Next Article in Journal
Enhanced Catalytic Reduction of 4-Nitrophenol over Porous Silica Nanospheres Encapsulating Pt-SnxOy Hybrid Nanoparticles
Previous Article in Journal
Tin Complexes Derived from the Acids Ph2C(X)CO2H (X = OH, NH2): Structure and ROP Capability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Degradation of Pollutants in Air Streams Using Luminous Textiles Under Ultraviolet Light Illumination: A Pilot-Scale Remediation Study

1
Laboratory Process Engineering and Industrial Systems (GPSI), ENIG, University of Gabes (UG), Zrig 6029, Gabes, Tunisia
2
College of Engineering, Imam Mohammad Ibn Saud Islamic University (IMSIU), P.O. Box 5701, Riyadh 11432, Saudi Arabia
3
College of Science, Imam Mohammad Ibn Saud Islamic University, Riyadh 11623, Saudi Arabia
4
Laboratoire de Génie des Procédés Chimiques, Department of Process Engineering, University of Ferhat Abbas, Setif 19000, Algeria
5
Department of Basic and Applied Sciences (DiSBA), University of Basilicata, Via Dell’ateneo Lucano 10, 85100 Potenza, Italy
6
Laboratoire de Gestion et Valorisation des Ressources Naturelles et Assurance Qualité, Faculté SNVST, Université de Bouira, Bouira 10000, Algeria
7
School of Engineering, Merz Court, Newcastle University, Newcastle upon Tyne NE1 7RU, UK
8
Ecole Nationale Supérieure de Chimie de Rennes, University Rennes, CNRS, ISCR-UMR6226, F-35000 Rennes, France
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(3), 262; https://doi.org/10.3390/catal15030262
Submission received: 2 February 2025 / Revised: 5 March 2025 / Accepted: 6 March 2025 / Published: 9 March 2025

Abstract

:
Air pollution from volatile organic compounds poses significant environmental and public health issues due to their toxicity and persistence in the environment. In this context, this experimental study explored photocatalytic degradation as a promising approach for the degradation of two polluting fatty acids, butyraldehyde (BUTY) and isovaleraldehyde, utilizing a TiO2 photocatalyst-supported nonluminous textile within a continuous planar reactor. The impact of varying airflow rates (2 to 6 m3/h), initial pollutant concentrations (10 to 60 mg/m3), and air relative humidity (5 to 90%) on oxidation performance and removal efficiency were systematically investigated. The following optimal conditions were identified: an inlet concentration of 10 mg/m3, an airflow rate of 2 m3/h, a catalyst mass of 25 g/m2, a UV intensity of 2 W/m2, and 50% RH. The luminous textile photocatalytic degradation exhibited notable effectiveness for BUTY removal. To enhance our understanding, a mass transfer model using the Langmuir–Hinshelwood approach as a kinetic model was developed. This modeling approach allowed us to determine kinetic adsorption and degradation constants, reasonably agreeing with the experimental data. This study provides valuable insights into applying nonluminous textile-supported TiO2 photocatalysts for environmental pollutant removal in continuous planar reactors.

1. Introduction

Air pollution, primarily caused by human activities, has become a serious health challenge to humans and other living organisms [1]. Several studies have shown that myriad volatile organic compounds (VOCs) with distinct concentrations are emitted by various industrial processes, agricultural activities, and animal husbandry industries [2]. VOCs have caused fatal diseases, including chronic respiratory and pulmonary problems, ischemic heart disease, and lung cancers. Emitted VOCs are also responsible for diverse environmental issues, such as global warming and ozone layer depletion [3,4]. Hence, their mineralization and/or removal has become vital. Several technologies, such as adsorption [5], thermal and catalytic incineration [6], ozonation, and biofiltration [6,7], have been employed to improve air quality and solve the problems caused by these pollutants. Nonetheless, these techniques have limitations, like secondary VOC emissions, solid waste generation, and considerably low energy efficiencies. Researchers are innovating a novel procedure to comprehensively remove volatile organic compounds (VOCs) from the air followed by efficient mineralization, ensuring high process-energy efficiency.
Advanced oxidation processes (AOPs), such as photocatalysis, non-thermal plasma (NTP) catalysis, and catalytic oxidation technologies, have been employed as promising alternatives for efficient and low-cost VOC removal from gaseous effluents to overcome the demerits of usual processes. These advanced oxidation processes (AOPs), generally employed in situ, generate highly reactive species such as hydroxyl (°OH) radicals at room temperature. Among these AOPs, photocatalysis is the most widely used air purification technique as it mineralizes VOCs into CO2 and H2O under ambient conditions [8]. It is based on UV irradiation on a semiconductor catalyst, mainly TiO2, to convert the pollutants into less harmful products with a series of reactions involving oxidizing agents like reactive oxygen species (O2° and °OH radicals) [9,10].
Table 1 summarizes the studies investigating the removal of volatile organic compounds (VOCs) using various advanced oxidation process (AOP) combinations. The synergy of AOPs with diverse physicochemical methods demonstrates their effectiveness at degrading VOCs. Notably, the research findings consistently highlight the superior efficacy of photocatalytic treatments for achieving comprehensive VOC removal. Photocatalytic oxidation-based air purification relies on semiconductors as photocatalysts, including TiO2, ZnO, ZrO2, and SnO2. Among these, due to its physicochemical properties, TiO2 is the most efficient and widely used catalyst for VOC removal in photocatalysis [11,12]. Hence, exploring novel TiO2 photocatalysts with varied textures and chemical characteristics is a thriving area of research. This dynamic field has seen the application of diverse, innovative catalysts, encompassing nanoparticle suspensions [13], nanotubes [14], and nanoparticles deposited on disparate substrates [13,14].
However, TiO2 powder is associated with many problems; for example, the photocatalytic efficiency of TiO2 increases with a decrease in particle size [15]. Notably, a decline in the TiO2 particle size causes difficulty in the post-treatment separation of particles from a solution [16,17]. Therefore, developing photocatalytic films is essential to avoid nanoparticle leaching in the environment. Many techniques, such as sol–gel and impregnation, have been adopted to prepare TiO2 films on various substrates. Materials exhibiting mechanical and chemical stability have been employed as TiO2 supports [18]. Various substrates, including carbonaceous materials [19,20], polymer supports [21], cellulosic paper [22], glass-fiber tissue [23,24], and polyester, have been investigated. However, the purification capacity of the effluent under visible light has not been proven to be significantly effective. Further enhancements are required to achieve more substantial levels of mineralization. Using an external UV lamp in conventional reactors diminishes catalyst efficiency due to the dissipation of UV radiation when traversing the gas layer. Employing an external lamp poses a challenge in reactor design, requiring substantial energy consumption and occupying a larger space. The development of a novel reactor utilizing optical fibers to support the photocatalyst has been introduced to address this issue. Placing the catalyst on the surface of the light source ameliorates their contact, overcoming material transfer limitations [25]. The specific treatment of the luminous textile triggers intense lighting thanks to the low energy consumption of LED technology. Nonetheless, a few studies on luminous textiles, concentrating on their characterization aspects, have been produced [17,26,27,28]. Our study focused on the photocatalytic oxidation of butyraldehyde (BUTY) and isovaleraldehyde (ISOVAL) using a groundbreaking catalytic planar reactor that integrated textile fibers (high-tech) and operated without an external light source. Factors such as the concentration of pollutants, air humidity, the amount of TiO2, airflow, and light intensity were investigated in terms of how they influenced the efficiency of the photocatalytic process. Additionally, this research delved into the impact of different pollutants on removal efficiency and the mineralization process. The experimental observations and theoretical modeling were integrated to comprehensively understand the process. This is the first investigation to enhance aldehyde treatment using TiO2-coated luminous textiles in a planar configuration for a pilot-scale setup.

