Next Article in Journal
Stabilization of b-Glucuronidase by Immobilization in Magnetic-Silica Hybrid Supports
Next Article in Special Issue
Eco-Friendly Cotton/Linen Fabric Treatment Using Aqueous Ozone and Ultraviolet Photolysis
Previous Article in Journal
Bimetallic Cu/Pt Oxygen Reduction Reaction Catalyst for Fuel Cells Cathode Materials
Previous Article in Special Issue
Bimetallic Catalysts for Volatile Organic Compound Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Development of Pharmaceutical VOCs Elimination by Catalytic Processes in China

1
College of Chemical and Pharmaceutical Engineering, Hebei University of Science and Technology, Shijiazhuang 050018, China
2
School of Chemical Engineering, The University of New South Wales, Sydney NSW 2052, Australia
3
Qingdao International Academician Park Research Institute, Qingdao 266000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(6), 668; https://doi.org/10.3390/catal10060668
Submission received: 26 May 2020 / Revised: 8 June 2020 / Accepted: 9 June 2020 / Published: 13 June 2020

Abstract

:
As a byproduct of emerging as one of the world’s key producers of pharmaceuticals, China is now challenged by the emission of harmful pharmaceutical VOCs. In this review, the catalogue and volume of VOCs emitted by the pharmaceutical industry in China was introduced. The commonly used VOC removal processes and technologies was recommended by some typical examples. The progress of catalytic combustion, photocatalytic oxidation, non-thermal plasma, and electron beam treatment were presented, especially the development of catalysts. The advantages and shortages of these technologies in recent years were discussed and analyzed. Lastly, the development of VOCs elimination technologies and the most promising technology were discussed.

1. Introduction

Recently, the Chinese public has become increasingly concerned about the levels of chemical air pollution present in the form of haze. A main contributing factor to this pollution is the release of volatile organic compounds (VOCs) from the industry. VOCs are organic compounds with boiling points in the range of 50–260 °C at atmospheric pressure or with a Reid vapor pressure of over 10.3 Pa at room temperature (293.15 K) and atmosphere pressure (101.325 kPa) [1,2]. The pharmaceutical industry is a major source of these VOCs [3,4], which can have serious ramifications, such as toxicity, carcinogenesis, mutagenesis, photochemical pollution, haze, and fog [5,6,7,8,9]. The pharmaceutical industry has developed greatly in China, as the production of bulk drug intermediates and Chinese patent drugs increased from 205,070 to 340,830 kilotons, and 112,890 to 374,600 kilotons in the past 10 years (2007–2016), respectively [10]. Meanwhile, the amounts of released VOCs increased from about 174.8 to 393.2 kt in the pharmaceutical industry [11,12]. The guiding emission standards of VOCs went into effect on 1 January, 2018 in China, which led to new imputes to upgrade VOCs elimination equipment and technologies in factories across China [13].

2. Catalogue and Emission Amounts of VOCs in the China Pharmaceutical Industry

2.1. Catalogue of Pharmaceutical VOCs in China

Medicines in China are mainly produced in six ways, biological fermentation, chemical synthesis, extraction, coagulation preparation, bioengineering, and treatment of traditional Chinese medicine [4,14]. The different methods produce varying levels of pollution, with the order of emitted VOCs being biological fermentation > chemical synthesis > extraction > bioengineering > treatment of traditional Chinese medicine > coagulation preparation [15]. All these processes can be plagued by the need for high volumes of solvents, the consumption of large quantities of complex organic precursors, or the production of volatile byproducts. Some of the volatile organic compounds used or produced in these processes may be released into the atmosphere, which causes air pollution. The released VOCs primarily include alkanes, alcohols, ketones, aromatic hydrocarbons, halohydrocarbons, amines, esters, ether, aldehyde, carboxylic acid, and sulfur containing organic compounds [4,11,12,14,15].
Biological fermentation is often used to produce antibiotics, vitamins, and amino acids, via processes which include fermentation, separation, purification, and refinement [16,17]. The primary source of emitted VOCs are the solvents used in these processes, especially in the separation and purification steps. Additionally, H2S is also produced as a byproduct in some fermentation processes.
The chemical synthesis technology is usually used for the production of medicine which can be used for prevention, cure, and diagnosis of the disease [18,19]. It contains the units to synthesize intermediates from raw materials, modify the structure of intermediates, purify the products, and dry the final products. The VOCs from this process comprise solvents and unreacted intermediates, which are more complex and often harder to be eliminated than those produced from the biological fermentation. These VOCs include heptane, toluene, xylene, methanol, n-propyl alcohol, isopropanol, phenol, aminomethane, dimethylamine, aminobenzene, cyclohexylamine, triethylamine, butyraldehyde, acetone, chloroform, chlorobenzene, etc.
Extraction involves the use of physical, chemical, and biological methods to separate a substance from a mixture. In the case of the pharmaceutical industry, extraction is carried out to separate organic compounds of interest from liquid solvents [16,20]. As a result, extraction processes require a large supply of solvents, including toluene, naphtha, methanol, ethanol, isopropanol, phenol, acetone, ether, acetic ether, diethylamine, dichloromethane, dichloloethane, chloroform, etc. VOCs are released during extraction mainly from the organic solvents used within the process.
The bioengineering method is a new way to produce some new medicines, via processes such as cloning antibodies, genetic engineering drugs, and genetic engineering vaccines [21]. The VOCs released from this process are similar to those of biological fermentation. These VOCs come from the solvents used within the process and as byproducts of certain reactions. They include n-hexane, methanol, ethanol, formaldehyde, acetaldehyde, formic acid, propanediol, acetone, aminoehtyl alcohol, acetonitrile, acetic acid, acetocaustin, N, N-dimethylformamide, phenol, butanone, 4-methyl-2-pentone, n-propyl alcohol, isopropanol, n-pentanol, isopropyl ether, isobutyl aldehyde, etc.
Another process of pharmaceutical production is the manufacture of traditional Chinese medicines or certain Chinese patent drugs [22]. As this is a traditional process which involves the use of only limited organic solvents to treat the natural animals and plants, the production of VOCs is limited. Hence, this process only emits a spot of VOCs, SO2, and smoke.
Coagulation preparation involves the formation of larger particle agglomerates from fine particle suspensions. In the synthesis of pharmaceuticals, it is the process of mixing the active ingredients with helper constituents (often called coagulants) to produce a drug which has the desired particle size. The physical nature of the coagulation process means that the waste produced via this method is primarily a solid particle with little notable VOC production [4,14,15,16,22].

2.2. The Guiding Emission Standards of VOCs in China

The guiding emission standards of VOCs include six kinds of VOCs on the boundary of factories (Table 1) and four pollutant classifications with a total of 16 compounds in the areas surrounding workshops and installations (Table 2) [13]. These parameters are determined in 1 h by 3~4 samples for the average value. They are detected by the portable instruments or GC. These target VOCs are carcinogenic, odorous, and harmful organic compounds, which are mainly used as solvents and raw materials.

2.3. VOCs Emissions in the Chinese Pharmaceutical Industry

China is the second largest producer of pharmaceutical products, only behind the United States of America. There are more than 1300 kinds of drug intermediates, 30 types of medicaments, and over 4500 pharmaceutical products made in China [23]. According to previous studies, the total VOCs emissions scaled linearly with the amount of final pharmaceutical products. [24]. The results showed that about 0.55 kg of VOCs were discharged to the atmosphere for the production of each 1 kg of final drug products.
The total VOCs emissions from the pharmaceutical industry increased by over 120% from about 174.8 kt in 2007 to 393.2 kt in 2016 (Figure 1) [12]. Although the VOCs emitted from the pharmaceutical industry only account for approximately 1.1% of China’s total VOCs emissions, the absolute emission amount is very large. VOCs emitted from the pharmaceutical industry are potentially more harmful to human beings and ecosystems than the VOCs emitted from other sources. Compared with other emission sources, such as decoration, oil extraction and refining, catering, shoemaking and furniture manufacturing, the VOCs from pharmaceutical industries are more diverse with higher local concentrations and are harder to be eliminated.

3. The Developing Technologies to Dispel VOCs

VOCs Elimination Technologies used in China’s Pharmaceutical Industry

VOCs elimination technologies used in China were originally developed with the aim of recycling organic compounds to reduce cost, but recently the use and implementation of these technologies are targeted at minimizing the environmental impacts of VOCs. Technologies applied in the elimination of VOCs in China include condensation, absorption, adsorption, membrane purification, incineration, catalytic combustion, and the non-thermal plasma process (Figure 2) [15,25,26,27]. These methods are applied according to different working conditions, such as temperature and pressure, depending upon the VOCs targeted for removal. They also have distinct advantages and disadvantages. The development of these VOCs elimination technologies will be briefly introduced and discussed in this review. It is important to note that a wide range of VOCs which are often produced in a single waste stream and the various treatment technologies have different efficacies for the removal of certain VOCs. In fact, multiple technologies are often combined to eliminate the VOC mixtures from the waste streams.
Condensation is the conversion of gas phase VOC mixtures with different vapor pressures to liquid via a decrease in temperature [28]. This technology is often used to recycle the solvents used in pharmaceutical production, with the key advantages that equipment requirements and operation of the condensation process are simple. It makes the cost of condensation lower than other technologies. Another advantage is that the gas produced from condensation is pure. Hence, condensation is often used as the first procedure to treat the waste gas from a pharmaceutical workshop before incineration and absorption to reduce the load on more complex and expensive downstream technologies. Water and air are the most commonly used cooling mediums for the condensation process, but ice, cold salt solutions and organic mediums have been used where cooling temperatures below 10 °C are required, such as the CaCl2 solution, NaCl solution, and ethylene glycol aqueous solution [4]. The condensation efficiency is sensitive to the temperature and pressure and is suitable for the removal of high concentration VOCs which exist as liquefied at moderate temperatures [29,30].
The removal of VOCs via adsorption involves the use of porous materials, while absorption utilizes solvents. The porous materials used in adsorption need a high absorption capacity, large surface area, good pore structure, stable chemical properties, high physical strength, and tolerance of acidic/basic conditions. The adsorption materials commonly used in this technology include activated carbon, porous silica, zeolite, and porous resin [31,32,33,34,35]. The solvents commonly used in the absorption of VOCs are water, acid solution, alkali solution, and other organic compounds [36]. The components used in the adsorption process include spray columns, filled towers, columns of trays, and washing apparatuses.
Both the adsorption and absorption methods have a high VOC removal efficiency and can almost completely remove VOCs from waste gas with low energy consumption (Figure 3). They can be used to recycle organic solvents and valuable compounds while remaining economically feasible. This technology is often used for the treatment of a large flow of waste gas with low VOC concentration in processes such as fix bed adsorption, moving bed adsorption, fluid-bed adsorption, and pressure swing adsorption [37,38,39]. Disadvantages of the adsorption and absorption process are also noteworthy, such as huge equipment requirements, complex procedures, and the need for desorption and regeneration of saturated absorbents. Due to its high overall VOC removal efficiency, this technology is commonly used in many pharmaceutical factories.
Incineration is another widely used technology in pharmaceutical factories to eliminate VOCs (Figure 4). If VOC recycling is not technologically or financially feasible, incineration is a suitable method to completely eliminate VOCs [40]. Incineration is carried out by burning VOCs in a stove or kiln. Ideally, incineration results in the conversion of VOCs into CO2 and H2O in an efficient, simple, and safe manner. However, it has multiple shortcomings. If the VOCs concentration is too low to support the incineration, additional fuel is needed, which increases the running cost. Additionally, some VOCs are less suitable for incineration, because the incomplete combustion of halogenated and other harmful VOCs can result in the release of toxic chemicals such as dioxin, NOx, and CO [40].
To overcome the shortages of the incineration process for VOC destruction, the catalytic combustion technology was developed. In this process, VOCs are decomposed over catalysts at a low temperature (lower than 500 °C, Figure 5 and Figure 6) [1,2,41,42,43,44]. The key factor which governs the catalytic combustion process is the activity of catalysts. Various kinds of catalysts have been used for the catalytic combustion of VOCs, such as noble metal catalysts, transition metal oxides, perovskite catalysts, and concentrated oxidation catalysts. The advantages for this technology are low operation temperature, decreased energy input requirements, high VOC removal efficiency, and minimal generation of toxic byproducts. Catalytic combustion is suitable for the treatment of waste streams containing VOCs across a wide range of concentrations. The main disadvantages of catalytic combustion are high investment requirements for equipment, short catalyst lifetime, and the need for process-specific designs, which are tailored to the waste stream. Nonetheless, the development of viable materials for the catalytic combustion process is still a hotspot for catalysis science.
The mechanism of catalytic combustion is considered by three types, with these being the Mars-Van Krevelen (MVK) model, Langmuir-Hinshelwood (LH) model, and Eley -Rideal (ER) model. In the Mars-Van Krevelen model, the VOCs molecules are initially adsorbed on the active sites, upon which they react with the oxygen species within the catalyst and are decomposed. Then, the reduced catalyst is re-oxidized by the supply of oxygen to the reactor. In the Langmuir-Hinshelwood model, the adsorbed VOCs molecules react directly with the adsorbed oxygen molecules, all occurring on the catalyst surface. In the Elay-Rideal mechanism, the adsorbed oxygen reacts with the VOCs molecules in the gas phase. The reaction pathway which follows depends on both the catalyst materials and the target VOCs in individual systems.
Biodegradation is a widely used process for the treatment of pharmaceutical wastewater (Figure 7) [45,46]. It also can be applied to the treatment VOCs in the gas phase, especially for low concentration VOCs which are suitable for the growth of microorganisms. This process works with using the VOCs as a feedstock for the microorganisms, where they are converted to cytoplasm, CO2, and H2O. The sulfur and nitrogen elements in VOCs can be transformed to H2S, nitrate, or N2 at moderate temperatures. However, VOCs emitted from the pharmaceutical process often contain aromatics or halogens, which would poison the microorganism, rendering this method largely unviable in the abatement of pharmaceutical VOCs.
Non-thermal plasma is also a commonly used technology for the elimination of VOCs [47,48,49]. Free electrons and radicals formed during the plasma process react with VOCs and lead to the degradation of VOCs to CO2 and H2O. There are some advantages of the non-thermal plasma technology, such as a low press drop across the reactor, compact size, and simple equipment structure. The process can be started immediately without warm up and can treat VOCs with solid particles and liquid drops. The non-thermal plasma technology can be combined with catalytic combustion to achieve a superior abatement of pharmaceutical VOCs.
The application of different VOCs elimination technologies depends on the factors such as the range and concentration of VOCs present in the waste stream, the volume of waste stream, and the funds available for installation and running costs. Jie Hao et al. summarized the scope of application for different VOCs elimination technologies (Table 3) [50]. Condensation and adsorption recycling are suitable for recycling VOCs. Catalytic combustion and incineration can remove the VOCs with a moderate concentration (3000–1/4 LEL) at high temperatures, while biodegradation and non-thermal plasma are suitable for abatement of low concentration VOCs at moderate temperatures. Hence, when selecting the efficient and economic VOCs abatement technologies, the scope of application for different technologies needs to be considered.
In a survey of 771 industrial applications of VOCs elimination processes, including 330 cases in China and 441 cases in various other countries (Figure 8), Jinying Xi et al. examined how often the various technologies were used [51]. The data showed that the most commonly used technology in China was adsorption (38%), followed by catalytic combustion (22%) and biodegradation (15%), while in other countries, the most often utilized were biodegradation (29%) and catalytic combustion (29%), followed by adsorption (16%). Due to its simple operation, low capital cost, and ability to recycle VOCs across a wide range of concentrations, adsorption was the most widely used technology in China. However, in some instances, the adsorption equipment was not well maintained and used correctly for recycling of VOCs. Adsorption, membrane purification, and condensation were the most effective methods, therefore most commonly applied, in instances where VOCs were present in concentrations above 10,000 mg·m−3. Catalytic combustion and incineration were used for the destruction of VOCs in the concentration range of 2000~10,000 mg·m−3, where recycling is not financially viable. Biodegradation and non-thermal plasma were applied for the treatment of VOCs in a lower concentration than 2000 mg·m−3. Some examples are introduced in the following section to illustrate the application of these technologies for the elimination of pharmaceutical VOCs in China [3,4,14,51,52].
In the production of cefuroxime axetil, cefuroxime sodium, and cefotaxime sodium, the emitted VOCs include methanol, acetone, dichloromethane, DMF, acetic ether, and cyclohexane. One reported setup for the removal of these VOCs via a combination of condensation and adsorption technologies is described in Figure 9. Firstly, a portion of the various solvents was removed across a three-stage condensing unit comprised of a single stage of recycled water condensation, followed by two stages of 7 °C water condensation. The VOCs which were unable to be removed via the condensation process were removed by a two-stage activated carbon adsorption tower. The VOC-rich waste gas feedstock had a total VOC concentration of 2400 mg·m−3, of which 800 mg·m−3 was attributed to methanol. It entered the first condensation unit at a rate of 2000 m3·h−1. More than 95% of the total VOC content was removed after the combined condensation and adsorption process. As a result, the emitted concentration and discharge rate of methanol was reduced to 18.9–29.3 mg·m−3 and 0.08 kg·h−1, respectively, while the concentration and discharge rate of the other VOCs present was 55–63 mg·m−3 and 0.16 kg·h−1, respectively. Both of these levels were in accordance with the emission standards of Hebei province where the factory was located.
Yan Li has investigated the VOC abatement process in four separate companies in Taizhou city, Zhejiang province [4]. The main products in these four companies were clindamycin, clindamycin phosphate; losartan potassium, valsartan, nevirapine; meropenem, imipenem; and ciprofloxacin, spirolactone, respectively. There were three different procedures used in these four companies. The various procedures included adsorption, catalytic combustion, and non-thermal plasma technologies. The author assessed the efficiencies of the processes by measuring the concentration of a range of VOCs including benzene, toluene, xylene, methanol, formaldehyde, dichloromethane, chloroform, acetic ether, butylene oxide, acetonitrile, dimethylformamide, dysodia, and isopropanol before and after treatment. The results showed that after these elimination processes, the concentrations of VOCs in emitted waste gas were lower than the required concentration in standards for emissions of atmospheric pollutants from the pharmaceutical industry in Zhejiang province [53].
An example of a catalytic VOC treatment process is shown in Figure 10 below, which uses a regenerative catalytic oxidizer. The waste gas from the workshop was collected and combined with the waste gas from the sewage station, storehouse, and solid waste pile. The gas mixture was first washed with a water and alkali solution, then dehydrated and defogged. The dry gas was filtered, then heated with a preheater with an attached regenerative heat transfer. The preheated gas mixture entered the catalytic reactor and was combusted. After catalytic combustion, the gas was washed again with a water and alkali solution and emitted to high altitude atmosphere by a fan. The removal efficiencies of dysodia and non-methane hydrocarbons were 68.07% and 94.33%, respectively. The running fee of the whole VOCs elimination system was about RMB 1 million per year.
Two companies used a process which combines a regenerative thermal oxidizer (RTO) with three condensation stages connected in a series (Figure 11). The waste gas was pretreated in the workshop below determining the lower explosion limit (LEL) of the mixture. Then, fresh air was added in accordance with the determined LEL, such that the proper fuel, oxygen ratio, would be present in the RTO. After that, the gas mixture was oxidized in the RTO at a temperature of 850 °C with 98% of the thermal energy recycled. The high temperature gas exiting the RTO was cooled in a cooling tower, then further cooled across a three-stage condensation setup before it was emitted to the atmosphere. The dysodia concentration in the total vent was 300 mg·m−3, while the concentration of non-methane hydrocarbons was lower than 85 mg·m−3. These concentrations correspond to removal rates of 89% and 92% for dysodia and non-methane hydrocarbons, respectively. The main drawback of this procedure was that the condensation step was not effective for dichloromethane recycling and a lot of HCl generated from the incineration of chlorinated organic compounds, which in turn led to the corrosion of the equipment.
A process which combines the non-thermal plasma and catalytic oxidation techniques (oxidation of VOCs by H2O2 in a low pH) to treat VOCs in waste gas is outlined in Figure 12. The waste gas, with a high VOC concentration, was pretreated, combined with exhaust gas, and washed with an alkali solution. Then, the washed gas entered the catalytic combustion/oxidation tower, where a portion of the VOCs content was oxidized to CO2 and H2O. The oxidized waste gas was washed with water and entered a dehydrator to remove humidity. After that, the waste gas from the sewage station was added to the treated gas, upon which the gas mixture was treated with a non-thermal plasma, to remove additional VOCs. The gas was dehydrated again and entered into the second catalytic oxidation tower. The gas was washed by the alkali solution again and emitted to the atmosphere. The dysodia concentration was reduced by 84% to no more than 250 mg·m−3. The removal efficiency of non-methane hydrocarbons content was approximately 92%, with a concentration of less than 85 mg·m−3 in the emitted gas.
These procedures (Figure 9, Figure 10, Figure 11 and Figure 12 Figure 9 Figure 10 Figure 11 Figure 12) are representative of the range of VOCs elimination technologies currently in use in China. Almost all kinds of VOCs abatement technologies have been applied in the treatment of waste from the Chinese pharmaceutical industry. While the application of these technologies has successfully decreased the emission of VOCs, the high cost of investment and low efficiency are still the main factors which hinder the application of these technologies in the industry. The adsorption, absorption, and biodegradation technologies face challenges concerning the production of secondary pollutants, the desorption of adsorbed VOCs, and the production of waste water and sludge. The incineration process is effective in removing the issue of secondary waste production, but it requires a large energy input and has safety risks posed by the high temperature and use of a flame within a factory. The catalytic combustion and non-thermal plasma technologies partially circumvent the issues of waste and high temperature, but they are currently costly techniques with a short equipment lifetime. Thus, the continuing improvement of these technologies and development of new technologies is needed. With the new emission standards/law coming into effect, VOC elimination processes need to be upgraded in many Chinese pharmaceutical factories. As such, there is a desire for novel, effective, and energy efficient technologies in the near future.

