Next Article in Journal
Machine Learning and Artificial Intelligence Techniques in Smart Grids Stability Analysis: A Review
Previous Article in Journal
Experimental Work to Investigate the Effect of Rooftop PV Panel Shading on Building Thermal Performance
Previous Article in Special Issue
Study on Swirling Flow and Spray Characteristics of Central Stage Direct Injection Combustor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regenerative Oxidation Technology for VOC Treatment: A Review

College of Energy Environment and Safety Engineering, China Jiliang University, Hangzhou 310018, China
*
Authors to whom correspondence should be addressed.
Energies 2025, 18(13), 3430; https://doi.org/10.3390/en18133430
Submission received: 28 February 2025 / Revised: 28 March 2025 / Accepted: 18 June 2025 / Published: 30 June 2025
(This article belongs to the Special Issue Experiments and Simulations of Combustion Process II)

Abstract

Regenerative combustion represents an efficient and energy-saving combustion technology that significantly enhances thermal efficiency, reduces energy consumption, and minimizes pollutant emissions by recovering and reusing heat energy. This technology has found extensive applications in traditional industries, such as chemical engineering, coating, and printing, as well as in contemporary fields, including food processing and pharmaceuticals. In recent years, advancements in the optimization of combustion devices and the development of efficient catalysts have successfully reduced the combustion temperature for treating organic waste gases while simultaneously improving pollutant removal efficiency. This paper reviews the current status of regenerative combustion technology, summarizes key achievements, analyzes the challenges faced in industrial applications, and anticipates future research directions.

1. Introduction

1.1. The Environmental and Human Health Impacts of VOC Emissions

Volatile organic compounds (VOCs) are a significant component of atmospheric pollution [1], as illustrated in Figure 1. With the rapid advancement of industrial activities, substantial amounts of VOCs are released into the atmosphere. Under sunlight, VOCs undergo photochemical reactions with nitrogen oxides (NOx), resulting in the formation of ozone [2]. Ozone poses considerable risks to both plant life and human health [3]. Elevated concentrations of ozone not only inhibit plant growth and diminish crop yields but also inflict severe damage on forest ecosystems. Furthermore, VOCs contribute to the formation of secondary organic aerosols (SOAs) through chemical reactions, which serve as key precursors to PM2.5 [4,5]. When VOCs and their oxidation products deposit onto soil and water bodies, they disrupt the structure of soil microbial communities and disturb the balance of aquatic ecosystems, thereby threatening ecosystem stability and biodiversity [6]. Consequently, enhancing the management of VOC emissions has emerged as a critical priority in contemporary air pollution control initiatives [7].
Exposure to VOCs is closely associated with various health issues [8], particularly concerning lung health. Inhaling high concentrations of VOCs can trigger acute respiratory symptoms, contribute to chronic respiratory diseases, decrease lung function, and even elevate the risk of lung cancer [9,10]. Yoon et al. [11] found that long-term exposure to toluene and xylene increases oxidative stress, which subsequently damages lung function. Additionally, Armenta-Reséndiz et al. [12] conducted experimental studies demonstrating that VOCs impact the neurobehavioral patterns of rats. A variety of VOCs exist, with the most common types listed in Table 1.
VOCs are primarily generated in industrial sectors such as chemical manufacturing, coating, printing, and pharmaceutical production. The typical components of VOCs include benzene, toluene, ethyl acetate, and formaldehyde. During industrial production processes, VOC emissions are characterized by several features: large exhaust flow rates, typically exceeding 5000 cubic meters per hour, complex gas compositions, and significant fluctuations in pollutant concentrations. Therefore, to achieve the effective control of and reduction in industrial VOC emissions, it is essential to employ treatment technologies that can simultaneously address multiple challenges, including high flow rates, complex components, and variations in concentration.

1.2. VOC Treatment Technologies

Currently, VOC treatment technologies are primarily categorized into two main types: separation and recovery, and decomposition and conversion. Separation and recovery technologies mainly encompass absorption [16], adsorption [14,17], condensation [18], and membrane separation. The absorption method employs solvents to transfer VOCs from the gas phase to the liquid phase, achieving a treatment efficiency of 80% to 90%. However, it necessitates specific absorption liquids for different VOCs [19], which leads to high costs and the potential for secondary pollution. The adsorption method utilizes porous materials, such as carbon-based [20] and oxygen-containing materials, to capture VOC molecules [21]; however, it is challenged by pore blockage. The condensation method separates VOCs by lowering the temperature or increasing the pressure to induce condensation. Meanwhile, the membrane separation method isolates VOCs from inert gases based on differences in their permeation rates through a polymer membrane, achieving an efficiency exceeding 90%. Nonetheless, this method incurs high equipment and operational costs [22,23].
Decomposition and conversion technologies mainly include a regenerative catalytic oxidizer (RCO) [24,25], regenerative thermal oxidation (RTO) [26,27], photocatalytic oxidation [28], and biodegradation [29]. The catalytic combustion method is suitable for low and medium concentrations of VOCs, with a treatment efficiency of 95–98%, but there are three main problems: First, it requires the use of precious metals (e.g., platinum and palladium) or transition metal oxides as catalysts, resulting in high initial investment and operation and maintenance costs. Second, the catalyst is easily inactivated by the poisoning of sulfur, halogens, and other impurities, not only to reduce the efficiency of the treatment but also to require frequent replacement or regeneration, which increases the operating costs, and the requirements for the feed components are strict; third, the treatment of catalyst deactivation may produce solid waste pollution, and some VOCs may also form harmful intermediate products during the catalytic process, with strict requirements on feed components. Third, the catalyst deactivation treatment may produce solid waste pollution, and some VOCs in the catalytic process may also form harmful intermediate products. In contrast, with the regenerative thermal oxidation method, although the initial equipment investment is larger, without the catalyst, there is no poisoning problem, the RTO system can directly oxidize the VOCs to CO2 and H2O at high temperatures, and the decomposition efficiency is more than 99%; while 95% of the heat recovery can significantly reduce the operation of the energy consumption [30], there is no catalyst waste generated, the risk of secondary pollution is low, and the RTO maintains a higher temperature environment. The RTO requires a large amount of fuel supplementation due to the maintenance of a high-temperature environment, and its operating costs are relatively high. Photocatalytic oxidation uses photocatalysts (such as TiO2) to generate electron–hole pairs under light, which produce highly oxidative hydroxyl radicals and superoxide anions, oxidizing VOCs into CO2 and H2O. This technology offers advantages, such as high chemical stability and non-toxicity, but it has high operational costs [31,32]. Biodegradation has a treatment efficiency of 60–90% and is suitable for low-concentration VOCs. However, it is highly affected by environmental conditions such as temperature, humidity, and pH [33]. The characteristics of commonly used VOC treatment methods are summarized in Table 2 below.
RTO technology has been extensively implemented in the field of VOC control [39], owing to its superior performance in treating atmospheric constituents, combined with notable advantages, including low production costs, the ease of installation, and operational reliability. However, the RTO needs to operate at high temperatures above 780 °C, resulting in high fuel consumption and high energy costs. To solve this problem, the regenerative catalytic oxidation (RCO) technology was developed by combining the RTO with catalytic oxidation technology. Matros et al. [40] have shown that RCO systems have multiple advantages over RTO systems: in terms of economics, they can reduce operating costs by about 30%, capital costs by 20%, and fuel consumption is only 1/3 to 1/5 of that of an RTO; in terms of performance, self-heating can be achieved to maintain operation even at very low VOC concentrations, and heat losses are smaller due to lower operating temperatures. In terms of equipment, due to the smaller size of the regeneration bed and reactor, more economical steel and insulation materials can be used, thus realizing significant capital investment savings.
This paper systematically reviews the working principle, reaction mechanism, and the latest progress of the catalyst research of RTO and RCO, aiming to provide a theoretical reference and technical guidance for further research in this field.

2. Regenerative Thermal Oxidation

2.1. Working Principle of RTO

RTO employs high-thermal-conductivity heat storage materials to facilitate heat recycling and waste gas purification. During operation, organic waste gas is directed through the heat storage material for preheating before entering the oxidation chamber, where it is combusted and transformed into carbon dioxide and water. Concurrently, the heat released is absorbed by the heat storage material and utilized to preheat the waste gas in the subsequent cycle. This heat recycling design not only improves combustion efficiency but also decreases energy consumption, rendering it an energy-efficient and environmentally friendly technology for industrial waste gas treatment.

2.2. Types of RTO

2.2.1. Two-Chamber RTO

In the 1970s, the two-chamber RTO was introduced. The two-chamber RTO adopts a dual-bed structure, consisting of two ceramic heat storage beds, an oxidation chamber, four interlocking high-temperature switching valves, and an intelligent control system (as shown in Figure 2). Its “one-in, one-out” process, combined with switching valves, enables the “heat storage” and “heat release” processes, laying the foundation for subsequent technological developments.
In the 1970s, the two-chamber RTO was introduced. This system employs a dual-bed structure, comprising two ceramic heat storage beds, an oxidation chamber, four interlocking high-temperature switching valves, and an intelligent control system (as illustrated in Figure 2). The “one-in, one-out” operational process, in conjunction with the switching valves, facilitates the “heat storage” and “heat release” mechanisms, thereby establishing a foundation for subsequent technological advancements.
Chou et al. [41] investigated the exhaust gas removal efficiency and nitrogen oxide formation characteristics of a two-chamber electrically heated RTO. Their findings revealed that at a valve switching time of 1.5 min, a flow velocity of 0.39 m/s, and an oxidation chamber temperature ranging from 750 to 950 °C, the VOC removal efficiency exceeded 96%, with no detectable nitrogen oxide emissions in the exhaust gas.
Iijima et al. [42] investigated the treatment of toluene waste gas utilizing a two-chamber RTO. Their findings revealed that at lower gas flow rates, radiative heat transfer predominated in the preheating process. Specifically, when the mass flow rate of the raw gas was 1810 kg/h and the honeycomb height was 1.5 m, toluene achieved self-ignition without the need for additional heating, resulting in a thermal efficiency of 95%. However, a reduction in honeycomb height to 0.9 m led to an increase in the minimum toluene concentration required for self-ignition from 340 ppm to 530 ppm, thereby indicating that honeycomb height significantly influences treatment efficiency.
Cannon et al. [43] studied the heat recovery of a two-chamber RTO for VOC treatment and found that when the VOC concentration exceeded 3% of the Lower Explosive Limit (LEL), the system could achieve a self-sustaining operation, with a heat recovery efficiency of up to 95%.
Due to the absence of a purge phase in the two-chamber RTO, a portion of the waste gas remains trapped in the heat storage material during each cycle. When the valve switching alters the flow direction of the waste gas, these residual gases are directly routed into the exhaust pipeline and subsequently discharged through the chimney. This process results in the release of partially unoxidized waste gas into the atmosphere, thereby negatively impacting the purification efficiency.