2. Results

2.1. Degradation of Butyraldehyde

2.1.1. Effect of Inlet Concentration and Flow Rate on the Photocatalytic Process Efficiency

The BUTY photocatalytic oxidation was studied using a planar reactor to investigate the effect of inlet concentrations and the feed flow rates on the process performance. Figure 1 shows the BUTY removal efficiency of the planar reactor deposited with 7 g/m2 of TiO2. The initial concentrations ranged between 10 and 60 mg/m3, the feed flow rate varied between 2 and 6 m3/h, and the relative humidity was 50%.
At a constant flow rate, when the BUTY initial concentration increased, the degradation rate decreased owing to a decrease in the number of available active sites on the catalyst surface [29]. Moreover, an increase in the feed flow led to a decrease in the pollutant removal performance. The reason for such a result was that at high flow rates, the contact time of the pollutants with the active species, such as °OH and O2°, which formed on the catalyst surface, became shorter, leading to a significant decline in the efficiency of the pollutant degradation [30]. Previous studies reported similar results [31]. The highest degradation efficiency occurred at a feed flow rate of 2 m3/h, which was attributed to an ideal equilibrium between adequate contact time and efficient pollutant diffusion onto the catalyst surface.
As shown in Table 1, various photocatalysts have been developed and used for the photocatalytic degradation of air pollution over the past five years. Mamaghani and collaborators [32] evaluated the effect of the operating parameters (inlet VOC concentration, residence time, and relative humidity) on pollutant degradation efficiency using various types of commercially available TiO2. They concluded that regardless of the catalyst type, the photocatalytic efficiency always declined with an increase in the airflow rate and/or inlet concentration [32,33]. El Falleh et al. [34] studied the photocatalytic degradation of BUTY and ISOVAL using TiO2-impregnated polyester (PES) and glass-fiber tissue (GFT) in a 1 L batch reactor. They found that TiO2/GFT exhibited more significant photocatalytic activity than TiO2/PES. Moreover, they found that BUTY removal on GFT was about twice as high as PES. The enhanced BUTY removal was ascribable to the accessibility of TiO2 at lower depths on the GFT, leading to a higher number of active sites per TiO2 nanoparticle on GFT than on PES.
Saoud et al. [35] investigated BUTY photocatalytic removal using a supported TiO2–glass-fiber tissue (GFT) continuous planar reactor. They reached a removal capacity of 25 mg/m3 (with 12 g/m2 of TiO2 and a UV intensity of 20 W/m2) of about 4.615 mg/(gTiO2) h−1.W−1. The removal capacity of the same concentration through a luminous textile (with a TiO2 mass of 7 g/m2 and UV intensity of 0.5 w/m2) was about 123.15 mg/(gTiO2) h−1.W−1. The removal capacity using the luminous textile was 26 times higher than that of TiO2 on glass fiber, with high energy efficiency. The photocatalytic removal efficiency of luminous textile composites improved under UV LED irradiation compared with TiO2 on glass-fiber tissue. They reported that the chloroform removal capacity of 25 mg/m3 was enhanced more than four times when using luminous textiles (Table 1).
Table 1. Comparison of VOC removal efficiency of various photocatalytic reactors.
Table 1. Comparison of VOC removal efficiency of various photocatalytic reactors.
Catalyst VOC Operating Parameters Studied Parameters Optimal Results Ref.
TiO2-deposited polyester (PES) and TiO2-deposited glass-fiber tissue (GFT)ButyraldehydeBatch reactor, UV intensity = 9 W, RH = 40–50%, and T = 20 °CInlet concentration,
RH, Q, and
residence time
90% removal in 30 min on GFT. BUTY removal on GFT was about twice as fast as PES[34]
Commercialized TiO2 (P25, PC500, and UV100)Toluene;
methyl ethyl ketone
Continuous reactor, UV at 5 mw/cm2, and T = 20 °CInlet concentration, Q, RH, and UVBetween 72% and 100% degradation[32]
TiO2/diatomiteFormaldehyde, acetone, and p-xyleneContinuous reactor, RH = 20%, I = 5 w/m2.
Composite dosage = 3.76 mg/cm2, light intensity = 1.33 mW/cm2, RH = 0%, and Q = 1 l.mn−1
Pollutant inlet concentration, RH, and Q,>90% of degradation for a concentration of 10 ppm[36]
TiO2 coating on GFTAmmonia and butyraldehydeContinuous reactor, UV lamp = 9 W, T = 20 °C, Q = 1 m3/h, and
RH = 50%
Inlet concentration,
RH, and flow rate
High selectivity for photodegradation at a rate of 72%[35]
TiO2-incorporated KIT6Ethylene and
propylene
A mixture of UVA (320–400 nm) and UVB (290–320 nm), 100 ppm, and RH 60%Pollutant conversion90% conversion for 30% TiO2 KIT6[37]
Glass-fiber tissue (GFT)Butyraldehyde
and dimethyl disulfide (DMDS)
Continuous reactor, UV intensity = 9 W, T = 20 °C, and Q = 4 m3/hInlet concentration
and flow rate
Around 70% removal[38]

2.1.2. Effect of VOC Inlet Concentration and Relative Humidity on the Photocatalytic Process Efficiency

The moisture in the air is an important parameter that can directly affect the efficiency of gas-phase photocatalytic oxidation. Numerous research studies have investigated the impact of relative humidity (RH) on photocatalytic oxidation [39,40,41]. In this context, the effect of relative humidity (5, 50, and 90%) on the elimination of BUTY was studied at a fixed flow rate of 2 m3/h for various BUTY inlet concentrations. The results are reported in Figure 2; the presence of water molecules in the airflow improved BUTY oxidation for the different inlet concentrations studied. By varying the relative humidity from 5 to 50%, the BUTY removal efficiency attained 27.8% and reached an optimum of 52.3% at an inlet concentration of 10 mg/m3. This finding indicated that low moisture in the gas stream promoted the BUTY photocatalytic degradation, probably due to OH° hydroxyl radical formation on the TiO2 surface [41,42] produced by the adsorbed water molecules, which are known to be very reactive chemical species for pollutant mineralization [43]. Moreover, as the humidity of the effluent to be treated continued to increase, reaching an additional increment in RH up to 50%, the removal efficiency exhibited a notable decline. It was noticed that at 60 mg/m3 and with an RH of 90%, the degradation rate decreased by around 40% compared with 50% RH. This decline was due to the competitive adsorption of BUTY and water molecules on TiO2 adsorption sites. Moreover, on the TiO2 surface, water molecules adsorbed and formed a thin film by creating strong hydrogen bonds with OH groups. Additionally, at high values of RH%, the adsorbed water acted as a physical barrier, blocking effective contact between the pollutant and the catalyst surface.
The BUTY molecules should have traversed the water layer to adhere to the photocatalyst surface [44]. Additionally, the relative humidity and catalyst properties affected the quantity and nature of the byproducts resulting from the reaction pathway. As can be seen in Figure 2, the best photocatalytic activities were achieved with 50% RH regardless of the inlet pollutant concentration. Consistent results were observed in previous studies. Our previous study demonstrated the impact of relative humidity on the photocatalytic degradation of BUTY and ammonia using GFT/TiO2, highlighting that moisture contributes to an increased degradation rate, particularly at an average humidity of 50%. Thus, the mechanism of reaction can be explained as follows: Under light irradiation, TiO2 deposed on optical fibers generates electron–hole pairs (e/h+). The trapped electrons can reduce molecular oxygen (O2) to generate reactive oxygen species and, in the presence of humid air, the H2O can be reduced to °OH. These ROS are highly reactive and can effectively degrade fatty acids, resulting in enhanced indoor air quality.
The effect of relative humidity on the photocatalytic degradation of chloroform and glutaraldehyde over a luminous textile with a UV LED lamp has been studied, reporting that better chloroform degradation efficiency varied between 50 and 60% relative humidity. Zhang et al. [36] studied the photocatalytic degradation of acetone and p-xylene on TiO2/diatomite composites. They showed that the highest degradation rate could be acquired when the relative humidity reached 70%. Mamaghani et al. [32] determined that the maximum removal efficiency for toluene photodegradation using various types of commercial TiO2 occurred at approximately 20% relative humidity.

