taylor vortex photocatalytic [34] Z. Cao, D.E. Wiley, A.G. Fane, CFD simulations of net-type turbulence promoters in
reactor for water purification, Chem. Eng. Sci. 59 (22) (2004) 5249–5259. a narrow channel, J. Membr. Sci. 185 (2) (2001) 157–176.
[18] K. Mehrotra, A.K. Ray, Performance enhancement of photocatalytic reactor [35] A.A. Assadi, et al., Modeling and simulation of VOCs removal by nonthermal
utilizing flow instability, Int. J. Chem. React. Eng. 1 (1) (2003). plasma discharge with photocatalysis in a continuous reactor: Synergetic effect and
[19] H.M. Kaci, et al., Effects of embedded streamwise vorticity on turbulent mixing, mass transfer, Chem. Eng. J. 258 (2014) 119–127.
Chemical Engineering Processing: Process Intensification 48 (10) (2009) [36] S. Ahmed, et al., Advances in Heterogeneous Photocatalytic Degradation of
1459–1476. Phenols and Dyes in Wastewater: A Review, Water Air Soil Pollut. 215 (1) (2011)
[20] M. Subramanian, A. Kannan, Photocatalytic degradation of phenol in a rotating 3–29.
annular reactor, Internat. J. Heat Fluid Flow 65 (9) (2010) 2727–2740. [37] S. Ahmed, et al., Heterogeneous photocatalytic degradation of phenols in
[21] A. Gordanshekan, et al., A comprehensive comparison of green Bi2WO6/g-C3N4 wastewater: A review on current status and developments, Desalination 261 (1)
and Bi2WO6/TiO2 S-scheme heterojunctions for photocatalytic adsorption/ (2010) 3–18.
degradation of Cefixime: Artificial neural network, degradation pathway, and [38] S. Ahmed, et al., Phenol degradation of waste and stormwater on a flat plate
toxicity estimation, Chem. Eng. J. 451 (2023), 139067. photocatalytic reactor with TiO2 on glass slide: An experimental and modelling
[22] T. Matiazzo, et al., CFD and radiation field modeling of the NETmix milli- investigation, J. Water Process Eng. 47 (2022), 102769.
photocatalytic reactor for n-decane oxidation at gas phase: Effect of LEDs number [39] P. Capel, Design of a photocatalytic water purification reactor for the testing of
and arrangement, Chem. Eng. J. 444 (2022), 136577. fixed films semiconductors, Queensland Universsity Of Technology, Australia,
[23] T.K. Sengupta, M.F. Kabir, A.K. Ray, A Taylor Vortex Photocatalytic Reactor for 2008.
Water Purification, Ind. Eng. Chem. Res. 40 (23) (2001) 5268–5281. [40] Fluent-Inc, FLUENT 6.3 User’s Guide. Modelling Turbulence. 2006: Lebanon.
[24] J.E. Duran, M. Mohseni, F. Taghipour, Computational fluid dynamics modeling of [41] S. Ahmed, et al., Influence of parameters on the heterogeneous photocatalytic
immobilized photocatalytic reactors for water treatment, AIChE J 57 (7) (2011) degradation of pesticides and phenolic contaminants in wastewater: A short
1860–1872. review, J. Environ. Manage. 92 (3) (2011) 311–330.
[25] J.E. Duran, F. Taghipour, M. Mohseni, CFD modeling of mass transfer in annular [42] J. Esteban Duran, F. Taghipour, M. Mohseni, CFD modeling of mass transfer in
reactors, International Journal of Heat Mass Transfer 52 (23–24) (2009) annular reactors, Int. J. Heat Mass Transf. 52 (23) (2009) 5390–5401.
5390–5401. [43] J.E. Duran, M. Mohseni, F. Taghipour, Modeling of annular reactors with surface
[26] A.M. Diez, et al., A step forward in heterogeneous photocatalysis: Process reaction using computational fluid dynamics (CFD), Chem. Eng. Sci. 65 (3) (2010)
intensification by using a static mixer as catalyst support, Chem. Eng. J. 343 (2018) 1201–1211.
597–606. [44] K. Tong, L. Yang, X. Du, Modelling of TiO2-based packing bed photocatalytic
[27] S.-R.-V. Castrillón, H. Ibrahim, H. De Lasa, Flow field investigation in a reactor with Raschig rings for phenol degradation by coupled CFD and DEM,
photocatalytic reactor for air treatment (Photo-CREC–air), Chem. Eng. Sci. 61 (10) Chem. Eng. J. 400 (2020), 125988.
(2006) 3343–3361. [45] J.O.D.B. Lira, et al., Fluid dynamics and mass transfer in curved reactors: A CFD
[28] D.A. Sozzi, F. Taghipour, Computational and experimental study of annular photo- study on Dean flow effects, J. Environ. Chem. Eng. 10 (5) (2022), 108304.
reactor hydrodynamics, Internat. J. Heat Fluid Flow 27 (6) (2006) 1043–1053. [46] U. Periyathamby, A.K. Ray, Computer simulation of a novel photocatalytic reactor
[29] G. Vincent, et al., CFD modelling of an annular reactor, application to the using distributive computing environment, Chem. Eng. Technol. 22 (10) (1999)
photocatalytic degradation of acetone, Process Safety Environ. Protect. 89 (1) 881–888.
(2011) 35–40. [47] S. Azizi, et al., Degradation of ofloxacin using the UV/ZnO/iodide process in an
[30] Y. Boyjoo, M. Ang, V. Pareek, Some aspects of photocatalytic reactor modeling integrated photocatalytic-biological reactor containing baffles, Ind. Eng. Chem.
using computational fluid dynamics, Chem. Eng. Sci. 101 (2013) 764–784. Res. 59 (52) (2020) 22440–22450.
[31] Queffeulou,, A.l,, et al., Kinetic study of acetaldehyde photocatalytic oxidation
with a thin film of TiO2 coated on stainless steel and CFD modeling approach,
Indust. Eng. Chem. Res. 49 (15) (2010) 6890–6897.

17

You might also like