2.2.2. Three-Chamber RTO

In the 1980s, the three-chamber RTO was introduced, incorporating a heat storage chamber in contrast to the two-chamber RTO for the treatment of residual waste gases (as illustrated in Figure 3). The residual waste gases are redirected into the oxidation chamber for secondary oxidation via a purge fan. The implementation of the “purging” function in the three-chamber RTO effectively mitigated the direct emission of residual waste gases during the switching process, thereby significantly enhancing the efficiency of waste gas purification.
Currently, experimental research on the three-chamber RTO is limited, with the majority of studies concentrating on numerical simulations. Wang et al. [26] conducted an investigation utilizing a three-chamber RTO model, revealing that the outlet temperature decreases as the apparent velocity of the inlet packing bed increases. Furthermore, an increase in the concentration of VOCs results in elevated temperatures in both the oxidation chamber and the outlet, which in turn leads to increased heat loss and reduced thermal efficiency. Additionally, extending the valve switching time causes the heat storage medium to accumulate heat, resulting in a decrease in the oxidation chamber temperature and an increase in the outlet temperature, further diminishing the system’s thermal efficiency. Extending the valve switching time led to heat accumulation in the heat storage medium, resulting in a decrease in the oxidation chamber temperature and an increase in the outlet temperature, ultimately reducing the system’s thermal efficiency. Lorenzo et al. [44] conducted a numerical flow field simulation study on a 20,000 m3/h three-chamber RTO system and found that the monolithic block structure resulted in uneven airflow distribution. By introducing random packing beneath the structure, they effectively resolved this imbalance. Hao et al. [45] conducted transient simulations to analyze the velocity and temperature distributions in a three-chamber RTO system. Their findings revealed that the unique structure of the ceramic honeycomb heat storage bed effectively ensured even airflow distribution, significantly improving the uniformity of the velocity and temperature. This enhancement played a crucial role in boosting the efficiency and stability of the RTO system. While the three-chamber RTO effectively addresses the issue of direct emissions of residual waste gases during the switching process in two-chamber RTO systems, it also has certain drawbacks, including a higher internal pressure (200–300 Pa) and relatively low space utilization (approximately 66%).

2.2.3. Rotary RTO

In the 1990s, the rotary RTO improved energy efficiency through a multi-chamber design and rotary valve switching. Compared to two-chamber and three-chamber RTO systems, the rotary RTO offers higher space utilization, smoother airflow switching, and more stable internal pressure (as shown in Figure 4).
The rotary RTO comprises three primary components: a ceramic heat exchange media, an oxidation chamber, and a rotary distribution valve. The furnace assembly is segmented into twelve distinct zones, consisting of five heating zones, five cooling zones, one purging zone, and one dead zone (illustrated in Figure 5). The rotary distribution valve, driven by a motor at a constant speed, facilitates the seamless transition of waste gas streams between zones, thereby minimizing pressure fluctuations and optimizing energy efficiency. The process begins as waste gases are introduced into the RTO via the main fan, where they traverse through the heating zones for thermal preconditioning before entering the oxidation chamber for high-temperature decomposition. Subsequently, the treated gases are temporarily retained in the cooling zones prior to atmospheric discharge. The thermal energy accumulated in the ceramic heat exchange media is utilized for preheating in the subsequent operational cycles, while a dedicated purge fan introduces clean gas to perform chamber cleaning operations, redirecting residual waste gases to the oxidation chamber for complete decomposition.
Amelio et al. [46] studied the impact of the switching valve system in RTO on the overall system and found that the switching valve system had a minimal effect on the thermal efficiency. However, the thermal efficiency of systems with valves was slightly lower than that of the valve-less systems. In the rotary RTO system, the energy demand was not affected by the cycle duration, and the valve speed could be adjusted freely based on actual needs.

2.2.4. Single-Tube Multi-Valve RTO

The single-tube multi-valve RTO represents an innovative integrated system that emerged in the 21st century, offering dual functionality for both pollutant treatment and heat recovery (as illustrated in Figure 6). This advanced configuration employs zero-leakage multi-rotary valve technology coupled with high-precision flow control mechanisms. The system’s distinctive feature lies in its ability to dynamically adjust the rotary valves according to the required treatment airflow, enabling the optimal distribution of the flow cross-sectional area ratio. This adaptive capability significantly enhances both the operational flexibility and overall system efficiency. Despite its technological advantages, the single-tube multi-valve RTO remains relatively understudied in the academic literature, with limited published research addressing its design and performance characteristics.
The RTO technology has undergone four distinct evolutionary stages, each marked by substantial advancements in key performance parameters, including pollutant removal efficiency, thermal efficiency, and operational longevity. Table 3 presents a comprehensive comparative analysis of RTO systems across these developmental phases, highlighting the progressive improvements in their performance metrics. This evolutionary trajectory demonstrates the systematic enhancement of RTO technology through iterative design refinements and technological innovations.
The data from Table 3 shows that RTO technology has undergone multiple generations of iterative optimization, resulting in significant improvements in waste gas treatment capacity and operational efficiency. The valve system has evolved from early push-down valves/lift valves to rotary valves and then to triple eccentric hard-sealing butterfly valves. This evolution has not only enhanced the sealing performance and service life but also optimized the gas flow characteristics. Currently, since the concentration of VOCs is usually highly volatile, when the concentration is at a low level, direct oxidation requires a large amount of additional energy, resulting in a significant increase in fuel costs. In order to solve this problem, the researchers proposed the treatment scheme of “concentration followed by oxidation”. Practice has shown that adsorption rotors and zeolite materials are ideal media for concentrating VOCs. Compared with the single oxidation process, the combined concentration–oxidation process has not only higher economy and treatment efficiency but can also effectively reduce the emissions of secondary pollutants [47].

3. Regenerative Catalytic Oxidizer

The RTO operates in a high-temperature environment of above 760 °C. During the heating phase and steady-state operation, a large amount of fuel is required to maintain the desired temperature, leading to high energy consumption and operational costs.
To effectively reduce energy consumption, researchers developed the RCO by integrating catalyst technology. Compared to the traditional RTO system, the RCO optimizes the heat storage body near the oxidation chamber in the heat storage chamber, replacing it with a material that has catalytic activity (as shown in the three-chamber RCO in Figure 7). This improvement allows the RCO system to operate within a lower temperature range, typically controlled between 250 and 400 °C. After heating to 250–300 °C, VOCs undergo flameless combustion under the action of the catalyst, significantly reducing the demand for fuel and lowering the overall operating costs.

3.1. The Working Principle of RCO

RCO combines catalytic oxidation with heat recovery technology, overcoming the limitations of traditional thermal oxidizers by using a catalyst to promote VOC oxidation at lower temperatures. The process is simplified as follows:
(1) Preheating stage: The exhaust gas flows through the heat storage medium for preheating. (2) Catalytic oxidation: The preheated gas passes through the catalytic bed, where the catalyst oxidizes the VOCs into CO2 and H2O at temperatures between 250 °C and 400 °C. (3) Heat recovery: High-temperature gases are recovered through heat accumulators and used to preheat cold exhaust gases. (4) Emissions control: The cooled gas is discharged, reducing VOCs and other pollutants.
The complete catalytic oxidation of VOCs is a key reaction in industrial waste gas treatment. The catalyst adsorbs reactant molecules and lowers the activation energy of the reaction, which not only reduces the reaction temperature but also accelerates the reaction rate. The mechanism is typically divided into three main categories (as shown in Figure 8): (1) Eley–Rideal (E-R) [48], (2) Mars–van Krevelen (MVK), and (3) Langmuir–Hinshelwood (L-H) [49].
The E-R model describes the direct reaction process between gas-phase reactants and surface-adsorbed species on the catalyst [50]. This model is particularly suitable for catalytic systems with weak adsorption and fast reactions, showing clear advantages under low-temperature and high-oxygen partial pressure conditions. The MVK model describes the redox cycling mechanism of VOCs with lattice oxygen on the catalyst surface [51]. It is primarily applicable to the catalyst systems of reducible metal oxides, such as cobalt oxide and cerium oxide. The L-H model describes the interaction between adsorbed VOCs and adsorbed oxygen species on the catalyst surface [52]. It can be further subdivided into single-center and dual-center L-H models based on the nature of the adsorption sites for the reactants. These three models describe the mechanism of VOC oxidation from different perspectives.
The nature of catalyst reactions is a redox process, in which reversible redox cycles (e.g., Ce3+/Ce4+, Co3+/Co2+) are realized through the electron transfer channels and active sites provided by the catalyst to maintain the long-lasting activity of the catalyst. The performance of combustion catalysts is mainly affected by the synergistic effects of oxygen vacancies, the surface structure, and redox properties: oxygen vacancies enhance the reaction efficiency by promoting oxygen activation, accelerating lattice oxygen migration (e.g., the Mars–van Krevelen mechanism for CeO2), and modulating the electronic structure; a moderate surface disorder exposes more active sites and enhances the catalyst’s contact with the reactants; and the catalysts’ reactivity is enhanced through cation doping (e.g., introduction of oxygen vacancies by La3+), the nanostructure design (increasing specific surface area), and the construction of metal–oxide interfaces (e.g., the enhancement of oxygen overflow by the SMSI effect of Pt/CeO2), which are some of the strategies that can further optimize the overall performance of the catalysts.
Currently, catalysts used for VOC purification are mainly categorized into three types: precious metal catalysts [53], non-precious metal oxide catalysts [54], and mixed-metal catalysts.

3.2. Precious Metal Catalysts

Precious metal catalysts have attracted significant attention in the field of VOC catalytic oxidation due to their excellent catalytic activity.
Rooke et al. [55] systematically investigated the effects of the pH value and non-ionic surfactants on the morphology of synthesized hierarchically porous Group Vb metal oxides. The study showed that acidic conditions favored the formation of Nb2O5 with a larger specific surface area, while Ta2O5 achieved the maximum specific surface area under alkaline conditions. Without noble metal loading, niobium oxide supports exhibited higher catalytic activity than tantalum oxide; however, after palladium loading, the tantalum oxide system showed more significant activity enhancement. Among them, supports synthesized at a pH = 6 demonstrated optimal catalytic performance, indicating that synthesis conditions influence catalytic activity by regulating the surface properties of the supports. Group Vb metal oxides show potential as noble metal catalyst supports, but the pH of the reaction medium significantly affects the catalytic performance of the Pd/support systems.
Liu et al. [56] prepared 0.5 wt% Pd/OMS-2 catalysts using the deposition–precipitation method (DP), the pre-incorporation method (PI), and ion-exchange (EX). Figure 9a–d show the TEM images and Pd particle size distributions of the different catalysts. It was found that 0.5 wt% Pd/OMS-2-DP exhibited a greater dispersion of PdOx particles on the surface. Figure 9c displays the XRD spectra of the different samples. Compared to OMS-2, the peaks at 2θ = 28.7° and 37.4° for 0.5 wt% Pd/OMS-2-PI and 0.5 wt% Pd/OMS-2-EX are narrower and stronger, indicating an improvement in crystallinity. Figure 9f shows the NH3-TPD spectra of different samples. For OMS-2, the spectrum features a low-temperature peak at 200 °C and high-temperature peaks at 250 °C and 350 °C, reflecting the distribution characteristics of the acidic sites on the catalyst surface. Figure 9g–i show the conversion efficiency of toluene, ethyl acetate, and carbon monoxide for catalysts prepared by different methods. It was found that the OMS-2 catalyst exhibited the lowest activity. The oxidation temperatures for 50% and 90% removal efficiency of toluene were 270 °C and 350 °C, respectively, while those for ethyl acetate were 190 °C and 250 °C, and for carbon monoxide, 128 °C and 140 °C. In comparison, Pd-loaded catalysts significantly reduced the reaction temperature and greatly enhanced the catalytic activity. The catalyst prepared by the DP method exhibited the best performance, attributed to its higher Pd loading, suitable acidity, excellent adsorption capacity, low-temperature reducibility, and high oxygen mobility.
Precious metal catalysts have widespread applications in the catalytic oxidation of VOCs. Among them, noble metals such as Au, Pt, and Pd have garnered significant attention due to their excellent catalytic activity and selectivity. By loading these precious metal active components onto ceramic supports (such as Al2O3, CeO2, TiO2, etc.) or metal support surfaces, the thermal stability, mechanical strength, and catalytic performance of the catalysts can be significantly enhanced. Based on a review of catalysts for the treatment of VOCs [13], this paper further extends the current progress of precious metal catalyst research results as shown in Table 4.
Precious metal catalysts exhibit excellent catalytic performance, but their application is limited due to the scarcity of precious metals, high costs, and deactivation issues such as sintering, aggregation, and poisoning during use. Precious metal catalysts are susceptible to chlorine poisoning and loss of activity. For example, Pt-Pd catalysts can suffer from reduced activity in the presence of CVOCs like TCE and trichloromethane, which are more difficult to oxidize than non-chlorinated VOCs. The chlorinated species can also inhibit the oxidation of co-fed non-chlorinated VOCs [72,73].
Recent studies have shown that non-precious metal-based catalysts not only possess good catalytic performance but also have lower costs, making them ideal alternatives to precious metals.