2.1.3. Effect of Pollutant Concentration and UV Radiation Intensity on the Photocatalytic Removal Efficiency

Numerous experiments were conducted using BUTY in a photocatalytic reactor involving varying input concentrations and UV radiation intensities to evaluate the impact of UV radiation on the performance of the planar reactor. Diverse light intensities were achieved through variations in the UV LED output, spanning values of 0.5, 1, and 2 W/m2. Figure 3 reports the BUTY removal efficiency for various concentrations under different UV radiation intensities. The results revealed that for a BUTY inlet concentration of 10 mg/m3, the removal efficiency was 28% when the light intensity was 0.5 W/m2. As the light intensity increased to 2 W/m2, the removal efficiency reached 65.7%. This increment occurred because raising the intensity generated more electrons and holes, produced more reactive species such as °OH and O2° on the TiO2 surface, and ultimately increased the BUTY degradation rate [36]. This non-linearity was due to the reduction in quantum efficiency at high intensities owing to the elevation of the recombination rate. When light intensities are too high, the recombination of electron–hole pairs become the dominant process, leading to a drop in the energy efficiency of the photoreaction. It is worth noting that, regardless of the UV radiation intensity, an increase in the inlet concentration of BUTY reduced the removal efficiency. Specifically, the efficiency of BUTY removal decreased by approximately 40% as the inlet concentration increased from 10 to 60 mg/m3.

2.1.4. Effect of Inlet Concentration and TiO2 Mass on the Photocatalytic Process

Several experiments were conducted, varying the amount of TiO2 from 7 to 25 g/m2 under different initial BUTY concentrations at a fixed flow rate of 2 m3/h, RH of 50%, and UV light intensity of 2 W/m2. The aim was to investigate the impact of TiO2 mass on BUTY degradation and reactor performance, and the result is reported in Figure 4. The trend of the removal efficiency vs. the catalyst mass remained consistent across different inlet BUTY concentrations. The experimental findings indicated that as the catalyst mass increased from 7 to 13 g/m2, the BUTY removal efficiency rose from 42% to 59% at an inlet concentration of 10 mg/m3. A subsequent 1.85-fold increase in the catalyst mass resulted in approximately a 1.5-fold enhancement in removal efficiency. The maximum removal efficiency was achieved at a catalyst mass of 13 g/m2, which was attributed to increased active sites on the catalyst surface [32,33]. The removal efficiency, however, dropped significantly when the catalyst mass increased to 25 g/m2. The degradation rate became 1.85-fold lower, reaching 32% when the catalyst mass was about 3.5 times higher. Similar trends were observed for the other inlet BUTY concentrations. Notably, the BUTY degradation rate decreased when the inlet concentration for the same catalyst mass was raised. This behavior could be attributed to the decrease in available catalyst active sites. Researchers have observed similar results for different reactor types and photocatalytic materials [33,36]. An increase in the catalyst mass can yield two counterproductive effects. Firstly, the mass transfer step becomes more significant, driven by enhanced site accessibility. Secondly, not all catalyst particles receive illumination owing to the aggregation of catalyst particles.

2.2. Degradation of a Mixture of Compounds

This section describes our aim to individually assess the removal efficiency of butyraldehyde (C4H8O) and isovaleraldehyde (C5H10O) and compare it with their degradation rates when both pollutants were mixed in equal proportions (50% butyraldehyde–50% isovaleraldehyde). The photocatalytic degradation of BUTY and ISOVAL and their mixture at different inlet pollutant concentrations (between 0.138 and 0.832 mmol/m3) were conducted at a flow rate of 2 m3/h, relative humidity of 50%, UV radiation intensity of 2 W/m2, and TiO2 mass of 25 g/m2. Figure 5 shows that the C4H8O and C5H10O photocatalytic removal efficiency decreased with the gas mixture (C4H8O and C5H10O). Moreover, for the different concentrations tested, the mineralization of BUTY and ISOVAL was affected when these two contaminants were mixed. When each contaminant was tested alone, the removal efficiency was 65.7% for BUTY and 58.7% for ISOVAL for an inlet pollutant concentration of 0.138 mmol/m3. The removal efficiency in the mixture was 53% for BUTY and 49% for ISOVAL because the mixed compounds seemed easier to eliminate than the pure compounds. Such a competition was traceable to the difference in volatility between the relevant ingredients. In our case, it was noticed that BUTY was the most volatile compound (Psat,vap = 12.2 KPa at 20 °C), which is why it was more vulnerable to the competition than ISOVAL (Psat,vap = 6.1 KPa at 20 °C). This result was very likely to be closely related to the competitive adsorption of these two compounds on the free active sites of the catalyst. Mixing the pollutants decreased the BUTY removal efficiency by 12% compared with a single VOC; however, mixed ISOVAL degradation declined by a factor of 1.2 compared with the single pollutant. Other works have reported similar behavior; for instance, Debono et al. (2017) noted an inhibitive effect when they investigated the photocatalytic degradation of the ternary mixture of toluene, decane, and trichloroethylene [45].

2.3. CO2 Selectivity

The CO2 selectivity of the photocatalytic removal of BUTY, ISOVAL, and their mixture was monitored and is reported in Figure 6. CO2 selectivity was negatively affected by increasing the initial concentration of the pollutants. The highest CO2 selectivity, 80.8%, was obtained for BUTY degradation at an inlet concentration of 0.14 mmol/m3. During the photocatalytic degradation of the contaminants, the carbonaceous intermediates could have accumulated on the catalyst surface, which could have blocked the active sites and prevented the access of substrate molecules. Even though the pollutant concentration in the reactor increased, the amount of photogenerated reactive species remained constant. Therefore, a small fraction of active species could be utilized to completely mineralize the adsorbed pollutants on the TiO2 surface.
The highest CO2 selectivity was obtained at a low input concentration because there was better accessibility to available active sites and, therefore, the intermediate byproducts could be mineralized entirely to CO2. When the inlet concentration increased, the ratio between the pollutant and the availability of catalyst active sites decreased, i.e., the byproducts produced could not be mineralized [46]. A competitive phenomenon, influenced by the representative concentrations of each entity and the affinity between the species and the catalytic support [47], manifested among BUTY, ISOVAL, and their respective byproducts. Figure 6 reveals that the CO2 selectivity of the mixture was lower than that obtained for each pollutant, likely due to the competition between the contaminants and their generated byproducts.