4. The Developing Technologies for VOCs Elimination

As outlined in the previous section, the VOC abatement technologies currently in use in China are hindered by some key limitations, such as high construction and running costs, low removal efficiencies for complex VOCs, and high temperature or energy requirements. To address these shortcomings, significant research is being carried out to improve the traditional methods of VOC removal, such as adsorption, catalytic combustion, and non-thermal plasma [1,2,33,41,43,44,47,48,49,54]. In addition, new technologies which can avoid the disadvantages of the traditional technologies have emerged, such as photocatalytic oxidation, condensation-oxidation, and electron beam treatment [2,42,48,55]. A summary of the various catalysts and conditions that have been investigated as active materials for the catalytic oxidation of VOCs is contained in Table 4 below.

4.1. Catalytic Combustion

Catalytic combustion has received attention recently, as it shows great potential to address the shortcomings of the incineration method. Catalytic combustion is suitable for waste gas streams with low VOC concentrations and a moderate flow rate. Compared with the incineration method, catalytic combustion has been shown to efficiently remove VOCs from waste streams with a wide range of VOC concentrations. In addition, it operates at a lower temperature (293 versus 673 K) than incineration and resistant to the production of undesirable byproducts, such as dioxins and NOx [105]. The diversity of VOC species necessitates the development of different kinds of catalysts for the combustion method, with commonly used catalysts including noble metals, non-noble metal oxides concentrated oxidation catalysts [1,2,41,42,43,44].
Previous reviews covering certain aspects of the catalytic combustion of VOCs have been published. For example, K. Everaert et al. reviewed, analyzed, and discussed the reaction kinetics, reactors, and reaction conditions of catalytic combustion research prior to 2004 [41]. Muhammad Shahzad Kamal et al. and Zhixiang Zhang et al. covered the recent progress in the development of combustion catalysts [2,44]. L.F. Liotta and W.B. Li et al. reviewed the mechanism of VOCs catalytic combustion over noble metal catalysts and non-noble metal catalysts, respectively [1,42]. In this section, we will briefly introduce the more recent development of new catalysts materials which have been used in this process, analyze the advantages and disadvantages they provide, and strategies for their successful implementation.

4.1.1. Noble Metal Catalysts

The general consensus of previous studies is that noble metal catalysts show the best catalytic performance in the combustion of non-halogenated VOCs. The noble catalysts which have been investigated include platinum, palladium, ruthenium, iridium, gold, and silver (Table 5) [2,41,42,43,44]. Due to their size-dependence catalytic properties and high price, noble metal catalysts are often supported on porous supports, such as γ-Al2O3, SiO2, zeolite, and other non-metal oxides to increase the dispersion of noble metal nanoparticles and surface area, which can improve the catalytic efficiency of noble metal catalysts [106,107,108,109,110,111]. Catalysts are often supported on the substrate, which is in the form of a monolith or honeycomb material, such as cordierite, aluminum, and stainless steel [112]. Noble catalysts display high efficiencies for VOC removal at a lower temperature than other kinds of catalysts. Pt shows the best catalytic combustion of VOCs, exhibiting equal removal efficiencies at operating temperatures of up to 100 K lower than those used for other noble metals [42,44]. Two key factors were found to influence the catalytic performance of alumina supported Pt, dispersion and loading amount, with an increase in either of these properties associated with an obvious improvement of catalytic activity [106,113,114]. The metal particle size is also an important factor that influences the catalytic activity. Changes to the active metal particle size can also enhance catalytic performance. For instance, when the crystalline size of alumina supported Pt increased from 1.0 to 15.5 nm, the oxidation rate was found to increase by a factor of 10 [115]. The same phenomenon was also found in the oxidation of propylene over alumina supported Pt and Pd catalysts.
An important factor in the performance of noble metal catalysts for VOC combustion is specificity; in other words, the target VOCs molecules. For instance, P. Papaefthimiou et al. found that Pt and Pd supported on alumina showed good performance for the oxidation of benzene and butanol, but not ethyl acetate, with Pd generally outperforming Pt [57]. M. J. Patterson et al. discovered that alumina supported rhodium is the most active noble metal for 1-hexene, but not for aromatics, while benzene can be decomposed the most easily on platinum, and palladium showed the best catalytic performance in abatement of toluene [116]. The impact of certain non-VOC species in the waste stream can also impact performance, where these species can potentially ‘poison’ the catalyst or cause side reactions to occur. One example is the presence of CO, which was found to have little effect on the performance of Pd catalysts, but significantly inhibit the activity over Pt [117].
Transition metal oxides can be utilized as both supports and promoters for noble metal catalysts. Pt, Pd, Ru, Au supported on MgO, SnO2, Co3O4, NiO, TiO2, CeO2, La2O3, ZrO2 or PrO2 have been explored for the oxidation of toluene, benzene, xylene, propene, light alkane, ethanol, propanol, butanol, formaldehyde, acetone, and acetic acid, respectively [42]. The role of the transition metal oxides is not only to supply a large surface area to disperse the noble metal particles, but in some cases, it can also improve the catalytic performance of noble metal particles by enhancing the mobility of lattice oxygen species. Previous studies showed that active oxygen species formed on cobalt oxide spinel-type crystallites can enhance the catalytic oxidation over PdO supported on alumina [118,119].
The reduction properties also influence the oxidation ability of supported noble metal catalysts. T. Mitsui et al. prepared SnO2, CeO2, and ZrO2 supported Pt and Pd catalysts for the abatement of acetaldehyde [120]. These prepared catalysts were treated in an H2/N2 flow and calcined in the atmosphere. The results showed that the SnO2 supported Pt and Pd showed the best catalytic performance among the calcined catalysts in the atmosphere, while after treatment in an H2/N2 flow, the catalytic activity of SnO2 supported Pt and Pd decreased due to the formation of inactive inter-metallic phases (PtSn and Pd3Sn2). In contrast, CeO2 and ZrO2 supported catalysts showed the improved catalytic activity after reduction. In the elimination of formaldehyde over a TiO2-supported catalyst, Pt/TiO2 showed a superior catalytic performance to Rh/TiO2, Pd/TiO2, Au/TiO2, and neat TiO2 [121]. Other research showed that a series of supported Pt, Pd, and Au catalysts can even partially eliminate formaldehyde at room temperature [122].
Recently, the use of MnOx based materials as catalyst supports has gained attention. In one instance, supporting Ag on NiO-doped MnO2 showed a high activity towards the combustion of o-xylene [89]. The improvement in xylene oxidation was attributed to the enhanced oxygen activation and mobility afforded to the catalyst support via the addition of NiO and Ag. Wenbo Pei et al. explored the use of ordered, mesoporous Mn2O3 supports with embedded Pt particles for the catalytic combustion of toluene [66]. They found that the strong interaction between Pt and Mn2O3 in the ordered structure improves the activity and stability of the catalyst.
The single noble metal catalysts cannot satisfy the requirements of VOCs combustion. Therefore, some mixed noble metal catalysts have been developed to combine the advantages of different noble metal catalysts, such as Pt-Au, Cu-Au, and Pd-Au. T. Tabakova et al. found that the Pd deposition on the deposited gold showed the best catalytic performance for benzene combustion, which was totally eliminated at 200 °C. It also showed good stability [57]. M. Hosseini et al. showed that the deposition of palladium on aurum supported on TiO2 (Pd(shell)-Au(core)/TiO2) can significantly improve the catalytic activity for oxidation of toluene and propylene [67]. Der Shing Lee et al. deposited Au-Pd bimetallic nanoparticles on CeO2 for toluene degradation, which showed a much better catalytic performance than Au/CeO2 and Pd/CeO2 catalysts due to the synergistic effect of gold and palladium [68]. The addition of non-noble metals also can improve the catalytic activity of noble metal catalysts. Roberto Fiorenza et al. prepared Au-Ag/CeO2 and Au-Cu/CeO2 bimetallic catalysts for alcohol oxidation and CO oxidation. These two catalysts showed higher selectivity for intermediate products higher CO conversion at a low temperature (100 °C) than Au/CeO2 [93]. The addition of Au also can improve the performance of Al2O3 supported Cu-Pt catalysts in DMDS oxidation [81].
Noble metal catalysts showed a high catalytic activity and remarkable thermal stability in catalytic elimination of VOCs. However, the use of noble metal catalysts is also associated with distinct disadvantages. Firstly, the high cost of noble metal limits their application in the industrial abatement of VOCs. Secondly, the presence of chlorine, sulfur, CO, and water can suppress the catalytic performance significantly [42,123,124,125]. The regeneration and recycling of noble metal catalysts poisoned by Cl and S is difficult, so they are not suited to the treatment chlorine and sulfur-containing VOCs. In fact, the release of chlorine and sulfur containing VOCs is particularly common in pharmaceutical production processes, so chlorine- and sulfur-containing VOCs need to be removed prior to the treatment by noble metal catalysts, which will further increase the cost of waste gas purification.

4.1.2. Non-Noble Metal Catalysts

To address the cost of noble metals, non-noble metal oxides catalysts were developed for the abatement of VOCs. The materials which have been studied as non-noble metals include the derivatives of transition metals and rare earth elements, such as Ti, Cu, Mn, Al, Ce, Co, Fe, Cr, and V (Table 6) [1,126,127,128,129]. Although transition metal oxides catalysts generally showed a lower catalytic activity than noble metal catalysts for the oxidation of VOCs, they have many advantages, such as a resistance to chlorine and sulfur poisoning, tunable material properties, low cost, long on-stream lifetime, easy regeneration, and low environmental impact. The non-noble metal oxides catalysts applied and studied in VOCs abatement include CuOx, MnO2, FeOx, NiOx, CrOx, and CoOx.
The non-noble metal oxides catalyst systems which we will discuss include ones in which the metal oxide is both supported or unsupported. Due to the presence of mobile oxide species in lattice, Co2O3 displays excellent reduction and oxidation abilities. As such, studies have shown Co2O3 to be one of the best catalysts used for the combustion of benzene, toluene, propane, 1,2-dichloroethane, and 1,2-dichlorobenzene [130,131,132,133]. The catalytic activity of Co2O3 is determined by the method of preparation, treatment conditions, and surface area.
MnO2 is another commonly used metal oxide catalyst, which has been applied and studied in the abatement of n-hexane, acetone, benzene, ethanol, toluene, propane, trichloroethene, ethyl acetate, and NOx [133,134,135,136,137,138,139]. The catalytic activity can be tuned by the preparation method and depends on the structure, surface area, support materials, and oxidation states of catalysts. In the catalytic combustion of ethyl acetate and hexane, MnO2 even achieved a better activity than Pt/TiO2 [140]. Yonghui Wei et al. removed the La atoms from the LaMnO3 perovskite to prepare MnO2 with a high surface area (>150 m2/g), upon which it showed excellent catalytic activity in the oxidation of toluene [96]. Zhang Kai et al. synthesized the nano-cubic MnOx which has a large specific surface area, many oxygen vacancies, and good low temperature reducibility. The conversion of toluene via combustion was more than 90% at 350 °C [95]. Xueqin Yang et al. found that the acid treatment did not change the morphology of the catalyst, but could improve the oxidation ability of the catalyst by increasing the number of Mn4+ species and structural defects on the surface of the catalyst [69]. A common theme throughout the implementation of Mn-based metal oxide catalysts for VOC oxidation is the availability and mobility of oxygen within the MnOx, which is attributed to the oxidation and reduction ability of the Mn afforded by the multiple oxidation states in which it can exist.
Copper oxides are another kind of efficient catalysts used in total oxidation of methane, methanol, ethanol, and acetaldehyde [141,142]. The main factors which influence the catalytic activity are the Cu oxidation state and the availability of lattice oxygen. The addition of other metal oxides, such as CeO2, can enhance the catalytic ability noticeably [133].
Chromium oxides are also promising oxidation catalysts, especially for the combustion of halogenated VOCs [124,133,143]. For chromium oxides, highly crystalline samples showed a better catalytic activity than amorphous ones [144]. Rotter et al.’s research showed that, when using TiO2 as a support material, chromium oxides achieved higher catalytic oxidation of trichloroethylene than manganese oxide, cobalt oxide, and iron oxide [82]. Chromium has also been successfully supported on silica, alumina, porous carbon, and clay to eliminate pollutants such as carbon tetrachloride, chloromethane, trichloroethylene, ethyl chloride, chlorobenzene, and perchloroethylene [142]. However, chromium oxides also suffer deactivation due to the reaction between chromium and chlorine to form Cr2Cl2 [145,146].
CeO2 is a widely used catalyst in oxidation reactions due to its strong interactions with other metals, high oxygen storage capacity, and ready shuttling between the Ce3+ and Ce4+ states [147,148,149]. Dai et al. compared the removal of chlorinated alkanes and alkylenes over CeO2 [83]. The results showed that CeO2 is more efficient when it comes to oxidizing chlorinated alkanes than chlorinated alkylenes. CeO2 also faced the deactivation problem due to the absorption of Cl2 and HCl on the surface [150], so the design of chlorine resistance metal oxide catalysts is still a challenge which must be overcome.
Vanadium oxides were also developed to decompose chlorinated VOCs, such as polychlorinated pollutants and dichlorobenzene due to its tolerance of chlorine and sulfur compounds [151]. The presence of water can enhance and suppress the catalytic activity of V2O5 via the removal of surface absorbed chlorine and reduction of active sites, respectively [152]. Other non-noble metal oxides were also investigated for abatement of VOCs, such as NiO and FeOx, which require further improvement of catalytic efficiency [153].
The above discussion suggests that the use of a single metal oxide as VOC oxidation catalysts is too often plagued by either a low catalytic activity or catalyst poisoning. Thus, focus has shifted to the development of mixed metal oxide catalysts such as Mn-Ce, Mn-Cu, Co-Ce, Sn-Ce, Mn-Co, and Ce-Cu oxides [119,154,155,156,157]. The logic here is that combining two metal oxides with different materials and catalytic properties allows for a synergistic enhancement in performance. The previous studies showed that the rate determining step of VOC catalytic combustion was the oxygen removal from the catalysts lattice [119], so the goal of mix-metal oxides catalysts design was the enhancement of the lattice oxygen species availability.
The addition of copper into CeO2 can promote the catalytic efficiency due to a synergistic effect, leading to a highly efficient decomposition of ethyl acetate, ethanol, propane, benzene, and toluene [158,159,160]. MnOx-CeO2 has been applied for the destruction of ethanol, formaldehyde, hexane, phenol, ethyl acetate, and toluene [155,161,162,163,164]. Mn-Co oxides catalysts also showed improved catalytic activity relative to either MnOx or Co2O3 in the combustion of ethyl acetate and n-hexane [100]. CeO2-CrOx showed excellent catalytic activity for the decomposition of chlorinated VOCs [165], while the removal of chlorobenzene over MnOx-TiO2 and MnOx-TiO2-SnOx showed much better catalytic performance than not only the individual oxides, but also achieved removal efficiencies on par with noble metal catalysts [166]. A three-dimensional ordered mesoporous material of mixed cerium-manganese oxide was prepared for the efficient catalytic combustion of chlorine-containing VOCs due to its large specific surface, enriched Ce3+ content, oxygen vacancies, active oxygen species, and acidic sites. It showed good water resistance and high airspeed applicability. However, catalyst deactivation caused by inorganic chlorine adsorption still occurred [90]. Layered copper manganese oxide has been prepared for the catalytic combustion of CO and VOCs, which showed efficient activity due to the interfacial structure of mixed phases and the formation of the Cu2+-O2−-Mn4+ entity [58]. Acidic sites can be provided by support to prevent the decrease of catalytic activity. CeO2@SiO2 was prepared to catalyze the combustion of 1,2-dichloroethane. SiO2 can provide weak acid sites, as well as promote the adsorption and activation of 1,2-dichloroethane and the desorption of generated HCl [85].
A vast number of preparation methods exist for the synthesis of mixed metal oxide catalysts, including thermal decomposition, impregnation, co-precipitation, and the sol-gel method [167,168,169]. The selection of preparation methods depends on the properties of catalysts and the application situation. Furthermore, as discussed above, complex ordered microporous, multilayer or core-shell structures have recently been applied to catalytic oxidation processes to access properties which come from having a highly controlled particle composition and morphology. The chance to alter not only the metal centers present within mixed metal oxide systems, but also the relative metal ratios.