3.3. Non-Precious Metal Oxide Catalysts

Non-precious metal catalysts perform excellently in VOC oxidation and can be classified into supported and unsupported types. Supported catalysts increase the number of active sites and the contact area by loading the active components onto a support. Among non-precious metal-based catalysts, transition metal and rare-earth metal oxides are ideal alternatives to precious metals due to their high efficiency and low cost. Oxides like CuO and MnO2, with their unique electronic structures and redox properties, can efficiently catalyze the conversion of VOCs into CO2 and H2O.
There is a wealth of research on non-precious metal oxide catalysts. He [74] prepared mesoporous structure ACeOx (A = Co, Cu, Fe, Mn, Zr) composite catalysts with large surface areas and high porosity using the spontaneous precipitation (IMSP) method. It was found that CuCeOx achieved a 99% removal rate of chlorobenzene at 328 °C, while other ACeOx catalysts required temperatures above 405 °C to reach the same level of performance. García et al. [75] investigated the catalytic combustion of naphthalene by metal oxide catalysts (CoOx, MnOx, CuO, ZnO, Fe2O3, CeO2, TiO2, A2O3, and CuZnO), prepared by the precipitation method. The cerium oxide catalysts were found to exhibit stable conversions and high conversion efficiencies at very low temperatures (175 °C), with a conversion efficiency of 90% at 210 °C for the catalytic combustion of naphthalene. Shie et al. [76] used the Pt/γ-Al2O3 catalyst for the catalytic oxidative decomposition of PAHS with naphthalene (the simplest and least toxic PAH) as the target compound. It investigated the relationship between conversion, operating parameters, and related factors such as processing temperature, catalyst size, and airspeed. The results showed that the decomposition reaction rate of naphthalene could be accelerated, and the reaction temperature could be reduced by using the Pt/γ-Al2O3 catalyst. A high conversion of more than 95% can be achieved at a moderate reaction temperature of 480 K and a space velocity lower than 35.000 h−1. Aparicio et al. [77] investigated the effect of CoOx/SiO2 catalysts on the naphthalene combustion reaction. The catalysts were prepared by impregnation with four different concentrations (1, 5, 10, and 15 wt%). It was found that all the loaded cobalt catalysts were found to reduce the combustion temperature of naphthalene during the catalytic combustion of naphthalene. The catalytic activity depended on the metal loading, with the higher loading catalysts being more active. The conversion of CoOx(15)/SiO2 to naphthalene was found to reach a conversion efficiency of 50% at a temperature of 228 °C. The conversion efficiency of CoOx(15)/SiO2 to naphthalene reached 100% at 270 °C.
Lv et al. [78] synthesized a series of M-Ce0.7Zr0.3O2 catalysts (M = S, P, Mo) by introducing Lewis and Brønsted acid sites onto the surface of a 3D-ordered macroporous (3DOM) CeO2 framework. Figure 10 shows the chlorobenzene conversion rate and CO2 yield curves for Ce1−xZrxO2 (Figure 10a,b) and M-Ce0.7Zr0.3O2 (Figure 10c,d). It was found that Ce0.7Zr0.3O2 exhibited the best catalytic activity for the combustion of chlorobenzene.
After modification with acid elements, the activity of Ce0.7Zr0.3O2 was further improved. Compared with P-Ce0.7Zr0.3O2, W-Ce0.7Zr0.3O2, and Mo-Ce0.7Zr0.3O2, S-Ce0.7Zr0.3O2 exhibited the best catalytic performance. As shown in Figure 10e, the conversion rate of chlorobenzene over ceria (CeO2) sharply decreased from 90% to 75% within 24 h of the reaction, whereas Ce0.7Zr0.3O2 and S-Ce0.7Zr0.3O2 maintained a chlorobenzene conversion rate of around 90% over the same period, demonstrating excellent resistance to deactivation.
In Figure 10f, the selectivity ratio of the catalyst for chlorine gas is shown. It was found that compared to CeO2 and Ce0.7Zr0.3O2, S-Ce0.7Zr0.3O2 exhibited lower chlorine selectivity, at only 1.2%. The large number of Brønsted acid sites on the catalyst led to the degradation of chlorobenzene mainly through a hydrolysis pathway, significantly reducing the selectivity for chlorine gas.
Figure 10g,h show the effect of water vapor on the catalytic performance of CB degradation and CO2 yield. After introducing 5% water vapor, the CB conversion rate and CO2 yield decreased for all three samples, as H2O competes with CB or its intermediates for adsorption sites. As the temperature increased, the activity curves for S–Ce0.7Zr0.3O2 became more consistent, indicating that the effect of water vapor diminished. However, for CeO2 and Ce0.7Zr0.3O2, even at 560 °C, the competitive adsorption effect remained significant.
Figure 11 shows the SEM images of CeO2 (Figure 11a,b), Ce0.7ZrO2 (Figure 11c,d), and S–Ce0.7Zr0.3O2 (Figure 11e,f), with Figure 11g presenting the EDS image of S–Ce0.7Zr0.3O2. The used samples still maintain a good three-dimensional ordered structure and mechanical stability. CeO2 and Ce0.7Zr0.3O2 exhibit a large specific surface area, and the EDS analysis indicates that sulfur is evenly dispersed on the surface of the framework. Figure 11h shows the XRD spectra of the three samples, which are similar, with the diffraction peaks weakening and shifting to higher angles after the addition of Zr ions. The HRTEM images (Figure 11i for CeO2, (Figure 11j) for Ce0.7Zr0.3O2, and (Figure 11k) for S–Ce0.7Zr0.3O2) display clear lattice fringes, and Ce0.7Zr0.3O2 shows an increase in lattice spacing due to the introduction of Zr ions.
Ismail et al. [79] synthesized three different morphologies of CeO2 catalysts: shuttle-shaped (CeO2(S)), nanorods (CeO2(R)), and nanoparticles (CeO2(P)). The schematic diagram of the catalyst synthesis process is shown in Figure 12a. The characterization by SEM and TEM (Figure 12b–i) reveals that CeO2(S) exhibits a uniform spindle-like morphology, with a large number of crystal defects, distortions, and significant oxygen vacancies on the surface. These oxygen vacancies play a key role in the catalytic reaction. The XRD results in Figure 12j show that CeO2(S) has broader and weaker diffraction planes compared to the other two morphologies. Additionally, Figure 12k displays a large number of 2–5 nm nanopores in its structure, indicating a higher surface porosity and surface area. Figure 12l shows that CeO2(S) achieves 90% toluene conversion at 225 °C, while CeO2(R) reaches 90% conversion at 283 °C, and CeO2(P) at 360 °C. CeO2(S) exhibits a clear advantage in toluene conversion.
Transition metal oxides have shown excellent catalytic performance in the oxidation of volatile organic compounds (VOCs), and this paper further extends the recent research advances of non-precious metal catalysts in the oxidation of VOCs based on relevant summaries in the literature [13], as shown in Table 5.
Compared to noble metal catalysts, non-noble metal oxides have gained significant attention due to their cost-effectiveness and strong resistance to poisoning. However, relevant studies have shown that constructing multi-component metal oxide composite systems can create synergistic effects at the molecular level, often leading to higher catalytic activity and selectivity than single-component metal oxides, thereby significantly enhancing the catalytic oxidation efficiency of VOCs.