2.4. Kinetic Modeling of the Degradation with the Contribution of Mass Transfer in the Planar Reactor

Both the mass transfer and chemical reactions must be considered during the photocatalytic reaction over an immobilized catalyst. Indeed, both external and internal mass transfers could play significant roles. Several researchers [48,49] have developed the Langmuir–Hinshelwood expression for modeling photocatalytic reactions on catalyst surfaces. The L–H ratio equation is widely recognized to deliver good correlations for heterogeneous reactions [50]. This work stands out, given that the results obtained when studying the impact of the air gap on the removal efficiency proved that the mass transfer stage should be considered.
The pertinent photocatalyst exhibited minimal intraparticle transport resistance, attributed to the non-porous configuration of the catalytic thin film. Consequently, only the mass transfer between the catalyst and the bulk fluid could be considered. The limitation of the internal mass transfer step was not considered in the mass transfer model because photocatalysis implies superficial reaction sites, i.e., the internal diffusion, either molecular or Knudsen, is negligible. Many researchers [51,52] have studied photocatalysis kinetics and mass transfer considering Equations (1) and (2).
The gaseous phase is as follows:
u d C m d z + k m × a v ( C m C s ) = 0
The solid phase is as follows:
k m a v c m C s = k o b s × k m × C s   1 + k m × C s
where u and kobs are gas flow velocities (m/s) and the observed kinetic constant, respectively; Cs and Cm are the bulk and the medium surface gas-phase concentrations of VOCs (mmol/m3), respectively; km is the mass transfer coefficient (m/s); and av is the medium area per unit volume of the reactor (m2/m3reactor). The BUTY and ISOVAL concentrations at the inlet were significantly low. Thus, their byproducts would also be minimal, and their competitive effect would be overlooked.
The viscosity and density of the fluid used to calculate the mass transfer were assumed to be like air. By altering the flow rate from 2 to 6 m3/h, the Reynolds number soared from 248.52 to 745.57, respectively, implying that the mass transfer step could play a role in the degradation process due to laminar flow (Re < 2000) [53]. Subsequently, it was suggested that the mass transfer stage be considered. The external coefficient of mass transfer, km, is a function of the gas phase, the fluid flow regime, and the nature of the gas. It can be set using a semi-empirical correlation (Equation (3)) [54].
S h = 0.664 R e 1 2 × S C 1 3
where Sh, Sc, and Re are the dimensionless numbers of Sherwood, Schmidt, and Reynolds, respectively.
The pollutant’s molecular diffusivity (Dm) of the pollutant is gauged using the Fuller, Schettler, and Giddings correlation (Equation (4)) [53].
D m = 10 3 T 1.75   P ( ( V A ) 1 3   + ( V B ) 1 3   ) 2 × ( 1 M A + 1 M B )   1 / 2
where P is the pressure of the gas stream (Pa), T is the absolute temperature (K), (Ʃν)A and (Ʃν )B are the molecular volumes of gases A and B, and MA and MB are the molecular weights of A and B (g/mol). The index “A” represents the air stream and “B” is the pollutant. This expression (4) permitted the evaluation of the diffusivity coefficients for both pollutants (Table 2).
From Equation (5), the Sherwood number, the diffusion coefficient Dm, and the mass transfer coefficient (km) can be determined.
S h = k m × d D m
Here, d is the equivalent diameter. Table 3 gives the Reynolds number and mass transfer coefficients in the planar reactor.
Maple software (2020.2) was used to develop a system using the Method series to solve Equations (1) and (2). Previous works showed that a second-order development is enough to correlate the experimental results [50]. The second-order development is as follows [55]:
C i n C o u t = L k m a v 2 u C i n + 1 K + k o b s k m × a v C i n + 1 K + k o b s k m × a v 2 + 4 C i n K 0.5
In Equation (6), the unknown parameters are K and kobs. Both were determined by a digital resolution using the Solver function in Excel. A target cell was made for each experimental point as the difference between the theoretical and empirical outlet concentrations. The sum of the squares of the target cell mean difference for each compound was minimized by optimizing the variable cell values of K and kobs, helping to find the best figures for both butyraldehyde and isovaleraldehyde. Table 4 summarizes the K and kobs values.
It is noteworthy that these constants were not flow-rate-dependent. The influence of the flow-rate regime was completely incorporated through the parameter km, and the mass transfer and the chemical reaction step were separated. The experimental results gave a more reasonable description than that delivered by this model for both VOCs. For instance, the theoretical removal efficiency (Figure 7a,b) correlated well with the experimental results for all the flow rates investigated.
Figure 7 illustrates the variation in the calculated removal efficiency for both pollutants concerning the removal efficiency achieved for different BUTY and ISOVAL inlet concentrations across various flow rates. The lower removal efficiency observed at high flow rates was attributed to the inadequate residence time for the pollutants to effectively transfer from the gaseous phase onto the catalytic surface. Notably, the removal efficiency declined when the pollutant inlet concentration rose at a constant airflow. This figure proves that this model was well suited to the experimental results. Thus, it was possible to estimate the removal efficiency at different inlet concentrations and distinct flow rates for both VOCs.

3. Materials and Methods

3.1. Reagents

In this study, two compounds, butyraldehyde (C4H8O) and isovaleraldehyde (C5H10O), were purchased from SIGMA-ALDRICH, Germany, and used without purification. TiO2 was supported on luminous textiles, i.e., a combination of textile fibers and optical fibers was used as a photocatalyst.

3.2. Luminous Textiles

The luminous textile used a combination of textile fibers made from polyester (Triwra CSTM fibers) and a luminous textile (optical fibers) made from polymethylmethacrylate resin (produced by PMMA CK-20 EskaTMfibers), interlaced and simultaneously woven according to a Jacquard loom with an average diameter of 480 µm, and covered with 10 µm thick fluoropolymer [6], which was supplied by Brochier Technologies (UVtex®). The luminous textile was collected to one extremity in the aluminum connector for one textile sample. The tissue was covered by an intermediate layer of silica, shielding the luminous textile degradation by photocatalysis. TiO2 Degussa P25 was used as a photocatalyst. The textile fiber was soaked for 60 min in a TiO2 suspension (50 g/L TiO2 powder) and then dried for 2 h at 70 °C. Eventually, 12 g/m2 of TiO2 was uniformly dispersed only on the textile fiber. These luminous textiles underwent treatment to enable the transmission and emission of light laterally from their external surface. The optical fiber and textile fiber consisted of polymethyl methacrylate, made by Mitsubishi (Tokyo, Japan) (PMMA CK-20 Eska™ fiber), and polyester (Trévira CS™ fiber), respectively. Figure 8a illustrates a scanning electron microscopy image characterizing the photocatalyst. Figure 8b highlights the luminous textiles placed in the planar reactor with the UV LEDs. Each UV LED illuminated the luminous textiles.