4.1.3. Perovskite Catalysts

Perovskite-type oxides are a kind of composite oxides which have a similar structure with CaTiO3, and can be expressed by ABO3. The common way to modify the perovskite catalysts is replacement of the cation B by B’ to tune the redox ability or enhance the stability [1]. With the replacement of the cation B, the crystal lattice would be distortion which leads to the enhancement of redox ability and improvement of stability. The most commonly used perovskite for catalytic combustion of VOCs is LaBO3, in which B can be Co, Fe, Ni, Mn, and Sr (Table 7) [60,97,170,171]. Huang et al. used Sr partially replaced La in LaCoO3 for catalytic combustion of propyl alcohol, toluene, and cyclohexane [70]. The results showed that the doped LaCoO3 showed a better catalytic performance than the undoped one, and the modified catalysts were stable in the reaction. R. Spinicci et al. compared the catalytic activity of LaMnO3 and LaCoO3 for catalytic combustion of acetone, isopropanol, and benzene [60]. They suggested that LaMnO3 showed a better performance than LaCoO3. In oxidation of isopropanol, acetone was the intermediate product. The surface oxygen species played a key role in this process. The increase of oxygen pressure is positive for the catalytic combustion of VOCs over these perovskite catalysts. G. Sinquin et al. applied LaMnO3 and LaCoO3 for the catalytic combustion of chlorinated VOCs, such as CH2Cl2 and CCl4. LaMnO3 showed a better chlorine resistance than LaCoO3 [97]. Mihai Alifanti et al. supported LaCoO3 on cerium-zirconium oxides (Ce1-xZrxO2, x = 0–0.3) for the catalytic combustion of benzene and toluene [172]. The results showed that all the supported catalysts showed a better performance than Ce1-xZrxO2 and 20% loaded LaCoO3 showed about 10 times higher catalytic activity than LaCoO3 for toluene oxidation due to its large surface area and good oxygen mobility. S. I. Suárez-Vázquez et al. synthesized SrTi1-xBxO3 (B = Mn, Cu) for toluene destruction [71]. Mn could replace Ti and enter the perovskite structure, while Cu could not. The Mn doped catalysts showed the highest catalytic activity and can completely decompose toluene to CO2 at a temperature lower than 350 °C. Perovskite also can be prepared from solid waste such as the obsoleting lithium battery. Mingming Guo et al. prepared manganese-based perovskite catalyst from the waste lithium battery for catalytic combustion of toluene, which showed a better catalytic activity than pure manganese perovskite catalyst due to more Mn4+ ions and lattice oxygen species, as well as high specific surface area [72]. In order to increase the amounts of active sites, Junxuan Yao et al. removed the La ions from LaCoO3 to obtain the disordered Co3O4. It showed a better catalytic activity for propane combustion than the one prepared by other methods [94]. To further improve the catalytic activity, γ-MnO2 was calcined on the surface of SmMnO3 which had a large specific surface area, high Mn4+/Mn3+ and Olatt/Oads. Compared with SmMnO3, it showed a better catalytic activity and stability (10 vol% water) in the process of catalytic reaction [61]. Jingsi Yang et al. assembled the LaMnO3 perovskite in MnO2 and adjusted La/Mn to 15. The redox ability of the catalyst was improved by enhancing the interaction between the active phase and the support [73]. The ratio of citric acid and metal ion (La3+Mn2+) was also tested to find out the best composite of perovskite catalysts. Zakaria Sihaib et al. prepared LaMnO3 with different ratios. The results show that the catalyst with the ratio of 0.5 to 1.5 has the best catalytic performance and the amount of citric acid affects the specific surface area of perovskite catalyst [74]. Li Wang et al. added Sr into the LaMnO3 to prepare La0.5Sr0.5MnO3. The amount of Mn4+ and the oxidation ability of vinyl chloride has been improved after HCl modification [104]. The perovskite catalysts showed a good catalytic activity for combustion of VOCs at low temperature due to their tunable redox property by replacing the B atom. However, they also have some disadvantages, such as low thermal stability. The catalytic activity and stability of perovskite catalysts need to be further improved.

4.1.4. Concentrated Oxidation Catalysts

The low concentration of VOCs can limit the catalytic efficiency of catalysts. Some porous materials were investigated to the concentration and decomposition of VOCs, such as zeolite, γ-Al2O3 (Table 8, Figure 13) [84,101,173,174,175]. R. Beauchet et al. tried to decompose the isopropanol and o-xylene mixture over the CsX, NaX, and HY zeolite [173]. O-xylene and isopropanol were totally decomposed at 250 °C over the NaX zeolite. The addition of Pt on the zeolite can significantly increase the catalytic activity. The main obstacle for the application of zeolite was the coke that formed during the reaction, which led to the short lifetime. Amir Ikhlaq et al. studied the mechanism and kinetics of decomposition of chlorinated VOCs by ozonation over the γ-Al2O3 and ZSM-5 zeolite [174]. The results suggest that ozone reacts with the absorbed VOCs on the surface of catalysts. However, the ozonation process will increase the cost in most of the industrial processes. Yuexin Peng et al. supported MnO2 on the Al-rich β-zeolite to degrade toluene [75]. The T90 is 285 °C, which is much lower than the MnO2 supported on γ-Al2O3 due to the lattice oxygen species in MnO2 and absorbed oxygen species on the zeolite. Cu and Co were also used to modify the β-zeolite for the destruction of toluene and trichloroethylene [76]. The modified zeolite showed good stability in the reaction. The catalytic activity is mainly from the supported metal oxides. The suitable acidity and strong oxidation stability can improve the CO2 selectivity. Dedong He et al. modified the HZSM-5 zeolite with a series of rare earth elements, including La, Ce, Pr, Nd, Sm, Y, and Er, for the catalytic decomposition of CH3SH [91]. Cu was also used to modify ZSM-5 for the combustion of acrylonitrile. The isolated Cu is the active center. The SiO2/Al2O3 ratio can affect the ion exchange capacity and the catalytic performance of the catalyst. When the ratio is 26, the catalyst shows the best catalytic activity [92]. SBA-15 was used to support MnOx for the combustion of toluene and showed good catalytic activity [176]. The La modified HZSM-5 zeolite showed much better activity and stability than HZSM-5 due to the tunable acidity, which can promote the adsorption and activation of the CH3SH molecule and inhibit the formation of coke deposit. The previous researches showed the potential application in the industry. However, to fulfill the requirements of the industry, the concentration-catalysis process to remove VOCs needs more research on lifetime, catalytic efficiency.

4.1.5. The Influence Factors on Catalytic Performance

In the real industrial process, a lot of factors influence the catalytic efficiency. Firstly, the kind of VOCs determines the selection of catalysts. For example, noble metal catalysts show the best VOCs eliminating efficiency, but they are not suitable for the destruction of chlorine and sulfur containing VOCs due to the poison of the catalysts. Secondly, the surface area of the catalysts is the main factor that influences the catalytic activity. Research shows that MnO2 with a higher surface area showed much better catalytic performance than the one with a lower surface area [133]. Thirdly, the crystal type of catalysts also influences the catalytic performance of the catalysts with the same content. For instance, the catalytic performance of TiO2 with different crystal types, namely rutile and anatase, showed a different catalytic activity [121,167]. The humidity is the common content in the industrial waste gas. In most of the reports, water molecules can suppress the catalytic activity due to the complete adsorption on active sites and destruction of catalysts [177]. In other researches, humidity plays a positive role in the oxidation. In the catalytic combustion of chlorobenzene over VOx/TiO2, VOx-WOx/TiO2, and VOx-MoOx/TiO2, water can remove the adsorbed Cl from the catalysts surface and react with chlorine to produce HCl [134]. The water vapor can show a different effect on the same VOCs over different catalysts. Kullavanijayam et al. reported that water enhanced the catalytic oxidation of cyclohexene over the ceria–alumina supported Pt and Rh, but it had a negative effect on the oxidation of cyclohexane over the ceria–alumina supported Pd catalyst [178]. CO is another poison for precious metal catalysts, so during the catalytic combustion, enough air is needed to avoid the generation of CO. The life of catalysts is also important for catalytic combustion. The main obstacle for the development of catalysts for the catalytic combustion of halogen and sulfur containing VOCs is the short life of common catalysts. Although V oxides show less catalytic efficiency than other catalysts, it is still used in the elimination of halogen and sulfur containing VOCs due to its long catalytic life [152]. Since after the installation of catalytic combustion equipment, it will be in operation for a long time, therefore, the life of the catalysts determines the cost of catalytic combustion.
The catalytic combustion process has been well developed in recent years. The diversity of the catalysts has been investigated. However, there are still some obstacles on the way to the industrial application, such as short lifetime, high cost, and no universality to different kinds of VOCs. More work should be done on these problems.

4.2. Photocatalytic Oxidation

Photocatalytic reactions have drawn a lot of attention and have been well developed in recent years since Fujishima found the splitting water to H2 and O2 over TiO2 [55,179,180]. Different kinds of photocatalysts have been developed to treat VOCs containing waste water, such as TiO2, WO3, ZnO2, CdS, g-C3N4, and BiOBr (Table 9) [56,180,181,182,183,184,185]. The mechanism of photocatalysis is that when the light with a suitable wavelength radio on catalysts (semiconductors), the electrons and holes were separated and generated on the surface of catalysts, then the radicals of •OH and O2− was formed on the surface of catalysts, the VOCs reacted with these radicals and decomposed to CO2 and H2O at last [186,187]. In the water solution, water can react with the catalysts and form •OH, which is positive for the decomposition of VOCs in water. The photocatalytic elimination of VOCs in the gas phase follows a similar mechanism, but the radicals are main O2- due to the shortage of humidity. In this section, the progress in the elimination of VOCs in the gas phase by photocatalytic methods was mainly discussed.
The most studied photocatalyst for the elimination of VOCs is TiO2 [44]. Wilson F. Jariam et al. degraded 17 kinds of VOCs with the concentration range of 400–600 ppmv over TiO2 under the radiation of ultraviolet light [98]. The results showed that trichloroethylene (99.9%), isooctane (98.9%), acetone (98.5%), methanol (97.9%), methyl ethyl ketone (97.1%), t-butyl methyl ether (96.1%), dimethoxymethane (93.9%), methylene chloride (90.4%), methyl isopropyl ketone (88.5%), isopropanol (79.7%), chloroform (69.5%), and tetrachloroethylene (66.6%) were decomposed efficiently over TiO2. The photodegradation of isopropylbenlene (30.3%), methyl chloroform (20.5%), and pyridine (15.8%) on TiO2 was not as efficient as other VOCs. The catalytic lifetime was also tested by toluene. The conversion of toluene decreased to 20.9% after a 150 min test, but the deactivated catalysts can be easily regenerated by washing with H2O2 and illumination. F. B. Li et al. prepared La ion doped TiO2 by the sol-gel method for photodegradation of benzene, toluene, ethylbenzene, and o-xylene in the gas phase [188]. The results showed that the La ion doped TiO2 performed much better than the pure TiO2. This was due to the improved adsorption ability and the enhanced electron–hole pairs separation by the presence of Ti3+ and the electron transfer between the conduction band/defect level and lanthanide crystal field state. Tânia M. Fujimoto et al. supported the palladium on TiO2 for the photocatalytic decomposition of octane, isooctane, n-hexane, and cyclohexane in a low concentration (100~120 ppmv) [99]. The modified catalysts showed excellent catalytic activity in the decomposition of VOCs rapidly. V. Héquet et al. used a closed-loop reactor to study the mixture effect over the P25 TiO2/SiO2 mixture (Figure 14) [189]. They have developed the accurate analytical methods to identify and quantify the majority of the potential formed intermediates, which provide an efficient way to study the reaction mechanism. Yajie Shu et al. used Mn doped TiO2 to degrade benzene by O3 under vacuum ultraviolet (VUV) irradiation [62]. The doped TiO2 showed better performance than the undoped one and P25 due to the formation of highly reactive oxidizing species. Jian Ji et al. showed that compared with the one without UV radiation, the UV radiation can improve the removal efficiency of benzene by about 10% (Figure 15) [63]. Marta Stucchi et al. developed a simultaneous photodegradation system for the VOC mixture elimination by TiO2 powders, which showed a good efficiency [86]. Huiling Huang et al. developed Mn modified ZSM-5 as catalysts for VUV photolysis combined with ozone-assisted catalytic oxidation and studied the mechanism [102]. The catalysts showed good efficiency. Although most of the catalysts showed excellent activity in the elimination of VOCs under UV light, the light utilization efficiency was still low due to the low percent of the UV light in nature light. Xufang Qian et al. designed mesoporous TiO2 films coated with carbon foam for photodegradation of acetone and toluene, which can converse more than 90% of VOCs to CO2 under visible light due to the plausible carbon doping and the strong interaction between the TiO2 precursor and the hydro-carbon foams [100]. The graphite-SiO2-TiO2 composite and BiOBr@SiO2 flower-like nanospheres were also used for photodegradation of VOCs under visible light and showed good catalytic activity [64]. The low solar utilization ratio was the main obstacle to improve photocatalytic efficiency. In order to improve the efficiency, materials with a good light adsorption ability was applied. Yun-En Lee et al. prepared black-TiO2 and LFO/black-TiO2 and they showed excellent photo catalytic activity for the removal of toluene and IPA due to their good light adsorption ability [77]. The modification of g-C3N4 by hydroxyl groups can enhance visible light-driven photocatalytic properties of g-C3N4 obviously which can improve the adsorption energy of g-C3N4 for water and phenol [88]. The structure of catalysts was also carefully tuned to improve the catalytic activity [190]. Bettini S et al. insulated a layer of SiO2 between zinc oxide and nano silver and controlled the thickness of the insulated layer, which enhanced the photocatalytic oxidation ability of the catalyst significantly [78]. The photocatalytic activity of ZnO can also be improved by doping CuO [191]. The photocatalytic elimination of VOCs is one of the most promising methods. However, the industrial application of this method is still a problem.
There is a large amount of reports on the photocatalytic elimination of VOCs. The advantages of this process are green, energy saving, and clean. However, the reported catalysts are not efficient enough for an industrial application at the present stage. Firstly, the reaction needs more time than other methods. Secondly, most of the reported catalysts could not use visible light as energy to start the reactions, which limited the improvement of quantum efficiency, so it still needs to develop efficient photocatalysts and design effective reaction systems.

4.3. Non-Thermal Plasma Process

In the non-thermal plasma process, electrons and their surroundings are not in a thermal equilibrium, so the electrons are heated by electric discharges instead of the gas itself, and produce the electrons with high energy, active radicals and ions which can promote numerous chemical reactions in the ionized zones. It can be used to treat the high flow for both low (<100 ppmv) and high (>1000 ppmv) concentrations of VOCs, including toluene, benzene, acetone, trichloroethylene, etc. (Table 10) [47]. The main bottlenecks for the commercialization of the technology are the formation of poison byproducts and high energy consumption. The discharge methods are important for the VOCs removal efficiency, which includes corona discharge, surface discharge, microwave discharge, dielectric barrier discharge, and packed bed dielectric barrier discharge [190,191,192,193,194]. Among these discharge types, the packed bed dielectric barrier discharge shows the most potential in the industry application [195]. Savita K. P. Veerapandian et al. reviewed the packed bed DBD [47]. The influences of dielectric constant, packing materials size, shape, surface properties, and the byproducts formation were discussed.
Using porous and catalytic materials as the packed bed can increase the resistant time of VOCs and decrease the unwanted byproducts, such as O3, NOx, and CO (Figure 16) [196]. The packed bed materials can be non-catalytic porous materials, such as activated carbon, porous Al2O3, glass, zeolite, graphene oxide, and the catalysts, such as metal oxides, noble metal loaded metal oxides, and catalytic porous materials [197]. For example, using Al2O3 as the packed bed can concentrate VOCs molecules on its surface and weaken the bond energy of VOCs, which can enhance the dissociation when these adsorbed molecules encounter the active species in the plasma and increase the collision probability and deep oxidation of VOCs [198]. Lee et al. used porous γ-Al2O3 as the packed bed to oxidize toluene to CO2. It showed a 100% conversion of toluene and high CO2 selectivity [199]. Gandhi’s research results showed that a large surface area and pore volume of alumina can not only increase the conversion of ethylene to CO2 or CO, but also decrease the selectivity of unwanted byproducts, such as acetaldehyde, acetylene, methane, N2O, and O3 [200]. Other researches also showed the same phenomenon in the abatement of benzene, acetone, formaldehyde, TCE, chlorobenzene in non-thermal plasma with the packed bed which has a large surface area and pore volume, such as porous alumina, TiO2, zeolite, and porous metal oxides [199,201,202,203]. Due to the large surface area, the retention time and concentration of VOCs molecules increased which led to the increase of collisions between VOCs and active species [200], the adsorption effect can weaken the chemical bond of VOCs [204], and form micro-discharges in the micro-pores in addition to the micro-discharges in the gas phase.
The addition of catalysts in the non-thermal plasma process as the packed bed can enhance the VOCs removal efficiency significantly due to the plasma assisted reactions on the surface of the catalyst. These plasma-assisted reactions can generate some active radical species, such as intermediates with electrons, O·, OH·, N2·, NO·, N2O· in the gas phase and O·, OH· on the surface of catalysts, which can improve the VOCs removal efficiency and increase the COx selectivity [80]. The catalysts include metal oxides, such as CuO, MnO2, CeO2, NiO, CoO2, and Fe2O3, and noble metals, such as Pt, Au, and Ag [65,196,203,205,206,207,208,209]. The different kinds of oxygen species in catalysts, namely lattice oxygen, surface oxygen, and absorbed oxygen, are very helpful for the oxidation of VOCs, which can activate the oxygen molecules in the gas phase. P.J. Asilevi et al. established a laboratory scale DBD reactor for the removal of toluene by ·O and ·OH which were generated from the reaction between O2 and H2O and the removal efficiency of VOCs can be improved by increasing the oxygen concentration and relative humidity [210]. Lu et al. used FeOx/SBA-15 as the packed bed to eliminate toluene in the non-thermal plasma process. The results showed that the presence of Fe2+ can increase the toluene removal efficiency and COx selectivity obviously and reduce the formation of unwanted and toxic byproducts [80]. Zhu et al. prepared 5 wt% CuO/-Al2O3 as the packed bed to the abatement acetone. It showed better performance than the one without CuO due to the better reducibility and abundant active oxygen species [87]. Li et al. prepared Pt/Al2O3 and it showed higher decomposition efficiency than the unloaded Al2O3 due to the presence of Pt, which increases the number of active sites and reduces the activation energy of the decomposition reaction and suppresses harmful NOx formation [65]. The research by Hua Song et al. showed that CoMnOx/TiO2 can obviously improve the plasma removal of VOCs [103].
There are other factors that influence the VOCs removal efficiency, such as humidity and the plasma structure. In industrial conditions, the humidity in the VOCs stream is the factor which must be considered. The humidity has both a positive and negative effect on the removal of VOCs in the non-thermal plasma process. The water molecules can modify the surface of the packed bed, especially the catalytic packed bed, quench the free electrons and active species, and produce OH radicals during the discharge process [103,202,207,211,212]. A lot of studies showed that the presence of humidity can reduce the VOCs removal efficiency by quenching the high energy electrons and reactive species, and modify the catalysts surface (BaTiO3) in the decomposition of toluene, benzene, xylene, and C2F2 [103,207,212]. On the other hand, the humidity can suppress the formation of toxic byproducts, such as CO and NOx by deactivation of some oxygen species in the catalysts [213,214]. The optimization of the plasma structure also can enhance the VOCs decomposition efficiency. Muhammad Farooq Mustafa et al. designed a double dielectric barrier discharge reactor, in which the conversion of tetrachloroethylene, toluene, trichloroethylene, benzene, ethyl acetate, and carbon disulfide can be 100% with BaTiO3 and HZSM-5 as catalysts, respectively [215].
There are also some disadvantages in the non-thermal plasma process, such as high energy consumption, low VOCs conversion (mostly about 30% to 70%), low COx selectivity, and produced unwanted byproducts. The other barrier for the application of non-thermal plasma in the industry is the high cost and the high energy consumption. The way to solve these problems is the development of a highly efficient catalytic packed bed and the optimization of an electrode structure.