3.4. Mixed-Metal Catalysts

The synergistic effects between metals in mixed-metal catalysts result in catalytic activity and oxidation performance that surpass those of single-metal catalysts. Currently, mixed-metal catalysts have become a research hotspot in the field of VOC treatment.
Wang et al. [89] investigated the effects of different Co loadings (1 wt%, 5 wt%, 10 wt%) on the low-temperature catalytic oxidation performance of CoCe bimetallic catalysts for toluene. Characterization using SEM, TEM, HRTEM, and elemental mapping (as shown in Figure 13a–j revealed that the Co species in the CoCe-5 catalyst were highly dispersed, and the morphology was similar to that of CeO2.
The Figure 13k XRD analysis shows that as the Co loading increased, the intensity of the CeO2 characteristic peaks weakened and shifted to higher angles, indicating that Co species were successfully incorporated into the CeO2 lattice. The introduction of an appropriate amount of Co significantly enhanced the catalyst’s activity in the low-temperature oxidation of toluene. The Figure 13l results demonstrate that the CoCe-5 catalyst exhibited the best catalytic performance, achieving a 90% toluene conversion rate at 192 °C. Wang et al. proposed and described the reaction mechanism of toluene oxidation over the CoCe-5 sample (as shown in Figure 13m). The introduction of Co promoted the Co2+/Co3+ and Ce3+/Ce4+ redox cycles, enhancing the formation of oxygen vacancies. Gas-phase oxygen was activated at oxygen vacancies to form active oxygen species, while the interaction between Co2+ and lattice oxygen facilitated the reduction of Ce4+ to Ce3+. Toluene underwent complete oxidation to CO2 and H2O through intermediates such as benzyl, benzyl alcohol, and benzoate species, under the synergistic effect of active oxygen and oxygen vacancies. According to the Mars–van Krevelen mechanism, the creation and annihilation of oxygen vacancies were key steps in toluene oxidation, while adsorbed oxygen species participated in the vacancy redox cycle involving the conversion between gaseous oxygen and adsorbed oxygen.
Ali [58] prepared Au/TOS, Mn/TOS, Au-Mn/TOS, and TOS catalysts and compared their propane oxidation performance (as shown in Figure 14a). It was found that Au-Mn/TOS exhibited the best catalytic performance, achieving a conversion rate of 95% at 375 °C and complete conversion at 400 °C. The TEM characterization of the Au-Mn/TOS catalyst (Figure 14b) showed that gold nanoparticles (approximately 5–10 nm) were uniformly dispersed on the surface of TOS, and Au and Mn formed a synergistic region on the surface of CeO2 and TiO2. Upon observing the XRD patterns of the catalysts in Figure 14c—(a) Au–Mn/TOS, (b) Au/TOS, (c) Mn/TOS, and (d) TOS—it was found that the XRD diffraction patterns of the Au/TOS and Mn/TOS catalysts were identical to that of TOS. In contrast, the Au-Mn/TOS catalyst exhibited a noticeable change in the 37–39° 2θ range, with the combination of Au and Mn generating new diffraction peaks. The enhanced catalytic activity of the Au-Mn catalysts may be related to the presence of Au5Mn2/Au2Mn. The XPS oxygen spectra of the catalyst before and after the catalytic reaction were studied, showing four main oxygen peaks (as shown in Figure 14d, A—after reaction; B—before reaction), corresponding to OTi (oxygen in TiO2), OCe (oxygen in CeO2), OZr (oxygen in ZrO2), and OH peaks. It was found that Au and Mn could jointly attract non-stoichiometric oxygen and lattice oxygen belonging to CeO2 and ZrO2, while some free lattice oxygen was attracted by TiO2, resulting in an increase in the OTi peak. Au and Mn jointly attract oxygen in a ratio of 1:6. Compared with monometallic catalysts, Au-Mn catalysts attract a higher total amount of oxygen, showing stronger synergistic effects. These phenomena can be explained by the Mars–Van Krevelen mechanism: after introducing Au into the Mn/TOS catalyst, strong synergistic effects occurred between Au and Mn, forming Au-Mn compounds, which promoted the mobility of non-stoichiometric and lattice oxygen in all the supporting oxides. The highly dispersed Au nanoparticles on the metal oxide support not only increased the migration rate of lattice oxygen but also led to the weakening of the metal–oxide bonds, collectively enhancing the catalytic activity of the Au-Mn/TOS catalyst.
Carabineiro [90] studied the structures and performances of the Ce-Co and La-Co mixed oxides. Figure 15a–f show the SEM images of the samples Ce-Co 1:2 CX, Ce-Co 1:2 EM, La-Co 1:2 CX, and La-Co 1:2 EM, along with the corresponding EDS spectra labeled as Z1 (Figure 15c), Z2 (Figure 15d), Z1 (Figure 15g), and Z2 (Figure 15h). The Ce-Co 1:2 CX sample displays a uniform dark-bright region structure with a similar composition, while the Ce-Co 1:2 EM sample shows an uneven distribution of cobalt and cerium. The La-Co 1:2 CX surface features a porous structure with some porosity, and the La-Co 1:2 EM sample exhibits an uneven structure and element distribution. NH3-TPD (Figure 15i) shows that the Ce-Co sample has a higher concentration of acidic sites in the low-temperature region (15.1 μmol/g), far exceeding that of the La-Co CX sample (6.1 μmol/g). The catalytic performance of the Ce-Co and La-Co samples is shown in Figure 15g and Figure 15k, respectively. The Ce-Co catalyst can completely convert toluene at around 250 °C, whereas the La-Co catalyst exhibits poorer activity. This is due to the strong interaction between Ce and Co, which provides more active sites and higher reducibility.
TANG et al. [91] prepared a porous-layered Mn-Ce composite oxide catalyst with a specific surface area of 176.2 m2/g using a simple precipitation/decomposition method. Due to the presence of a large amount of Mn4+ and Ce3+, as well as abundant surface adsorbed oxygen, the catalyst demonstrated excellent catalytic performance in VOC oxidation. The catalyst achieved 90% conversion efficiency for benzene, toluene, and ethyl acetate at temperatures of 260 °C, 245 °C, and 180 °C, respectively.
Based on the review of the existing catalysts for VOCs [13], this paper further summarizes the latest research progress of mixed-metal catalysts in the oxidation of VOCs, and the related research results are shown in Table 6.
In the field of mixed-metal catalyst research, researchers construct bimetallic or multimetallic catalysts by introducing different metal elements (such as Co, Ce, Mn, Au, etc.) to fully utilize the synergistic effects between metals. Numerous studies have shown that by optimizing the composition and structure of such catalysts, their catalytic performance can be significantly enhanced, achieving the complete combustion of single-component VOCs at relatively low temperatures (400 °C).
In recent years, regarding chalcogenide-based catalysts during the reaction process, Liu et al. [104] found that the denitrification performance of MnFeTiOx catalysts significantly decreased after sulfidation treatment, which was mainly attributed to two aspects. Firstly, the presence of SO2 inhibited the adsorption of NH3 on the Lewis acid sites, which led to a reduction in the nitrate species on the surface of the catalysts, which in turn impeded the redox reaction between NH3 and its intermediates and the NOx redox reaction between NH3 and its intermediates and NOx, which ultimately reduced the NOx conversion. Secondly, SO2 reacts with NH3 adsorbed on the catalyst surface to generate sulfur-containing compounds such as NH4HSO4 or (NH4)2SO4, which are deposited on the catalyst surface and cover the active sites, further reducing the catalytic activity. The catalyst deactivation poisoning mechanism mainly includes four aspects. First, harmful substances (e.g., sulfur, chlorine, etc.) react with the active components of the catalyst to generate stable compounds, resulting in the irreversible deactivation of the active sites. Second, impurities or reaction by-products (e.g., ammonium sulfate, chloride, etc.) are deposited on the surface of the catalyst as crystals, which cover the active sites and impede the contact of the reactants [105]. Third, the poisonous substance destroys the pore structure of the catalyst or causes sintering, resulting in a lower specific surface area and fewer active sites [106]. Finally, toxicant molecules compete with reactants for adsorption on the active sites and occupy the catalytic reaction sites [107]. These mechanisms can act individually or synergistically, ultimately leading to a significant degradation in catalyst performance.
In recent years, perovskite-based catalysts have gradually attracted the attention of researchers as a new type of catalytic material. Although its research in the field of VOC degradation is still in its infancy, it has been demonstrated that this type of catalyst has unique advantages and broad application prospects in VOC pollution control.
Liu et al. [108] investigated the performance of hollow spherical LaCoO3 catalysts in toluene oxidation. They found that the LaCoO3 catalyst exhibited excellent low-temperature activity, achieving 50% toluene conversion at 220 °C and 90% conversion at 237 °C. This superior performance was attributed to its large specific surface area, which provided abundant active sites and favorable oxygen adsorption and reduction capabilities. Rezlescu et al. [109] conducted a comparative study between iron spinel (CuFe2O4, MgFe2O4, Ni0.5Co0.5Fe2O4) and perovskite-type (SrMnO3, FeMnO3, La0.6Pb0.2Ca0.2MnO3) catalysts. The results showed that perovskite-type SrMnO3 and La0.6Pb0.2Ca0.2MnO3 demonstrated superior performance in acetone oxidation (>95% conversion at 300 °C), significantly outperforming iron spinel catalysts, which achieved approximately 70% conversion. The excellent performance of perovskite-type catalysts was primarily attributed to their higher concentration of oxygen vacancies. These oxygen vacancies could accommodate more adsorbed oxygen species (O, O2, O2−), thereby enhancing the oxidation efficiency of VOCs. Tarjomannejad et al. [110] examined the catalytic performance of various perovskite-type catalysts LaMn1−xBxO3 (B = Cu, Fe; x = 0, 0.3, 0.7) and La0.8A0.2Mn0.3B0.7O3 (A = Sr, Ce; B = Cu, Fe) in toluene oxidation reactions. They discovered that Fe doping was more effective than Cu doping in LaMn1−xBxO3 perovskite-type catalysts, and the introduction of Sr or Ce at the A site further enhanced the catalytic activity. Among these, La0.8Ce0.2Mn0.3Fe0.7O3 exhibited the best performance, achieving complete toluene conversion at 200 °C.

4. Summary and Outlook

RTO technology is a mainstream method for treating VOCs, offering high removal efficiency and stable operation. However, its high-temperature operation results in increased fuel reheating costs. To reduce energy consumption, catalytic combustion technology was introduced, enabling the efficient oxidation of VOCs into water and carbon dioxide at temperatures between 250 and 400 °C. This approach lowers reheating demands, reduces operating costs, and effectively suppresses the formation of harmful by-products such as dioxins.
Catalysts used in catalytic combustion include precious metals, non-precious metal oxides, and mixed-metal oxides. Precious metal catalysts exhibit high efficiency but are expensive and susceptible to sintering and poisoning. Non-precious metal oxide catalysts are cost-effective, resistant to poisoning, and have long lifespans, but their efficiency is relatively lower. Mixed-metal oxide catalysts offer superior activity and stability. However, all catalysts tend to deactivate over time due to carbon deposition, poisoning, sintering, and water vapor adsorption.
In current VOC catalytic oxidation research, catalysts for single-component VOCs have received significant attention. However, industrial emissions typically consist of complex mixtures of multiple VOCs. Therefore, research should expand into the area of mixed VOC catalytic oxidation, focusing on the development of efficient catalysts and the optimization of multi-site active designs to enhance treatment efficiency and adaptability for complex waste gases. Moreover, the preparation methods of catalysts and the selection of support materials are crucial. Optimizing preparation processes, reducing costs, and improving catalyst performance—particularly through doping techniques to increase the specific surface area and expose more active sites—can further enhance catalytic efficiency and stability. Lastly, improving catalyst thermal stability is also a key focus for future research. By designing new active components or introducing promoters, thermal stability can be enhanced, catalyst deactivation can be reduced, and economic benefits can be improved. In conclusion, future research should focus on the in-depth studies of multi-component VOC catalytic oxidation, combining advanced catalyst design, preparation process optimization, and enhanced thermal stability to provide more efficient, economical, and sustainable solutions for industrial waste gas treatment.

Author Contributions

Investigation, writing—original draft, P.Y.; formal analysis, T.Z.; data curation, investigation, methodology, X.Z.; resources, project administration, M.L.; supervision, writing—review and editing, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the “Pioneer” and “Leading Goose” R&D Program of Zhejiang (2023C03125).