3.3. Experimental Setup

Figure 9 portrays the schematic of the experimental setup, which consisted of the following three sections: (i) gas stream preparation, (ii) photocatalytic reactor, and (iii) analytical systems. The VOC removal process operated using a single-pass tangential flow configuration. Before beginning the photocatalytic process, the photoreactor (the glass chamber + the photocatalyst) was air-cleaned under UV light for 60 min to eliminate contamination. The VOCs were continuously injected into the gas stream using syringe pumps (Kd Scientific Model 100, Holliston, MA, USA) and thoroughly mixed before entering the reactor. The heating tape was wrapped around the pipe from the injection port to the reactor inlet to guarantee adequate evaporation of the pollutant. The airflow rate was controlled using a calibrated flow meter (Bronkhorst, Montigny-lès-Cormeilles, France), and the airflow varied between 2 and 6 m3/h.
The experimental reactor was a continuous-operation rectangular chamber. It had the flexibility to be supplied with either an airflow containing a single pure compound (BUTY or ISOVAL) or an equimolar binary mixture of contaminants. An analytical system facilitated the measurement of both the initial and final concentrations of the pollutants. Regular gas samples were extracted throughout the experiment and promptly subjected to gas chromatography (GC) analyses. The gas stream was directed towards a CO2 analyzer at the installation outlet, enabling continuous monitoring of the mineralization rate. The reactor, depicted in Figure 8, comprised a rectangular glass chamber (length = 1000 mm; square section = 145 mm × 145 mm). Two parallel glass plates, each with a length of 800 mm and a thickness of 1 mm, were installed within the chamber. Luminous textiles, equipped with an ultraviolet LED, showed even radiation distribution across the catalyst’s surface. All experiments were conducted under ambient conditions, maintaining atmospheric pressure and room temperature. A column humidifier was employed to achieve relative humidity (RH) values of 5%, 50%, and 90%.

3.4. Apparatus and Analysis

The concentrations of BUTY and ISOVAL were studied using gas chromatography (GC) equipped with a flame ionization detector (FID). The analysis was performed with a chromatograph (Fisons GC9000) fitted with a 25 m Chrompact FFAP-CB capillary column. Next, 500 μL of the sample was manually injected at least three times using a gas-tight syringe. The analysis conditions were as follows: injector and detector temperatures were fixed at 200 and 190 °C, respectively. The temporal evolution of CO2 during the photocatalytic treatment was monitored using an infrared (IR) multi-gas CO2 analyzer (MIR 9000 Environment SA Gas Filters Correlation).
The photocatalytic process efficiency was measured based on the VOC removal efficiency (RE) and CO2 selectivity (SCO2) using Equations (7) and (8).
R E % = C i n l e t C o u t l e t C i n l e t × 100
where [C]inlet and [C]outlet are the pollutant inlet and outlet concentrations, respectively.
S C O 2 ( % ) = [ C O 2 ] o u t l e t [ C O 2 ] i n l e t n c , v o c x R E % × [ V O C ] i n l e t × 10 4
where [VOC]inlet and nc,voc are the VOC inlet concentration and the number of carbon atoms in the pollutant, respectively.

4. Conclusions

The degradation of butyraldehyde and isovaleraldehyde through photocatalysis was investigated in a TiO2-supported luminous textile planar reactor. The degradation rates of BUTY and ISOVAL decreased with high pollutant inlet concentrations and feed flow rates, while an increase in relative humidity initially enhanced pollutant photodegradation. The presence of water vapor proved beneficial at low concentrations due to the formation of OH° radicals. However, at high relative humidity (90%), degradation diminished due to competition between the water and pollutants for active sites.
The impact of the inlet concentration and UV intensity on the removal efficiency was explored, revealing that regardless of the UV light intensity, an increase in inlet concentration led to a 40% reduction in the BUTY degradation rate, particularly from 10 to 60 mg/m3. The influence of the catalyst quantity was also studied for various BUTY inlet concentrations, and high removal efficiency was achieved under the following specific conditions: a flow rate of 2 m3/h, a concentration of 10 mg.m−3, relative humidity of 50%, and a TiO2 amount of 2 g/m2 on the luminous textile. Notably, the BUTY removal capacity using the luminous textile surpassed that achieved for the same TiO2 concentration on fiberglass more than 26 times, with reduced energy consumption. Additionally, photocatalysis of the BUTY and ISOVAL mixture and the effect of the inlet concentration on removal efficiency and overall CO2 selectivity were studied. A mass transfer model was developed using the Langmuir–Hinshelwood approach to determine kinetic adsorption and degradation constants. The external mass transfer was estimated using semi-empirical correlations, yielding results consistent with the experimental findings.

Author Contributions

Conceptualization, A.A.A. and A.A.; validation, L.K., A.A.A. and M.G.; formal analysis M.A. and H.T.; investigation, M.A.; writing—original draft preparation, O.B., L.K., M.A., H.T. and M.G.; writing—review and editing, O.B., L.M., H.T., J.Z., W.E., M.G., A.A.A. and A.A.; visualization, M.G., A.A.A. and A.A.; supervision, M.G., and A.A.A.; project administration, M.G., A.A.A., and A.A.; funding acquisition, A.A.A. and M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2502).

