Radiation Models for Computational Fluid Dynamics Simulations of Photocatalytic Reactors
Abstract
The literature on computational fluid dynamics (CFD) simulations applied to photocatalytic systems is reviewed. CFD simulations referring to three models, namely, P-1, discrete ordinates (DO), and Monte Carlo (MC) models, for simulating radiation distribution in annular photoreactors are addressed and previous works using these models are reviewed. The cases are a three-dimensional (3D) annular photoreactor and a two-dimensional (2D) rectangular enclosure with black walls and the results are compared with the literature. The DO model was selected to solve the radiation transfer equation (RTE) as the most reliable model to fit the radiation distribution inside the reactor. The setting-up of mesh, angular discretization, and boundary conditions for CFD simulation of photocatalytic reactors are addressed. When correctly set, the CFD models show an excellent agreement with the literature results.
1 Introduction
In recent years, photo-assisted technologies have received increasing attention primarily due to the possibility of using the sun as light source, a renewable and economical source of energy that can be used for the photo-activation of several processes, including water treatment 1, 2, chemical synthesis 3, 4, energy production 5, 6, and air treatment 7, 8. The optimization of the photoreactor configuration is crucial for the applicability of this technology. Computational fluid dynamics (CFD) is an efficient design tool to predict and enhance the performance of photocatalytic reactors because it enables the integration of hydrodynamics, mass transfer, reaction kinetics, and photon flux distribution. The models used in the simulations can affect the results obtained. So, the selection of the models is critical in the computational design or numerical studies of photocatalytic reactors 9.
The first works on photocatalysis are dated back to the 20th century 10. Although, only from the 1990s onward the field experienced a boom in the number of publications. Yet, the first research using “CFD” dates to the end of the last year of the 1990s. Fig. 1 shows the annual number of papers until 2022, retrieved from Scopus with the term photocatalysis and CFD in the title, abstract or keywords.

Tab. 1 summarizes the CFD models used for the simulation of radiation in different types of photoreactors. Most works consider annular, pipe or plate reactors. Regarding the simulation of radiation, the discrete ordinates (DO) method is the most used. This survey shows the relative novelty of the radiation simulation in the study of photocatalytic reactors with the first work dating from 2003. The number of papers worldwide on this topic is still below 500 (Scopus, data), although presenting a steady yearly increase. The novelty of the field justifies a short tutorial on the implementation of the radiation models along with their review.
Photocatalytic reactor type |
Application |
CFD models |
Radiation source |
Ref. |
||
---|---|---|---|---|---|---|
Hydrodynamics |
Radiation |
Reaction |
||||
Pilot-scale annular bubble column |
Photodegradation of Bayer liquor |
Three-phase flow |
DO |
|
UV |
11 |
Turbulent k-ε |
||||||
Annular slurry pilot-plant photoreactor |
Photoxidation of pollutants |
|
Analytical solution of RTE |
|
UV |
12 |
Annular photoreactor |
Gas phase vinyl chloride (VC) oxidation |
Laminar |
|
Surface reaction |
|
13 |
Annular photoreactor |
Photochemical reactions (no application specifically) |
|
DO |
|
|
14 |
Annular dual-lamp photoreactor |
No application specifically |
|
DO |
|
UV |
15 |
Annular |
Air treatment |
Laminar |
|
Surface reaction for trichloroethylene (TCE) oxidation |
UV |
16 |
Plate |
Photocatalytic oxidation (PCO) of indoor air pollutants |
Laminar |
LSSE |
Oxidation of trichloroethylene |
UV-A |
17 |
Photo-CREC-air reactor |
Air purification |
Turbulent k-ω based Shear Stress Transport (SST) |
|
Mineralization of acetone to carbon dioxide and water |
UV |
18 |
Multi-annular reactor |
Air Ttreatment |
Laminar |
|
Perchloroethylene (PCE) degradation |
UV |
19 |
Externally irradiated bubble tank |
Pollutant decomposition |
Turbulent k-ε |
DO |
– |
UV-A |
20 |
Multiphase system |
||||||
Flat plate |
Degradation of TCE in a serpentine flow field |
Laminar |
|
Degradation of TCE |
|
21 |
Externally illuminated immobilized catalyst photoreactor |
Water treatment |
Multiphase |
|
Benzoic acid dissolution |
|
22 |
Turbulent k-ε |
||||||
Tubular optical fibers |
Water Treatment |
Laminar |
DO |
Photodegradation of oxalic acid |
Optical fibers |
23 |
Annular |
Photocatalytic treatment of bioaerosols |
Turbulent k-ω |
|
|
UV-A |
24 |
Annular |
Air Treatment |
Laminar |
|
Acetaldehyde photocatalytic oxidation |
UV |
25 |
Plate |
Photocatalytic inactivation of spores of Bacillus subtilis |
|
DO |
Photocatalytic inactivation of spores of Bacillus subtilis |
UV-A |
26 |
Plate |
Air treatment |
Turbulent k-ε |
RTE solution in SPEOS |
Acetaldehyde removal |
UV |
27 |
Annular |
Study of radiation models |
|
ESVE |
|
|
28 |
ESDE |
||||||
LSSE |
||||||
LSDE |
||||||
ESVEA |
||||||
ESVERA |
||||||
Tubular (immobilized catalyst) |
Water treatment |
Laminar |
DO |
Degradation of formic acid |
UV |
29 |
Turbulent RSM |
||||||
Turbulent |
||||||
Realizable k-ε |
||||||
Annular |
Water treatment |
Turbulent AKN |
|
Degradation of benzoic acid |
UV |
30 |
Turbulent RSM |
||||||
Turbulent R k-ε |
||||||
Turbulent S k-ε |
||||||
Laminar |
||||||
Annular slurry |
Wastewater treatment |
Turbulent k-ε |
DT model (discrete transfer) |
Degradation of pollutants in CO2 and H2O |
UV |
31 |
Multiphase-flow |
||||||
Corrugated reactor |
Air treatment |
Laminar |
|
Degradation of formaldehyde |
UV |
32 |
CPC (continuous flow reactor with compound parabolic collectors) |
Water disinfection |
Turbulent S k-ε |
|
Inactivation of E. coli particles |
UV |
33 |
Flat plate |
Air pollution remediation |
Laminar |
|
Degradation of a target pollutant HCOC |
UV |
34 |
Arrow-slit, plat-plate flow-through |
Photocatalytic deodorization process |
Laminar |
Radiation field model of LED |
Photocatalytic degradation of dimethyl sulfide (DMS) |
UV-LED |
35 |
Microchannel reactor |
Photocatalytic degradation of salicylic acid (SA) |
Laminar |
|
Photocatalytic degradation of salicylic acid (SA) |
UV |
36 |
Multiplate reactor |
Air/water purification |
|
Monte Carlo model |
|
|
37 |
Fluidized-bed reactor |
Degradation of pollutants |
|
Photon modeling |
|
|
38 |
Annular photocatalytic reactor |
Water treatment |
Laminar |
|
Mineralization of Rhodamine B |
UV |
39 |
Photo-CREC water II annular reactor |
Wastewater treatment |
|
MC model |
|
UV |
40 |
Annular photocatalytic reactor |
Water purification |
Laminar |
|
Mineralization of Rhodamine B |
UV |
41 |
Slit-shaped flat-bed photocatalytic reactor |
Air treatment |
Laminar |
|
Photocatalytic degradation of gaseous acetaldehyde |
UVA |
42 |
Circular flow cell |
Water treatment |
Laminar |
|
Degradation of phenol |
UV LEDs |
43 |
RVE (representative volume element) of textile |
Degradation of pollutants |
Laminar |
|
– |
UV |
44 |
Turbulent |
||||||
Multitube photoreactor |
Air treatment |
Laminar |
|
Photocatalytic degradation of gaseous acetaldehyde |
UV |
45 |
New reactor based on a stagnation point geometry |
VOCs removal applications |
Turbulent k-ε |
|
Photocatalytic degradation of acetaldehyde |
UV |
46 |
Stacked frame photocatalysis reactor (with impeller) |
Production of biofuels from cellulose |
Turbulent k-ε |
|
|
UV |
47 |
Cascade disc Reactor |
Environmental purification |
Laminar |
|
Benzoic acid dissolution |
|
48 |
Turbulent |
||||||
Bubbling twin Reactor |
Production of valued-added fuels |
Turbulent |
|
CO2 reduction |
UV |
49 |
Two-phase |
||||||
Silicone microreactors |
Photocatalytic production of hydrogen |
Laminar |
|
Mixture of water-ethanol to produce hydrogen |
UV |
50 |
Internal air lift circulating photoreactor |
Pollutants degradation |
Turbulent S k-ε |
|
|
|
51 |
Three-phase flow |
||||||
Annular slurry bubble column reactor |
Wastewater Treatment |
|
Monte Carlo |
|
UV |
52 |
Annular |
Degradation of organic pollutants |
Laminar |
|
Mineralization of toluene |
UV |
53 |
Gas-phase photocatalytic multitube reactor |
Indoor air treatment |
Laminar |
Ray optics module of Comsol v5.3a |
Acetaldehyde degradation |
UV-A |
54 |
Turbulent S k-ε |
||||||
Lab-scale multitube reactor |
Air VOCs treatment |
Laminar |
Ray optics module of Comsol v5.3a |
Acetaldehyde degradation |
UV |
55 |
Continuous flat-plate photoreactor |
Air purification |
Laminar |
DO |
Photocatalytic NOx removal |
UV |
56 |
Flat plate |
NOx abatement |
Laminar |
|
NO2 consumption |
|
57 |
CPC-PR |
Water and air treatment |
Turbulent S k-ε |
DO |
Formaldehyde formation |
Solar radiation |
58 |
Production of fine chemicals |
||||||
Lamellae |
Air purification |
Laminar |
|
VOCs degradation |
UV |
59 |
Photocatalytic multitube reactor |
Air pollution remediation |
Turbulent k-ε |
Ray optics module of Comsol v5.3a |
Acetaldehyde degradation |
UV |
60 |
Self-made airtight reactor |
Air treatment |
Turbulent RNG k-ε |
|
Degradation of formaldehyde |
UV |
61 |
PCO (photocatlytic oxidation) reactor |
Air purification applications |
Turbulent k-ω |
|
|
UV |
62 |
1. Annular reactor |
Advanced oxidation processes |
|
DO |
|
UV-A (isotropic emission) |
63 |
2. Tubular with CPC |
||||||
3. Tubular with LEDs |
CPC (parallel emission) |
|||||
UV-A LEDs (cone-shaped and power-cosine emission) |
||||||
Photocatalytic microreactor |
(Waste)water treatment |
Laminar |
|
Photodegradation of 4-nitrophenol |
UV-visible |
64 |
Urban Street Canyon |
Air pollution treatment |
Turbulent RNG k-ω |
|
Photocatalytic oxidation of NOx |
UV |
65 |
Flat-plate photocatalytic reactor |
Air treatment |
Laminar |
DO |
NOx abatement |
|
8 |
Annular photoreactor |
Water treatment |
|
P-1 |
MO (methyl orange) dye degradation |
UV |
66 |
Parallel-channel microreactor |
Water treatment |
Laminar |
|
Photodegradation of 4-nitrophenol |
UV |
67 |
An urban area |
Air treatment |