4.4. Electron Beam Treatment

In the electron beam technology, an electron beam generated from an EB accelerator and absorbed by the background gas to ionize the air and form some active oxidizing radicals, such as ·OH and ·HO2 and excited ions as O (3P), which is critical for the destruction of VOCs [216,217,218]. The oxidizing radicals reacted with VOCs and decomposed them to inorganic compounds. There are three types of accelerators used for VOCs decomposition, including transformer accelerator, ultrahigh frequency accelerator, and linear microwave accelerator [217]. This technology has been used for the elimination of acetaldehyde, benzene, dodecane, ethylbenzene, hexadecane, pentane, styrene, tetradecane, toluene, trimethylamine, turpentine, xylene, dimethyl sulfide, dimethyl disulfide, methyl mercaptan, and chlorinated VOCs in a lab or pilot scale [218,219,220,221,222,223,224,225,226,227,228]. The electron beam can treat the waste gas with a low concentration of VOCs (10 to hundreds ppm). The initial electrons from an electron beam can hardly decompose the VOCs molecules, while it can react with background gases and form active radicals. The way to improve the abatement rate of aromatic VOCs includes the addition of humidity (formation of OH·), ammonia, chlorinated VOCs, ozone, and using an absorbed dose [220,229]. The reaction temperature, VOCs molecular structure, and background gas also influence the decomposition efficiency. However, there are also some drawbacks of the electron beam technology, the generation of active oxidizing radicals, e.g., ·OH and ·HO2 can lead to the creation of toxic intermediates, especially in the case of the VOCs emission with an unknown composition, such as the production of toxic byproducts (aerosol, ozone, CO, and other trace organic compounds). The equipment is also too complex. It still needs more research for the industrial application.

5. Outlook of the Different Kinds of Technologies

Different kinds of VOCs were emitted from different parts of pharmaceutical production processes. With the new VOCs discharge standards coming into force, it is critical for the pharmaceutical companies to eliminate VOCs exhaustively in China. According to the production processes and conditions, the suitable VOCs elimination technologies must be selected with the highest efficiency and at the lowest cost. Adsorption, absorption, incineration, catalytic combustion, biodegradation, and non-thermal plasma technologies have been applied in factories for the abatement of VOCs. Adsorption and absorption are commonly the first used methods to recycle useful VOCs. Then, the rest of the VOCs were decomposed by other technologies, such as incineration, catalytic combustion, non-thermal plasma, biodegradation, and non-thermal plasma. The key point for the improvement of catalytic combustion and non-thermal plasma is the development of efficient catalysts or packed bed materials. Some new technologies are emerging for the abatement of VOCs, such as photocatalysis and electron beam radiation. These new technologies have shown good potential for the elimination of VOCs with a high efficiency and low energy cost. However, they still have some engineering problems for the industrial application. It is important to develop VOCs elimination technologies to decrease the amounts of VOCs in the discharged waste gas. The final solution for the VOCs is avoiding the emission of VOCs from the whole pharmaceutical production processes by developing new chemical and engineering technologies for the production of medicine.

Author Contributions

L.Z. has written the main content of this paper. C.M. has added some new research papers of this topic and summarized the catalysts used in different researches. J.H. has improved the English writing of the whole paper and written parts of photocatalysis parts. R.L. has written the parts of VOCs elimination policies in China and improve the structure of the paper. J.Y. has designed the whole structure of the paper and improved the manuscript a lot. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Scientific Technology Research Program for the University in Hebei Province Youth Project (No. QN2020126).