Data Availability Statement

The data presented in this study are available within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Edo, G.I.; Itoje-akpokiniovo, L.O.; Obasohan, P.; Ikpekoro, V.O.; Samuel, P.O.; Jikah, A.N.; Nosu, L.C.; Ekokotu, H.A.; Ugbune, U.; Oghroro, E.E.A.; et al. Impact of environmental pollution from human activities on water, air quality and climate change. Ecol. Front. 2024, 44, 874–889. [Google Scholar] [CrossRef]
  2. Lee, S.; Hung, R.; Bennett, J.W. An Overview of Fungal Volatile Organic Compounds (VOCs); Springer: Berlin/Heidelberg, Germany, 2024. [Google Scholar]
  3. Lavaine, E.; Majerus, P.; Treich, N. Health, air pollution, and animal agriculture. Rev. Agric. Food Environ. Stud. 2020, 101, 517–528. [Google Scholar] [CrossRef]
  4. Fetisov, V.; Gonopolsky, A.M.; Davardoost, H.; Ghanbari, A.R.; Mohammadi, A.H. Regulation and impact of VOC and CO2 emissions on low-carbon energy systems resilient to climate change: A case study on an environmental issue in the oil and gas industry. Energy Sci. Eng. 2023, 11, 1516–1535. [Google Scholar] [CrossRef]
  5. Kiani, S.S.; Faiz, Y.; Farooq, A.; Ahmad, M.; Irfan, N.; Nawaz, M.; Bibi, S. Synthesis and adsorption behavior of activated carbon impregnated with ASZM-TEDA for purification of contaminated air. Diam. Relat. Mater. 2020, 108, 107916. [Google Scholar] [CrossRef]
  6. Masi, M.; Nissim, W.G.; Pandolfi, C.; Azzarello, E.; Mancuso, S. Modelling botanical biofiltration of indoor air streams contaminated by volatile organic compounds. J. Hazard. Mater. 2022, 422, 126875. [Google Scholar] [CrossRef] [PubMed]
  7. Wu, X.; Lin, Y.; Wang, Y.; Wu, S.; Yang, C. Volatile organic compound removal via biofiltration: Influences, challenges, and strategies. Chem. Eng. J. 2023, 471, 144420. [Google Scholar] [CrossRef]
  8. Zhang, G.; Liu, J.; Tan, Z.; Yu, H. Multivariate modeling of intrinsic kinetics for gas-solid heterogeneous photocatalytic reaction: A general method for different pollutant-photocatalyst systems. Chem. Eng. J. 2024, 479, 147651. [Google Scholar] [CrossRef]
  9. Tomić, J.; Malinović, N. Titanium dioxide photocatalyst: Present situation and future approaches. AIDASCO Rev. 2023, 1, 26–30. [Google Scholar] [CrossRef]
  10. Armaković, S.J.; Savanović, M.M.; Armaković, S. Spray-Deposited TiO2 Layers on Aluminum Foil for Sustainable Water Remediation. Crystals 2024, 14, 875. [Google Scholar] [CrossRef]
  11. Ghosh, R.; Sahu, R.P.; Ganguly, R.; Zhitomirsky, I.; Puri, I.K. Photocatalytic activity of electrophoretically deposited TiO2 and ZnO nanoparticles on fog harvesting meshes. Ceram. Int. 2020, 46, 3777–3785. [Google Scholar] [CrossRef]
  12. Jaison, A.; Mohan, A.; Lee, Y.C. Recent Developments in Photocatalytic Nanotechnology for Purifying Air Polluted with Volatile Organic Compounds: Effect of Operating Parameters and Catalyst Deactivation. Catalysts 2023, 13, 407. [Google Scholar] [CrossRef]
  13. Chen, Z.; Pan, M.; Cheng, C.; Luo, J.; Deng, X. Water disinfection: Advances in photocatalysis and piezo/triboelectric catalysis with progressively enhanced energy utilization. SusMat 2024, 4, e232. [Google Scholar] [CrossRef]
  14. Dhiflaoui, H.; Hajjaji, M.A.; Hajjaji, A.; Khezami, L.; Karrech, A.; Bessais, B.; Larbi, A.B.C.; Amlouk, M. Enhanced Interfacial Adhesion of TiO2 Nanotubes Decorated With Ag Silver Nanoparticles Prepared by Photo-Reduction Process. J. Tribol. 2023, 145, 091106. [Google Scholar] [CrossRef]
  15. Wang, Y.; Yang, C.; Chen, A.; Pu, W.; Gong, J. Influence of yolk-shell Au@TiO2 structure induced photocatalytic activity towards gaseous pollutant degradation under visible light. Appl. Catal. B Environ. 2019, 251, 57–65. [Google Scholar] [CrossRef]
  16. Xu, T.; Zheng, H.; Zhang, P. Isolated Pt single atomic sites anchored on nanoporous TiO2 film for highly efficient photocatalytic degradation of low concentration toluene. J. Hazard. Mater. 2020, 388, 121746. [Google Scholar] [CrossRef]
  17. Udayabhanu; Lakshmana Reddy, N.; Shankar, M.V.; Sharma, S.C.; Nagaraju, G. One-pot synthesis of Cu–TiO2/CuO nanocomposite: Application to photocatalysis for enhanced H2 production, dye degradation & detoxification of Cr (VI). Int. J. Hydrogen Energy 2020, 45, 7813–7828. [Google Scholar] [CrossRef]
  18. Sun, Y.; Feng, B.; Li, Q.; Tian, C.; Ma, L.; Li, Z. The Application of Bi-Doped TiO2 for the Photocatalytic Oxidation of Formaldehyde. Cryst. Res. Technol. 2022, 57, 2100231. [Google Scholar] [CrossRef]
  19. Noureen, L.; Wang, Q.; Humayun, M.; Shah, W.A.; Xu, Q.; Wang, X. Recent advances in structural engineering of photocatalysts for environmental remediation. Environ. Res. 2023, 219, 115084. [Google Scholar] [CrossRef]
  20. Azami, M.S.; Jalil, A.A.; Hassan, N.S.; Hussain, I.; Fauzi, A.A.; Aziz, M.A.A. Green carbonaceous material—fibrous silica-titania composite photocatalysts for enhanced degradation of toxic 2-chlorophenol. J. Hazard. Mater. 2021, 414, 125524. [Google Scholar] [CrossRef]
  21. Ramasundaram, S.; Balasankar, A.; Arokiyaraj, S.; Sumathi, P.; Hwan Oh, T. Multi-usable titanium dioxide and poly(vinylidene fluoride) composite foam photocatalyst for degradation of organic pollutants. Appl. Surf. Sci. 2023, 609, 155264. [Google Scholar] [CrossRef]
  22. Kwon, M.; Kim, J.; Kim, J. Photocatalytic activity and filtration performance of hybrid tio2-cellulose acetate nanofibers for air filter applications. Polymers 2021, 13, 1331. [Google Scholar] [CrossRef] [PubMed]
  23. Amor, N.; Tayyab Noman, M.; Petru, M.; Sebastian, N.; Balram, D. Machining performance of TiO2 embedded-glass fiber reinforced composites with snake optimizer. Meas. J. Int. Meas. Confed. 2024, 227, 114253. [Google Scholar] [CrossRef]
  24. Ma, X.; Wang, H.; Chen, Y.; Fu, L.; Zhou, J.; Zhang, L.; Xing, Z.; Zhang, Q.; Xia, L. Application of Ag@g-C3N4/TiO2 cotton fabric flexible substrate with dual functionality: Photocatalytic reusability and SERS signal amplification for food safety detection. Appl. Surf. Sci. 2024, 661, 160068. [Google Scholar] [CrossRef]
  25. Yang, J.; Zeng, X.; Tebyetekerwa, M.; Wang, Z.; Bie, C.; Sun, X.; Marriam, I.; Zhang, X. Engineering 2D Photocatalysts for Solar Hydrogen Peroxide Production. Adv. Energy Mater. 2024, 14, 2400740. [Google Scholar] [CrossRef]
  26. O’Neal Tugaoen, H.; Garcia-Segura, S.; Hristovski, K.; Westerhoff, P. Compact light-emitting diode optical fiber immobilized TiO2 reactor for photocatalytic water treatment. Sci. Total Environ. 2018, 613–614, 1331–1338. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, C.J.; Wu, C.C.; Rahman, K.H.; Chen, K.C. A study on photodegradation of trichloroethylene using an optical fiber coated with different photocatalysts. Mater. Sci. Semicond. Process 2023, 163, 107538. [Google Scholar] [CrossRef]
  28. Behineh, E.S.; Solaimany Nazar, A.R.; Farhadian, M.; Karimi-Alavijeh, H.R. Embedded ZnO nanorod array/TiO2/GO coated optical fiber in a photocatalytic microreactor for Cefixime degradation: Diffused or focused light sources effect. J. Ind. Eng. Chem. 2024, 130, 623–637. [Google Scholar] [CrossRef]
  29. Shojaei, A.; Ghafourian, H.; Yadegarian, L.; Lari, K.; Sadatipour, M.T. Removal of volatile organic compounds (VOCs) from waste air stream using ozone assisted zinc oxide (ZnO) nanoparticles coated on zeolite. J. Environ. Heal. Sci. Eng. 2021, 19, 771–780. [Google Scholar] [CrossRef]
  30. Xu, P.; Ding, C.; Li, Z.; Yu, R.; Cui, H.; Gao, S. Photocatalytic degradation of air pollutant by modified nano titanium oxide (TiO2)in a fluidized bed photoreactor: Optimizing and kinetic modeling. Chemosphere 2023, 319, 137995. [Google Scholar] [CrossRef]
  31. Verma, S.; Vikrant, K.; Kim, K.H. Optimization of process variables for the concurrent removal of aliphatic and aromatic volatile organic compounds over a copper impregnated titanium dioxide photocatalyst. Environ. Sci. Nano 2023, 10, 2035–2052. [Google Scholar] [CrossRef]
  32. Mamaghani, A.H.; Haghighat, F.; Lee, C.S. Photocatalytic oxidation technology for indoor environment air purification: The state-of-the-art. Appl. Catal. B Environ. 2017, 203, 247–269. [Google Scholar] [CrossRef]
  33. Mamaghani, A.H.; Haghighat, F.; Lee, C.S. Photocatalytic degradation of VOCs on various commercial titanium dioxides: Impact of operating parameters on removal efficiency and by-products generation. Build. Environ. 2018, 138, 275–282. [Google Scholar] [CrossRef]
  34. Elfalleh, W.; Assadi, A.A.; Bouzaza, A.; Wolbert, D.; Kiwi, J.; Rtimi, S. Innovative and stable TiO2 supported catalytic surfaces removing aldehydes under UV-light irradiation. J. Photochem. Photobiol. A Chem. 2017, 343, 96–102. [Google Scholar] [CrossRef]
  35. Abou Saoud, W.; Assadi, A.A.; Guiza, M.; Bouzaza, A.; Aboussaoud, W.; Soutrel, I.; Ouederni, A.; Wolbert, D.; Rtimi, S. Abatement of ammonia and butyraldehyde under non-thermal plasma and photocatalysis: Oxidation processes for the removal of mixture pollutants at pilot scale. Chem. Eng. J. 2018, 344, 165–172. [Google Scholar] [CrossRef]
  36. Zhang, G.; Liu, Y.; Hashisho, Z.; Sun, Z.; Zheng, S.; Zhong, L. Adsorption and photocatalytic degradation performances of TiO2/diatomite composite for volatile organic compounds: Effects of key parameters. Appl. Surf. Sci. 2020, 525, 2400740. [Google Scholar] [CrossRef]