Turbulen Realizable k-ε |
|
Degradation of NOx |
UV |
68 |
NETmix reactor |
Water/air treatment |
|
MC |
|
LEDs |
69 |
The modeling of photoreactors requires the study of (i) hydrodynamics of the reactor, (ii) radiation transfer equation (RTE), (iii) reaction kinetics, and (iv) mass balance in the light and dark zones of the photoreactors. At first, the hydrodynamics of the system is modeled and simulated, enabling the full spatial description of the flow inside the reactor. The characterization of the radiation field of a photochemical or photocatalytic reactor involves the incorporation of optical phenomena (reflection, refraction, and shadow zones), lamp characteristics, and its position in the photoreactor configuration 70.
The kinetics of the reaction is incorporated in the simulation, which also depends on the radiation distribution inside the reactor. Finally, the mass balance is the heart of the model, integrating the hydrodynamics model, radiation model (emission and scattering-absorption), and kinetic model which includes a quantum yield model. The sequential steps for the modeling of photocatalytic systems are presented in Fig. 2 71, 72. Some studies in Tab. 1 consider only one or two of these components; e.g., some use the CFD to simulate the RTE, while others focus only on the flow in the photoreactor.

The steps (i) and (ii) can be simulated separately. If the thermal contribution from radiation causes a negligible impact on the flow, the radiation field can be simulated over the hydrodynamics. If the medium is homogeneous regarding light absorption and refraction index, the flow field will have no impact on radiation. The latter is generally not the case in chemical reactors where the concentration distribution of chemical species evolves in space or time 9, 73. The final step is the simulation of mass balance, which requires all the previous steps.
Radiation distribution is a crucial parameter in the simulation of photochemical and photocatalytic systems 74-76. The light intensity distribution inside the reactor is simulated by solving the RTE and it depends on parameters such as type and position of the lamp, reactor geometry, radiation wavelength, the optical properties of the medium, specifications of reactor walls, and operating conditions (inflow and outflow rate, catalyst type and concentrations) 77, 78.

where is the position vector,
is the direction vector,
is the scattering direction vector, s (m) is the path length, α (m−1) is the absorption coefficient, n is the refractive index, σs (m−1) is the scattering coefficient, σ (5.669 × 10−8 W m−2K−4) is the Stefan-Boltzmann constant, I (W m−2sr−1) is the radiation intensity which depends on position (
) and direction (
), T (K) is the local temperature, ϕ is the phase function, and
(sr) is the solid angle. (α + σs) is the optical thickness or opacity of the medium 80.
Therefore, the RTE describes the radiation transport phenomena: the attenuation by absorption (), the scattering (out scattering) (
), and source terms by emission (
) and by radiation in-scattered from other directions along a path (
), as shown in Fig. 3
79.

An exact analytical solution of the RTE is only possible for one-dimension ideal situations, after simplifying assumptions such as uniform radiative properties of the medium and homogeneous boundary conditions 79, 81.
Radiation is present in many engineering problems such as photocatalysis for water treatment and fine chemicals production and most of these systems are multidimensional, the media is heterogeneous, and radiative properties are spectral. The solution of RTE requires the evaluation of scattering and absorption coefficients of the media and of the phase function, which depends on wavelength, fluid composition, type and content of suspended particles, temperature, and pressure 79.
The radiative transfer models can also be applied for immobilized coatings. The immobilized catalyst system presents some advantages when compared with the suspension systems, since at the end of the reaction a separation process is not needed and is particularly useful for scale-up efforts. Claes et al. report the study of a translucent structured photocatalytic reactor, presenting the results of the influence of the base structure on absorption efficiency and mass transfer limitations 82. These reactors are an effective design to scale up photocatalytic processes, enabling the increase of the catalyst loading without increasing the catalyst layer thickness, leading to fewer diffusion limitations 83.
It is assumed that scattering of light and radiation absorption in the catalyst coating can be neglected for some immobilized systems such as optical fibers, coated plates, and monoliths. For photochemical systems carried out at low or medium temperatures, the energy gain due to emission in reaction media can be also neglected 84. The scattering effects can be ignored in homogeneous reaction mixtures because they are rather weak 84. In these cases, the RTE reduces to the Beer-Lambert law 71.
Numerical simulation has gradually become a powerful technique to study the radiative heat transfer in absorbing, emitting, and scattering media because radiation models are capable of simplifying the integro-differential equations and predicting the radiation properties of the media with accuracy 79, 81. CFD is an efficient design tool for photoreactors because it enables the integration of hydrodynamics, reaction kinetics, and photon flux distribution 28, 75. The knowledge of the radiant energy distribution, particularly of the UV fraction, is fundamental in designing and optimizing photoreactors.
Annular geometry is the most commonly used in photoreactors for high-throughput and continuous operation 71. In this configuration, one or more UV lamps are placed in the middle of a cylindrical vessel. Each lamp is enclosed in its quartz sleeve to isolate it from the reacting medium. This photoreactor promotes a non-uniform distribution of radiation, characterized by higher radiation intensities close to the source and lower intensities near the reactor wall, because of the increased photon flux area and the optical thickness. This problem can be minimized if the annular region is reduced or if highly reflective reactor walls are employed instead 71.
The present critical review of radiation modeling in a CFD package is based on a case with an annular geometry proposed by Quan et al. 85.
2 State-of-the-Art
The modeling of photoreactors involves the simulation of fluid flows and the solution of the RTE within the medium. Several methods have been developed to solve the RTE, such as the P-1 model, the discrete ordinates model (DOM), which are approximations of the solution of the RTE, and the Monte Carlo (MC) model, which involves a statistical treatment 50.
In addition to accurately accounting for catalyst activity, the primary motivation for modeling flow and radiation in photocatalytic reactors is to identify regions of poor catalyst illumination 83. The rate of reaction in an immobilized system is related to incident radiation. In this paper, the incident radiation is simulated to evaluate the light intensity distribution in an annular photocatalytic reactor and assess the reliability and accuracy of radiation models. This showcase takes a tutorial approach to show the implementation of radiation modeling, in the present case using ANSYS/Fluent that is a widespread commercial CFD package distributed with built-in radiation models.
2.1 Radiation Models for CFD
The Radiation Models described in this work are those implemented in the commercial software package ANSYS/Fluent. Table 2 summarizes the advantages and limitations of the radiation models.
Advantages |
Disadvantages |
---|---|
P-1 Model |
|
Simple diffusion equation for the incident radiation (G) |
All surfaces are considered diffuse |
Low computational effort |
Loss of accuracy depending on the complexity of geometry |
Accurate for large optical thickness cases (αL > 1) |
Loss of accuracy if participating media are optically thin |
Includes particulate (and anisotropic) scattering |
Loss of accuracy at localized heat sources/sinks (overprediction of the radiative heat flux). |
Non-gray radiation modeled using a gray-band model |
Assumes gray gases |
Easily applied to complex geometries with curvilinear coordinates |
|
DO Model |
|
Applicable to all optical thicknesses |
High computational effort |
Includes particulate and anisotropic scattering |
Finite number of radiation directions causes numerical smearing |
Radiation in semi-transparent media |
|
Diffuse and specular reflection |
|
High accuracy |
|
MC Model |
|
Wide range of optical thicknesses |
High computational cost |
Allows quasi-exact solutions |
The physical quantities are calculated as surface or volume averages |
|
Some cases are not supported by this model. |
|
Fluent only supports Monte Carlo Model for 3D geometries |
2.1.1 P-1 Model
The P-1 radiation model is a widely used numerical method to solve the RTE in heat transfer. The P-1 approximation is the simplest case of the more general P-N model, which is based on the expansion of the radiation intensity into an orthogonal series of spherical harmonics 79, 80, 86. The direction dependence in RTE is integrated out, transforming the RTE into a diffusion equation, which is easy to solve and compatible with the finite volume method 80, 87.


Where σs is the scattering coefficient and C is the linear-anisotropic phase function.



Therefore, incident radiation can be obtained by solving the simpler diffusion equation 87. The non-gray radiation can be modeled using the P-1 model, with a gray-band model.
The first reported work of P-1 model was in 1996, when Sazhin et al. presented the advantages and limitations of the P-1 model for thermal radiation, demonstrating that the P-1 model is expected to make reliable predictions in optically thick and thin media in a simple geometry 86. The P-1 approximation was used to model radiative transfer in a rectangular enclosure in participating media 88, in an annular photoreactor 77, 87, and a closed-conduit reactor 89.
2.1.2 Discrete Ordinates (DO) Model
The DO model transforms the integro-differential form of the RTE into a system of algebraic equations 72. The RTE is solved for a finite number of discrete solid angles, each associated with a vector direction () 80. Thus, the DOM solves the RTE through directional and spatial discretization, giving a set of linear simultaneous equations for the radiation intensity at various points 78. The DO model solves for as many transport equations as there are directions (
) 79, 80.

where the variables are: the radiation intensity,
the position vector,
the direction vector,
the scattering direction vector, α the absorption coefficient of the medium, σs the scattering coefficient, n the refractive index of the medium, σ the Stefan-Boltzmann constant (5.67 × 10−8 W m−2K−4), T the local temperature, ϕ the phase function, and
the solid angle.
As for the P-1 model, the non-gray radiation can be performed using the DO model, with a gray-band model, which divides the radiation spectrum into N wavelength bands.
In the DO model, the RTE is integrated over the control volume (existing mesh) using a finite volume method 80, 84. For the case of anisotropic scattering at complex surfaces, the DO model does not conserve radiant energy. Some CFD codes implemented a conservative variant of the DO to tackle this issue 91.
Through the years, the DOM was used in the simulation of several types of photochemical or photocatalytic (immobilized or slurry) reactors due to its accuracy. Romero et al. proposed the DOM to solve the RTE in a heterogeneous reaction medium in a photocatalytic system with an annular geometry 92. Subsequently, many researchers studied the radiation field using the DOM in annular photoreactors 11, 78, 90, 93. The radiation field in a flat-plate reactor using DOM was investigated by Brandi et al. 94.
Sgalari et al. computed the radiative transfer with DOM within an annular photoreactor for typical situations of photocatalytic processes, investigating the effects of the parameters on the local rate of radiant energy absorption and the angular distribution of the radiation intensity 93. The radiation transport equation was solved using a finite-volume-based DO method for an annular bubble-column photocatalytic reactor with a three-phase flow for photodegradation of Bayer liquor 11. Pareek et al. applied the MC approach to estimate light intensity distribution (LVREA) in an annular photocatalytic reactor and the DO model to assess the effect of wall reflectivity, catalyst loading, and phase function parameter on the light intensity distribution 78.
Bagheri and Mohseni developed a CFD model for simulating annular VUV/UV photoreactors considering reflection, absorption, and refraction of photons by the wall and slurry media in radiation modeling using the DO method 90. Li et al. used the DOM to study the wall reflection on an annular single-lamp UV reactor, which directly affects the radiation distribution 95. Asadollahfardi et al. presented a comparison of the reliability of the DOM with MC predictions for solving RTE 76. The DOM is widely applied to model steady-state radiative transfer in media 96.
2.1.3 Monte Carlo Model
The Monte Carlo (MC) radiation model generates photons randomly in a stochastic method and will, therefore, produce speckled results if the target number of histories is relatively small. Increasing the target number of histories produces a smoother and more accurate solution, but at the expense of higher computation effort 79. The MC model considers that the intensity is proportional to the differential angular flux of the photons. Then, the mean radiation intensity is obtained by computing the distance travelled for a typical selection of photons, in each volume cell 80.
An MC approach is used to solve the radiant energy balance equation and determine scattering phenomena in photosensitized reactions for continuous processes by Spadoni et al. 97. The radiation field inside an annular photocatalytic reactor was performed using an MC model in a specific case of radiative transfer in absorbing-scattering media 98. Imoberdorf et al. presented the MC approach to simulate the radiation field in single and multi-lamp homogeneous photoreactors with an annular geometry, showing the importance of the reflection, refraction, and absorption of photons by the lamp and the wall 99. The MC model was employed to model and optimize the radiation field in a multiplate photocatalytic reactor (MPPR – parallel photocatalytic plates irradiated by cylindrical UV lamps orthogonal to the plates) for air or water purification 100.
3 Case Study
The models for simulation of radiation in CFD are assessed using a case study with a tutorial approach to the models' implementation. In this tutorial, the UV-intensity fields were simulated. In the base case, a 3D annular photoreactor and a 2D rectangular enclosure with axisymmetry (axial cut the 3D annular geometry) with black walls were considered as showcases.
3.1 Geometry Model and Computational Grid
The case study reactor has an annular geometry with 645 mm of length, 15 mm of external radius, and 2.25 mm of internal radius. In this configuration of the photoreactor, a UV lamp was placed in the middle of the cylindrical vessel. In the laboratory case, the lamp is enclosed in a quartz sleeve to isolate it from the reacting medium, but in the simulated system the quartz medium was not considered. This reactor has been used in several experimental and numerical works 77, 85, 87. The photoreactor shown in Fig. 4a was simulated in 2D and 3D. The inlet and outlet correspond to the circular crowns at the ends of the cylindrical reactor. The geometry model was generated using ANSYS DesignModeler®.

After the generation of the geometry, the computational grid was generated using ANSYS Meshing®. The geometric model was discretized, subdividing it into small finite volumes that define the computational grid and where the equations are solved. The generation of the grid is one of the most critical steps of the CFD analysis because the size of the elements influences the accuracy of the results due to numerical approximations. The mesh structure is quite relevant for the simulation of radiation. Thus, there is a compromise between the accuracy of the results and the computational effort required.
Figs. 4b and 4c represent the computational grid of the 3D model and 2D model, respectively. The 2D meshes are made of regular quadrilateral elements. Refinement in this mesh consists of the division of the quadrilaterals. The 3D mesh is made mainly of hexahedral elements, which start as quadrilaterals in the top face and are repeated in layers along the heigh. This repetition creates a regular mesh.
Tab. 3 summarizes the properties of the studied meshes. The 2D meshes were generated using several element sizes (meshes 1, 2, and 3). The lamp wall was identified as the most critical region of the mesh. Thus, meshes 4 and 5 were generated to refine the region near the lamp wall by adding several layers. The 2D domain was generated using an axisymmetric model due to the symmetric condition of the photoreactor. The 3D mesh 7 was refined in Ansys Fluent using different levels of refinement near the lamp wall.
Mesh |
Dimension |
Element size [mm] |
Layers refinement on lamp wall |
Number of elements |
---|---|---|---|---|
1 |
2D |
3.0 |
– |
9.7 × 103 |
2 |
2D |
1.0 |
– |
9.0 × 104 |
3 |
2D |
0.1 |
– |
9.0 × 106 |
4 |
2D |
3.0 |
20 |
1.4 × 104 |
5 |
2D |
3.0 |
67 |
2.5 × 104 |
6 |
3D |
3.0 |
– |
9.7 × 105 |
7 |
3D Fluent Level 4 |
3.0 |
4 |
1.3 × 106 |
8 |
3D Meshing inflation |
10.0 |
40 |
3.8 × 105 |
3.2 Selection of Radiation Model
The DO model using the finite volume method (FVM) of ANSYS Fluent R2020® was chosen to simulate the RTE. This is a promising method for modeling radiative transfer due to its extensive applicability and high accuracy.
The P-1 model was not employed because the chosen case had low optical thicknesses. The medium which defines the optical thickness is air and the maximum length in the geometry of the problem is approximately 0.15 m. The MC model was excluded due to its ergodic treatment of the problem. To obtain good approximations of the physical quantities, a high number of histories of the photons was necessary, which results in a high computational cost.
3.3 Lamp Characteristics
The presence of radiation is the distinguishing factor of the photoreactors, so it is necessary to consider the characteristics of the radiation source (lamp). The reactor simulated has a single lamp concentric to the body reactor.

Where dlamp (m) is the lamp diameter and L (m) is the lamp length.
3.4 Boundary Conditions
In the CFD analysis, air was considered as the fluid in the annular space, and its properties were set using the standard models available in Fluent. For the external reactor wall and the wall of the lamp, the properties of fused quartz were considered. The properties of the materials are provided as Supporting Information.
The simulation of the radiation distribution implies the calculation of irradiation, which corresponds to the radiation emitted by the lamp to set as boundary condition, considering the lamp a semi-transparent wall. Then, the simulations were carried out simulating an infinite cylinder, defining as radiation boundary condition the intensity of the radiation emitted by the lamp.
For both 2D and 3D cases, the flow was set as stagnant. For 2D cases, air inlet and outlet were considered as interface boundaries linked by a periodic condition. Because the pressure gradient between the boundaries was set to ΔP/x = 0 Pa m−1, there is no flow field in this case. In 3D cases, air inlet and outlet were considered as walls, with the three velocity components set to zero. The boundary conditions used in the 2D and 3D models are summarized in Tabs. 4 and 5, respectively.
Boundary |
Boundary conditions |
---|---|
Name: Air inlet |
Periodic BC – linked to outlet ΔP/x = 0 |
Type: Interface boundary |
|
Name: Air outlet |
Periodic BC – linked to inlet ΔP/x = 0 |
Type: Interface boundary |
|
Name: Lamp wall |
Hflux = 0 W m−2 |
Type: Wall |
Ilamp = 127.3 W m−2 |
Semi-transparent wall, direct irradiation, ε = 1 |
|
Velocity non-slip condition |
|
Name: Reactor wall |
Hflux = 0 W m−2 |
Type: Wall |
Ilamp = 0 W m−2 |
Semi-transparent wall, direct irradiation, ε = 1 |
|
Velocity non-slip condition |
Boundary |
Boundary conditions |
---|---|
Name: Air inlet |
Velocity components set to 0 |
Type: Wall |
Opaque wall, ε = 1 |
Name: Air outlet |
Velocity components set to 0 |
Type: Wall |
Opaque wall, ε = 1 |
Name: Lamp wall |
Hflux = 0 W m−2 |
Type: Wall |
Ilamp = 127.3 W m−2 |
Semi-transparent wall, direct irradiation, ε = 1 |
|
Velocity non-slip condition |
|
Name: Reactor wall |
Hflux = 0 W m−2 |
Type: Wall |
Ilamp = 0 W m−2 |
Semi-transparent wall, direct irradiation, ε = 1 |
|
Velocity non-slip condition |
To obtain the radiation field from the RTE without the effect of thermal radiation, the temperature of the domain was set to1 K (minimal temperature supported in ANSYS Fluent R2020®). The flow was stagnant, so although all sets of equations were defined, as well as the boundary conditions (BCs), the actual flow equations were not solved, and so the flow variables were all set to zero. This enables to focus solely on the issues pertaining to the simulation of radiation in CFD, particularly with ANSYS/Fluent.
3.5 Numerical Solution and Convergence Criteria
In the chosen case the flow is stagnant and the boundary conditions for the radiation and the energy are constant. So, the solvers were set to steady state and only the sets of equations for radiation and energy are solved. The discretization scheme is second-order upwind for the energy and second-order upwind for discrete ordinates. The convergence criterion for the numerical solution was residuals should be less than 10−20. The numerical mesh validity is checked in Sect. 4.1, for the several meshes described in Sect. 3.1.
4 Validation and Evaluation of Radiation Models
First, the mesh quality was evaluated to ensure the lowest numerical diffusion and the most reliable results for the 2D axisymmetric and 3D configurations. Then, angular discretization and absorption coefficient were investigated to determine their influence in the radiation field. Finally, a comparison between simulation results and previous works is made to assess the validity of the results obtained with 2D axisymmetric and 3D models.
4.1 Study of the Computational Grid
Assuming a case with a medium that does not absorb light, the mesh element size and the number of layers in the lamp wall were assessed. The 2D meshes were refined using ANSYS Meshing®, one of the 3D meshes was refined in Ansys Fluent using the refinement tool and another in ANSYS Meshing®. In 3D cases, the simulations were carried out using 20 × 20 directions and 10 × 10 pixilation of angular discretization Figs. 5 and 6 present the light distribution along the radius at a line in the mid-axial position, for 2D and 3D models, respectively.


It is observed that the refinement of the grid in the lamp wall influences the result of the radiation distribution. Only the more refined meshes around the lamp sleeve can describe accurately the radiation intensity. Away from the lamp, the local values in the neighbouring of the sleeve do not influence the overall radiation profile. For the simulation of radiative transfer, it is used a mesh able to accurately describe the full radiation profile. The selected mesh has elements with a size of 3.0 mm where refinement of the 67-layers in the lamp wall was considered for the 2D axisymmetric model. In the 3D model, a grid with an element size of 10 mm and refinement of the 40-layers shows the best results.
These findings indicate that the 2D model describes the same radiation profile behavior as the 3D model. This accordance between geometric models was expected because the 3D case is axisymmetric, with complete homogeneity through the angular direction.
4.2 Study of the Angular Discretization
Angular discretization and pixelation play a major role to obtain a reliable solution for RTE. At any spatial location, each octant of the angular space 4π is divided into solid angles (Nθ × Nϕ), in which θ and ϕ represent polar and azimuthal angles. In 2D calculations, there are only 4Nθ × Nϕ directions since only four octants are solved due to symmetry. In 3D cases, 8Nθ × Nϕ directions are solved. For Cartesian meshes, it is possible to align the global angular discretization with the control volume face. However, when generalized unstructured meshes are used, control volume faces do not align with the angular discretization, leading to control angle overhang. Control angle overhang can be corrected using pixelation. The overhanging control angle is discretized into Nθp × Nϕp 80.
The RTE is simulated for two angular discretizations: (i) 5 × 5 directions (solid angles) and pixilation of 3 × 3 and (ii) 20 × 20 directions and pixilation of 10 × 10 were performed considering the non-absorption case (α = 0 m−1) for 2D and 3D models. Figs. 7 and 8 represent the radiation profile and radiation intensity results for each case of angular discretization for 2D and 3D models, respectively.


The results of angular discretization indicate that for the 2D case the angular discretization does not have a role in the distribution of radiation results, because radial homogeneity is already implicit when simulating only a slice of the reactor. For the 3D case, an increase in angular discretization enables more uniform distribution of light, namely in the lamp wall.
Fig. 8 displays the radiation intensity map for two angular discretization values. The visual effect of the discretization is striking even at the lamp surface with the occurrence of stripes of larger intensity. This heterogeneity is propagated into the medium, as seen in the axial cut of the reactor. The effect of angular discretization is quantified in Fig. 8, where the angular distribution of the radiation intensity is shown at five radial locations. For the less dense angular discretization, the radiation has a wavy pattern originating at the source boundary. Thus, the simulations for 3D cases were performed using an angular discretization of 20 × 20 directions and 10 × 10 pixilation.
4.3 Effect of Absorption Coefficients in Radiation Profiles
For the study of absorption coefficients, only 2D simulation was performed, since for the non-absorption case the 2D and 3D results were very similar. The RTE was solved by discretizing each control volume into 5 × 5 directions (solid angles) and pixelation of 3 × 3 for the 2D model. After the numerical validation of the simulation conditions, the models were used for absorption coefficients in the range of α = 0 to 357.6 m−1. The 2D light distribution profiles for the several absorption coefficients are depicted in Fig. 9.

The incident radiation map inside the annular reactor shows the overall profile from the lamp towards the outer wall. The distribution presented in Fig. 9 can be expected for the annular configuration of the photoreactor due to the radial increase of the section for radiation associated with the medium absorption. The radiation profile is flat throughout the reactor length due to the periodic BCs, so a representative profile can be obtained along any axial position. The dimensionless light distribution profiles along the radius are depicted in Fig. 9, which conveys a more detailed picture of the effect of the absorption coefficient.
The profiles presented in Fig. 9 show that the light intensity decreases rather sharply with the increase in the absorption coefficient. For the case with zero absorption coefficient, the decrease of incident radiation is due to the increment of the photonic flow area. When the absorption coefficient is equal to 357.6 m−1, 90 % of the incident light is absorbed in the initial ≈ 1.5 mm away from the lamp. This indicates that the exposure of the catalyst to radiation, when there are heterogeneities of absorption in the medium, will depend on the flow patterns and degree of mixing within the reactor in the photocatalytic systems. If the flow is very segregated promoting large gradients in the light absorption throughout the reactor, the illumination of the catalyst in the outer wall will be very heterogeneous.
4.4 Comparison with Reported Data

where I is the radiation intensity (W m−2), which depends on initial radiation intensity (W m−2), I0, the reactor radius (m), r0, the local radius (m), r, and absorption coefficient (m−1), α. The results are shown in Fig. 10.

From Fig. 10 it was concluded that the models to simulate radiation distribution inside an annular reactor described the same behavior as reported in the literature. Also, the proposed method to solve the RTE by the Lambert-Beer law proposed by Abu-Ghararah 103 displays the same curve shape for all absorption coefficients. Thus, this review presents a reliable CFD method to simulate the radiation distribution inside an annular reactor.
5 Conclusions
A review of radiation models is presented. The discrete ordinate model was the most reliable model to predict the radiation distribution inside a reactor. Using this model, a base case analysis was carried out and sequential steps of simulation radiation inside an annular reactor were described. Several critical parameters to obtain reliable results were investigated such as mesh refinement and angular discretization.
The model results were compared with the ones reported in the literature and with the Lambert-Beer law. The radiation distribution of 2D and 3D models proposed describe the same behavior of radiation distribution as in the literature, namely, Quan et al. (2004) 85, Yu et al. 87, and Huang et al. 77. Therefore, this review represents a tutorial for radiation simulation inside an annular reactor.
Supporting Information
Supporting Information for this article can be found under DOI: https://doi.org/10.1002/ceat.202200551.
Acknowledgements
This work was financially supported by LA/P/0045/2020 (ALiCE), UIDB/50020/2020, and UIDP/50020/2020 (LSRE-LCM), and by the project SuN2Fuel (2022.04682.PTDC), funded by national funds through FCT/MCTES (PIDDAC). I. S. O. Barbosa acknowledges her FCT grant UI/BD/151092/2021.
The authors have declared no conflict of interest.
Symbols used
-
- a [m2s−1]
-
thermal diffusivity
-
- C [–]
-
linear-anisotropic phase function
-
- CP [J kg−1K−1]
-
heat capacity
-
- dlamp [m]
-
lamp diameter
-
- G [W m−2]
-
incident radiation
-
- Hflux [W m−2]
-
heat flux
-
- I [W m−2sr−1]
-
radiation intensity
-
- Ilamp [W m−2]
-
lamp irradiance
-
- I0 [W m−2]
-
initial radiation intensity
-
- k [W m−1K−1]
-
thermal conductivity
-
- n [–]
-
refractive index
-
- N [–]
-
number
-
- Plamp [W]
-
power output of the lamp
-
- qr [W m−2]
-
radiation flux
-
- r [m]
-
local radius
-
- r0 [m]
-
reactor radius
-
[–]
-
unit position vector
-
- s [m]
-
path length
-
[–]
-
unit direction vector
-
[–]
-
unit scattering direction vector
-
- SG [W m−3]
-
user-defined radiation source
-
- T [K]
-
local temperature
Greek letters
-
- α [m−1]
-
absorption coefficient
-
- Γ [m]
-
diffusivity of the incident radiation
-
- ε [–]
-
emissivity
-
- θ [sr]
-
polar angle
-
- μ [Pa s]
-
viscosity of the fluid
-
- ρ [kg m−3]
-
density of the fluid
-
- σ [W m−2K−4]
-
Stefan-Boltzmann constant
-
- σs [m−1]
-
scattering coefficient
-
- φ [–]
-
phase function
-
- ϕ [sr]
-
azimuthal angle
-
[sr]
-
solid angle
Sub- and superscripts
-
- Lamp
-
Lamp
-
- 0
-
Initial
-
- Flux
-
Flux
Abbreviations
-
- 2D
-
two-dimensional
-
- 3D
-
three-dimensional
-
- AKN
-
Abe-Nagano-Kondoh
-
- CFD
-
computational fluid dynamics
-
- CPC
-
compound parabolic collectors
-
- CPC-PR
-
compound parabolic collectors photocatalytic reactor
-
- DO
-
discrete ordinates
-
- DOM
-
discrete ordinates model
-
- DT
-
discrete transfer
-
- ESDE
-
extensive source superficial diffusive emission
-
- ESVE
-
extensive source volumetric emission
-
- ESVEA
-
modified ESVE with photon absorption
-
- ESVERA
-
modified ESVEA with reflection, refraction, and absorption
-
- FVM
-
finite volume method
-
- LED
-
light-emitting diode
-
- LSDE
-
line source diffuse emission
-
- LSSE
-
line source spherical emission
-
- LVREA
-
local volumetric rate energy absorption
-
- MC
-
Monte Carlo
-
- MO
-
methyl orange
-
- MPPR
-
multiplate photocatalytic reactor
-
- PCO
-
photocatalytic oxidation
-
- R
-
realizable
-
- RNG
-
renormalization group
-
- RSM
-
Reynolds stress model
-
- RTE
-
radiation transfer equation
-
- S
-
standard
-
- SA
-
salicylic acid
-
- SST
-
shear stress transport
-
- TCE
-
trichloroethylene
-
- UV
-
ultraviolet
-
- UV-A
-
ultraviolet A
-
- UV-Vis
-
ultraviolet visible
-
- VC
-
vinyl chloride
-
- VOC
-
volatile organic compound