Acknowledgments

The authors gratefully acknowledge the help of Jing Wang and Ping Chen from Tianjushi Engineering&Technology Group Co., Ltd. who gave us a lot of suggestions about the VOCs treatment technologies and policies in China. The authors gratefully acknowledge the help of Rose Amal for the help of improvement the parts of catalytic processes for the treatment of VOCs.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, W.B.; Wang, J.X.; Gong, H. Catalytic combustion of VOCs on non-noble metal catalysts. Catal. Today 2009, 148, 81–87. [Google Scholar] [CrossRef]
  2. Muhammad, S.K.; Shaikh, A.R.; Mohammad, M.H. Catalytic oxidation of volatile organic compounds (VOCs): A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar]
  3. Cheng, G. Study on VOCs Emissions Research Present Situation and the Emission Reduction Potential of Key Industries in Hebei Province. Master’s Thesis, Hebei University of Science and Technology, Shijiangzhuang, China, 2016. [Google Scholar]
  4. Li, Y. Study on Emission Standard of Air Pollutants for Pharmaceutical Industry Chemical Synthesis Products Category. Master’s Thesis, Zhejiang University of Technology, Hangzhou, China, 2016. [Google Scholar]
  5. Atkinson, R. Gas-phase tropospheric chemistry of volatile organic compounds: 1. Alkanes and alkenes. J. Phys. Chem. Ref. Data 1997, 26, 215–229. [Google Scholar] [CrossRef] [Green Version]
  6. Fontane, H.; Veillerot, M.; Gallo, J.C.; Guillermo, R. 8th International Symposium on Transport and Air Pollution; Verlag der Technischen Universität Graz: Graz, Austria, 1999. [Google Scholar]
  7. Rivière, E. CITEPA Report; Fieldwork Inhabitants: Paris, France, 1998. [Google Scholar]
  8. Duncan, B.N.; Yoshida, Y.; Olson, J.R.; Sillman, S.; Martin, R.V.; Lamsal, L.; Hu, Y.; Pickering, K.E.; Retscher, C.; Allen, D.J.; et al. Application of OMI observations to a space-based indicator of NOx and VOC controls on surface ozone formation. Atmos. Environ. 2010, 44, 2213–2223. [Google Scholar] [CrossRef] [Green Version]
  9. Atkinson, R. Atmospheric chemistry of VOCs and NOx. Atmos. Environ. 2000, 34, 2063–2101. [Google Scholar]
  10. National Bureau of Statistics of the People’s Republic of China. Available online: http://www.stats.gov.cn/tjsj/ (accessed on 26 May 2020).
  11. Chen, Y. Study on Current and Future Industrial Emission of Volatile Organic Compounds in China. Master’s Thesis, South China University of Technology, Guangzhou, China, 2011. [Google Scholar]
  12. Huang, W. Characteristics of Industrial VOCs Emissions and Evaluation of Control Technology in China. Master’s Thesis, Zhejiang University, Hangzhou, China, 2016. [Google Scholar]
  13. Ministry of Ecological Environment of the People’s Republic of China. Standards for Emissions of Atmospheric Pollutants from Pharmaceutical Industry (Draft for Comments); Ministry of Ecological Environment of the People’s Republic of China: Beijing, China, 2017.
  14. Lu, Y. Establishment and Application on Assessment System of VOCs’s Control Technology of Pharmaceutical Industry. Master’s Thesis, Hebei University of Science and Technology, Shijiazhuang, China, 2016. [Google Scholar]
  15. Hu, G. Research of Emission and Control Technology of VOCs and Odorous Gas from Pharmacy Industry in China. Master’s Thesis, Nankai University, Tianjing, China, 2013. [Google Scholar]
  16. Li, X. The System of Discharge Standards of Pollutants for Pharmaceutical Industry and a case Study. Master’s Thesis, The Chinese Research Academy of Environmental Science, Beijing, China, 2006. [Google Scholar]
  17. He, J. Fermentology; China Medical Science Press: Beijing, China, 2009. [Google Scholar]
  18. Wang, Y. Chemical Pharmaceutical Technology; Chemical Industry Press: Beijing, China, 2008. [Google Scholar]
  19. Zhou, T.; Ma, Q.; Chen, H. Treatment of chemical synthesized pharmaceutical wastewater by hybrid biological reactor. Ind. Water Wastewater 2010, 41, 42–45. [Google Scholar]
  20. Chen, P. Pharmaceutical Technology; Hubei Science and Technology Press: Wuhan, China, 2008. [Google Scholar]
  21. Qi, X. Modern Biopharmaceutical Technology; Chemical Industry Press: Beijing, China, 2009. [Google Scholar]
  22. Cao, G. Pharmaceutical Engineering of Traditional Chinese Medicine; Chemical Industry Press: Beijing, China, 2004. [Google Scholar]
  23. Wu, C. The thinking and Countermeasures for the Structuring of Learning Seals Team in Chengdu Zhonghui Pharmaceuticals Company. Master’s Thesis, Southwestern University of Finance and Economics, Chengdu, China, 2006. [Google Scholar]
  24. Chen, Y.; Ye, D.; Liu, X.; Wu, J.; Huang, B.; Zheng, Y. Source tracing and characteristics of industrial VOCs emissions in China. Chin. Environ. Sci. 2012, 32, 48–55. [Google Scholar]
  25. Song, J. Studies on the Adsorption of VOCs by Activated Carbons and the Structure-Function Relationship. Ph.D. Thesis, Central South University, Changsha, China, 2014. [Google Scholar]
  26. Khan, F.I.; Ghoshal, A.K. Removal of Volatile Organic Compounds from polluted air. J. Loss Prevent. Proc. Ind. 2000, 13, 527–545. [Google Scholar] [CrossRef]
  27. Parmar, G.R.; Rao, N.N. Emerging Control Technologies for Volatile Organic Compounds. Crit. Rev. Environ. Sci. Technol. 2009, 39, 41–78. [Google Scholar] [CrossRef]
  28. Shah, R.; Thonon, B.; Benforado, D. Opportunities for heat exchanger applications in environmental systems. Appl. Therm. Eng. 2000, 20, 631–650. [Google Scholar] [CrossRef]
  29. Dunn, R.F.; El-Halwagi, M.M. Selection of Optimal VOC-condensation Systems. Waste Manag. 1994, 14, 103–113. [Google Scholar] [CrossRef]
  30. Huang, W.; Shi, L.; Hu, Z.; Zheng, Z. Integrated technology of condensation and adsorption for volatile organic compounds recovery. Chem. Eng. 2012, 6, 13–17. [Google Scholar]
  31. Zhang, X.; Gao, B.; Creamer, A.E.; Cao, C.; Li, Y. Adsorption of VOCs onto engineered carbon materials: A review. J. Hazard. Mater. 2017, 338, 102–123. [Google Scholar] [CrossRef]
  32. Wang, S.; Zhang, L.; Long, C.; Li, A. Enhanced adsorption and desorption of vocs vapor on novel micro-mesoporous polymeric adsorbents. J. Coll. Interface Sci. 2014, 428, 185–190. [Google Scholar] [CrossRef] [PubMed]
  33. Serna-Guerrero, R.; Sayari, A. Applications of pore-expanded mesoporoussilica 7, Adsorption of volatile organic compounds. Environ. Sci. Technol. 2007, 41, 4761–4766. [Google Scholar] [CrossRef] [PubMed]
  34. U.S. Environmental Protection Agency Clean Air Technology Center. Choosing an adsorption system for VOC: Carbon, zeolite, or polymers? In Proceedings of the SPIE—The International Society for Optical Engineering, Research Triangle Park, NC, USA, 1 May 1999. [Google Scholar]
  35. Kujawa, J.; Cerneaux, S.; Kujawski, W. Removal of hazardous volatile organic compounds from water by vacuum pervaporation with hydrophobic ceramic membranes. J. Membr. Sci. 2015, 474, 11–19. [Google Scholar] [CrossRef]
  36. Li, Y.X.; Chen, J.Y.; Sun, Y.H. Adsorption of multicomponent volatile organic compounds on semi-coke. Carbon 2008, 46, 858–863. [Google Scholar]
  37. Komori. Preparation of N-acetylmorpholine. J. Chem. Soc. Jpn. Ind. Chem. 1959, 62, 220–225. [Google Scholar]
  38. Chang, F.T.; Lin, Y.C.; Bai, H.; Pei, B.S. Adsorption and desorption characteristics of semiconductor volatile organic compounds on the thermal swing honeycomb zeolite concentrator. J. Air Waste Manage. Assoc. 2003, 53, 1384–1390. [Google Scholar] [CrossRef] [Green Version]
  39. Liu, Y. Study on Ceramic Honeycomb Monolithic Adsorbents for VOCs Adsorption. Master’s Thesis, South China University of Technology, Guangzhou, China, 2015. [Google Scholar]
  40. Moretti, E.C. Reduce VOC and HAP emissions. Chem. Eng. Prog. 2002, 98, 30–40. [Google Scholar]
  41. Everaert, K.; Baeyens, J. Catalytic combustion of volatile organic compounds. J. Hazard. Mater. 2004, B109, 113–139. [Google Scholar] [CrossRef]
  42. Liotta, L.F. Catalytic oxidation of volatile organic compounds on supported noble metals. Appl. Catal. B Environ. 2010, 100, 403–412. [Google Scholar] [CrossRef]
  43. Spivey, J.J. Complete Catalytic Oxidation of Volatile Organics. Ind. Eng. Chem. Res. 1987, 26, 2165–2180. [Google Scholar] [CrossRef]
  44. Zhang, Z.; Jiang, Z.; Shangguan, W. Low-temperature catalysis for VOCs removal in technology andapplication: A state-of-the-art review. Catal. Today 2016, 264, 270–278. [Google Scholar] [CrossRef]
  45. Barbusinski, K.; Kalemba, K.; Kasperczyk, D.; Urbaniec, K.; Kozik, V. Biological methods for odor treatment: A review. J. Clean. Prod. 2017, 152, 223–241. [Google Scholar] [CrossRef]
  46. Mudliar, S.; Giri, B.; Padoley, K.; Satpute, D.; Dixit, R.; Bhatt, P.; Pandey, R.; Juwarkar, A.; Vaidya, A. Bioreactors for treatment of VOCs and odours—A review. J. Environ. Manage. 2010, 91, 1039–1054. [Google Scholar] [CrossRef]
  47. Veerapandian, S.K.P.; Leys, C.; Geyter N., D.; Morent, R. Abatement of VOCs using packed bed non-thermal plasma reactors: A review. Catalysts 2017, 7, 113. [Google Scholar] [CrossRef]
  48. Son, Y.S. Decomposition of VOCs and odorous compounds by radiolysis: A critical review. Chem. Eng. J. 2017, 316, 609–622. [Google Scholar] [CrossRef]
  49. Wang, H. Study on Removal of Miced VOCs in Air by Dielectric Barrier Discharge. Ph.D. Thesis, Dalian University of Technology, Dalian, China, 2009. [Google Scholar]
  50. Hao, J. Control Study of VOCs Regioned Joint Prevention and Control of Atmospheric Pollution. Master’s Thesis, Hebei University of Technology, Shijiazhuang, China, 2012. [Google Scholar]
  51. Xi, J.; Wu, J.; Hu, H.; Wang, C. Application status of industrial VOCs gas treatment techniques. China Environ. Sci. 2012, 32, 1955–1960. [Google Scholar]
  52. Hao, Y. Study on The Pharmaceutical and Chemical Industry VOCs and Odor Pollution Characteristics. Master’s Thesis, Hebei University of Science and Technology, Shijiazhung, China, 2014. [Google Scholar]
  53. Department of Environmental Protection of Zhejiang Province. Standards for Emissions of Atmospheric Pollutants from Pharmaceutical Industry in Zhejiang Province; Department of Environmental Protection of Zhejiang Province: Hangzhou, Chain, 2015–2016.
  54. Salar-García, M.J.; Ortiz-Martínez, V.M.; Hernández-Fernández, F.J.; de los Ríos, A.P.; Quesada-Medina, J. Ionic liquid technology to recover volatile organic compounds (VOCs). J. Hazard. Mater. 2017, 321, 484–499. [Google Scholar] [CrossRef]
  55. Minella, M.; Baudino, M.; Minero, C. A revised photocatalytic transformation mechanism for chlorinated VOCs: Experimental evidence from C2Cl4 in the gas phase. Catal. Today 2018, 313, 114–121. [Google Scholar] [CrossRef]
  56. Papaefthimiou, P.; Ioannides, T.; Verykios, X.E. Combustion of non-halogenated volatile organic compounds over group VIII metal catalysts. Appl. Catal. B Environ. 1997, 13, 175–184. [Google Scholar] [CrossRef]
  57. Tabakova, T.; Ilieva, L.; Petrova, P. 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]
  58. Wang, Y.; Yang, D.; Li, S.; Zhang, L.; Zheng, G.; Guo, L. Layered copper manganese oxide for the efficient catalytic CO and VOCs oxidation. Chem. Eng. J. 2018, 357, 258–268. [Google Scholar] [CrossRef]
  59. Alifanti, M.; Florea, M.; Pârvulescu, V.I. Ceria-based oxides as supports for LaCoO3 perovskite catalysts for total oxidation of VOC. Appl. Catal. B Environ. 2007, 70, 400–405. [Google Scholar] [CrossRef]
  60. Spinicci, R.; Faticanti, M.; Marini, P.; De Rossi, S.; Porta, P. Catalytic activity of LaMnO3 and LaCoO3 perovskites towards VOCs combustion. J. Mol. Catal. A Chem. 2003, 197, 147–155. [Google Scholar] [CrossRef]
  61. Liu, L.; Li, J.; Zhang, H.; Li, L.; Zhou, P.; Meng, X.; Guo, M.; Jia, J.; Sun, T. In situ fabrication of highly active γ-MnO2/SmMnO3 catalyst for deep catalytic oxidation of gaseous benzene, ethylbenzene, toluene, and o-xylene. J. Hazard. Mater. 2019, 362, 178–186. [Google Scholar] [CrossRef]
  62. Shu, Y.; Ji, J.; Xu, Y.; Deng, J.; Huang, H.; He, M.; Leung, D.Y.C.; Wu, M.; Liu, S.; Liu, S.; et al. Promotional role of Mn doping on catalytic oxidation of VOCs over mesoporous TiO2 under vacuum ultraviolet (VUV) irradiation. Appl. Catal. B Environ. 2018, 220, 78–87. [Google Scholar] [CrossRef]
  63. Ji, J.; Xu, Y.; Huang, H.; He, M.; Liu, S.; Liu, G.; Xie, R.; Feng, Q.; Shu, Y.; Zhan, Y.; et al. Mesoporous TiO2 under VUV irradiation: Enhanced photocatalytic oxidation for VOCs degradation at room temperature. Chem. Eng. J. 2017, 327, 490–499. [Google Scholar] [CrossRef]
  64. Yadav, H.M.; Jung, S.C.; Kim, J.S. Visible light photocatalyticperformance of in situ synthesized graphite-SiO2–TiO2 composite towards degradation of benzene gas. J. Nanosci. Nanotechnol. 2018, 18, 2032–2036. [Google Scholar] [CrossRef]
  65. Li, J.; Han, S.; Bai, S.; Han, S.; Song, H.; Pu, Y.; Zhu, X.; Chen, W. Effect of Pt/gamma-Al2O3 catalyst on nonthermal plasma decomposition of benzene and byproducts. Environ. Eng. Sci. 2011, 28, 395–403. [Google Scholar] [CrossRef]
  66. Pei, W.; Liu, Y.; Deng, J.; Zhang, K.; Hou, Z.; Zhao, X.; Dai, H. Partially embedding Pt nanoparticles in the skeleton of 3DOM Mn2O3: An effective strategy for enhancing catalytic stability in toluene combustion. Appl. Catal. B Environ. 2019, 256, 117814–117824. [Google Scholar] [CrossRef]
  67. Hosseini, M.; Barakat, T.; Cousin, R. 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, 218–224. [Google Scholar] [CrossRef]
  68. Lee, D.S.; Chen, Y.W. The mutual promotional effect of Au–Pd/CeO2 bimetallic catalysts on destruction of toluene. J. Taiwan Ins. Chem. Eng. 2013, 44, 40–44. [Google Scholar] [CrossRef]
  69. Yang, X.; Yu, X.; Lin, M.; Ma, X.; Ge, M. Enhancement effect of acid treatment on Mn2O3 catalyst for toluene oxidation. Catal. Today 2019, 327, 254–261. [Google Scholar] [CrossRef]
  70. Huang, H.; Liu, Y.; Tang, W.; Chen, Y. Catalytic activity of nanometer La1-xSrxCoO3 (x = 0, 0.2) perovskites towards VOCs combustion. Catal. Commun. 2008, 9, 55–59. [Google Scholar] [CrossRef]
  71. Suárez-Vázquez, S.I.; Gil, S.; García-Vargas, J.M.; Cruz-López, A.; Giroir-Fendler, A. Catalytic oxidation of toluene by SrTi1-XBXO3 (B = Cu and Mn) with dendritic morphology synthesized by one pot hydrothermal route. Appl. Catal. B Environ. 2018, 223, 201–208. [Google Scholar] [CrossRef]
  72. Guo, M.; Li, K.; Liu, L.; Zhang, H.; Hu, X.; Min, X.; Jia, J.; Sun, T. Resource utilization of spent ternary lithium-ions batteries: Synthesis of highly active manganese-based perovskite catalyst for toluene oxidation. J. Taiwan Inst. Chem. Eng. 2019, 102, 268–275. [Google Scholar] [CrossRef]
  73. Yang, J.; Li, L.; Yang, X.; Song, S.; Li, J.; Jing, F.; Chu, W. Enhanced catalytic performances of in situ-assembled LaMnO /δ-MnO2, hetero-structures for toluene combustion. Catal. Today 2019, 327, 19–27. [Google Scholar] [CrossRef]
  74. Sihaib, Z.; Puleo, F.; Pantaleo, G.; Parola, V.L.; Valverde, J.L.; Gil, S.; Liotta, L.F.; Giroir-Fendler, A. The effect of citric acid concentration on the properties of LaMnO3 as a catalyst for hydrocarbon oxidation. Catalysts 2019, 9, 226. [Google Scholar] [CrossRef] [Green Version]
  75. Peng, Y.; Zhang, L.; Chen, L.; Yuan, D.; Wang, G.; Meng, X.; Xiao, F.S. Catalytic performance for toluene abatement over Al-rich Beta zeolite supported manganese oxides. Catal. Today 2017, 297, 182–187. [Google Scholar] [CrossRef]
  76. Qin, Y.; Qu, Z.; Dong, C.; Wang, Y.; Huang, N. Highly catalytic activity of Mn/SBA-15 catalysts for toluene combustion improved by adjusting the morphology of supports. J. Environ. Sci. 2019, 76, 208–216. [Google Scholar] [CrossRef]
  77. Lee, Y.E.; Chung, W.C.; Chang, M.B. Photocatalytic oxidation of toluene and isopropanol by LaFeO3/black-TiO2. Environ. Sci. Pollut. Res. 2019, 26, 20908–20919. [Google Scholar] [CrossRef]
  78. Bettini, S.; Pagano, R.; Semeraro, P.; Ottolini, M.; Salvatore, L.; Marzo, F.; Lovergine, N.; Giancane, G.; Valli, L. SiO2-Coated ZnO Nanoflakes Decorated with Ag Nanoparticles for Photocatalytic Water Oxidation. Chem. Eur. J. 2019, 25, 14123–14132. [Google Scholar] [CrossRef] [PubMed]
  79. Li, Z.; Pan, X.; Yi, Z. Photocatalytic oxidation of methane over CuO-decorated ZnO nanocatalysts. J. Mater. Chem. A 2019, 7, 469–475. [Google Scholar] [CrossRef]
  80. Lu, M.; Huang, R.; Wu, J.; Fu, M.; Chen, L.; Ye, D. On the performance and mechanisms of toluene removal by FeOx/SBA-15-assisted non-thermal plasma at atmospheric pressure and room temperature. Catal. Today 2015, 242, 274–286. [Google Scholar] [CrossRef]
  81. Nevanperä, T.K.; Ojala, S.; Laitinen, T.; Pitkäaho, S.; Saukko, S.; Keiski, R.L. Catalytic Oxidation of Dimethyl Disulfide over Bimetallic Cu–Au and Pt–Au Catalysts Supported on γ-Al2O3, CeO2, and CeO2–Al2O3. Catalysts 2019, 9, 603. [Google Scholar] [CrossRef] [Green Version]
  82. Rotter, H.; Landau, M.; Herskowitz, M. Combustion of chlorinated VOC on nanostructured chromia aerogel as catalyst and catalyst support. Environ. Sci. Technol. 2005, 39, 6845–6850. [Google Scholar] [CrossRef]
  83. Dai, Q.; Wang, X.; Lu, G. Low-temperature catalytic destruction of chlorinated VOCs over cerium oxide. Catal. Commun. 2007, 8, 1645–1649. [Google Scholar] [CrossRef]
  84. Ikhlaq, A.; Kasprzyk-Hordern, B. Catalytic ozonation of chlorinated VOCs on ZSM-5 zeolites and alumina: Formation of chlorides. Appl. Catal. B Environ. 2017, 200, 274–282. [Google Scholar] [CrossRef] [Green Version]
  85. Fei, Z.; Cheng, C.; Chen, H.; Li, L.; Yang, Y.; Liu, Q.; Chen, X.; Zhang, Z.; Tang, J.; Cui, M.; et al. Construction of uniform nanodots CeO2 stabilized by porous silica matrix for 1,2-dichloroethane catalytic combustion. Chem. Eng. J. 2019, 370, 916–924. [Google Scholar] [CrossRef]
  86. Stucchi, M.; Galli, F.; Bianchi, C.L.; Pirola, C.; Boffito, D.C.; Biasioli, F.; Capucci, V. Simultaneous photodegradation of VOC mixture by TiO2 powders. Chemosphere 2018, 193, 198–206. [Google Scholar] [CrossRef] [PubMed]
  87. Zhu, X.; Tu, X.; Mei, D.; Zheng, C.; Zhou, J.; Gao, X.; Luo, Z.; Ni, M.; Cen, K. Investigation of hybrid plasma-catalytic removal of acetone over CuO/-Al2O3 catalysts using response surface method. Chemosphere 2016, 155, 9–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Li, Z.; Meng, X.; Zhang, Z. Fabrication of surface hydroxyl modified g-C3N4 with enhanced photocatalytic oxidation activity. Catal. Sci. Technol. 2019, 9, 3979–3993. [Google Scholar] [CrossRef]
  89. Wu, Y.; Shi, S.; Yuan, S.; Yuan, S.; Bai, T.; Xing, S. Insight into the enhanced activity of Ag/NiOx-MnO2 for catalytic oxidation of o-xylene at low temperatures. Appl. Surf. Sci. 2019, 479, 1262–1269. [Google Scholar] [CrossRef]
  90. Yang, S.; Zhao, H.; Dong, F.; Zha, F.; Tang, Z. Highly efficient catalytic combustion of o-dichlorobenzene over three-dimensional ordered mesoporous cerium manganese bimetallic oxides: A new concept of chlorine removal mechanism. Mol. Catal. 2019, 463, 119–129. [Google Scholar] [CrossRef]
  91. He, D.; Zhao, Y.; Yang, S.; Mei, Y.; Yu, J.; Liu, J.; Chen, D.; He, S.; Luo, Y. Enhancement of catalytic performance and resistance to carbonaceous deposit of lanthanum (La) doped HZSM-5 catalysts for decomposition of methyl mercaptan. Chem. Eng. J. 2018, 336, 579–586. [Google Scholar] [CrossRef]
  92. Liu, N.; Shi, D.; Zhang, R.; Li, Y.; Chen, B. Highly selective catalytic combustion of acrylonitrile towards nitrogen over Cu-modified zeolites. Catal. Today 2018, 332, 201–213. [Google Scholar] [CrossRef]
  93. Fiorenza, R.; Crisafulli, C.; Condorelli, G.G.; Lupo, F.; Scire, S. Au–Ag/CeO2 and Au–Cu/CeO2 Catalysts for Volatile Organic Compounds Oxidation and CO Preferential Oxidation. Catal. Lett. 2015, 145, 1691–1702. [Google Scholar] [CrossRef]
  94. Yao, J.; Lu, H.; Xiao, Y.; Hou, B.; Li, D.; Jia, L. Sub-molten salt-acid treatment of LaCoO3 for a highly active catalyst towards propane combustion. Catal. Commun. 2019, 128, 10578–10583. [Google Scholar] [CrossRef]
  95. Zhang, K.; Peng, X.; Yang, H.; Wang, X.; Zhang, Y.; Zheng, Y.; Xiao, Y.; Jiang, L. Effect of MnO2 morphology on its catalytic performance in lean methane combustion. Mater. Res. Bull. 2018, 111, 338–341. [Google Scholar] [CrossRef]
  96. Wei, Y.; Ni, L.; Li, M.; Zhao, J. A template-free method for preparation of MnO2 catalysts with high surface areas. Catal. Today 2017, 297, 188–192. [Google Scholar] [CrossRef]
  97. Sinquin, G.; Petit, C.; Hindermann, J.P.; Kiennemann, A. Study of the formation of LaMO3 (M = Co, Mn) perovskites by propionates precursors: Application to the catalytic destruction of chlorinated VOCs. Catal. Today 2001, 70, 183–196. [Google Scholar] [CrossRef]
  98. Alberici, R.M.; Jardim, W.F. Photocatalytic destruction of VOCs in the gas-phase using titanium dioxide. Appl. Catal. B Environ. 1997, 14, 55–68. [Google Scholar] [CrossRef]
  99. Fujimoto, T.M.; Ponczek, M.; Rochetto, U.L.; Landers, R.; Tomaz, E. Photocatalytic oxidation of selected gas-phase VOCs using UV light, TiO2, and TiO2/Pd. Environ. Sci. Pollut. Res. 2017, 24, 6390–6396. [Google Scholar] [CrossRef]
  100. Qian, X.; Ren, M.; Yue, D.; Zhu, Y.; Han, Y.; Bian, Z.; Zhao, Y. Mesoporous TiO2 films coated on carbon foam based on waste polyurethane for enhanced photocatalytic oxidation of VOCs. Appl. Catal. B Environ. 2017, 212, 1–6. [Google Scholar] [CrossRef]
  101. Zhang, L.; Peng, Y.; Zhang, J.; Chen, L.; Meng, X.; Xiao, F.S. Adsorptive and catalytic properties in the removal of volatile organic compounds over zeolite-based materials. Chin. J. Catal. 2016, 37, 800–809. [Google Scholar] [CrossRef]
  102. Huang, H.; Huang, H.; Zhan, Y.; Liu, G.; Wang, X.; Lu, H.; Xiao, L.; Feng, Q.; Leung, D.Y.C. Efficient degradation of gaseous benzene by VUV photolysis combined with ozone-assisted catalytic oxidation: Performance and mechanism. Appl. Catal. B Environ. 2016, 186, 62–68. [Google Scholar] [CrossRef]
  103. Song, H.; Peng, Y.; Liu, S.; Bai, S.; Hong, X.; Li, J. The Roles of Various Plasma Active Species in Toluene Degradation by Non-thermal Plasma and Plasma Catalysis. Plasma Chem. Plasma Process. 2019, 39, 1469–1482. [Google Scholar] [CrossRef]
  104. Wang, L.; Wang, C.; Xie, H.; Zhan, W.; Guo, Y. Catalytic combustion of vinyl chloride over Sr doped LaMnO3. Catal. Today 2018, 327, 190–195. [Google Scholar] [CrossRef]
  105. Kołodziej, A.; Łojewska, J. Optimization of structured catalyst carriers for VOC combustion. Catal. Today 2005, 105, 378–384. [Google Scholar] [CrossRef]
  106. 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]
  107. Kim, S.C.; Shim, W.G. Properties and performance of Pd based catalysts for catalytic oxidation of volatile organic compounds. Appl. Catal. B Environ. 2009, 92, 429–436. [Google Scholar] [CrossRef]
  108. Bedi, J.; Rosas, J.M.; Rodríguez-Mirasol, J.; Cordero, T. Pd supported on mesoporous activated carbons with high oxidation resistance as catalysts for toluene oxidation. Appl. Catal. B Environ. 2010, 94, 8–18. [Google Scholar] [CrossRef]