  37. Hussain, M.; Akhter, P.; Iqbal, J.; Ali, Z.; Yang, W.; Shehzad, N.; Majeed, K.; Sheikh, R.; Amjad, U.E.S.; Russo, N. VOCs photocatalytic abatement using nanostructured titania-silica catalysts. J. Environ. Chem. Eng. 2017, 5, 3100–3107. [Google Scholar] [CrossRef]
  38. Abou Saoud, W.; Assadi, A.A.; Guiza, M.; Bouzaza, A.; Aboussaoud, W.; Ouederni, A.; Soutrel, I.; Wolbert, D.; Rtimi, S. Study of synergetic effect, catalytic poisoning and regeneration using dielectric barrier discharge and photocatalysis in a continuous reactor: Abatement of pollutants in air mixture system. Appl. Catal. B Environ. 2017, 213, 53–61. [Google Scholar] [CrossRef]
  39. Payan, A.; Charchi Aghdam, N.; Soltan, J. Novel insight into the effect of relative humidity on the acetone photocatalytic decomposition using Ag−CeO2 under vacuum ultraviolet illumination: Catalytic synergies, performance, and mechanistic interpretation. J. Photochem. Photobiol. A Chem. 2023, 443, 114847. [Google Scholar] [CrossRef]
  40. Jin, H.; Lee, T.M.; Choi, H.; Kim, K.S. Effects of process variables for NO conversion by double-layered photocatalytic mortar with TiO2 nanoparticles. J. Ind. Eng. Chem. 2023, 117, 461–472. [Google Scholar] [CrossRef]
  41. Zhong, L.; Brancho, J.J.; Batterman, S.; Bartlett, B.M.; Godwin, C. Experimental and modeling study of visible light responsive photocatalytic oxidation (PCO) materials for toluene degradation. Appl. Catal. B Environ. 2017, 216, 122–132. [Google Scholar] [CrossRef]
  42. Masresha, G.; Jabasingh, S.A.; Kebede, S.; Doo-Arhin, D.; Assefa, M. A review of prospects and challenges of photocatalytic decomposition of volatile organic compounds (VOCs) under humid environment. Can. J. Chem. Eng. 2023, 101, 6905–6918. [Google Scholar] [CrossRef]
  43. Tu, L.N.Q.; Nhan, N.V.H.; Van Dung, N.; An, N.T.; Long, N.Q. Enhanced photocatalytic performance and moisture tolerance of nano-sized Me/TiO2–zeolite Y (Me=Au, Pd) for gaseous toluene removal: Activity and mechanistic investigation. J. Nanoparticle Res. 2019, 21, 194. [Google Scholar] [CrossRef]
  44. Zhang, J.; Vikrant, K.; Kim, K.-H.; Dong, F.; Won Chung, M.; Weon, S. Unveiling the Collective Effects of Moisture and Oxygen on the Photocatalytic Degradation of m-Xylene using a Titanium Dioxide Supported Platinum Catalyst. Chem. Eng. J. 2022, 439, 135747. [Google Scholar] [CrossRef]
  45. Debono, O.; Hequet, V.; Le Coq, L.; Locoge, N.; Thevenet, F. VOC ternary mixture effect on ppb level photocatalytic oxidation: Removal kinetic, reaction intermediates and mineralization. Appl. Catal. B Environ. 2017, 218, 359–369. [Google Scholar] [CrossRef]
  46. Sun, X.; Li, C.; Yu, B.; Wang, J.; Wang, W. Removal of gaseous volatile organic compounds via vacuum ultraviolet photodegradation: Review and prospect. J. Environ. Sci. 2023, 125, 427–442. [Google Scholar] [CrossRef] [PubMed]
  47. Munnik, P.; De Jongh, P.E.; De Jong, K.P. Recent Developments in the Synthesis of Supported Catalysts. Chem. Rev. 2015, 115, 6687–6718. [Google Scholar] [CrossRef]
  48. Chong, M.C.; Narindri Rara Winayu, B.; Chu, H. Advancement of visible light-driven photocatalytic degradation of dimethyl sulfide by Zn doped rGO/TiO2. Appl. Surf. Sci. 2024, 648, 159048. [Google Scholar] [CrossRef]
  49. Malayeri, M.; Lee, C.S.; Niu, J.; Zhu, J.; Haghighat, F. Kinetic modeling and reaction mechanism of toluene and by-products in photocatalytic oxidation reactor. Chem. Eng. J. 2022, 427, 131536. [Google Scholar] [CrossRef]
  50. Seo, B.; Ko, E.H.; Kim, B.; Park, N.K.; Kang, S.B.; Kang, D.; Kim, M. Computational screening-based development in VOC removal catalyst: Methyl ethyl ketone oxidation over Pt/TiO2. Chem. Eng. J. 2023, 452, 139466. [Google Scholar] [CrossRef]
  51. Malayeri, M.; Lee, C.S.; Haghighat, F.; Klimes, L. Modeling of gas-phase heterogeneous photocatalytic oxidation reactor in the presence of mass transfer limitation and axial dispersion. Chem. Eng. J. 2020, 386, 124013. [Google Scholar] [CrossRef]
  52. Ješić, D.; Lašič Jurković, D.; Pohar, A.; Suhadolnik, L.; Likozar, B. Engineering photocatalytic and photoelectrocatalytic CO2 reduction reactions: Mechanisms, intrinsic kinetics, mass transfer resistances, reactors and multi-scale modelling simulations. Chem. Eng. J. 2021, 407, 126799. [Google Scholar] [CrossRef]
  53. Thevenet, F.; Sivachandiran, L.; Guaitella, O.; Barakat, C.; Rousseau, A. Plasma-catalyst coupling for volatile organic compound removal and indoor air treatment: A review. J. Phys. D. Appl. Phys. 2014, 47, 224011. [Google Scholar] [CrossRef]
  54. Reisener, J.; Reuter, M.A.; Krüger, J. Modelling of the mass transfer in gas-sparged electrolysers with neural nets. Chem. Eng. Sci. 1993, 48, 1089–1101. [Google Scholar] [CrossRef]
  55. Romkes, S.J.P.; Dautzenberg, F.M.; van den Bleek, C.M.; Calis, H.P.A. CFD modelling and experimental validation of particle-to-fluid mass and heat transfer in a packed bed at very low channel to particle diameter ratio. Chem. Eng. J. 2003, 96, 3–13. [Google Scholar] [CrossRef]
Figure 1. Effect of feed flow rate and BUTY initial concentration on removal efficiency under similar operating conditions. Experimental conditions: HR = 50%, UV intensity = 0.5 w/m2, TiO2 loading = 7 g/m2, and T = 20 °C.
Figure 1. Effect of feed flow rate and BUTY initial concentration on removal efficiency under similar operating conditions. Experimental conditions: HR = 50%, UV intensity = 0.5 w/m2, TiO2 loading = 7 g/m2, and T = 20 °C.
Catalysts 15 00262 g001
Figure 2. BUTY removal efficiency as a function of relative humidity. At a fixed airflow rate, the BUTY concentration varied between 10 and 60 mg/L. Q = 2 m3/h, UV intensity = 0.5 W/m2, mass of TiO2 = 25 g/m2, and T = 20 °C.
Figure 2. BUTY removal efficiency as a function of relative humidity. At a fixed airflow rate, the BUTY concentration varied between 10 and 60 mg/L. Q = 2 m3/h, UV intensity = 0.5 W/m2, mass of TiO2 = 25 g/m2, and T = 20 °C.
Catalysts 15 00262 g002
Figure 3. The effect of BUTY concentration and UV radiation intensity on removal efficiency. Q = 2 m3/h, RH = 50%, mass TiO2 = 25 g/m2, and T = 20 °C.
Figure 3. The effect of BUTY concentration and UV radiation intensity on removal efficiency. Q = 2 m3/h, RH = 50%, mass TiO2 = 25 g/m2, and T = 20 °C.
Catalysts 15 00262 g003
Figure 4. Effect of amount of catalyst loading and BUTY inlet concentration on removal efficiency. Q = 2 m3 h−1, HR = 50%, UV intensity = 2 W/m2, and T = 20 °C.
Figure 4. Effect of amount of catalyst loading and BUTY inlet concentration on removal efficiency. Q = 2 m3 h−1, HR = 50%, UV intensity = 2 W/m2, and T = 20 °C.
Catalysts 15 00262 g004
Figure 5. Development of the removal efficiency of butyraldehyde in a mixture with isovaleraldehyde. Q = 2 m3 h−1, HR = 50%, UV intensity= 2 W/m2, and mass TiO2 = 25 g/m2.
Figure 5. Development of the removal efficiency of butyraldehyde in a mixture with isovaleraldehyde. Q = 2 m3 h−1, HR = 50%, UV intensity= 2 W/m2, and mass TiO2 = 25 g/m2.
Catalysts 15 00262 g005
Figure 6. Effect of pollutant concentration on CO2 selectivity of the photocatalytic removal of BUTY, ISOVAL, and their mixture. Q = 2 m3/h, RH = 50%, UV intensity = 2 W/m2, mass TiO2 = 25 g/m2, and T = 20 °C.
Figure 6. Effect of pollutant concentration on CO2 selectivity of the photocatalytic removal of BUTY, ISOVAL, and their mixture. Q = 2 m3/h, RH = 50%, UV intensity = 2 W/m2, mass TiO2 = 25 g/m2, and T = 20 °C.
Catalysts 15 00262 g006
Figure 7. Theoretical removal efficiency of butyraldehyde (a) and isovaleraldehyde (b) as a function of experimental removal efficiency at different flow rates. RH = 50%, UV intensity = 0.5 W.m−2, mass TiO2 = 7 g/m2, and T = 20 °C.
Figure 7. Theoretical removal efficiency of butyraldehyde (a) and isovaleraldehyde (b) as a function of experimental removal efficiency at different flow rates. RH = 50%, UV intensity = 0.5 W.m−2, mass TiO2 = 7 g/m2, and T = 20 °C.
Catalysts 15 00262 g007
Figure 8. (a) SEM image of the textile and (b) the catalyst in the reactor.
Figure 8. (a) SEM image of the textile and (b) the catalyst in the reactor.
Catalysts 15 00262 g008
Figure 9. Schematic illustration of the experimental pilot setup.
Figure 9. Schematic illustration of the experimental pilot setup.
Catalysts 15 00262 g009
Table 2. Diffusivity values of BUTY and ISOVAL in air.
Table 2. Diffusivity values of BUTY and ISOVAL in air.
PollutantFormulaDiffusivity in the Air (Dm) (cm2/s)
ButyraldehydeC4H8O0.0892
IsovaleraldehydeC5H10O0.0797
Table 3. Reynolds number and mass transfer coefficients.
Table 3. Reynolds number and mass transfer coefficients.
Flow Rate (Q) (m3/h)Reynolds NumberMass Transfer Coefficient (km) (m/s)
ButyraldehydeIsovaleraldehyde
22487.77.2
449710.910.1
674513.412.4
Table 4. kobs and K values for butyraldehyde and isovaleraldehyde.
Table 4. kobs and K values for butyraldehyde and isovaleraldehyde.
ButyraldehydeIsovaleraldehyde
kobs (mmol/(m 3.s))0.0180.081
K (m3/mmol)1.1590.246
R2 (%)98.692
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abdelkader, M.; Assadi, A.A.; Guiza, M.; Elfalleh, W.; Khezami, L.; Tahraoui, H.; Baaloudj, O.; Mouni, L.; Zhang, J.; Amrane, A. Photocatalytic Degradation of Pollutants in Air Streams Using Luminous Textiles Under Ultraviolet Light Illumination: A Pilot-Scale Remediation Study. Catalysts 2025, 15, 262. https://doi.org/10.3390/catal15030262