Data Availability Statement

The data are available on reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Anav, A.; De Marco, A.; Friedlingstein, P.; Savi, F.; Sicard, P.; Sitch, S.; Vitale, M.; Paoletti, E. Growing season extension affects ozone uptake by European forests. Sci. Total Environ. 2019, 669, 1043–1052. [Google Scholar] [CrossRef] [PubMed]
  2. He, Z.; Wang, X.; Ling, Z.; Zhao, J.; Guo, H.; Shao, M.; Wang, Z. Contributions of different anthropogenic volatile organic compound sources to ozone formation at a receptor site in the Pearl River Delta region and its policy implications. Atmos. Chem. Phys. 2019, 19, 8801–8816. [Google Scholar] [CrossRef]
  3. Bachheti, A.J.; Bhalla, P.; Bachheti, R.K.; Husen, A. Growth and development of medicinal plants, and production of secondary metabolites under ozone pollution. In Environmental Pollution and Medicinal Plants; CRC Press: Boca Raton, FL, USA, 2022; pp. 25–38. [Google Scholar]
  4. Xu, J.; Hu, W.; Liang, D.; Gao, P. Photochemical impacts on the toxicity of PM2.5. Crit. Rev. Environ. Sci. Technol. 2022, 52, 130–156. [Google Scholar] [CrossRef]
  5. Hui, L.; Liu, X.; Tan, Q.; Feng, M.; An, J.; Qu, Y.; Zhang, Y.; Cheng, N. VOC characteristics, sources and contributions to SOA formation during haze events in Wuhan, Central China. Sci. Total Environ. 2019, 650, 2624–2639. [Google Scholar] [CrossRef]
  6. Ma, W.; Hu, J.; Li, J.; Gao, P.; Okoli, C.P.; Wang, P.; Li, J. Distribution, sources, and health risk assessment of VOCs/SVOCs in soils obtained from petrochemical-contaminated sites in Guangzhou, a subtropical coastal megacity in southern China. J. Clean. Prod. 2023, 426, 139198. [Google Scholar] [CrossRef]
  7. Hao, R.; Xue, S.; Sun, H.; Yang, T.; Wang, H. Emission Characteristics and the Environmental Impact of VOCs from Typical FRP Manufacture Industry. Atmosphere 2022, 13, 1274. [Google Scholar] [CrossRef]
  8. Liu, R.; Chen, J.; Li, G.; An, T. Using an integrated decontamination technique to remove VOCs and attenuate health risks from an e-waste dismantling workshop. Chem. Eng. J. 2017, 318, 57–63. [Google Scholar] [CrossRef]
  9. Zhao, Q.; Li, Y.; Chai, X.; Xu, L.; Zhang, L.; Ning, P.; Huang, J.; Tian, S. Interaction of inhalable volatile organic compounds and pulmonary surfactant: Potential hazards of VOCs exposure to lung. J. Hazard. Mater. 2019, 369, 512–520. [Google Scholar] [CrossRef]
  10. Mozaffar, A.; Zhang, Y.-L. Atmospheric volatile organic compounds (VOCs) in China: A review. Curr. Pollut. Rep. 2020, 6, 250–263. [Google Scholar] [CrossRef]
  11. Yoon, H.-I.; Hong, Y.-C.; Cho, S.-H.; Kim, H.; Kim, Y.H.; Sohn, J.R.; Kwon, M.; Park, S.-H.; Cho, M.-H.; Cheong, H.-K. Exposure to volatile organic compounds and loss of pulmonary function in the elderly. Eur. Respir. J. 2010, 36, 1270–1276. [Google Scholar] [CrossRef]
  12. Armenta-Reséndiz, M.; Ríos-Leal, E.; Rivera-García, M.T.; López-Rubalcava, C.; Cruz, S.L. Structure-activity study of acute neurobehavioral effects of cyclohexane, benzene, m-xylene, and toluene in rats. Toxicol. Appl. Pharmacol. 2019, 376, 38–45. [Google Scholar] [CrossRef]
  13. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs)—A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  14. Yang, C.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of various types of VOCs by adsorption/catalytic oxidation: A review. Chem. Eng. J. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  15. Wang, H.; Nie, L.; Li, J.; Wang, Y.; Wang, G.; Wang, J.; Hao, Z. Characterization and assessment of volatile organic compounds (VOCs) emissions from typical industries. Chin. Sci. Bull. 2013, 58, 724–730. [Google Scholar] [CrossRef]
  16. Dwivedi, P.; Gaur, V.; Sharma, A.; Verma, N. Comparative study of removal of volatile organic compounds by cryogenic condensation and adsorption by activated carbon fiber. Sep. Purif. Technol. 2004, 39, 23–37. [Google Scholar] [CrossRef]
  17. Li, X.; Zhang, L.; Yang, Z.; Wang, P.; Yan, Y.; Ran, J. Adsorption materials for volatile organic compounds (VOCs) and the key factors for VOCs adsorption process: A review. Sep. Purif. Technol. 2020, 235, 116213. [Google Scholar] [CrossRef]
  18. Davis, R.J.; Zeiss, R.F. Cryogenic condensation: A cost-effective technology for controlling VOC emissions. Environ. Prog. 2002, 21, 111–115. [Google Scholar] [CrossRef]
  19. Wu, L.; Zhu, Y.; Yuan, J.; Guo, X.; Zhang, Q. Advances in Adsorption, Absorption, and Catalytic Materials for VOCs Generated in Typical Industries. Energies 2024, 17, 1861. [Google Scholar] [CrossRef]
  20. Xu, X.; Wang, H.; Yu, S.; Chen, H.; Guo, Y.; Zhou, C.; Zeng, Z.; Li, L. Designing activated carbon and porous carbon nanofibers for insight into their differences in adsorption affinity mechanisms of VOCs. Appl. Surf. Sci. 2024, 659, 159961. [Google Scholar] [CrossRef]
  21. Wang, X.; Chai, X.; Huang, W.; Li, X.; Zhu, B.; Li, X.; Zhou, Y.; Yang, Z.; Sun, X.; Fu, L. Green synthesis of biomass-derived porous carbon with hierarchical pores and enhanced surface area for superior VOCs adsorption. Mater. Today Commun. 2024, 39, 108906. [Google Scholar] [CrossRef]
  22. Zhang, L.; Weng, H.-X.; Gao, C.-J.; Chen, H.-L. Remove volatile organic compounds (VOCs) with membrane separation techniques. J. Environ. Sci. 2002, 14, 181–187. [Google Scholar]
  23. Yan, X.; Anguille, S.; Bendahan, M.; Moulin, P. Ionic liquids combined with membrane separation processes: A review. Sep. Purif. Technol. 2019, 222, 230–253. [Google Scholar] [CrossRef]
  24. Li, K.; Luo, X. Research Progress on Catalytic Combustion of Volatile Organic Compounds in Industrial Waste Gas. Catalysts 2023, 13, 268. [Google Scholar] [CrossRef]
  25. Wang, Q.; Yeung, K.L.; Bañares, M.A. Ceria and its related materials for VOC catalytic combustion: A review. Catal. Today 2020, 356, 141–154. [Google Scholar] [CrossRef]
  26. Wang, F.; Lei, X.; Hao, X. Key factors in the volatile organic compounds treatment by regenerative thermal oxidizer. J. Air Waste Manag. Assoc. 2020, 70, 557–567. [Google Scholar] [CrossRef]
  27. Morrone, P.; Di Maio, F.P.; Di Renzo, A.; Amelio, M. Modeling process characteristics and performance of fixed and fluidized bed regenerative thermal oxidizer. Ind. Eng. Chem. Res. 2006, 45, 4782–4790. [Google Scholar] [CrossRef]
  28. Almaie, S.; Vatanpour, V.; Rasoulifard, M.H.; Koyuncu, I. Volatile organic compounds (VOCs) removal by photocatalysts: A review. Chemosphere 2022, 306, 135655. [Google Scholar] [CrossRef]
  29. Yoshikawa, M.; Zhang, M.; Toyota, K. Biodegradation of volatile organic compounds and their effects on biodegradability under co-existing conditions. Microbes Environ. 2017, 32, 188–200. [Google Scholar] [CrossRef]
  30. IIjima, S.; IIjima, S.; Nakayama, K.; Kubota, M.; Matsuda, H. Study on High-efficiency Heat Storage and Heat Release Conditions in a Pilot-scale Regenerative Thermal Oxidizer for Toluene Decomposition. J. CJpn. Soc. Mater. Cycles Waste Manag. 2009, 20, 252–261. [Google Scholar] [CrossRef]
  31. Huang, Y.; Ho, S.S.H.; Lu, Y.; Niu, R.; Xu, L.; Cao, J.; Lee, S. Removal of Indoor Volatile Organic Compounds via Photocatalytic Oxidation: A Short Review and Prospect. Molecules 2016, 21, 56. [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. Wei, Q.; Yang, J.; Tian, C.; Zhang, Z.; Zhang, X.; Zhang, Z.; Li, D. Research on the progress of VOCs adsorption by biomass nanocomposites. J. Phys. Conf. Ser. 2022, 2022, 012023. [Google Scholar] [CrossRef]
  34. Belaissaoui, B.; Le Moullec, Y.; Favre, E. Energy e29fficiency of a hybrid membrane/condensation process for VOC (Volatile Organic Compounds) recovery from air: A generic approach. Energy 2016, 95, 291–302. [Google Scholar] [CrossRef]
  35. Strots, V.O.; Bunimovich, G.A.; Roach, C.R.; Matros, Y.S. Chapter 10—Regenerative catalytic oxidizer technology for VOC control. In Reaction Engineering for Pollution Prevention; Abraham, M.A., Hesketh, R.P., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2000; pp. 113–126. [Google Scholar]
  36. Wu, L.; Zhu, Y.; Yuan, J.; Guo, X.; Zhang, Q. Volatile Organic Compounds (VOCs) in China: Progress and Prospects of Research on Treatment Technologies and Policy Provisions. J. Mater. Sci. Chem. Eng. 2024, 12, 1–43. [Google Scholar] [CrossRef]
  37. Lin, L.; Chai, Y.; Zhao, B.; Wei, W.; He, D.; He, B.; Tang, Q. Photocatalytic oxidation for degradation of VOCs. Open J. Inorg. Chem. 2013, 3, 14–25. [Google Scholar] [CrossRef]
  38. Datta, A.; Philip, L. Biodegradation of Volatile Organic Compounds from Paint Industries. Appl. Biochem. Biotechnol. 2012, 167, 564–580. [Google Scholar] [CrossRef]
  39. Liu, J.; Peng, Z. Experimental and Numerical Investigations into Temperature Distributions and VOC Conversion Rate of RTO. IOP Conf. Ser. Earth Environ. Sci. 2021, 943, 012014. [Google Scholar] [CrossRef]
  40. Matros, Y.S.; Bunimovich, G.A.; Patterson, S.E.; Meyer, S.F. Is it economically feasible to use heterogeneous catalysts for VOC control in regenerative oxidizers? Catal. Today 1996, 27, 307–313. [Google Scholar] [CrossRef]
  41. Chou, M.-S.; Hei, C.-M.; Huang, Y.-W. Regenerative Thermal Oxidation of Airborne N, N-Dimethylformamide and Its Associated Nitrogen Oxides Formation Characteristics. J. Air Waste Manag. Assoc. 2007, 57, 991–999. [Google Scholar] [CrossRef]
  42. Iijima, S.; Nakayama, K.; Kuchar, D.; Kubota, M.; Matsuda, H. Optimum conditions for effective decomposition of toluene as VOC gas by pilot-scale regenerative thermal oxidizer. Int. J. Energy Power Eng. 2008, 2, 1589–1594. [Google Scholar]
  43. Cannon, B.J. Dual-chamber RTO oxidizers provide cost-effective VOC compliance for metal finishers and coaters. Met. Finish. 2003, 101, 53–56. [Google Scholar] [CrossRef]
  44. Giuntini, L.; Bertei, A.; Tortorelli, S.; Percivale, M.; Paoletti, E.; Nicolella, C.; Galletti, C. Coupled CFD and 1-D dynamic modeling for the analysis of industrial Regenerative Thermal Oxidizers. Chem. Eng. Process.-Process Intensif. 2020, 157, 108117. [Google Scholar] [CrossRef]
  45. Hao, X.; Li, R.; Wang, J.; Yang, X. Numerical simulation of a regenerative thermal oxidizer for volatile organic compounds treatment. Environ. Eng. Res. 2018, 23, 397–405. [Google Scholar] [CrossRef]
  46. Amelio, M.; Florio, G.; Morrone, P.; Senatore, S. The influence of rotary valve distribution systems on the energetic efficiency of regenerative thermal oxidizers (RTO). Int. J. Energy Res. 2008, 32, 24–34. [Google Scholar] [CrossRef]
  47. Yang, J.; Chen, Y.; Cao, L.; Guo, Y.; Jia, J. Development and Field-Scale Optimization of a Honeycomb Zeolite Rotor Concentrator/Recuperative Oxidizer for the Abatement of Volatile Organic Carbons from Semiconductor Industry. Environ. Sci. Technol. 2012, 46, 441–446. [Google Scholar] [CrossRef]
  48. Ruaud, M.; Loison, J.C.; Hickson, K.M.; Gratier, P.; Hersant, F.; Wakelam, V. Modelling complex organic molecules in dense regions: Eley–Rideal and complex induced reaction. Mon. Not. R. Astron. Soc. 2015, 447, 4004–4017. [Google Scholar] [CrossRef]
  49. Zhang, Z.; Jiang, Z.; Shangguan, W. Low-temperature catalysis for VOCs removal in technology and application: A state-of-the-art review. Catal. Today 2016, 264, 270–278. [Google Scholar] [CrossRef]
  50. Weinberg, W.H. Eley− Rideal surface chemistry: Direct reactivity of gas phase atomic hydrogen with adsorbed species. Acc. Chem. Res. 1996, 29, 479–487. [Google Scholar] [CrossRef]
  51. Vannice, M.A. An analysis of the Mars–van Krevelen rate expression. Catal. Today 2007, 123, 18–22. [Google Scholar] [CrossRef]
  52. Baxter, R.; Hu, P. Insight into why the Langmuir–Hinshelwood mechanism is generally preferred. J. Chem. Phys. 2002, 116, 4379–4381. [Google Scholar] [CrossRef]
  53. Liotta, L.F. Catalytic oxidation of volatile organic compounds on supported noble metals. Appl. Catal. B Environ. 2010, 100, 403–412. [Google Scholar] [CrossRef]
  54. Lou, B.; Shakoor, N.; Adeel, M.; Zhang, P.; Huang, L.; Zhao, Y.; Zhao, W.; Jiang, Y.; Rui, Y. Catalytic oxidation of volatile organic compounds by non-noble metal catalyst: Current advancement and future prospectives. J. Clean. Prod. 2022, 363, 132523. [Google Scholar] [CrossRef]
  55. Rooke, J.C.; Barakat, T.; Brunet, J.; Li, Y.; Finol, M.F.; Lamonier, J.-F.; Giraudon, J.-M.; Cousin, R.; Siffert, S.; Su, B.L. Hierarchically nanostructured porous group Vb metal oxides from alkoxide precursors and their role in the catalytic remediation of VOCs. Appl. Catal. B Environ. 2015, 162, 300–309. [Google Scholar] [CrossRef]
  56. Liu, L.; Song, Y.; Fu, Z.; Ye, Q.; Cheng, S.; Kang, T.; Dai, H. Effect of preparation method on the surface characteristics and activity of the Pd/OMS-2 catalysts for the oxidation of carbon monoxide, toluene, and ethyl acetate. Appl. Surf. Sci. 2017, 396, 599–608. [Google Scholar] [CrossRef]
  57. Gennequin, C.; Lamallem, M.; Cousin, R.; Siffert, S.; Aïssi, F.; Aboukaïs, A. Catalytic oxidation of VOCs on Au/Ce-Ti-O. Catal. Today 2007, 122, 301–306. [Google Scholar] [CrossRef]
  58. Ali, A.M.; Daous, M.A.; Khamis, A.A.M.; Driss, H.; Burch, R.; Petrov, L.A. Strong synergism between gold and manganese in an Au–Mn/triple-oxide-support (TOS) oxidation catalyst. Appl. Catal. A Gen. 2015, 489, 24–31. [Google Scholar] [CrossRef]
  59. Ousmane, M.; Liotta, L.F.; Carlo, G.D.; Pantaleo, G.; Venezia, A.M.; Deganello, G.; Retailleau, L.; Boreave, A.; Giroir-Fendler, A. Supported Au catalysts for low-temperature abatement of propene and toluene, as model VOCs: Support effect. Appl. Catal. B Environ. 2011, 101, 629–637. [Google Scholar] [CrossRef]
  60. Tabakova, T.; Ilieva, L.; Petrova, P.; Venezia, A.M.; Avdeev, G.; Zanella, R.; Karakirova, Y. Complete benzene oxidation over mono and bimetallic Au–Pd catalysts supported on Fe-modified ceria. Chem. Eng. J. 2015, 260, 133–141. [Google Scholar] [CrossRef]
  61. Tabakova, T.; Kolentsova, E.; Dimitrov, D.; Ivanov, K.; Manzoli, M.; Venezia, A.; Karakirova, Y.; Petrova, P.; Nihtianova, D.; Avdeev, G. CO and VOCs catalytic oxidation over alumina supported Cu–Mn catalysts: Effect of Au or Ag deposition. Top. Catal. 2017, 60, 110–122. [Google Scholar] [CrossRef]
  62. Carabineiro, S.A.C.; Chen, X.; Martynyuk, O.; Bogdanchikova, N.; Avalos-Borja, M.; Pestryakov, A.; Tavares, P.B.; Órfão, J.J.M.; Pereira, M.F.R.; Figueiredo, J.L. Gold supported on metal oxides for volatile organic compounds total oxidation. Catal. Today 2015, 244, 103–114. [Google Scholar] [CrossRef]
  63. Papaefthimiou, P.; Ioannides, T.; Verykios, X.E. Performance of doped Pt/TiO2 (W6+) catalysts for combustion of volatile organic compounds (VOCs). Appl. Catal. B Environ. 1998, 15, 75–92. [Google Scholar] [CrossRef]
  64. Joung, H.-J.; Kim, J.-H.; Oh, J.-S.; You, D.-W.; Park, H.-O.; Jung, K.-W. Catalytic oxidation of VOCs over CNT-supported platinum nanoparticles. Appl. Surf. Sci. 2014, 290, 267–273. [Google Scholar] [CrossRef]
  65. Xu, Z.; Mo, S.; Li, Y.; Zhang, Y.; Wu, J.; Fu, M.; Niu, X.; Hu, Y.; Ye, D. Pt/MnOx for toluene mineralization via ozonation catalysis at low temperature: SMSI optimization of surface oxygen species. Chemosphere 2022, 286, 131754. [Google Scholar] [CrossRef]
  66. Gaálová, J.; Topka, P.; Kaluža, L.; Šolcová, O. Gold versus platinum on ceria–zirconia mixed oxides in oxidation of ethanol and toluene. Catal. Today 2011, 175, 231–237. [Google Scholar] [CrossRef]
  67. Pérez-Cadenas, A.F.; Morales-Torres, S.; Kapteijn, F.; Maldonado-Hódar, F.J.; Carrasco-Marín, F.; Moreno-Castilla, C.; Moulijn, J.A. Carbon-based monolithic supports for palladium catalysts: The role of the porosity in the gas-phase total combustion of m-xylene. Appl. Catal. B Environ. 2008, 77, 272–277. [Google Scholar] [CrossRef]
  68. Bi, F.; Zhang, X.; Chen, J.; Yang, Y.; Wang, Y. Excellent catalytic activity and water resistance of UiO-66-supported highly dispersed Pd nanoparticles for toluene catalytic oxidation. Appl. Catal. B Environ. 2020, 269, 118767. [Google Scholar] [CrossRef]
  69. Hosseini, M.; Barakat, T.; Cousin, R.; Aboukaïs, A.; Su, B.L.; De Weireld, G.; Siffert, S. Catalytic performance of core–shell and alloy Pd–Au nanoparticles for total oxidation of VOC: The effect of metal deposition. Appl. Catal. B Environ. 2012, 111–112, 218–224. [Google Scholar] [CrossRef]
  70. Ren, S.; Liang, W.; Li, Q.; Zhu, Y. Effect of Pd/Ce loading on the performance of Pd–Ce/γ-Al2O3 catalysts for toluene abatement. Chemosphere 2020, 251, 126382. [Google Scholar] [CrossRef] [PubMed]
  71. Wang, Y.; Zhang, C.; Liu, F.; He, H. Well-dispersed palladium supported on ordered mesoporous Co3O4 for catalytic oxidation of o-xylene. Appl. Catal. B Environ. 2013, 142–143, 72–79. [Google Scholar] [CrossRef]
  72. Musialik-Piotrowska, A.; Syczewska, K. Catalytic oxidation of trichloroethylene in two-component mixtures with selected volatile organic compounds. Catal. Today 2002, 73, 333–342. [Google Scholar] [CrossRef]
  73. Musialik-Piotrowska, A. Destruction of trichloroethylene (TCE) and trichloromethane (TCM) in the presence of selected VOCs over Pt-Pd-based catalyst. Catal. Today 2007, 119, 301–304. [Google Scholar] [CrossRef]
  74. He, C.; Xu, B.-T.; Shi, J.-W.; Qiao, N.-L.; Hao, Z.-P.; Zhao, J.-L. Catalytic destruction of chlorobenzene over mesoporous ACeOx (A=Co, Cu, Fe, Mn, or Zr) composites prepared by inorganic metal precursor spontaneous precipitation. Fuel Process. Technol. 2015, 130, 179–187. [Google Scholar] [CrossRef]
  75. García, T.; Solsona, B.; Taylor, S.H. Naphthalene total oxidation over metal oxide catalysts. Appl. Catal. B Environ. 2006, 66, 92–99. [Google Scholar] [CrossRef]
  76. Shie, J.-L.; Chang, C.-Y.; Chen, J.-H.; Tsai, W.-T.; Chen, Y.-H.; Chiou, C.-S.; Chang, C.-F. Catalytic oxidation of naphthalene using a Pt/Al2O3 catalyst. Appl. Catal. B Environ. 2005, 58, 289–297. [Google Scholar] [CrossRef]
  77. Aparicio, M.S.L.; Lick, I.D. Total oxidation of propane and naphthalene from emission sources with supported cobalt catalysts. React. Kinet. Mech. Catal. 2016, 119, 469–479. [Google Scholar] [CrossRef]
  78. Lv, X.; Cai, S.; Chen, J.; Yan, D.; Jiang, M.; Chen, J.; Jia, H. Tuning the degradation activity and pathways of chlorinated organic pollutants over CeO2 catalyst with acid sites: Synergistic effect of Lewis and Brønsted acid sites. Catal. Sci. Technol. 2021, 11, 4581–4595. [Google Scholar] [CrossRef]
  79. Ismail, A.; Zahid, M.; Hu, B.; Khan, A.; Ali, N.; Zhu, Y. Effect of Morphology-Dependent Oxygen Vacancies of CeO2 on the Catalytic Oxidation of Toluene. Catalysts 2022, 12, 1034. [Google Scholar] [CrossRef]
  80. Assebban, M.; Tian, Z.-Y.; El Kasmi, A.; Bahlawane, N.; Harti, S.; Chafik, T. Catalytic complete oxidation of acetylene and propene over clay versus cordierite honeycomb monoliths without and with chemical vapor deposited cobalt oxide. Chem. Eng. J. 2015, 262, 1252–1259. [Google Scholar] [CrossRef]
  81. de Rivas, B.; López-Fonseca, R.; Jiménez-González, C.; Gutiérrez-Ortiz, J.I. Highly active behaviour of nanocrystalline Co3O4 from oxalate nanorods in the oxidation of chlorinated short chain alkanes. Chem. Eng. J. 2012, 184, 184–192. [Google Scholar] [CrossRef]
  82. Ma, C.; Mu, Z.; He, C.; Li, P.; Li, J.; Hao, Z. Catalytic oxidation of benzene over nanostructured porous Co3O4-CeO2 composite catalysts. J. Environ. Sci. 2011, 23, 2078–2086. [Google Scholar] [CrossRef]
  83. Kim, S.C.; Shim, W.G. Catalytic combustion of VOCs over a series of manganese oxide catalysts. Appl. Catal. B Environ. 2010, 98, 180–185. [Google Scholar] [CrossRef]
  84. Solsona, B.; Davies, T.E.; Garcia, T.; Vázquez, I.; Dejoz, A.; Taylor, S.H. Total oxidation of propane using nanocrystalline cobalt oxide and supported cobalt oxide catalysts. Appl. Catal. B Environ. 2008, 84, 176–184. [Google Scholar] [CrossRef]
  85. Dai, Q.; Wang, X.; Lu, G. Low-temperature catalytic combustion of trichloroethylene over cerium oxide and catalyst deactivation. Appl. Catal. B Environ. 2008, 81, 192–202. [Google Scholar] [CrossRef]
  86. Genuino, H.C.; Dharmarathna, S.; Njagi, E.C.; Mei, M.C.; Suib, S.L. Gas-Phase Total Oxidation of Benzene, Toluene, Ethylbenzene, and Xylenes Using Shape-Selective Manganese Oxide and Copper Manganese Oxide Catalysts. J. Phys. Chem. C 2012, 116, 12066–12078. [Google Scholar] [CrossRef]
  87. Behar, S.; Gonzalez, P.; Agulhon, P.; Quignard, F.; Świerczyński, D. New synthesis of nanosized Cu–Mn spinels as efficient oxidation catalysts. Catal. Today 2012, 189, 35–41. [Google Scholar] [CrossRef]
  88. Wang, L.; Wang, Y.; Zhang, Y.; Yu, Y.; He, H.; Qin, X.; Wang, B. Shape dependence of nanoceria on complete catalytic oxidation of o-xylene. Catal. Sci. Technol. 2016, 6, 4840–4848. [Google Scholar] [CrossRef]
  89. Wang, L.; Li, Y.; Liu, J.; Tian, Z.; Jing, Y. Regulation of oxygen vacancies in cobalt-cerium oxide catalyst for boosting decontamination of VOCs by catalytic oxidation. Sep. Purif. Technol. 2021, 277, 119505. [Google Scholar] [CrossRef]
  90. Carabineiro, S.A.C.; Chen, X.; Konsolakis, M.; Psarras, A.C.; Tavares, P.B.; Órfão, J.J.M.; Pereira, M.F.R.; Figueiredo, J.L. Catalytic oxidation of toluene on Ce–Co and La–Co mixed oxides synthesized by exotemplating and evaporation methods. Catal. Today 2015, 244, 161–171. [Google Scholar] [CrossRef]
  91. Tang, W.; Wu, X.; Liu, G.; Li, S.; Li, D.; Li, W.; Chen, Y. Preparation of hierarchical layer-stacking Mn-Ce composite oxide for catalytic total oxidation of VOCs. J. Rare Earths 2015, 33, 62–69. [Google Scholar] [CrossRef]
  92. Xie, S.; Deng, J.; Zang, S.; Yang, H.; Guo, G.; Arandiyan, H.; Dai, H. Au–Pd/3DOM Co3O4: Highly active and stable nanocatalysts for toluene oxidation. J. Catal. 2015, 322, 38–48. [Google Scholar] [CrossRef]
  93. Wu, Z.; Zhang, L.; Guan, Q.; Fu, M.; Ye, D.; Wu, T. Catalytic oxidation of toluene over Au–Co supported on SBA-15. Mater. Res. Bull. 2015, 70, 567–572. [Google Scholar] [CrossRef]
  94. Torrente-Murciano, L.; Solsona, B.; Agouram, S.; Sanchis, R.; López, J.M.; García, T.; Zanella, R. Low temperature total oxidation of toluene by bimetallic Au–Ir catalysts. Catal. Sci. Technol. 2017, 7, 2886–2896. [Google Scholar] [CrossRef]
  95. Li, J.; Zhao, P.; Liu, S. SnOx–MnOx–TiO2 catalysts with high resistance to chlorine poisoning for low-temperature chlorobenzene oxidation. Appl. Catal. A Gen. 2014, 482, 363–369. [Google Scholar] [CrossRef]
  96. Li, S.; Wang, H.; Li, W.; Wu, X.; Tang, W.; Chen, Y. Effect of Cu substitution on promoted benzene oxidation over porous CuCo-based catalysts derived from layered double hydroxide with resistance of water vapor. Appl. Catal. B Environ. 2015, 166–167, 260–269. [Google Scholar] [CrossRef]
  97. Zuo, S.; Liu, F.; Tong, J.; Qi, C. Complete oxidation of benzene with cobalt oxide and ceria using the mesoporous support SBA-16. Appl. Catal. A Gen. 2013, 467, 1–6. [Google Scholar] [CrossRef]
  98. de Rivas, B.; Sampedro, C.; García-Real, M.; López-Fonseca, R.; Gutiérrez-Ortiz, J.I. Promoted activity of sulphated Ce/Zr mixed oxides for chlorinated VOC oxidative abatement. Appl. Catal. B Environ. 2013, 129, 225–235. [Google Scholar] [CrossRef]
  99. Zhou, G.; He, X.; Liu, S.; Xie, H.; Fu, M. Phenyl VOCs catalytic combustion on supported CoMn/AC oxide catalyst. J. Ind. Eng. Chem. 2015, 21, 932–941. [Google Scholar] [CrossRef]
  100. Tang, W.; Wu, X.; Li, S.; Li, W.; Chen, Y. Porous Mn–Co mixed oxide nanorod as a novel catalyst with enhanced catalytic activity for removal of VOCs. Catal. Commun. 2014, 56, 134–138. [Google Scholar] [CrossRef]
  101. Chen, X.; Carabineiro, S.A.C.; Tavares, P.B.; Órfão, J.J.M.; Pereira, M.F.R.; Figueiredo, J.L. Catalytic oxidation of ethyl acetate over La-Co and La-Cu oxides. J. Environ. Chem. Eng. 2014, 2, 344–355. [Google Scholar] [CrossRef]
  102. Genty, E.; Cousin, R.; Capelle, S.; Gennequin, C.; Siffert, S. Catalytic Oxidation of Toluene and CO over Nanocatalysts Derived from Hydrotalcite-Like Compounds (X62+ Al23+): Effect of the Bivalent Cation. Eur. J. Inorg. Chem. 2012, 2012, 2802–2811. [Google Scholar] [CrossRef]
  103. Wang, H.; Yang, W.; Tian, P.; Zhou, J.; Tang, R.; Wu, S. A highly active and anti-coking Pd-Pt/SiO2 catalyst for catalytic combustion of toluene at low temperature. Appl. Catal. A Gen. 2017, 529, 60–67. [Google Scholar] [CrossRef]
  104. Liu, F.; He, H. Selective catalytic reduction of NO with NH3 over manganese substituted iron titanate catalyst: Reaction mechanism and H2O/SO2 inhibition mechanism study. Catal. Today 2010, 153, 70–76. [Google Scholar] [CrossRef]
  105. Zhan, S.; Zhang, H.; Zhang, Y.; Shi, Q.; Li, Y.; Li, X. Efficient NH3-SCR removal of NOx with highly ordered mesoporous WO3(χ)-CeO2 at low temperatures. Appl. Catal. B Environ. 2017, 203, 199–209. [Google Scholar] [CrossRef]
  106. Pan, H.; Jian, Y.; Yu, Y.; He, C.; Shen, Z.; Liu, H. Regeneration and sulfur poisoning behavior of In/H-BEA catalyst for NOx reduction by CH4. Appl. Surf. Sci. 2017, 401, 120–126. [Google Scholar] [CrossRef]
  107. Oudar, J. Sulfur Adsorption and Poisoning of Metallic Catalysts. Catal. Rev. 1980, 22, 171–195. [Google Scholar] [CrossRef]
  108. Liu, Y.; Dai, H.; Deng, J.; Zhang, L.; Zhao, Z.; Li, X.; Wang, Y.; Xie, S.; Yang, H.; Guo, G. Controlled Generation of Uniform Spherical LaMnO3, LaCoO3, Mn2O3, and Co3O4 Nanoparticles and Their High Catalytic Performance for Carbon Monoxide and Toluene Oxidation. Inorg. Chem. 2013, 52, 8665–8676. [Google Scholar] [CrossRef]
  109. Rezlescu, N.; Rezlescu, E.; Dorin Popa, P.; Doroftei, C.; Ignat, M. Some nanograined ferrites and perovskites for catalytic combustion of acetone at low temperature. Ceram. Int. 2015, 41 Pt B, 4430–4437. [Google Scholar] [CrossRef]
  110. Tarjomannejad, A.; Farzi, A.; Niaei, A.; Salari, D. An experimental and kinetic study of toluene oxidation over LaMn1−xBxO3 and La0.8A0.2Mn0.3B0.7O3 (A=Sr, Ce and B=Cu, Fe) nano-perovskite catalysts. Korean J. Chem. Eng. 2016, 33, 2628–2637. [Google Scholar] [CrossRef]
Figure 1. Sources and hazards of VOCs.
Figure 1. Sources and hazards of VOCs.
Energies 18 03430 g001
Figure 2. Schematic diagram of the two-chamber RTO.
Figure 2. Schematic diagram of the two-chamber RTO.
Energies 18 03430 g002
Figure 3. Schematic diagram of the three-chamber RTO.
Figure 3. Schematic diagram of the three-chamber RTO.
Energies 18 03430 g003
Figure 4. Schematic diagram of the rotary RTO.
Figure 4. Schematic diagram of the rotary RTO.
Energies 18 03430 g004
Figure 5. Distribution of the 12 chambers of the rotating RTO furnace (a) and distribution of the four zones of the rotating RTO (b).
Figure 5. Distribution of the 12 chambers of the rotating RTO furnace (a) and distribution of the four zones of the rotating RTO (b).
Energies 18 03430 g005
Figure 6. Schematic diagram of a single-cylinder multi-valve RTO.
Figure 6. Schematic diagram of a single-cylinder multi-valve RTO.
Energies 18 03430 g006
Figure 7. Three-chamber RCO schematic diagram.
Figure 7. Three-chamber RCO schematic diagram.
Energies 18 03430 g007
Figure 8. (a) Eley–Rideal, (c) Langmuir–Hinshelwood, and (b) Mars–van Krevelen [48,49].
Figure 8. (a) Eley–Rideal, (c) Langmuir–Hinshelwood, and (b) Mars–van Krevelen [48,49].
Energies 18 03430 g008
Figure 9. The TEM images and Pd particle size distributions for MOS–2, 0.5 wt% Pd/OMS–2–DP, 0.5 wt% Pd/OMS–2–PI, and 0.5 wt% Pd/OMS–2–EX are shown in Figure (ad), respectively. Figure (e,f) show the NH3–TPD spectra and XRD characterization of the catalysts. Figure (gi) show the catalytic performance of the catalyst samples for the oxidation of toluene, ethyl acetate, and carbon monoxide, respectively [56].
Figure 9. The TEM images and Pd particle size distributions for MOS–2, 0.5 wt% Pd/OMS–2–DP, 0.5 wt% Pd/OMS–2–PI, and 0.5 wt% Pd/OMS–2–EX are shown in Figure (ad), respectively. Figure (e,f) show the NH3–TPD spectra and XRD characterization of the catalysts. Figure (gi) show the catalytic performance of the catalyst samples for the oxidation of toluene, ethyl acetate, and carbon monoxide, respectively [56].
Energies 18 03430 g009
Figure 10. The chlorobenzene conversion rates for Ce1−xZrxO2 and M-Ce0.7Zr0.3O2 are shown in Figure (a,c), respectively, while the CO2 yields are presented in Figure (b,d). Figure (e) illustrates the chlorobenzene conversion rates of the catalyst samples over a 24 h reaction time. Figure (f) compares the selectivity of the catalyst samples for chlorine gas. Figure (g,h) demonstrate the effects of water vapor on the CB degradation catalytic performance and CO2 yield [78].
Figure 10. The chlorobenzene conversion rates for Ce1−xZrxO2 and M-Ce0.7Zr0.3O2 are shown in Figure (a,c), respectively, while the CO2 yields are presented in Figure (b,d). Figure (e) illustrates the chlorobenzene conversion rates of the catalyst samples over a 24 h reaction time. Figure (f) compares the selectivity of the catalyst samples for chlorine gas. Figure (g,h) demonstrate the effects of water vapor on the CB degradation catalytic performance and CO2 yield [78].
Energies 18 03430 g010
Figure 11. SEM images of CeO2 (a,b), Ce0.7ZrO2 (c,d), and S–Ce0.7Zr0.3O2 (e,f), and the EDS image of S–Ce0.7Zr0.3O2 (g). (h) shows the XRD characterization of the catalyst samples. (ik) are HRTEM images of CeO2, Ce0.7Zr0.3O2, and S–Ce0.7Zr0.3O2 catalyst samples, respectively [78].
Figure 11. SEM images of CeO2 (a,b), Ce0.7ZrO2 (c,d), and S–Ce0.7Zr0.3O2 (e,f), and the EDS image of S–Ce0.7Zr0.3O2 (g). (h) shows the XRD characterization of the catalyst samples. (ik) are HRTEM images of CeO2, Ce0.7Zr0.3O2, and S–Ce0.7Zr0.3O2 catalyst samples, respectively [78].
Energies 18 03430 g011
Figure 12. Schematic illustration of the preparation of CeO2(S), CeO2(R), and CeO2(P) (a); SEM and TEM images of CeO2(S) (b,c); SEM and TEM images of CeO2(R) (d,e); SEM and TEM images of CeO2(P) (f,g); HRTEM images of CeO2(S) (h,i); XRD characterization patterns (j); N2 adsorption–desorption isotherms (k); and catalytic performance of the catalysts for toluene conversion (l) [79].
Figure 12. Schematic illustration of the preparation of CeO2(S), CeO2(R), and CeO2(P) (a); SEM and TEM images of CeO2(S) (b,c); SEM and TEM images of CeO2(R) (d,e); SEM and TEM images of CeO2(P) (f,g); HRTEM images of CeO2(S) (h,i); XRD characterization patterns (j); N2 adsorption–desorption isotherms (k); and catalytic performance of the catalysts for toluene conversion (l) [79].
Energies 18 03430 g012
Figure 13. The SEM images of CeO2, Co3O4, and CoCe-5 (a,d,g), TEM images (b,e,h), HRTEM images (c,f,i), the elemental distribution of CoCe-5 (j), the XRD characterization of the sample catalyst (k), the effect of the sample catalyst on the toluene conversion rate (l), and the reaction mechanism of toluene oxidation on the CoCe-5 sample (m) [89].
Figure 13. The SEM images of CeO2, Co3O4, and CoCe-5 (a,d,g), TEM images (b,e,h), HRTEM images (c,f,i), the elemental distribution of CoCe-5 (j), the XRD characterization of the sample catalyst (k), the effect of the sample catalyst on the toluene conversion rate (l), and the reaction mechanism of toluene oxidation on the CoCe-5 sample (m) [89].
Energies 18 03430 g013
Figure 14. The propane conversion rates of Au/TOS, Mn/TOS, Au-Mn/TOS, and TOS catalysts (a), TEM image of Au-Mn/TOS (b), XRD images of Au/TOS, Mn/TOS, Au-Mn/TOS, and TOS catalysts (c), and XPS oxygen spectra (d) [58].
Figure 14. The propane conversion rates of Au/TOS, Mn/TOS, Au-Mn/TOS, and TOS catalysts (a), TEM image of Au-Mn/TOS (b), XRD images of Au/TOS, Mn/TOS, Au-Mn/TOS, and TOS catalysts (c), and XPS oxygen spectra (d) [58].
Energies 18 03430 g014
Figure 15. The SEM images of the samples Ce-Co 1:2 CX (a), Ce-Co 1:2 EM (b), La-Co 1:2 CX (e), and La-Co 1:2 EM (f); the corresponding EDS spectra marked as Z1 (c), Z2 (d), Z1 (g), and Z2 (h); and the NH3-TP profiles (i), as well as the toluene conversion efficiency of the Ce-Co and La-Co samples (j,k) [90].
Figure 15. The SEM images of the samples Ce-Co 1:2 CX (a), Ce-Co 1:2 EM (b), La-Co 1:2 CX (e), and La-Co 1:2 EM (f); the corresponding EDS spectra marked as Z1 (c), Z2 (d), Z1 (g), and Z2 (h); and the NH3-TP profiles (i), as well as the toluene conversion efficiency of the Ce-Co and La-Co samples (j,k) [90].
Energies 18 03430 g015
Table 1. Common VOC pollutant classifications [13,14,15].
Table 1. Common VOC pollutant classifications [13,14,15].
Types of VOCsMain Representative Objects
HydrocarbonsToluene, N-hexane, benzene, xylene, cyclohexane, etc.
Aldehydes and ketonesFormaldehyde, acetaldehyde, acetone, cyclohexanone, etc.
EthersMethyl ether, ethyl ether, dimethyl ether, benzyl ether, benzyl ether, epoxyethane, etc.
EstersEthyl acetate, butyl acetate, methyl acetate, ethyl formate, etc.
CyanideHydrogen cyanide, acrylonitrile, acetonitrile, sodium cyanide, potassium cyanide, benzonitrile, etc.
AmideDimethylformamide, diethylformamide, methylacetamide, formamide, etc.
Halogenated hydrocarbonChloroethylene, dichloromethane, trichloroethylene, bromomethane, freon, etc.
Table 2. Common technologies for treating VOCs.
Table 2. Common technologies for treating VOCs.
Processing TechnologyProcessing MethodsSpecificities
Separation and recycling technologyAbsorption methodSimple equipment, easy to operate, low cost, but low processing efficiency, poor effect on non-polar or insoluble in water exhaust gas [19].
Adsorption methodHigh processing efficiency, large equipment volume, no secondary pollution, low cost, easy to clog, and humidity and temperature affect the treatment effect [17].
Condensing methodSimple process, easy to operate, high energy consumption, large operating costs, high treatment costs [34].
Membrane separation methodSimple process, high processing efficiency, low energy consumption, no secondary pollution, high cost of membrane material [22], more suitable for high-concentration gas treatment.
Decomposition and conversion technologiesRCO methodLow energy consumption, high purification efficiency, no secondary pollution, expensive catalyst cost, with a need to be replaced regularly [35].
RTO methodLarge initial investment, large footprint, high removal rate, applicable to various types and concentrations of VOCs [36].
Photocatalytic methodPhotocatalytic device equipment investment is small, simple structure, easy operation and maintenance, high requirements for light source and catalyst, no secondary pollution [37].
Biodegradation methodSlow reaction speed, large floor space, processing efficiency greatly affected by environmental conditions, low operating costs [38].
Table 3. Comparison of different RTO performance.
Table 3. Comparison of different RTO performance.
TypeTwo-Chamber RTOThree-Chamber RTORotary RTOSingle-Tube Multi-Valve RTONotes
Removal efficiency95%99%99.5%99.5%
Thermal efficiency90%95%95%95%
Maximum concentration handling<1 g/m3<5 g/m3<5 g/m3<10 g/m350 mg/m3 is emission standard
Number of valves49115
Valve formPush-down valve/lift valvePush-down valve/lift valveRotary valveTriple eccentric hard-sealing butterfly valve
Valve life1–2 years1–2 years0.5–1 years3–5 years
Number of heat storage chambers23125
Land occupation100%130%65%65%Using a two-chamber RTO as the baseline
Table 4. Research findings on noble metal catalysts for VOC treatment.
Table 4. Research findings on noble metal catalysts for VOC treatment.
CatalystSupportVOCsTemperature (°C)Conversion (%)Reference
AuCeZrEthanol22290[55]
AuCeZrToluene32290
AuCe0.3Ti0.7O2Propylene293100[57]
AuCe0.1Ti0.9O2Propylene300100
AuTiO2Propylene400100
AuCeO2/ZrO2/TiO2Propylene45095[58]
AuCeO2Propene230100[59]
AuCe7.5/Al2O3Propene250100
AuTiO2Propene320100
AuAl2O3Propene341100
AuCeO2Toluene293100
AuCe7.5/Al2O3Toluene330100
AuTiO2Toluene400100
AuAl2O3Toluene450100
AuCeO2/Fe2OBenzene200100[60]
AuCu:Mn 2:1Methanol18092[61]
AuCu:Mn 2:1Dimethyl ether36095
AuCuOEthyl acetate289100[62]
AuFe2O3Ethyl acetate354100
AuLa2O3Ethyl acetate325100
AuMgOEthyl acetate290100
AuNiOEthyl acetate345100
AuCuOToluene315100
AuFe2O3Toluene345100
AuLa2O3Toluene>400100
AuMgOToluene387100
AuNiOToluene320100
Ptγ-Al2O3Ethyl acetate310>90[63]
PtTiO2Ethyl acetate260>90
PtTiO2(0.45% W6+)Ethyl acetate220>90
Ptγ-Al2O3Benzene18090
PtTiO2Benzene17090
PtTiO2(0.45% W6+)Benzene16090
PtCarbon nanotubesBenzene112100[64]
PtCarbon nanotubesToluene109100
PtCarbon nanotubesEthylbenzene106100
PtCarbon nanotubesMeta xylene104100
PtMnOx-TToluene3098[65]
PtMnOx-LTToluene7099
PtMnOx-BToluene9099
PtMnOx-HBToluene9099
PtCeZrEthanol19390[66]
PtCeZrToluene20790
PdCarbonm-Xylene170100[67]
PdUiO-66Toluene200100[68]
PdTiO2Toluene260100[69]
Pdγ-Al2O3Toluene20090[70]
PdNb2O5Toluene28590[55]
PdTa2O5Toluene28190
PdCo3O4(3D)O-xylene19690[71]
PdCo3O4(3DL)O-xylene25490
Table 5. Research findings on non-precious metal oxides catalysts for VOC treatment.
Table 5. Research findings on non-precious metal oxides catalysts for VOC treatment.
CatalystVOCsT90 (°C)T100 (°C)Reference
CuOEthyl acetate290311[62]
Fe2O3Ethyl acetate355370
La2O3Ethyl acetate367384
MgOEthyl acetate>400>400
NiOEthyl acetate344365
CuOToluene309330
Fe2O3Toluene394-
La2O3Toluene>400>400
MgOToluene>400>400
NiOToluene324330
Co3O4Acetylene-360[80]
Co3O4Propylene-460
Co3O41,2-Dichloroethane-350[81]
Co3O4Benzene263-[82]
Nb2O5Toluene400 [55]
Mn3O4Toluene270-[83]
Mn2O3Toluene295-
MnO2Toluene375-
Co3O4Propane 250[84]
CeO2Trichloroethylene205 [85]
Mn2O3Ethylbenzene374 [86]
Mn3O4Toluene325100[87]
CuOToluene-535
CeO2O-xylene240-[88]
CeO2Naphthalene195-
Table 6. Research findings on mixed metal catalysts for VOC treatment.
Table 6. Research findings on mixed metal catalysts for VOC treatment.
CatalystVOCsTemperature (°C)Conversion (%)Reference
Au-PdToluene16890[92]
Au-CoToluene300100[93]
Au-IrToluene250100[94]
Mn-TiChlorobenzene17790[95]
Sn-Mn-TiChlorobenzene22597
Cu-CoBenzene29090[96]
Co-Ce (6:1)Benzene295100[97]
Co-Ce (12:1)Benzene285100
Co-Ce (18:1)Benzene268100
Co-Ce (24:1)Benzene310100
Ce-Zr1–2 Dichloroethane12090[98]
Mn-Cotoluene25098.7[99]
Mn-CoEthyl acetate19490[100]
Mn-Con-hexane21090
Mn-CeBenzene26090[91]
Mn-CeToluene24590
Mn-CeEthyl acetate18090
La-Cu (1:1 CX)Ethyl acetate310100[101]
La-Cu (1:2 CX)Ethyl acetate309100
La-Cu (1:1 EM)Ethyl acetate319100
La-Cu (1:2 EM)Ethyl acetate304100
La-Co (1:1 CX)Ethyl acetate245100
La-Co (1:2 CX)Ethyl acetate243100
La-Co (1:1 EM)Ethyl acetate242100
La-Co (1:2 EM)Ethyl acetate239100
Cu-CeChlorobenzene32899[74]
Mn-AlToluene260100[102]
Co-AlToluene310100
Cu-AlToluene345100
Fe-AlToluene385100
Ni-AlToluene380100
Pd-PtToluene16098[103]
Cu-Mn (8:2)Toluene377100[87]
Cu-Mn (86:14)Toluene390100
Cu-Mn (94:6)Toluene380100
Cu-Mn (98:2)Toluene389100
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

Yang, P.; Zhang, T.; Ling, Z.; Liu, M.; Zeng, X. Regenerative Oxidation Technology for VOC Treatment: A Review. Energies 2025, 18, 3430. https://doi.org/10.3390/en18133430

AMA Style

Yang P, Zhang T, Ling Z, Liu M, Zeng X. Regenerative Oxidation Technology for VOC Treatment: A Review. Energies. 2025; 18(13):3430. https://doi.org/10.3390/en18133430

Chicago/Turabian Style

Yang, Peng, Tao Zhang, Zhongqian Ling, Maosheng Liu, and Xianyang Zeng. 2025. "Regenerative Oxidation Technology for VOC Treatment: A Review" Energies 18, no. 13: 3430. https://doi.org/10.3390/en18133430

APA Style

Yang, P., Zhang, T., Ling, Z., Liu, M., & Zeng, X. (2025). Regenerative Oxidation Technology for VOC Treatment: A Review. Energies, 18(13), 3430. https://doi.org/10.3390/en18133430

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