  109. Usón, L.; Colmenares, M.G.; Hueso, J.L.; Sebastián, V.; Balas, F.; Arruebo, M.; Santamaría, J. VOCs abatement using thick eggshell Pt/SBA-15 pellets with hierarchical porosity. Catal. Today 2014, 227, 179–186. [Google Scholar] [CrossRef]
  110. Carrillo, A.M.; Carriazo, J.G. Cu and Co oxides supported on halloysite for the total oxidation of toluene. Appl. Catal. B Environ. 2015, 164, 443–452. [Google Scholar] [CrossRef]
  111. Kucherov, A.V.; Sinev, I.M.; Ojala, S.; Keiski, R.L.; Kustov, M. Adsorptive-catalytic removal of CH3OH, CH3SH, and CH3SSCH3 from air over the bifunctional system noble metals/HZSM-5. Stud. Surf. Sci. Catal. 2007, 170, 1129–1136. [Google Scholar]
  112. Wang, J. Study on Supported Ruthenium Catalysts for the Catalytic Oxidation of VOCs. Ph.D. Thesis, Institute of Process Engineering, CAS, China, 2016. [Google Scholar]
  113. Abdelouahab-Reddam, Z.; Mail, R.E.; Coloma, F.; Sepúlveda-Escribano, A. Platinum supported on highly-dispersed ceria on activated carbon for the total oxidation of VOCs. Appl. Catal. A Gen. 2015, 494, 87–94. [Google Scholar] [CrossRef] [Green Version]
  114. Piumetti, M.; Fino, D.; Russo, N. Mesoporous manganese oxides prepared by solution combustion synthesis as catalysts for the total oxidation of VOCs. Appl. Catal. B Environ. 2015, 163, 277–287. [Google Scholar] [CrossRef]
  115. Grbic, B.; Radic, N.; Terlecki-Baricevic, A. Kinetics of deep oxidation of n-hexane and toluene over Pt/Al2O3 catalysts: Oxidation of mixture. Appl. Catal. B Environ. 2004, 50, 161–166. [Google Scholar] [CrossRef]
  116. Patterson, M.J.; Angove, D.E.; Cant, N.W. The effect of carbon monoxide on the oxidation of four C6 to C8 hydrocarbons over platinum, palladium and rhodium. Appl. Catal. B Environ. 2000, 26, 47–57. [Google Scholar] [CrossRef]
  117. McCabe, R.W.; Mitchell, P.J. Exhaust-catalyst development for methanol-fueled vehicles: 1. A comparative study of methanol oxidation over alumina-supported catalysts containing group 9, 10, and 11 metals. Appl. Catal. 1986, 27, 83–98. [Google Scholar] [CrossRef]
  118. Liotta, L.F.; Ousmane, M.; Di Carlo, G.; Pantaleo, G.; Deganello, G.; Marcì, G.; Retailleau, L.; Giroir-Fendler, A. Total oxidation of propene at low temperature over Co3O4–CeO2 mixed oxides: Role of surface oxygen vacancies and bulk oxygen mobility in the catalytic activity. Appl. Catal. A Gen. 2008, 347, 81–88. [Google Scholar] [CrossRef] [Green Version]
  119. Liotta, L.F.; Ousmane, M.; Di Carlo, G.; Pantaleo, G.; Deganello, G.; Boreave, A.; Giroir-Fendler, A. Catalytic removal of toluene over Co3O4–CeO2 mixed oxide catalysts: Comparison with Pt/Al2O3. Catal. Lett. 2009, 127, 270–276. [Google Scholar] [CrossRef]
  120. Mitsui, T.; Tsutsui, K.; Matsui, T.; Kikuchi, R.; Eguchi, K. Catalytic abatement of acetaldehyde over oxide-supported precious metal catalysts. Appl. Catal. B Environ. 2008, 78, 158–165. [Google Scholar] [CrossRef]
  121. Zhang, C.; He, H. A comparative study of TiO2 supported noble metal catalysts for the oxidation of formaldehyde at room temperature. Catal. Today 2007, 126, 345–350. [Google Scholar] [CrossRef]
  122. Zhang, C.B.; He, H.; Tanaka, K. Catalytic performance and mechanism of a Pt/TiO2 catalyst for the oxidation of formaldehyde at room temperature. Appl. Catal. B Environ. 2006, 65, 37–43. [Google Scholar] [CrossRef]
  123. Agarwal, S.K.; Spivey, J.J.; Butt, J.B. Catalyst deactivation during deep oxidation of chlorohydrocarbons. Appl. Catal. A Gen. 1992, 82, 259–275. [Google Scholar] [CrossRef]
  124. Petrosius, S.C.; Drago, R.S.; Young, V.; Grunewald, G.C. Low-temperature decomposition of some halogenated hydrocarbons using metal oxide/porous carbon catalysts. J. Am. Chem. Soc. 1993, 115, 6131–6137. [Google Scholar] [CrossRef]
  125. Sedjame, H.J.; Fontaine, C.; Lafaye, G.; Barbier, J.J. On the promoting effect of the addition of ceria to platinum based alumina catalysts for VOCs oxidation. Appl. Catal. B Environ. 2014, 144, 233–242. [Google Scholar] [CrossRef]
  126. Carabineiro, S.; Chen, X.; Konsolakis, M.; Psarras, A.; Tavares, P.; Orf~ao, J.; Pereira, M.; Figueiredo, J. 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] [Green Version]
  127. Carabineiro, S.; Chen, X.; Martynyuk, O.; Bogdanchikova, N.; Avalos-Borja, M.; Pestryakov, A.; Tavares, P.; Orf~ao, J.; Pereira, M.; Figueiredo, J. Gold supported on metal oxides for volatile organic compounds total oxidation. Catal. Today 2015, 244, 103–114. [Google Scholar] [CrossRef]
  128. Castano, M.H.; Molina, R.; Moreno, S. Catalytic oxidation of VOCs on MnMgAlOx mixed oxides obtained by auto-combustion. J. Mol. Catal. A Chem. 2015, 398, 358–367. [Google Scholar] [CrossRef]
  129. Chen, H.; Zhang, H.; Yan, Y. Fabrication of porous copper/manganese binary oxides modified ZSM-5 membrane catalyst and potential application in the removal of VOCs. Chem. Eng. J. 2014, 254, 133–142. [Google Scholar] [CrossRef]
  130. 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]
  131. Liu, Q.; Wang, L.C.; Chen, M.; Cao, Y.; He, H.Y.; Fan, K.N. Dry citrate precursor synthesized nanocrystalline cobalt oxide as highly active catalyst for total oxidation of propane. J. Catal. 2009, 263, 104–113. [Google Scholar] [CrossRef]
  132. Tseng, T.K.; Wang, L.; Ho, C.T.; Chu, H. The destruction of dichloroethane over a g-alumina supported manganese oxide catalyst. J. Hazard. Mater. 2010, 178, 1035–1040. [Google Scholar] [CrossRef]
  133. Krishnamoorthy, S.; Rivas, J.A.; Amiridis, M.D. Catalytic oxidation of 1, 2-dichlorobenzene over supported transition metal oxides. J. Catal. 2000, 193, 264–272. [Google Scholar] [CrossRef]
  134. Lahousse, C.; Bernier, A.; Grange, P.; Delmon, B.; Papaefthimiou, P.; Ioannides, T.; Verykios, X. Evaluation of g-MnO2 as a VOC removal catalyst: Comparison with a noble metal catalyst. J. Catal. 1998, 178, 214–225. [Google Scholar] [CrossRef]
  135. Parida, K.; Samal, A. Catalytic combustion of volatile organic compounds on Indian Ocean manganese nodules. Appl. Catal. A Gen. 1999, 182, 249–256. [Google Scholar] [CrossRef]
  136. Luo, J.; Zhang, Q.; Huang, A.; Suib, S.L. Total oxidation of volatile organic compounds with hydrophobic cryptomelane-type octahedral molecular sieves. Microporous Mesoporous Mater. 2000, 35–36, 209–217. [Google Scholar] [CrossRef]
  137. Aguero, F.N.; Scian, A.; Barbero, B.P.; Cadús, L.E. Influence of the support treatment on the behavior of MnOx/Al2O3 catalysts used in VOC combustion. Catal. Lett. 2009, 128, 268–280. [Google Scholar] [CrossRef]
  138. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Manganese oxide catalysts for NOx reduction with NH3 at low temperatures. Appl. Catal. A Gen. 2007, 327, 261–269. [Google Scholar] [CrossRef]
  139. Miranda, B.; Díaz, E.; Ordonez, S.; Vega, A.; Díez, F.V. Oxidation of trichloroethene over metal oxide catalysts: Kinetic studies and correlation with adsorption properties. Chemosphere 2007, 66, 1706–1715. [Google Scholar] [CrossRef]
  140. Sun, H.; Liu, Z.; Chen, S.; Quan, X. The role of lattice oxygen on the activity and selectivity of the OMS-2 catalyst for the total oxidation of toluene. Chem. Eng. J. 2015, 270, 58–65. [Google Scholar] [CrossRef]
  141. Cordi, E.M.; O’Neill, P.J.; Falconer, J.L. Transient oxidation of volatile organic compounds on a CuO/Al2O3 catalyst. Appl. Catal. B Environ. 1997, 14, 23–36. [Google Scholar] [CrossRef]
  142. Hutchings, G.J.; Taylor, S.H. Designing oxidation catalysts. Catal. Today 1999, 49, 105–113. [Google Scholar] [CrossRef]
  143. Heynderickx, P.M.; Thybaut, J.W.; Poelman, H.; Poelman, D.; Marin, G.B. The total oxidation of propane over supported Cu and Ce oxides: A comparison of single and binary metal oxides. J. Catal. 2010, 272, 109–120. [Google Scholar] [CrossRef]
  144. Sinha, A.K.; Suzuki, K. Novel mesoporous chromium oxide for VOCs elimination. Appl. Catal. B Environ. 2007, 70, 417–422. [Google Scholar] [CrossRef]
  145. Padilla, A.M.; Corella, J.; Toledo, J.M. Total oxidation of some chlorinated hydrocarbons with commercial chromia based catalysts. Appl. Catal. B Environ. 1999, 22, 107–121. [Google Scholar] [CrossRef]
  146. Yim, S.D.; Chang, K.-H.; Nam, I.S. Deactivation of chromium oxide catalyst for the removal of perchloroethylene (PCE). Stud. Surf. Sci. Catal. 2001, 139, 173–180. [Google Scholar]
  147. Gorte, R.J. Ceria in catalysis: From automotive applications to the water-gas shift reaction. AlChE J. 2010, 56, 1126–1135. [Google Scholar] [CrossRef]
  148. Zimmer, P.; Tschope, A.; Birringer, R. Temperature-programmed reaction spectroscopy of ceria-and Cu/ceria-supported oxide catalyst. J. Catal. 2002, 205, 339–345. [Google Scholar] [CrossRef]
  149. Huang, H.; Xu, Y.; Feng, Q.; Leung, D.Y.C. Low temperature catalytic oxidation of volatile organic compounds: A review. Catal. Sci. Technol. 2015, 5, 2649–2669. [Google Scholar] [CrossRef]
  150. Li, H.; Lu, G.; Dai, Q.; Wang, Y.; Guo, Y.; Guo, Y. Hierarchical organization and catalytic activity of high-surface-area mesoporous ceria microspheres prepared via hydrothermal routes. ACS Appl. Mater. Interfaces 2010, 2, 838–846. [Google Scholar] [CrossRef]
  151. Jones, J.; Ross, J.R. The development of supported vanadia catalysts for the combined catalytic removal of the oxides of nitrogen and of chlorinated hydrocarbons from flue gases. Catal. Today 1997, 35, 97–105. [Google Scholar] [CrossRef]
  152. Delaigle, R.; Debecker, D.P.; Bertinchamps, F.; Gaigneaux, E.M. Revisiting the behaviour of vanadia-based catalysts in the abatement of (chloro)-aromatic pollutants: Towards an integrated understanding. Top. Catal. 2009, 52, 501–516. [Google Scholar] [CrossRef]
  153. Solsona, B.; Garcia, T.; Aylon, E.; Dejoz, A.M.; Vazquez, I.; Agouram, S.; Davies, T.E.; Taylor, S.H. Promoting the activity and selectivity of high surface area Ni-Ce-O mixed oxides by gold deposition for VOC catalytic combustion. Chem. Eng. J. 2011, 175, 271–278. [Google Scholar] [CrossRef]
  154. Delimaris, D.; Ioannides, T. VOC oxidation over MnOx-CeO2 catalysts prepared by a combustion method. Appl. Catal. B Environ. 2008, 84, 303–312. [Google Scholar] [CrossRef]
  155. Morales, M.R.; Barbero, B.P.; Cadús, L.E. Total oxidation of ethanol and propane over Mn-Cu mixed oxide catalysts. Appl. Catal. B Environ. 2006, 67, 229–236. [Google Scholar] [CrossRef]
  156. Vasile, A.; Bratan, V.; Hornoiu, C.; Caldararu, M.; Ionescu, N.I.; Yuzhakova, T.; Redey, A. Electrical and catalytic properties of cerium etin mixed oxides in CO depollution reaction. Appl. Catal. B Environ. 2013, 140, 25–31. [Google Scholar] [CrossRef]
  157. 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]
  158. Larsson, P.O.; Andersson, A. Complete oxidation of CO, ethanol, and ethyl acetate over copper oxide supported on titania and ceria modified titania. J. Catal. 1998, 179, 72–89. [Google Scholar] [CrossRef]
  159. Hu, C.; Zhu, Q.; Jiang, Z.; Zhang, Y.; Wang, Y. Preparation and formation mechanism of mesoporous CuOeCeO2 mixed oxides with excellent catalytic performance for removal of VOCs. Microporous Mesoporous Mater. 2008, 113, 427–434. [Google Scholar] [CrossRef]
  160. Delimaris, D.; Ioannides, T. VOC oxidation over CuO-CeO2 catalysts prepared by a combustion method. Appl. Catal. B Environ. 2009, 89, 295–302. [Google Scholar] [CrossRef]
  161. Rao, T.; Shen, M.; Jia, L.; Hao, J.; Wang, J. Oxidation of ethanol over Mn-Ce-O and Mn-Ce-Zr-O complex compounds synthesized by solegel method. Catal. Commun. 2007, 8, 1743–1747. [Google Scholar] [CrossRef]
  162. Tang, X.; Li, Y.; Huang, X.; Xu, Y.; Zhu, H.; Wang, J.; Shen, W. MnOx-CeO2 mixed oxide catalysts for complete oxidation of formaldehyde: Effect of preparation method and calcination temperature. Appl. Catal. B Environ. 2006, 62, 265–273. [Google Scholar] [CrossRef]
  163. Picasso, G.; Gutierrez, M.; Pina, M.; Herguido, J. Preparation and characterization of Ce-Zr and Ce-Mn based oxides for n-hexane combustion: Application to catalytic membrane reactors. Chem. Eng. J. 2007, 126, 119–130. [Google Scholar] [CrossRef]
  164. Chen, H.; Sayari, A.; Adnot, A.; Larachi, F. Composition activity effects of Mn-Ce-O composites on phenol catalytic wet oxidation. Appl. Catal. B Environ. 2001, 32, 195–204. [Google Scholar] [CrossRef]
  165. Yang, P.; Yang, S.; Shi, Z.; Meng, Z.; Zhou, R. Deep oxidation of chlorinated VOCs over CeO2-based transition metal mixed oxide catalysts. Appl. Catal. B Environ. 2015, 162, 227–235. [Google Scholar] [CrossRef]
  166. 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]
  167. Rao, G.R.; Sahu, H.R.; Mishra, B.G. Surface and catalytic properties of Cu-Ce-O composite oxides prepared by combustion method. Coll. Surf. A Physicochem. Eng. Asp. 2003, 220, 261–269. [Google Scholar] [CrossRef]
  168. Jiang, X.; Lu, G.; Zhou, R.; Mao, J.; Chen, Y.; Zheng, X. Studies of pore structure, temperature-programmed reduction performance, and micro-structure of CuO/CeO2 catalysts. Appl. Surf. Sci. 2001, 173, 208–220. [Google Scholar]
  169. Zheng, X.C.; Wu, S.H.; Wang, S.P.; Wang, S.R.; Zhang, S.M.; Huang, W.P. The preparation and catalytic behavior of copperecerium oxide catalysts for low-temperature carbon monoxide oxidation. Appl. Catal. A Gen. 2005, 283, 217–223. [Google Scholar] [CrossRef]
  170. Pecchi, G.; Reyes, P.; Zamora, R.; Cadus, L.E.; Fierro, J.L.G. Surface properties and performance for VOCs combustion of LaFe1−yNiyO3 perovskite oxides. J. Solid State Chem. 2008, 181, 905–912. [Google Scholar] [CrossRef]
  171. Wei, T. Study on Catalytic Combustion of VOCs over Perovskite Catalysts. Master’s Thesis, Zhejiang University of Technology, Zhejiang, China, 2005. [Google Scholar]
  172. Beauchet, R.; Magnoux, P.; Mijoin, J. Catalytic oxidation of volatile organic compounds (VOCs) mixture (isopropanol/o-xylene) on zeolite catalysts. Catal. Today 2007, 124, 118–123. [Google Scholar] [CrossRef]
  173. Rachapudi, R.; Chintawar, P.S.; Greene, H.L. Aging and Structure/Activity Characteristics of CR–ZSM-5 Catalysts during Exposure to Chlorinated VOCs. J. Catal. 1999, 185, 58–72. [Google Scholar] [CrossRef]
  174. Muniandy, L.; Adam, F.; Mohamed, A.R.; Iqbal, A.; Rahman, N.R.A. Cu2+ coordinated graphitic carbon nitride (Cu-g-C3N4) nanosheets from melamine for the liquid phase hydroxylation of benzene and VOCs. Appl. Surf. Sci. 2017, 398, 43–55. [Google Scholar] [CrossRef]
  175. Blanch-Raga, N.; Palomares, A.E.; Martínez-Triguero, J.; Valencia, S. Cu and Co modified beta zeolite catalysts for the trichloroethylene oxidation. Appl. Catal. B Environ. 2016, 187, 90–97. [Google Scholar] [CrossRef]
  176. Lopez-Fonseca, R.; Aranzabal, A.; Steltenpohl, P.; Gutierrez-Ortiz, J.I.; Gonzalez Velasco, J.R. Performance of zeolites and product selectivity in the gas-phase oxidation of 1,2-dichloroethane. Catal. Today 2000, 62, 367–377. [Google Scholar] [CrossRef]
  177. Kullavanijayam, E.; Trimm, D.L.; Cant, N.W. Adsocat: Adsorption/catalytic combustion for VOC and odour control. Stud. Surf. Sci. Catal. 2000, 130, 569–574. [Google Scholar]
  178. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductorelectrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  179. Wang, D.; Hou, P.; Yang, P.; Cheng, X. BiOBr@SiO2 flower-like nanospheres chemically-bonded on cement-based materials for photocatalysis. Appl. Surf. Sci. 2018, 430, 539–548. [Google Scholar] [CrossRef]
  180. Li, Y.; Wu, X.; Li, J.; Wang, K.; Zhang, G. Z-scheme g-C3N4@CsxWO3 heterostructure as smart window coating for UV isolating, Vis penetrating, NIR shielding and full spectrum photocatalytic decomposing VOCs. Appl. Catal. B Environ. 2018, 229, 218–226. [Google Scholar] [CrossRef]
  181. Mishra, A.; Mehta, A.; Kainth, S.; Basu, S. Effect of different plasmonic metals on photocatalytic degradation of volatile organic compounds (VOCs) by bentonite/M-TiO2 nanocomposites, under UV/visible light. Appl. Clay Sci. 2018, 153, 144–153. [Google Scholar] [CrossRef]
  182. Song, S.; Lu, C.; Wu, X.; Jiang, S.; Sun, C.; Le, Z. Strong base g-C3N4 with perfect structure for photocatalytically eliminating formaldehyde under visible-light irradiation. Appl. Catal. B Environ. 2018, 227, 145–152. [Google Scholar] [CrossRef]
  183. Bhatkhande, D.S.; Pangarkar, V.G.; Beenackers, A.A.C.M. Photocatalytic degradation for environmental applications—A review. J. Chem. Technol. Biotechnol. Int. Res. Process Environ. Clean Technol. 2001, 77, 107–116. [Google Scholar] [CrossRef]
  184. Choi, H.; Stathatos, E.; Dionysiou, D.D. Sol–gel preparation of mesoporous photocatalytic TiO2 films and TiO2/Al2O3 composite membranes for environmental applications. Appl. Catal. B Environ. 2006, 63, 60–67. [Google Scholar] [CrossRef]
  185. Fujishima, A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  186. Minero, C. Surface modified photocatalysts, in environmental photochemistry part III. In The Handbook of Environmental Chemistry; Bahnemann, D.W., Robertson, K.J., Eds.; Springer: Berlin/Heidelberg, Germany, 2015; Volume 35, pp. 23–44. [Google Scholar]
  187. Tejasvi, R.; Sharma, M.; Upadhyay, K. Passive photo-catalytic destruction of air-borne VOCs in high traffic areas using TiO2-coated flexible PVC sheet. Chem. Eng. J. 2015, 262, 875–881. [Google Scholar] [CrossRef]
  188. Li, F.B.; Li, X.Z.; Ao, C.H.; Lee, S.C.; Hou, M.F. Enhanced photocatalytic degradation of VOCs using Ln3+–TiO2 catalysts for indoor air purification. Chemosphere 2005, 59, 787–800. [Google Scholar] [CrossRef] [Green Version]
  189. Héqueta, V.; Raillarda, C.; Debonoa, O.; Thévenet, F.; Locoge, N.; Le Coq, L. Photocatalytic oxidation of VOCs at ppb level using a closed-loop reactor: The mixture effect. Appl. Catal. B Environ. 2018, 226, 473–486. [Google Scholar] [CrossRef] [Green Version]
  190. Huang, L.; Nakajo, K.; Ozawa, S.; Matsuda, H. Decomposition of dichloromethane in a wire-in-tube pulsed corona reactor. Environ. Sci. Technol. 2001, 35, 1276–1281. [Google Scholar] [CrossRef]
  191. Norberg, A. Modeling current pulse shape and energy in surface discharges. IEEE Trans. Ind. Appl. 1992, 28, 498–503. [Google Scholar] [CrossRef]
  192. Rousseau, A.; Dantier, A.; Gatilova, L.; Ionikh, Y.; Röpcke, J.; Tolmachev, Y. On NOx production and volatile organic compound removal in a pulsed microwave discharge in air. Plasma Sour. Sci. Technol. 2005, 14, 70–75. [Google Scholar] [CrossRef]
  193. Kogelschatz, U. Dielectric-barrier discharges: Their history, discharge physics, and industrial applications. Plasma Chem. Plasma Process. 2003, 23, 1–46. [Google Scholar] [CrossRef]
  194. Yamamoto, T.; Ramanathan, K.; Lawless, P.A.; Ensor, D.S.; Newsome, J.R.; Plaks, N.; Ramsey, G.H. Control of volatile organic compounds by an ac energized ferroelectric pellet reactor and a pulsed corona reactor. IEEE Trans. Ind. Appl. 1992, 28, 528–534. [Google Scholar] [CrossRef]
  195. Ding, H.-X.; Zhu, A.-M.; Yang, X.-F.; Li, C.-H.; Xu, Y. Removal of formaldehyde from gas streams via packed-bed dielectric barrier discharge plasmas. J. Phys. D App. Phys. 2005, 38, 4160–4167. [Google Scholar] [CrossRef]
  196. Kim, H.H.; Ogata, A.; Futamura, S. Oxygen partial pressure-dependent behavior of various catalysts for the total oxidation of VOCs using cycled system of adsorption and oxygen plasma. Appl. Catal. B Environ. 2008, 79, 356–367. [Google Scholar] [CrossRef]
  197. Futamura, S.; Yamamoto, T.; Lawless, P.A. Towards understanding of VOC decomposition mechanisms using nonthermal plasmas. In Proceedings of the 1995 Thirtieth IAS Annual Meeting IEEE Conference Record of Industry Applications Conference, Orlando, FL, USA, 8–12 October 1995; Volume 2, pp. 1453–1458. [Google Scholar]
  198. Zheng, C.; Zhu, X.; Gao, X.; Liu, L.; Chang, Q.; Luo, Z.; Cen, K. Experimental study of acetone removal by packed-bed dielectric barrier discharge reactor. J. Ind. Eng. Chem. 2014, 20, 2761–2768. [Google Scholar] [CrossRef]
  199. Lee, B.Y.; Park, S.H.; Lee, S.C.; Kang, M.; Choung, S.J. Decomposition of benzene by using a discharge plasma-photocatalyst hybrid system. Catal. Today 2004, 93–95, 769–776. [Google Scholar] [CrossRef]
  200. Gandhi, M.S.; Ananth, A.; Mok, Y.S.; Song, J.I.; Park, K.H. Effect of porosity of α-alumina on non-thermal plasma decomposition of ethylene in a dielectric-packed bed reactor. Res. Chem. Intermed. 2014, 40, 1483–1493. [Google Scholar] [CrossRef]
  201. Hu, J.; Jiang, N.; Li, J.; Shang, K.; Lu, N.; Wu, Y. Degradation of benzene by bipolar pulsed series surface/packed-bed discharge reactor over MnO2-TiO2/zeolite catalyst. Chem. Eng. J. 2016, 293, 216–224. [Google Scholar] [CrossRef]
  202. Ogata, A.; Yamanouchi, K.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Oxidation of dilute benzene in an alumina hybrid plasma reactor at atmospheric pressure. Plasma Chem. Plasma Process. 1999, 19, 383–394. [Google Scholar] [CrossRef]
  203. Zhu, X.; Gao, X.; Qin, R.; Zeng, Y.; Qu, R.; Zheng, C.; Tu, X. Plasma-catalytic removal of formaldehyde over Cu–Ce catalysts in a dielectric barrier discharge reactor. Appl. Catal. B Environ. 2015, 170–171, 293–300. [Google Scholar] [CrossRef] [Green Version]
  204. Bradford, M.C.J.; Vannice, M.A. Estimation of CO heats of adsorption on metal surfaces from vibrational spectra. Ind. Eng. Chem. Res. 1996, 35, 3171–3178. [Google Scholar] [CrossRef]
  205. An, H.T.Q.; Huu, T.P.; Le Van, T.; Cormier, J.M.; Khacef, A. Application of atmospheric non thermal plasma-catalysis hybrid system for air pollution control: Toluene removal. Catal. Today 2011, 176, 474–477. [Google Scholar]
  206. Futamura, S.; Einaga, H.; Kabashima, H.; Hwan, L.Y. Synergistic effect of silent discharge plasma and catalysts on benzene decomposition. Catal. Today 2004, 89, 89–95. [Google Scholar] [CrossRef]
  207. Wu, J.; Huang, Y.; Xia, Q.; Li, Z. Decomposition of toluene in a plasma catalysis system with NiO, MnO2, CeO2, Fe2O3, and CuO catalysts. Plasma Chem. Plasma Process. 2013, 33, 1073–1082. [Google Scholar] [CrossRef]
  208. Zhao, D.-Z.; Li, X.-S.; Shi, C.; Fan, H.-Y.; Zhu, A.-M. Low-concentration formaldehyde removal from air using a cycled storage-discharge (CSD) plasma catalytic process. Chem. Eng. Sci. 2011, 66, 3922–3929. [Google Scholar] [CrossRef]
  209. Ran, L.; Wang, Z.; Wang, X. The effect of Ce on catalytic decomposition of chlorinated methane over RuOx catalysts. Appl. Catal. A gen. 2014, 470, 442–450. [Google Scholar] [CrossRef]
  210. Asilevi, P.J.; Yi, C.W.; Li, J.; Nawaz, M.I.; Wang, H.J.; Yin, L.; Junli, Z. Decomposition of formaldehyde in strong ionization non-thermal plasma at atmospheric pressure. Int. J. Environ. Sci. Technol. 2020, 17, 765–776. [Google Scholar] [CrossRef]
  211. Zhu, T.; Li, J.; Jin, Y.; Liang, Y.; Ma, G. Decomposition of benzene by non-thermal plasma processing: Photocatalyst and ozone effect. Int. J. Environ. Sci. Technol. 2008, 5, 375–384. [Google Scholar] [CrossRef] [Green Version]
  212. Urashima, K.; Kostov, K.G.; Chang, J.S.; Okayasu, Y.; Iwaizumi, T.; Yoshimura, K.; Kato, T. Removal of C2F6 from a semiconductor process flue gas by a ferroelectric packed-bed barrier discharge reactor with an adsorber. IEEE Trans. Ind. Appl. 2001, 37, 1456–1463. [Google Scholar] [CrossRef]
  213. Ogata, A.; Ito, D.; Mizuno, K.; Kushiyama, S.; Gal, A.; Yamamoto, T. Effect of coexisting components on aromatic decomposition in a packed-bed plasma reactor. Appl. Catal. A Gen. 2002, 236, 9–15. [Google Scholar] [CrossRef]
  214. Ogata, A.; Shintani, N.; Yamanouchi, K.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Effect of water vapor on benzene decomposition using a nonthermal-discharge plasma reactor. Plasma Chem. Plasma Process. 2000, 20, 453–467. [Google Scholar] [CrossRef]
  215. Mustafa, M.F.; Fu, X.; Liu, Y.; Abbas, Y.; Wang, H.; Lu, W. Volatile organic compounds (VOCs) removal in non-thermal plasma double dielectric barrier discharge reactor. J. Hazard. Mater. 2018, 347, 317–324. [Google Scholar] [CrossRef] [PubMed]
  216. Hirota, K.; Sakai, H.; Washio, M.; Kojima, T. Application of electron beams for the treatment of VOC streams. Ind. Eng. Chem. Res. 2004, 43, 1185–1191. [Google Scholar] [CrossRef]
  217. Zimek, Z. High power accelerators and processing systems for environmental application. Radiation treatment of gaseous and liquid effluents for contaminant removal. IAEA-TECDOC 2005, 1473, 125–137. [Google Scholar]
  218. Son, Y.S.; Kim, J.; Kim, J.C. Decomposition of acetaldehyde using an electron beam. Plasma Chem. Plasma Process. 2014, 34, 1233–1245. [Google Scholar] [CrossRef]
  219. Son, Y.S.; Kim, K.H.; Kim, K.J.; Kim, J.C. Ammonia Decomposition Using Electron Beam. Plasma Chem. Plasma Process. 2013, 33, 617–629. [Google Scholar] [CrossRef]
  220. Han, D.H.; Stuchinskaya, T.; Won, Y.S.; Park, W.S.; Lim, J.K. Oxidative decomposition of aromatic hydrocarbons by electron beam irradiation. Radiat. Phys. Chem. 2003, 67, 51–60. [Google Scholar] [CrossRef]
  221. Hashimoto, S.; Hakoda, T.; Hirata, K.; Arai, H. Low energy electron beam treatment of VOCs. Radiat. Phys. Chem. 2000, 57, 485–488. [Google Scholar] [CrossRef]
  222. Sun, Y.; Chmielewski, A.G.; Licki, J.; Bułka, S.; Zimek, Z. Decomposition of organic compounds in simulated industrial off-gas by using electron beam irradiation. Radiat. Phys. Chem. 2009, 78, 721–723. [Google Scholar] [CrossRef]
  223. Son, Y.S.; Kim, P.; Park, J.H.; Kim, J.; Kim, J.C. Decomposition of trimethylamine by an electron beam. Plasma Chem. Plasma Process. 2013, 33, 1099–1109. [Google Scholar] [CrossRef]
  224. Son, Y.S.; Park, J.H.; Kim, P.; Kim, J.C. Oxidation of gaseous styrene by electron beam irradiation. Radiat. Phys. Chem. 2012, 81, 686–692. [Google Scholar] [CrossRef]
  225. Auslender, V.L.; Ryazantsev, A.A.; Spiridonov, G.A. The use of electron beam for solution of some ecological problems in pulp and paper industry. Radiat. Phys. Chem. 2002, 63, 641–645. [Google Scholar] [CrossRef]
  226. Son, Y.S.; Kim, J.C. Decomposition of sulfur compounds by radiolysis: I. Influential factors. Chem. Eng. J. 2015, 262, 217–223. [Google Scholar] [CrossRef]
  227. Son, Y.S.; Jung, I.H.; Lee, S.J.; Kim, J.C. Decomposition of sulfur compounds by a radiolysis: III. A hybrid system and field application. Chem. Eng. J. 2015, 274, 9–16. [Google Scholar] [CrossRef]
  228. Son, Y.S.; Jung, I.H.; Lee, S.J.; Koutrakis, P.; Kim, J.C. Decomposition of sulfur compounds by radiolysis: II. By-products and mechanisms. Chem. Eng. J. 2015, 269, 27–34. [Google Scholar] [CrossRef]
  229. Kim, J.C. Factors affecting aromatic VOC removal by electron beam treatment. Radiat. Phys. Chem. 2002, 65, 429–435. [Google Scholar] [CrossRef]
Figure 1. The production of bulk drugs and Chinese patent drugs and VOCs emissions from 2007 to 2016.
Figure 1. The production of bulk drugs and Chinese patent drugs and VOCs emissions from 2007 to 2016.
Catalysts 10 00668 g001
Figure 2. Classification of VOCs control techniques. Reprinted with permission from [26], 2000, Elsevier Ltd.
Figure 2. Classification of VOCs control techniques. Reprinted with permission from [26], 2000, Elsevier Ltd.
Catalysts 10 00668 g002
Figure 3. A typical activated carbon VOC removal (solvent recovery) plant. Reprinted with permission from [26], 2000, Elsevier Ltd.
Figure 3. A typical activated carbon VOC removal (solvent recovery) plant. Reprinted with permission from [26], 2000, Elsevier Ltd.
Catalysts 10 00668 g003
Figure 4. Schemes of thermal oxidation. (a) Regenerative thermal oxidation; (b) recuperative thermal oxidation. Reprinted with permission from [26], 2000, Elsevier Ltd.
Figure 4. Schemes of thermal oxidation. (a) Regenerative thermal oxidation; (b) recuperative thermal oxidation. Reprinted with permission from [26], 2000, Elsevier Ltd.
Catalysts 10 00668 g004
Figure 5. Scheme of catalytic oxidation. Reprinted with permission from [26], 2000, Elsevier Ltd.
Figure 5. Scheme of catalytic oxidation. Reprinted with permission from [26], 2000, Elsevier Ltd.
Catalysts 10 00668 g005
Figure 6. Schematic diagram of a reverse flow reactor. Reprinted with permission from [26], 2000, Elsevier Ltd.
Figure 6. Schematic diagram of a reverse flow reactor. Reprinted with permission from [26], 2000, Elsevier Ltd.
Catalysts 10 00668 g006
Figure 7. A simple schematic sketch of a bio-filtration system. Reprinted with permission from [26], 2000, Elsevier Ltd.
Figure 7. A simple schematic sketch of a bio-filtration system. Reprinted with permission from [26], 2000, Elsevier Ltd.
Catalysts 10 00668 g007
Figure 8. The ratios of VOCs elimination technologies used in China (a) and other countries (b) (the numbers in the brackets are the numbers of companies using the related technologies). Reprinted with permission from [51], copyright 2012, CNKI.
Figure 8. The ratios of VOCs elimination technologies used in China (a) and other countries (b) (the numbers in the brackets are the numbers of companies using the related technologies). Reprinted with permission from [51], copyright 2012, CNKI.
Catalysts 10 00668 g008
Figure 9. Representation of the VOC abatement process in a factory which produced cefuroxime axetil, cefuroxime sodium, and cefotaxime sodium. Reprinted with permission from [4], 2016, CNKI.
Figure 9. Representation of the VOC abatement process in a factory which produced cefuroxime axetil, cefuroxime sodium, and cefotaxime sodium. Reprinted with permission from [4], 2016, CNKI.
Catalysts 10 00668 g009
Figure 10. VOCs abatement process in the A factory. Reprinted with permission from [4], 2016, CNKI.
Figure 10. VOCs abatement process in the A factory. Reprinted with permission from [4], 2016, CNKI.
Catalysts 10 00668 g010
Figure 11. VOCs abatement process in the B and C factory, utilizing a regenerative thermal oxidizer (RTO) and condensation. Reprinted with permission from [4], 2016, CNKI.
Figure 11. VOCs abatement process in the B and C factory, utilizing a regenerative thermal oxidizer (RTO) and condensation. Reprinted with permission from [4], 2016, CNKI.
Catalysts 10 00668 g011
Figure 12. VOCs abatement process in the D factory. Reprinted with permission from [4], 2016, CNKI.
Figure 12. VOCs abatement process in the D factory. Reprinted with permission from [4], 2016, CNKI.
Catalysts 10 00668 g012
Figure 13. Scheme of adsorption/oxidation apparatus. Reprinted with permission from [1], 2009, Elsevier Ltd.
Figure 13. Scheme of adsorption/oxidation apparatus. Reprinted with permission from [1], 2009, Elsevier Ltd.
Catalysts 10 00668 g013
Figure 14. A 420-L of continuous closed-loop photocatalytic reactor: (1) Photocatalytic unit containing the TiO2 photocatalytic medium and the ultraviolet (UV) lamps, (2) fan, (3) air input and sampling points for analysis, (4) air tranquilization chamber, (5) flow rate measurement. Reprinted with permission from [176], 2018, Elsevier Ltd..
Figure 14. A 420-L of continuous closed-loop photocatalytic reactor: (1) Photocatalytic unit containing the TiO2 photocatalytic medium and the ultraviolet (UV) lamps, (2) fan, (3) air input and sampling points for analysis, (4) air tranquilization chamber, (5) flow rate measurement. Reprinted with permission from [176], 2018, Elsevier Ltd..
Catalysts 10 00668 g014
Figure 15. The schematic diagram of the VUV-PCO system. Reprinted with permission from [63], 2017, Elsevier Ltd.
Figure 15. The schematic diagram of the VUV-PCO system. Reprinted with permission from [63], 2017, Elsevier Ltd.
Catalysts 10 00668 g015
Figure 16. Schematic diagram of the plasma-driven catalyst (PDC) reactor. Reprinted with permission from [196], 2007, Elsevier Ltd.
Figure 16. Schematic diagram of the plasma-driven catalyst (PDC) reactor. Reprinted with permission from [196], 2007, Elsevier Ltd.
Catalysts 10 00668 g016
Table 1. The extreme emission value of volatile organic compounds (VOCs) on the boundary of factories and the highest concentration of selected VOCs allowed in 1 h for employees [13].
Table 1. The extreme emission value of volatile organic compounds (VOCs) on the boundary of factories and the highest concentration of selected VOCs allowed in 1 h for employees [13].
PollutantsLimiting Value (mg·m−3)The Highest Concentration of Selected VOCs Allowed in 1 h for Employees (mg·m−3)
Benzene0.46
Formaldehyde0.20.1
Trichloroethylene a0.130
Dimethyl sulfate a0.50.5
Dichloromethane a4.00.5
Non-methane organic compounds4.0-
Ozone200.26
a. They will be put into effect after the publication of national standards.
Table 2. The limiting value of specific pollutants in workshops and installations [13].
Table 2. The limiting value of specific pollutants in workshops and installations [13].
ClassificationPollutantThe Limiting Values (mg·m−3)
General AreaKey Area
CarcinogensTrichloroethylene a11
Benzene44
Formaldehyde55
Toxic substancesPhosgene0.50.5
HCN1.91.9
Acrolein33
Methyl sulfate a55
Cl255
Photochemically active substancesToluene2515
xylene4020
Dimethyl sulfoxide a10050
Butylene oxide a10050
OtherNH32010
HCl2010
CH3OH5030
CH2Cl2 a7545
a. They will be put into effect after the publication of national standards.
Table 3. VOCs elimination techniques and their operating conditions [50].
Table 3. VOCs elimination techniques and their operating conditions [50].
Abatement TechnologiesVOC Concentration (mg·m−3)Discharge Rate (m3·h−1)Temperature (°C)
Adsorption recycling100–1.5×104<6 × 104<45
Preheated catalytic combustion3000–1/4 LEL *<4 × 104<500
Thermal storage catalytic combustion1000–1/4 LEL<4 × 104<500
Preheated incineration3000–1/4 LEL<4 × 104>700
Thermal storage incineration1000–1/4 LEL<4 × 104>700
Adsorption concentration<1500104–1.2 × 105<45
Biodegradation<1000<1.2 × 104<45
Condensation104–105<104<150
Non-thermal<500<3 × 104<80
* Lower explosive limit (LEL).
Table 4. Reported catalysts for VOCs elimination.
Table 4. Reported catalysts for VOCs elimination.
VOCsReferCatalystCatalytic PerformanceRemarksCatalytic Mechanism
Benzene[56]0.3% Pt/γ-Al2O3T95: 220 °C--
[57]Pd-AuFeCeIMT100: 200 °C--
[58]Layered copper manganese oxide (LCMO)T50: 218 °C
T90: 240 °C
The introduction of water vapor had no significant effect on the catalyst.
The catalyst activity decreased about 30% after SO2 was added.
The addition of CO2 has no effect on the catalytic activity of the catalyst.
The larger surface area and the formation of Cu2+-O2−-Mn4+ entities at the interface between CuO and layered MnO2 promoted the CO and VOCs oxidation.
[59]LaCoO3T50: 330 °CWater vapor has no significant effect on the catalytic performance of catalyst.Factors improve activity: (i) The larger exposed surface, (ii) the composition of the support provides increased oxygen mobility.
[60]LaMnO3
LaCoO3
T50: 301 °C
T50: 323 °C
--
[61]γ-MnO2/SmMnO3T50: 213 °C
T90: 226 °C
-Toluene-Mn ions+ surface reactive oxygen → benzaldehyde → benzoic acid → chain carboxylic acids → CO2.
[62]Mn/meso-TiO2--Benzene → (splitting of benzene, hydrogenation, H abstraction by OH radicals) six-member ring cyclitols+aldehydes ketones etc. → a class of fulvene (isomerization) →CO2 + H2O.
[63]Mesoporous TiO2--Benzene is mainly degraded by photo-generated electron–hole pairs and hydroxyl radicals.
[64]Graphite-SiO2-TiO2 The graphite-SiO2-TiO2 composites exhibited higher photocatalytic activity for degradation of benzene gas under visible light irradiation than that of pure TiO2.Optimum concentration of graphite facilitates the separation of photogenerated electron–hole pairs for graphite-SiO2-TiO2 composites by visible light.
[65]Pt/Al2O3---
Toluene[66]2.3 wt% Pt/3DOM-Mn2O3T50: 165 °C
T90: 194 °C
-A first-order reaction mechanism. Toluene → cat. Surface → benzyl alcohol → benzoic acid and benzaldehyde → (rise temperature) maleic anhydride → CO2 and H2O.
[67]Pd(shell)-Au(core)/TiO2T50: 220 °C
T90: 230 °C
-Langmuir-Hinshelwood mechanism.
[68]Au-Pd/CeO2T50: 120 °C
T90: 150 °C
-Due to the synergistic effect between Au and Pd nanoclusters, Au-Pd/CeO2 bimetallic catalysts are much better than Au and Pd single metal catalysts.
[69]Mn2O3T50: 231 °C
T90: 239 °C
- Toluene +cat. → benzyl species + active oxygen species → aromatic alkoxide → enzaldehyde → benzoate species+active oxygen specie → maleic anhydrides→CO2.
[58]Layered copper manganese oxide (LCMO)T50: 187 °C
T90: 207 °C
--
[59]LaCoO3T50: 244 °C--
[70]La0.8Sr0.2CoO3T50: < 160 °C
T99: 300 °C
-Partial substitution of strontium (Sr) for lanthanum (La) greatly increased the oxygen vacancies in the surface regions to enhance catalytic activity.
[71]SrTi1−XCuXO3
SrTi1−XMnXO3
T50: 343 °C
T90: 398 °C
T50: 302 °C
T90: 335 °C
-Incorporation of Mn attributed a higher amount of oxygen vacancies in the perovskite surface to promote the toluene conversion. Supra facial mechanism.
[72]LaMn1−xBxO3 (B = Co, Ni, Cu, Al)T50: 250 °C
234 °C
226 °C
242 °C
T90: none
278 °C
271 °C
318 °C
-Addition of nickel into LaMnO3 can improve the catalytic oxidation of toluene by the generation of more Mn4+ species and oxygen vacancies and the enhancing reducibility at a low temperature.
[61]γ-MnO2/SmMnO3T50: 187 °C
T90: 208 °C
--
[73]LaMnO3/δ-MnO2T90: 258 °C--
[74]LaMnO3T50: 258 °C
T90: 275 °C
--
[75]MnOx/H-Beta-SDS
MnOx/K-Beta-SDS
MnOx/Si-Beta
T50: 253, 262, 280 °C
T90: 285, 295, 312 °C
-Mars-Van Krevelen mechanism. Organic molecules + lattice oxygen→oxygen vacancy + CO2 + H2O, oxygen vacancy + O2→lattice oxygen.
[76]Mn/R-SBA-15T98: 240 °C-MVK mechanism. Mn2O3-MnO2/R-SBA-15 supply more lattice oxygen species.
[77]LaFeO3/black-TiO2The removal efficiency of black TiO2 and LaFeO3 for toluene was 89% and 98%, respectively, and the removal efficiency for IPA was 90% and 94%, respectively.-Cat.+photon (the wavelength shorter than 440 nm)→ electrons + O2 →O2−, O2+toluene and IPA → CO2 + H2O(g).
[78]Pd/CeO2/γ-Al2O3T98: 205 °C--
[79]CuO/ZnO Nanocomposite photocatalysts--Photooxidative activity and stability over ZnO are improved by loading CuO.
[80]FeOx/SBA-15-Under the condition of 3% Fe loading, the oxidation of toluene is the best one.(1) Direct removal caused by the collision of electrons or oxidation caused by the gas-phase radicals (O•, OH•, N2•, NO•, NO2•) in gas phase; (2) the reaction between adsorbed toluene or other intermediates and the active species (O•, OH•) on the catalyst surface. O2 might be fixed on the catalyst surface via facile interconversion between Fe2+ and Fe3+ states and then be transported to the toluene or intermediates leading to CO2 formation.
Butanol[56]0.3% Pt/γ-Al2O3T95: 200 °C--
CH3SSCH3[81]Pt-Au/Ce-AlT50: 425 °C
T90: 480 °C
-The addition of Au improved the selectivity of ceria-containing catalysts by decreasing the formation of byproducts. This may have had a connection to a lower amount of reactive oxygen after the Au addition.
[81]Cu-Au/Ce-AlT50: 275 °C
T90: 370 °C
---
Chlorobenzene[82]Pt/CrOOHT50: 340 °C
T90: 378 °C
-Platinum accelerates the hydrolysis of Cr-Cl bonds formed at the CrOOH surface and determines the catalyst stability.
Dichloroethane[82]Pt/CrOOHT50: 283 °C
T90: 317 °C
--
Dichloromethane[83]CeO2T90: 260 °C
T90: 160 °C
The addition of 3% (v/v) water can obviously inhibit the catalytic decomposition of VOCs on CeO2.Trichloroethylene + CeO2 →C2HCl→HCl, Cl2, CO2 and trace CO.
Dichlorobenzene[84]H-ZSM-5 and Na-ZSM-5---
Trichloro benzene[84]H-ZSM-5 and Na-ZSM-5---
DCE[85]CeO2@SiO2-400T50: 219 °C
T90: 275 °C
The conversion of 1 vol% H2O and 3 vol% H2O decreased by 6% and 19%, respectively.DCE+acid sites→HCl+ VC. (1) VC→reactive carbonium ion→adsorbed alcohol species→acetate species. Or (2) VC→1,1,2-trichloroethane→dichloroethylene formed (subsequent chlorination reactions) →H2O, COx and HCl.
Ethylbenzene[61]γ-MnO2/SmMnO3T50: 201 °C
T90: 217 °C
--
Formaldehyde[86]TiO2TiO2 degraded almost 100% of formaldehyde or acetaldehyde at a starting concentration of 400–500 ppb with a relative humidity of 40%.-The rate-determining step is the adsorption (external diffusion) on the catalysts active sites, thus the higher surface area, the higher the degradation.
Acetaldehyde[86]TiO2---
Acetone[78]Pd/CeO2/γ-Al2O3T98:220 °C--
[87]CuO/g-Al2O3-5.0 wt% CuO/g-Al2O3 catalyst has the highest removal rate of acetone, reaching 67.9%.Both short-lived radicals and acetone/intermediates can be adsorbed on the catalyst surfaces to initiate a series of surface oxidation reactions, forming CO, CO2, H2O, and byproducts.
Ethyl acetate[78]Pd/CeO2/γ-Al2O3T98:275 °C--
Benzoquinone[88]g-C3N4--OH radicals+phenol→dihydroxycyclohexadienyl radical adducts→phenoxy radicals→H2O(a very slow process), adducts+dissolved O2→dihydroxy intermediates (-HO2) →CO2+H2O.
Hydroquinone Catechol[88]g-C3N4---
O-xylene[89]Ag/NiOx-MnO2T50: 145 °C
T90: 190 °C
The catalytic activity decreased with the addition of water vapor, but recovered after the removal of water vapor.Mars-Van Krevelen mechanism. (i) Electrophilic Ol +aromatic ring→maleate; (ii) O2+ Cat.→electrophilic oxygen species (O22−, O2−, O), electrophilic oxygen species+aromatic ring→maleate, carbonate+(O2. O22−, O2−, O or O2) +Ag→nucleophilic oxygen, nucleophilic oxygen + maleate and carboxylates→CO2 and H2O.
[58]Layered copper manganese oxide (LCMO)T50: 213 °C
T90: 227 °C
--
[61]γ-MnO2/SmMnO3T50: 232 °C
T90: 250 °C
--
O-dichlorobenzene[90]CeMn30GHSV = 15,000 mL/(g·h)
T50: 291 °C
T90: 347 °C
GHSV= 7500 mL/(g·h)
T90: 224 °C
GHSV = 30,000 mL/(g·h).
T90: 360 °C
The addition of low concentration of water vapor is beneficial to the early reaction, but will weaken the catalytic activity of catalyst.O-dichlorobenzene molecules+basic oxygen (Mn-O-Ce-Vö, nucleophilic substitution)→HCl-H-O-Ce-Vö+phenyl. Mn-O-Ce weakened HCl-H-O-Ce-Vö and then prevented Deacon reaction and chlorine poisoning. Phenyl+ active lattice oxygen→organic intermediates →CO2+H2O. (Mars-Van Krevelen mechanism).
CO[58]Layered copper manganese oxide (LCMO)T50: 62 °C
T90: 76 °C
--
Cyclohexane[70]La0.8Sr0.2CoO3T50: 180 °C
T99: 260 °C
--
CH3SH[91]La(13)/HZSM-5T100: 40 °C-The La-modified HZSM-5 increased basic sites and displayed better adsorption ability to CH3SH, decreased in strong acid sites and suppressed the formation of coke deposit.
C2H3CN[92]Cu-ZSM-5(SiO2/Al2O3 = 26)T90: 325 °CThe catalyst has good resistance to steam poisoning.Without H2O: C2H3CN-SCC+O2→NCO→oxidation products; with H2O: C2H3CN-SCC+O2 + H2O→NH3→products.
Cumene[84]H-ZSM-5 and Na-ZSM-5---
Isopropanol[60]LaMnO3
LaCoO3
T50: 216 °C
237 °C
--
[79]CuO/ZnO Nanocomposite photocatalysts--Photooxidative activity and stability over ZnO are improved by loading CuO.
[93]Au-Ag/CeO2T50: 105 °C
T90: 158 °C
-MVK redox mechanism.
Propyl alcohol[60]LaMnO3
LaCoO3
T50: 203 °C
222 °C
--
[70]La0.8Sr0.2CoO3T90: 160 °C--
Propane[94]LaCoO3 T50: 208 °C
T90: 238 °C
-MVK mechanism.
Methane[95]Nanocubic MnO2T50: 293 °C
T90: 350 °C
-CH4+lattice oxygen or surface oxygen vacancies (MnO2-C)→carboxylate species (Langmuir-Hinshelwood route), + active oxygen species→CO2 and H2O.
VOCs[96]MnO2T50: 233 °C
T90: 256 °C
--Toluene + Mn cations→CO2 and H2O.
[97]LaMnO3 and LaCoO3---
[98]La3+-TiO2 and Nd3+-TiO21.2% La3+-TiO2 had the highest photocatalytic activity.--
[99]TiO2/PdThe conversion rate of VOCs reached 90% when the residence time was 27 s.--
[100]MesoTiO2/hydro-CF--Promotion effects on degradation of gaseous polar acetone come from well crystallized anatase nanocrystals, hydro-CF skeleton for adsorption, and fast mass transportation within the hierarchical frameworks.
[101]CsX, NaX, and HYT100: 200 °C-Mars-Van Krevelen mechanism, involving several redox steps.
O3[102]Mn/ZSM-5O3 can be efficiently decomposed by the Mn/ZSM-5 and used for benzene degradation through the OZCO.-Benzene+•OH→phenol→benzoquinone→CO2 + H2O.
[103]CoMnOx/TiO2-When the temperature is 320 °C, the decomposition efficiency of O3 is 98%.-
Vinyl chloride[104]HCl modified La0.5Sr0.5MnO3T100: 300 °C-Doping and acid treatment obviously promote the active Mn4+ species amount and oxygen activation ability, and affect the chloric by-product distribution. Lower temperature inhibits the Deacon reaction and chlorination.
Table 5. Noble metal catalysts for catalytic combustion of VOCs.
Table 5. Noble metal catalysts for catalytic combustion of VOCs.
CatalystVOCsCatalytic PerformanceRemarksCatalytic Mechanism
0.3% Pt/γ-Al2O3Butanol
Benzene
T95: 200 °C
220 °C
--
Pd-AuFeCeIMBenzeneT100: 200 °C--
Pd(shell)-Au(core)/TiO2TolueneT50: 220 °C
T90: 230 °C
-Langmuir-Hinshelwood mechanism.
Au-Pd/CeO2TolueneT50: 120 °C
T90: 150 °C
-Due to the synergistic effect between Au and Pd nanoclusters, Au-Pd/CeO2 bimetallic catalysts are much better than Au and Pd single metal catalysts.
Au-Ag/CeO2IsopropanolT50: 105 °C
T90: 158 °C
-MVK redox mechanism.
Pt-Au/Ce-AlCH3SSCH3T50: 425 °C
T90: 480 °C
-The addition of Au improved the selectivity of ceria-containing catalysts by decreasing the formation of byproducts due to lower amount of reactive oxygen after the Au addition.
Cu-Au/Ce-AlCH3SSCH3T50: 275 °C
T90: 370 °C
---
Ag/NiOx-MnO2O-xyleneT50: 145 °C
T90: 190 °C
The catalytic activity decreased with the addition of water vapor, but recovered after the removal of water vapor.Mars-Van Krevelen mechanism. (i) Electrophilic Ol +aromatic ring→maleate; (ii) O2+ Cat.→electrophilic oxygen species (O22−, O2−, O), electrophilic oxygen species + aromatic ring→maleate, carbonate+(O2, O22−, O2−, O or O2) +Ag→nucleophilic oxygen, nucleophilic oxygen + maleate and carboxylates→CO2 and H2O.
2.3 wt% Pt/3DOM-Mn2O3TolueneT50: 165 °C
T90: 194 °C
-A first-order reaction mechanism. Toluene →cat. surface→benzyl alcohol→benzoic acid and benzaldehyde→(rise temperature) maleic anhydride→CO2 and H2O.
Table 6. Non-noble metal catalysts for catalytic combustion of VOCs.
Table 6. Non-noble metal catalysts for catalytic combustion of VOCs.
ReferCatalystVOCsCatalytic PerformanceRemarks
[69]Mn2O3TolueneT50: 231 °C
T90: 239 °C
-
[82]Pt/CrOOHDichloroethane

Chlorobenzene
T50: 283 °C
T90: 317 °C
T50: 340 °C
T90: 378 °C
-
[83]CeO2Tetrachloroethylene dichloromethaneT90: 260 °C
T90: 160 °C
The addition of 3% (v/v) water can obviously inhibit the catalytic decomposition of VOCs on CeO2.
[90]CeMn30O-dichlorobenzeneGHSV = 15,000 mL/(g·h)
T50: 291 °C
T90: 347 °C
GHSV= 7500 mL/(g·h)
T90: 224 °C
GHSV = 30,000 mL/(g·h)
T90: 360 °C
The addition of low concentration of water vapor is beneficial to the early reaction, but will weaken the catalytic activity of the catalyst.
[96]MnO2VOCsT50: 233 °C
T90: 256 °C
-
[95]Nanocubic MnO2MethaneT50: 293 °C
T90: 350 °C
-
Table 7. Perovskite catalysts for catalytic combustion of VOCs.
Table 7. Perovskite catalysts for catalytic combustion of VOCs.
ReferCatalystVOCsCatalytic PerformanceRemarksCatalytic Mechanism
[58]Layered copper manganese oxide (LCMO)CO

Benzene

Toluene

O-xylene
T50: 62 °C
T90: 76 °C
T50: 218 °C
T90: 240 °C
T50: 187 °C
T90: 207 °C
T50: 213 °C
T90: 227 °C
The introduction of water vapor had no significant effect on the catalyst.
The catalyst activity decreased about 30% after SO2 was added.
The addition of CO2 has no effect on the catalytic activity of the catalyst.
The larger surface area and the formation of Cu2+-O2−-Mn4+ entities at the interface between CuO and layered MnO2 promoted the CO and VOCs oxidation.
[59]LaCoO3 Benzene
Toluene
T50: 330 °C
T50: 244 °C
Water vapor has no significant effect on the catalytic performance of catalyst.Factors improve activity: (i) The larger exposed surface, (ii) the composition of the support provides increased oxygen mobility.
[60]LaMnO3



LaCoO3

Isopropanol
Benzene
Propyl alcohol

Isopropanol
Benzene
Propyl alcohol
T50:
216 °C
301 °C
203 °C
T90:
237 °C
323 °C
222 °C
--
[61]γ-MnO2/SmMnO3Toluene

Benzene

O-xylene

Ethylbenzene
T50: 187 °C
T90: 208 °C
T50: 213 °C
T90: 226 °C
T50: 232 °C
T90: 250 °C
T50: 201 °C
T90: 217 °C
-Toluene-Mn ions+ surface reactive oxygen → benzaldehyde → benzoic acid → chain carboxylic acids → CO2.
[70]La0.8Sr0.2CoO3toluene

Cyclohexane

Propyl alcohol
T50:<160 °C
T99: 300 °C
T50: 180 °C
T99: 260 °C
T90: 160 °C
-Partial substitution of strontium (Sr) for lanthanum (La) greatly increased the oxygen vacancies in the surface regions to enhance the catalytic activity.
[71]SrTi1-XCuXO3
SrTi1-XMnXO3
TolueneT50: 343 °C
T90: 398 °C
T50: 302 °C
T90: 335 °C
-Incorporation of Mn attributed a higher amount of oxygen vacancies in the perovskite surface to promote the toluene conversion. Supra facial mechanism.
[72]LaMn1-xBxO3 (B = Co, Ni, Cu, Al)Toluene
LMLi
LMNi
LMCo
LMAl

LMLi
LMNi
LMCo
LMAl
T50:
250 °C
234 °C
226 °C
242 °C
T90:
none
278 °C
271 °C
318 °C
-Addition of nickel into LaMnO3 can improve the catalytic oxidation of toluene by generation of more Mn4+ species and oxygen vacancies and the enhancing reducibility at low temperature.
[73]LaMnO3/δ-MnO2TolueneT90: 258 °C--
[74]LaMnO3TolueneT50: 258 °C
T90: 275 °C
--
[85]CeO2@SiO2-400DCET50: 219 °C
T90: 275 °C
The conversion of 1 vol% H2O and 3 vol% H2O decreased by 6% and 19%, respectively.DCE+acid sites→HCl+ VC. (1) VC→reactive carbonium ion→adsorbed alcohol species→acetate species. Or (2) VC→1,1,2-trichloroethane→dichloroethylene formed (subsequent chlorination reactions) →H2O, COx and HCl.
[94]LaCoO3PropaneT50: 208 °C
T90: 238 °C
-MVK mechanism.
[97]LaMnO3 and LaCoO3Chlorinated VOCs---
[104]HCl modified La0.5Sr0.5MnO3Vinyl chlorideT100: 300 °C-Doping and acid treatment obviously promote the active Mn4+ species amount and oxygen activation ability, and affect the chloric byproduct distribution. The lower temperature inhibits the Deacon reaction and chlorination.
Table 8. Concentrated oxidation catalysts for catalytic combustion of VOCs.
Table 8. Concentrated oxidation catalysts for catalytic combustion of VOCs.
ReferCatalystVOCsCatalytic PerformanceRemarksCatalytic Mechanism
[75]MnOx/H-Beta-SDS
MnOx/K-Beta-SDS
MnOx/Si-Beta
TolueneT50: 253, 262, 280 °C
T90: 285, 295, 312 °C
-Mars-Van Krevelen mechanism. Organic molecules+lattice oxygen→oxygen vacancy+CO2+H2O, oxygen vacancy+O2→lattice oxygen.
[84]H-ZSM-5 and Na-ZSM-5Cumene
Dichlorobenzene
Trichloro benzene
---
[91]La(13)/HZSM-5CH3SHT100: 40 °C-The La-modified HZSM-5 increased basic sites and displayed better adsorption ability to CH3SH, decreased in strong acid sites and suppressed the formation of coke deposit.
[92]Cu-ZSM-5(SiO2/Al2O3 = 26)C2H3CNT90: 325 °CThe catalyst has good resistance to steam poisoning.Without H2O: C2H3CN-SCC+O2→NCO→oxidation products; with H2O: C2H3CN-SCC+O2+ H2O→NH3→products.
[101]CsX, NaX, and HYVOCs mixtureT100: 200 °C-Mars-Van Krevelen mechanism, involving several redox steps.
[176]Mn/R-SBA-15TolueneT98: 240 °C-MVK mechanism. Mn2O3-MnO2/R-SBA-15 supply more lattice oxygen species.
Table 9. Photocatalysts for elimination of VOCs.
Table 9. Photocatalysts for elimination of VOCs.
ReferCatalystVOCsCatalytic PerformanceRemarksCatalytic Mechanism
[62]Mn/meso-TiO2Benzene--Benzene→(splitting of benzene, hydrogenation, H abstraction by OH radicals) six-member ring cyclitols+aldehydes ketones, etc. →a class of fulvene (isomerization)→CO2 +H2O.
[63]Mesoporous TiO2Benzene--Benzene is mainly degraded by photo-generated electron-hole pairs and hydroxyl radicals.
[64]Graphite-SiO2-TiO2BenzeneThe graphite-SiO2-TiO2 composites exhibited higher photocatalytic activity for degradation of benzene gas under visible light irradiation than that of pure TiO2.-Optimum concentration of graphite facilitates the separation of photogenerated electron–hole pairs for graphite-SiO2-TiO2 composites by visible light.
[77]LaFeO3/black-TiO2TolueneThe removal efficiency of black-TiO2 and LaFeO3 for toluene was 89% and 98%, respectively, and the removal efficiency for IPA was 90% and 94%, respectively.-Cat.+photon (the wavelength shorter than 440 nm)→ electrons +O2→O2−, O2+toluene and IPA →CO2+H2O(g).
[78]Pd/CeO2/γ-Al2O3
Toluene
Acetone
Ethyl acetate
T98:
205 °C
220 °C
275 °C
--
[79]CuO/ZnO nanocomposite photocatalystsToluene
Iopropanol
--Photooxidative activity and stability over ZnO are improved by loading CuO.
[86]TiO2Formaldehyde
acetaldehyde
TiO2 degraded almost 100% of formaldehyde or acetaldehyde at a starting concentration of 400–500 ppb with a relative humidity of 40%.-The rate-determining step is the adsorption (external diffusion) on the catalysts active sites, thus the higher the surface area, the higher the degradation.
[88]g-C3N4Benzoquinone
Hydroquinone catechol
--OH radicals+phenol→dihydroxycyclohexadienyl radical adducts→phenoxy radicals→H2O (a very slow process), adducts+dissolved O2→dihydroxy intermediates (-HO2) →CO2+H2O.
[98]La3+-TiO2 and Nd3+-TiO2VOCs mixture1.2% La3+-TiO2 had the highest photocatalytic activity.--
[99]TiO2/PdVOCsThe conversion rate of VOCs reached 90% when the residence time was 27 s.--
[102]Mn/ZSM-5O3O3 can be efficiently decomposed by the Mn/ZSM-5 and used for benzene degradation through the OZCO.-Benzene+•OH→phenol→benzoquinone→CO2 +H2O.
[100]Meso-TiO2/hydro-CFVOCs--Promotion effects on degradation of gaseous polar acetone come from well crystallized anatase nanocrystals, hydro-CF skeleton for adsorption, and fast mass transportation within the hierarchical frameworks.
Table 10. Catalysts for VOCs elimination through the non-thermal plasma process.
Table 10. Catalysts for VOCs elimination through the non-thermal plasma process.
ReferCatalystVOCsCatalytic PerformanceRemarksCatalytic Mechanism
[65]Pt/Al2O3Benzene
byproducts
---
[80]FeOx/SBA-15Toluene-Under the condition of 3% Fe loading, the oxidation of toluene is the best one.(1) Direct removal caused by the collision of electrons or oxidation caused by the gas-phase radicals (O•, OH•, N2•, NO•, NO2•) in the gas phase; (2) the reaction between adsorbed toluene or other intermediates and the active species (O•, OH•) on the catalyst surface. O2 might be fixed on the catalyst surface via a facile interconversion between Fe2+ and Fe3+ states and then be transported to the toluene or intermediates leading to CO2 formation.
[87]CuO/g-Al2O3Ccetone-5.0 wt% CuO/g-Al2O3 catalyst has the highest removal rate of acetone, reaching 67.9%.Both short-lived radicals and acetone/intermediates can be adsorbed on the catalyst surfaces to initiate a series of surface oxidation reactions, forming CO, CO2, H2O, and byproducts.
[103]CoMnOx/TiO2O3-When the temperature is 320 °C, the decomposition efficiency of O3 is 98%.-

Share and Cite

MDPI and ACS Style

Zhou, L.; Ma, C.; Horlyck, J.; Liu, R.; Yun, J. Development of Pharmaceutical VOCs Elimination by Catalytic Processes in China. Catalysts 2020, 10, 668. https://doi.org/10.3390/catal10060668

AMA Style

Zhou L, Ma C, Horlyck J, Liu R, Yun J. Development of Pharmaceutical VOCs Elimination by Catalytic Processes in China. Catalysts. 2020; 10(6):668. https://doi.org/10.3390/catal10060668

Chicago/Turabian Style

Zhou, Lilong, Chen Ma, Jonathan Horlyck, Runjing Liu, and Jimmy Yun. 2020. "Development of Pharmaceutical VOCs Elimination by Catalytic Processes in China" Catalysts 10, no. 6: 668. https://doi.org/10.3390/catal10060668

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