AMA Style

Abdelkader M, Assadi AA, Guiza M, Elfalleh W, Khezami L, Tahraoui H, Baaloudj O, Mouni L, Zhang J, Amrane A. Photocatalytic Degradation of Pollutants in Air Streams Using Luminous Textiles Under Ultraviolet Light Illumination: A Pilot-Scale Remediation Study. Catalysts. 2025; 15(3):262. https://doi.org/10.3390/catal15030262

Chicago/Turabian Style

Abdelkader, Meriem, Amine Aymen Assadi, Monia Guiza, Walid Elfalleh, Lotfi Khezami, Hichem Tahraoui, Oussama Baaloudj, Lotfi Mouni, Jie Zhang, and Abdeltif Amrane. 2025. "Photocatalytic Degradation of Pollutants in Air Streams Using Luminous Textiles Under Ultraviolet Light Illumination: A Pilot-Scale Remediation Study" Catalysts 15, no. 3: 262. https://doi.org/10.3390/catal15030262

APA Style

Abdelkader, M., Assadi, A. A., Guiza, M., Elfalleh, W., Khezami, L., Tahraoui, H., Baaloudj, O., Mouni, L., Zhang, J., & Amrane, A. (2025). Photocatalytic Degradation of Pollutants in Air Streams Using Luminous Textiles Under Ultraviolet Light Illumination: A Pilot-Scale Remediation Study. Catalysts, 15(3), 262. https://doi.org/10.3390/catal15030262

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop