Next Article in Journal
Adaptive Composite Fault Diagnosis of Rolling Bearings Based on the CLNGO Algorithm
Next Article in Special Issue
Visible Fenton Degradation of Bisphenol A by Boron-Doped, Graphene-Oxide-Coated Nano-Fe3O4
Previous Article in Journal
Special Issue on Modelling, Simulation and Control in Combustion Processes of Renewable Fuels
Previous Article in Special Issue
Biochemical Study of Bacillus stearothermophilus Immobilized Lipase for Oily Wastewater Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Removal of Volatile Organic Compounds (VOCs) from Air: Focus on Biotrickling Filtration and Process Modeling

Department of Process Engineering and Chemical Technology, Faculty of Chemistry, Gdańsk University of Technology, Narutowicza 11/12 Street, 80-233 Gdańsk, Poland
Processes 2022, 10(12), 2531; https://doi.org/10.3390/pr10122531
Submission received: 19 October 2022 / Revised: 16 November 2022 / Accepted: 23 November 2022 / Published: 29 November 2022
(This article belongs to the Special Issue Advanced Liquid Waste and Gas Waste Treatment Processes)

Abstract

:
Biotrickling filtration is a well-established technology for the treatment of air polluted with odorous and volatile organic compounds (VOCs). Besides dozens of successful industrial applications of this technology, there are still gaps in a full understanding and description of the mechanisms of biotrickling filtration. This review focuses on recent research results on biotrickling filtration of air polluted with single and multiple VOCs, as well as process modeling. The modeling offers optimization of a process design and performance, as well as allows deeper understanding of process mechanisms. An overview of the developments of models describing biotrickling filtration and conventional biofiltration, as primarily developed and in many aspects through similar processes, is presented in this paper.

1. Air Pollution with VOCs: Problem and Treatment Strategies

Air pollution with volatile organic compounds, such as hydrocarbons, ethers, esters or alcohols that resultfrom both anthropogenic and industrial activities, poses a broad scope of health hazards to living organisms [1]. Volatile organic compounds easily evaporate, even at room temperature. Because these are chemicals with different chemical properties, including small molecular size with no electric charge, variable lipophilicity, and volatility, inhalation is the major route of their absorption by living organisms [2]. Inhalation-related hazards are not only related to short-term exposure and possible feelings of discomfort, but mainly to long-term exposure and possible unknown health effects. This has also become evident from the perspective of COVID-19 epidemics [3]. For example, number of reported infections and death cases related to COVID-19 may be related to the severity of atmospheric pollution in a given territory [4]. It is obvious, therefore, that efficient, reliable, and economic methods of air treatment need to be developed and improved.
Various treatment methods of air polluted with VOCs have been established worldwide. These methods include combustion [5], absorption [6], adsorption [7], condensation [8,9], membrane processes [10], oxidative catalysis [11,12], UV-oxidation [13], and non-thermal plasma processes [14]. Besides the fact that such gas treatment methods have plenty of advantages, like rapid start-up periods and efficient removal rates of the pollutants, they are usually costly, especially in operation, and may consume considerable amounts of chemicals that generate hazardous by-products. Thus, some of these methods may be considered unsustainable. Due to high efficiency, their possibility to economically treat large volumes of gases, their limited or nonexistent use of chemicals, and their capability of operating at ambient temperatures (10–40 °C) and under atmospheric pressure, biological methods find promising applications as deodorization methods [15,16,17,18]. These methods mainly include biofiltration, bioscrubbing, biotrickling filtration, and other bioreactor configurations, and sometimes include combinations of these methods with other physico-chemical methods [19,20,21,22].
Biological methods of VOC removal from air have been known for about one hundred years, and constitute the prevailing share of conventional biofiltration processes. Initially, microbial reactions in soil were employed for air treatment [18,23]. Biofiltration finds applications in odor control for different waste gas streams, e.g., in rendering plants, chemical factories, agriculture, and animals farming. Biofilters, i.e., bioreactors in which biofiltration is carried out, are fixed-bed reactors that are inhabited by active microorganisms. These microorganisms are able to degrade VOCs present in air, and use them as carbon and energy sources. Biofiltration mechanisms of VOC removal include a mix of various phenomena, like absorption, adsorption, diffusion, and biodegradation. As a result of biofiltration, VOCs are converted into biomass, CO2, and water [18,24]. A distinctive feature of biofiltration processes, also termed conventional biofiltration (abbreviated BF), is the lack of mobile liquid phases inside the biofilter. BF can be realized in open or closed bioreactors.
This paper focuses on biotrickling filtration (abbreviated BTF). According to a definition by Deshuses, Gabriel, and Cox [25,26] biotrickling filters differ from conventional biofilters in two main areas i.e., the packing type and liquid phase occurrence. Biotrickling filters are packed with inert or manufactured media, over which a liquid phase is trickled; thus, biotrickling filters are by definition bioreactors that contain gas, solid and liquid phases. This is contrary to conventional biofilters, which are packed with organic materials, are naturally inhabited by various types of microorganisms, and which humidify the polluted air in a separate chamber prior to entering the biofilter [27].
This paper aims to provide an update for the removal of VOCs using biological methods, with a focus on biotrickling filtration research on the removal of single VOCs and multiple VOCs, as well as a review of the modeling of biotrickling filtration as a core of this work. An overview of conventional biofiltration modeling is also presented as a basis for the development of BTF modeling. To the best knowledge of the author, no paper has yet presented a comprehensive outlook on the mentioned aspects of biotrickling filtration up to this point.

2. Review Methodology

This review was prepared on the basis of the literature research using well-established scientific databases (e.g., Science Direct) with over 250 references, out of which 174 were finally referred to. The references on the biotrickling filtration research were limited to the last ten years in order to present the most updated results. The exception is the review of treating multiple VOCs, for which the timespan is from 2004 to 2022. The review of the development of models for biofilters and biotrickling filters covered the period from about 1990 to 2022, to show the broadest possible perspective on BF and BTF modeling aspects. The paper is focused on biotrickling filtration; however, attention is also paid to conventional biofiltration as primarily developed and in many similar aspects to BTF.

3. Biotrickling Filtration for the Removal of VOCs from Air: Background, Current Problems, and Research Overview

Biotrickling filtration is a biotreatment method that enables the removal of gaseous contaminants from the air [16,23,28]. The process consists of passing a polluted gas through a packed layer. Elements of this layer are usually inert. Prior to starting the process, these elements are inoculated with either a single microorganism (usually for model investigations or for testing a new species with high expected potential to degrade VOCs) or a mixed microbial inoculum (a consortium of selected microbial species or an inoculum originated from the activated sludge of a wastewater treatment plant). As a result of supplying the biotrickling filter with polluted air (containing oxygen as well as pollutants, which serve carbon sources for microbes), as well as spraying the packed layer with a so-called trickling liquid, a biofilm is formed over the elements of a packed layer [16,29]. Within this biofilm, a biodegradation of selected/all air pollutants is possible, which results in the emission of cleaned air from the bioreactor. When the biofiltration process is realized effectively, the treated air leaving the biotrickling filter contains only a fraction or practically no VOCs, but contains slightly elevated CO2 concentrations, due to biodegradation, when compared to the inlet stream of air. When it comes to the liquid phase i.e., the trickling solution, it is usually an aqueous solution of mineral salts and micronutrients, e.g., vitamins. When used in the biotrickling filtration process, the liquid absorbs by-products of biodegradation, flushes fractions of dead microbial matter, and may contain variable concentrations of VOCs, depending on their solubility and biodegradation rate. The trickling liquid is usually recirculated unless a certain level of its evaporation or absorption/intake of the above-mentioned substances is reached. Then, the liquid must be regenerated e.g., by either removing and discarding the remaining volume and introducing a fresh liquid, or by daily liquid make-ups to maintain a fixed level.
There are recent review papers on controlling air pollution using biotrickling filters [16,23,30] that also underlie the possibility of using the biotrickling filtration of VOCs in wastewaters treatment [31], in which the process parameters and their effects on the biofiltration mechanisms are thoroughly discussed. Another excellent recent review paper in the field of biotrickling filtration deals with the effects of the liquid phase on the gas contaminant removal [32], as well as on the enhanced removal of hydrophobic air contaminants as a result of the addition of a surfactant or a hydrophilic gas-phase component [28]. An interesting review paper [33] concerns mass transfer aspects for biofilters and biotrickling filters, among others. The paper discusses the main mechanisms for the resistance of component transfers from gas to biofilm phases, which mainly include mass transfer limitations or kinetic limitations. The paper discusses the methods of determining mass transfer rates, as well as the modes of influencing and enhancing the mass transfer properties. These aspects are especially important in the case of modeling any process occurring as a result of contact between at least two different phases. Yang et al. [34] provided an interesting paper on the possible interactions between VOCs when multiple gas pollutants were treated in biofilters. A recent review paper by Danila et al. [35] discusses the issues of packing humidity and irrigation patterns for biofilters, which is an important, though seldom discussed, process parameter. In a recent review paper by Marycz et al. [36] the advantages of fungi utilization in biotrickling filters are discussed.
Interestingly, biotrickling filters for industrial applications can be obtained by retrofitting the existing chemical scrubbers, as discussed by Gabriel and Deshusses [37].
The conditions for conducting the biotrickling filtration process are mainly determined by the volumetric flow rate of treated gas (Q, m3), the concentration of VOCs in the inlet gas stream (Cin, g m−3 or ppmv), the biofilter dimension (cross-surface area represented by internal diameter, din, and height of packing, H), the resulting effective volume of the biofilter packing (V, m3), and the packing characteristics (porosity, surface area) [16,23,30]. These parameters are usually represented by the values of specific inlet loading (IL, g m−3 h−1) and empty bed residence time (EBRT, s). Process efficiency is usually reported using the values of removal efficiency (RE, dimensionless or %) and elimination capacity (EC, g m−3 h−1). While both RE and EC are dependent upon the inlet concentrations of VOCs in the gas phase, EC gives a clear and comparable value representing the actual performance capacity of a given biofilter (what mass of VOCs can be removed using a given volume of biofilter per unit of time). Additionally, the value of EBRT is also an important parameter for the comparison of different processes, especially from a practical viewpoint. However, when referring to the process efficiency with the use of RE, one should always present the inlet VOC concentrations, because sometimes a high RE can be reached only for low Cin or IL values, and thus may be misleading when evaluating biofiltration performance. The formulae for calculating the above-mentioned parameters are given below.
I L = Q · C i n · V 1 ,
E B R T = V · Q 1 ,
R E = ( C i n C o u t ) · C i n 1 ,
E C = Q · ( C i n C o u t ) · V 1 ,
where Cout is the VOC concentration in the outlet gas stream (g m−3).
For a given biotrickling filter set-up, including microorganisms and fixed process conditions (i.e., VOC concentrations in the gas phase and its flow rate), biotrickling filtration performance can be regulated by manipulating the trickling liquid velocity and frequency, the process temperature, the pH and composition of the trickling liquid, the packing material, and the feeding and biomass growth control strategies. However, it should be taken into account that the frequency of the trickling process cannot be implemented in all applications, as agricultural applications strongly require continuous irrigation. A brief summary of the effects of these parameters on biofiltration performance is given in Table 1.
The main body of evidence of research on biotrickling filtration covers the treatment of single air pollutants. Such an approach allows for the definition of process mechanisms with respect to particular compounds, and may be profitable for practical application when a single compound is dominating in the gas mixture. Table 2 presents selected research results on the biotrickling filtration of single VOCs with a focus on the enhancement of hydrophobic VOC biotrickling filtration, as well as presents advances in process performance, design, and construction modifications.
However, in the industrial practice and in various processes of municipal waste stream treatments, including wastewaters and solid waste processing, mixtures of various VOCs are present. The co-existence of a number of compounds to be treated simultaneously affects the process mechanism and performance, and requires a different design approaches than for single compounds. Additionally, the process modeling requires a different assumptions and approaches; e.g., Baltzis et al. [47] considered the case that different compounds underwent biodegradation carried out by different biofilm types, i.e., different types of microorganisms. Table 3 presents selected resent research on the biotrickling filtration of multiple VOCs. The data was selected from the perspective of the evaluation of process mechanisms for the effects of substrate interactions (e.g., co-metabolism and various types of synergistic effects) on the process performance.
Besides considerable research on the biotrickling filtration of VOC mixtures and the increasing number of scientific papers being published on the topic [48], the interactions between the mixed pollutants and their effects on the biofiltration performance and mechanisms still need to be explored [49,50,51,52].
For example, when the co-treatment of multiple VOCs takes place in a biotrickling filter, several aspects of the process mechanism may lead to either improved or decreased process efficiency [34]. The decrease in biofiltration performance may be a result of substrate inhibition or toxic effects. For example, methanol at low concentrations may promote the removal of hexane, while at higher concentrations may lead to the decreased removal of hexane, because it becomes the dominant carbon source and can become toxic to microorganisms [50]. An increase of BTF performance may, on the other hand, result from improving the mass transfer of the hydrophobic substrates as a result of the increased solubility in the more hydrophilic compounds, which thus enhances biodegradation and leads to increased microbial growth and the promotion of co-substrate degradation [49]. Some VOCs, like toluene, can also modify the production and composition of extracellular polymeric substances (abbreviated as EPS) [40,51]. EPS are mainly composed of proteins and sugars, and the relative ratio of these can affect the biofilm hydrophobicity i.e., biofilm hydrophobicity increases with the increased ration of proteins to sugars. It is evident that the higher hydrophobicity of biofilm contributes to the mass transfer of hydrophobic VOCs from gas to biofilm, thus resulting in their enhanced biodegradation. It should also be mentioned that the degradation of hydrophobic VOCs can be accelerated by the use of packing materials with enhanced adsorptive capacities, i.e., hydrophobic VOCs can adsorb and accumulate on these carriers, and mass transfer to microbial cells and final biodegradation is enhanced by higher VOC gradients between the carrier and the biofilm [53,54,55,56].
Research results presented in Table 2 show a growing interest in the investigations of biofiltration performance for gas mixtures that mimic the real waste gas compositions. Various configurations and process conditions have been studied from the laboratory pilot scale of a BTF. A considerable problem during the design and operation of biotrickling filters is their limited efficiency in treating hydrophobic volatile organic compounds [16,28]. This problem is mainly related to the low solubility of such compounds in the aqueous solutions, thus resulting in their low availability for microorganisms, as well as their usually high toxicity and resistance to biodegradation. Thus, modes of enhancing the removal of hydrophobic VOCs are of special interest nowadays, and present promising results when the utilization of two-phase biotrickling filters is considered [57].
The overview of current research on biotrickling filtration (Table 2 and Table 3) includes selected process conditions and parameters, like bioreactor dimensions, packing type, inoculum, process performance, and important outcomes for selected treated VOCs. Due to the dominance of papers discussing the removal of single air pollutants (Table 2), the selected recent research results from the period from 2011 to 2022 are revised. Due to a lower number of papers on investigations of the biofiltration of mixed VOCs, Table 3 presents selected results from a longer time perspective, i.e., 2004–2022. The results indicate that research has been mainly performed on laboratory-scale biofilters (with packing volumes of about several liters); however, there are also reports on pilot and full-scale applications. Dominant packing materials include ceramics, minerals, and polymeric materials, such as polyurethane foam, while bio-based materials, structured packing, or other engineered and novel materials are in minority. The most common inocula originate from activated sludge from wastewater treatment plants; however, research has also been conducted on the identification and testing of new microbial species that are effective in degrading VOCs. The majority of EBRT values range between 30-100 s, while the total VOCs concentrations in the inlet gas streams, irrespective of the number of compounds, range from about several tens to thousands of mg m−3, which are approximately equivalent to hundreds of ppm (v/v). Current research goals include the improvement of biotrickling filtration performance for hydrophobic VOCs by using surface active substances, modified liquid phases, two-phase bioreactors, the addition of hydrophilic compounds, or the application of fungi. Other areas of development include scaling up processes, the management of secondary waste, and the incorporation of pre- or post-treatment processes to increase process efficiency or obtain broader utilitarian purposes, such as coupling technologies to cope with environmental issues. Another novel approach includes the design of biotrickling filters as microbial fuel cells for energy generation during waste air treatment [58,59,60,61,62].
Table 2. Selected recent research on biotrickling filtration of single VOCs.
Table 2. Selected recent research on biotrickling filtration of single VOCs.
VOCBioreactor
Dimensions:
din × H [m]; V [dm3]
Packing
Material
InoculumProcess
Parameters
Interesting OutcomesReference
n-Butanol0.08 × 0.68; 2.5Peat & perlite, ceramic Raschig ringsNo additional inoculationEBRT: 60 s, Cin: 100–800 ppmThe extent of reduction in VOCs concentration in gas phase is not quantitatively equivalent to odor reduction[63]
Triethylamine0.21 × 1; 52Lava rocksSludge from WWTPEBRT: 31–312 s
Cin: 600 mg m−3
Increasing of the gas flow rate is more cost-effective than increasing EBRT[64]
Trimethylamine0.077 × 1.7; 5.6Polyurethane ringsAminobacter aminovoransEBRT: 85, 170 s IL: 0.2–12 g·m−3·h−1H2S only slightly affected the removal efficiency of trimethylamine[65]
Ethyl acetate0.2 × 0.14 (two sections)Anode materials: carbon coke or carbide porous ceramic rings (CPCR)Activated sludge from WWTPEBRT: 60–120 s
Cin: 0.54–3.23 g m−3
The use of CPCR resulted in higher removal efficiency of ethyl acetate that carbon coke[61]
Methyl acrylate0.12 × 1.1; 6.44Ceramic particlesActivated sludge from WWTPEBRT: 200–400 s
Cin: 120–7505 mg m−3
Three-layer BTF was applied and it was the 1st packing layer that mainly reduced methyl acrylate concentrations in air[1]
Chloroform0.076 × 1.3; 5.89Celite (pelletized diatomaceous earth)Filamentous fungiEBRT: 344 s
Cin = 5 ppm
The application of ethanol as a co-metabolite; acidic conditions enhanced the fungi growth[66]
0.076 × 1.3; 5.89Pelletized diatomaceous earthBacteria from WWTPEBRT: 344 s
Cin: 200 ppm
Aerobic and anaerobic conditions (with methanogenic bacteria) were investigated[67]
Ethylbenzene0.1 × 0.78; 6.1Polyurethane spongeActivated sludge from WWTPEBRT: 30 s
Cin: 2.5 g m−3
The addition of saponins to the liquid phase resulted in enhanced removal of ethylbenzene[52]
0.1 × 0.8 (2 sections); 12.56Polyurethane spongeFresh biological from WWTPEBRT: 30 s
Cin: 1 g m−3
The high removal of ethylbenzene due to addition of biosurfactant (from piggery wastewater) to liquid phase[68]
Styrene0.144 × 1.63; 20Polypropylene Ralu ringsPseudomonas sp. E-93486EBRT: 62 s
IL: 1 g m−3 h−1
A three-phase dynamic mathematical model was proposed[69]
Six rectangular sections with total packing volume of 50 dm3Fern or plastic chipsActivated sludge from WWTPEBRT: 21 s
Cin: 2.39 g m−3
The study provides a comparison of packing materials, presenting the superior performance of plastic over fern chips during styrene removal from air[70]
0.08 × 1.5; 3.77Ceramic Raschig ringsActivated sludge from cooking plantEBRT: 68 s
IL = 180 g m−3 h−1
The removal of styrene from air is enhanced by the acclimatization of a biotrickling filter to toluene-styrene mixture[71]
Toluene0.12 × 0.78; 4.27Ceramic pelletsFungi from activated sludge (e.g., Fusarium)EBRT: 55 s
IL = 100.3 g m−3 h−1
A fungi-based biotrickling filter was applied to efficiently remove toluene from air[72]
0.1 × 0.6 (approx.); 4.56Ceramsite particlesFungi (Fusarium)EBRT: 59 s
Cin: 1.053 g m−3
The effects of bed porosity on pressure drop and its maintenance during biotrickling filtration process were studied[73]
Differential biotrickling filter with a packing volume of 0.4 dm3Glass beadsInoculum from previously working biotrickling filterEBRT: 28 s
IL: 472 g m−3 h−1
A differential biotrickling filter was proposed in opposition to popular integral-column bioreactors[74]
Cyclohexane0.318 × 1; 52Polyurethane foamAcidovorax sp. CHX 100EBRT: 37 s
Cin: 720 mg dm−3
A proposal for the removal of cyclohexane from air in a biotrickling filter using Acidovorax sp.[75]
Hexane0.07 × 1; 2.4PerliteFusarium solaniEBRT: 78 s
Cin: 7400 mg dm−3
The addition of different carbon sources to the trickling liquid may shorten the start-up period[76]
0.076 × 1.3; 2.7Pellets of diatomaceous earth (Celite)Fungi with dominant Cladosporium and Rhodotorula species (genus)EBRT: 120 s
Cin: 125 ppmv
The effects of pH and methanol addition on the removal of hexane were studied; low pH was favorable for enhanced removal of hexane[77]
0.1 × 0.65; 5.1Reticular polyurethane spongeFilamentous bacteriaEBRT: 30 s
IL: 124 g m−3 h−1
The intermittent mode of spraying and reticulated packing configuration, enhancing the mass transfer, enabled high removal of hexane[78]
Table 3. Selected research on biotrickling filtration of VOCs mixtures.
Table 3. Selected research on biotrickling filtration of VOCs mixtures.
VOCs MixtureBioreactor Dimensions (din x H; V)Packing
Material
InoculumProcess
Parameters
OutcomesReference
Acetone, toluene, trichloroethylene0.07 × 0.6; 1.9 (bioreactor organized in 2 segments)Granular activated carbonAcclimated microbial seeds from industrial WWTPEBRT: 155 s
Cin (of each VOC): 10–800 ppmv
Relative concentrations of VOCs in treated gas mixture as a crucial parameter for the functioning of microbes during biofiltration [79]
Trichloroethylene, perchloroethylene0.07 × 0.6; 3Granular activated carbonMixed microbial consortium from soil and activated sludgeEBRT: 6–36s
Cin: 35.4 ppmv (PCE) and 46.7 ppmv (TCE)
The application of photooxidation enhanced the removal of target pollutants in biotrickling filter, and various removal mechanisms in a biofilter were identified (hydrolysis, adsorption, biodegradation)[80]
Hydrogen sulfide, methyl mercaptan, ammonia, VOCs mixture, dimethyl sulfide3.2 × 13OdourTeQ vesselsThiobacillus for H2S removal, heterotrophic microbes for VOCs removalEBRT: 17 sIndustrial application[81]
Methanol, ethanol, acetone, toluene, chloroform0.054 × 1; 1.7Polyvinyl chloride particlesMicrobial consortium taken from previously operating bioreactorsEBRT: 25–68.6 s
Cin: 1–4 g m−3 (VOCs); 0.05–1 g m−3 (chloroform)
High removal of target air pollutants and resistance to intermittent loadings[82]
Methyl mercaptan, toluene, alpha-pinene and hexane0.08 × 1; 4Polyurethane foam cubes Acclimated activated sludgeRE up to about 99% for hydrophilic and about 80% for hydrophobic VOCs at inlet concentrations of order of mg m−3 for EBRT = 7 sThe addition of silicone oil enhanced and stabilized biofiltration performance; high overall mass transfer coefficients reported for polyurethane foam[83]
Ethyl acetate, toluene, ethyl benzene, xylene, ethyl toluene, trimethylbenzene2 × 2 × 4 (rectangular bioreactor); 6 m3Ceramic particles, hollow plastic balls, Raschig ringsEnriched activated sludge from WWTPEBRT: 7.2 s
Cin: up to 150 mg m−3 (sum of VOCs)
BTF combined with photocatalytic oxidation revealed high performance for VOC removal from air[84]
Acetone, methyl ethyl ketone (MEK), toluene, styrene Field-scale stainless steel BTF: 1.5 × 1.5 × 1.8 [m]Pall ringsActivated sludge from the local paint and coating plant wastewater treatment plantEBRT: 14 s
Total inlet loading of VOCs mixture: 30.5 m m−3 h−1 RE from 67 to 99 %
Mixed VOCs from paint and coating plant; higher removal of acetone and MEK than toluene and styrene; shutdown periods diminished BTF performance, especially for hydrophobic VOCs[85]
Methanol, n-hexane0.076 × 1.3; 2.72Pelletized diatomaceous earth (Celite)Aerobic microbial culture from secondary clarifier of an activated sludge system at Cincinnati municipal wastewater treatment plantEBRT: 120 s
RE of methanol about 99% regardless process conditions; RE of n-hexane decreased from about 98% to 60% when inlet loading increased from 1 to 13.2 g m−3 h−1
Methanol significantly increased the removal of hexane in the BTF; switching of flow directions (co- and counter-current) was applied to ensure uniform biofilm formation across the packing; decrease in BTF performance was observed after bed backwashing due to high biomass loss; high methanol concentrations may have inhibited n-hexane biodegradation, as it is a more accessible carbon source[50]
Hydrogen sulfide, methanol, α-pinene0.094 × 0.7; 4.55Polypropylene pall ringsAutotrophic H2S-degrading culture and pure strains of Candida boidinii, Rhodococcus erythropolis, and Ophiostoma stenocerasEBRT: 26 s
Maximum elimination capacities of methanol, α-pinene and H2S were 302, 175, and 191 g m−3 h−1, respectively
The α-pinene degraders developed more slowly than degraders for methanol and H2S; diversified microbial population within the filter bed offered promising VOCs removal for industrial applications[57]
Butanone, toluene, α-pinene, hexane0.08 × 1; 4Kaldness K1 ringsActivated sludge from Valladolid WWTPEBRT: 6 s
RE for inlet concentrations of butanone (3 mg m−3); toluene (1.4 mg m−3); α-pinene (1.4 mg m−3) and hexane (1.3 mg m−3) were about 99%, 98%, 97% and 65%, respectively
A two-phase BTF showed better performance and supported a richer as well as more uniform microbial community in the biofilter bed than a single-phase BTF [86]
Methanol, n-hexane0.076 × 1.3; 2.72Pelletized diatomaceous earth (Celite)Activated sludge from WWTP as reported in [50]EBRT: 120 s; 3:1 and 5:1 ratios of methanol to n-hexane; maximum elimination capacity for n-hexane was 11.2 g m−3 h−1 for its inlet loading of 13.2 g m−3 h−1 when lower ratio of methanol was testedRemoval of methanol and hexane was favored in acidic conditions and offered high removal efficiencies; degradation of methanol was not affected by n-hexane[87]
Pelletized diatomaceous earth support media (Celite)
From previously working BTF
From previously working BTF
EBRT: 120 s
Acclimation with intermittent loading of methanol
Inlet loading of hexane from 21.5 to 47.7 g m−3 h−1
Acidic conditions (pH = 4) with a dominant fungi consortium offered higher performance than the process realized at pH = 7; a change in the inlet loading of VOCs may affect the microbial community, especially for high VOC loads[88]
EBRT: 120 s
Inlet loading: hexane 13.2, methanol 37.7 g m−3 h−1
Methanol introduction ameliorated the removal of hexane from air, a high reaction rate constant for hexane biofiltration was obtained for the alternate strategy of treatment air polluted with either hexane or its mixture with methanol; the reintroduction of methanol increased the removal of hexane, due to an increase in hexane bioavailability[89]
n-Hexane, benzene, methanol0.076 × 1.3; 2.72Pelletized diatomaceous earth support mediaFrom previously working BTF, pH = 4 favors fungal growthEBRT: 120 s
Total inlet loading of VOCs between 96.4 and 117.7 g m−3 h−1;
Results showed that the top section of BTF was the most active part of a biofilter where competition among VOC degraders took place; VOCs concentrations played a role in the recovery of biofilter performance after shut-down periods, as well as affected the removal efficiency of resistant hydrophobic VOCs, such as hexane[90]
Benzene, toluenePilot plant; 2.76 × 1; 6000Blue mussel shellsNatural from blue mussel shells and WWTP effluent used as a trickling liquidGas flow:
0.9 m3 h−1; inlet concentrations of benzene and toluene were 0.4–56 and 1.6–22.8 mg m−3, respectively
The two-stage process combining a water scrubber and BTF resulted in a high purification degree of air from WWTP [91]
Styrene, acetone0.15 × 1.5; 17.6Polypropylene pall ringsMixed culture from previously working biofilterEBRT: 53 sInvestigated system combined a BTF with a downstream conventional biofilter; acetone accumulated in the trickling liquid and hindered the biodegradation; recovery of the BTF performance was possible when acetone in liquid was degraded; a proposed two-stage biofiltration system is promising for the smoothing variation of concentrations or for the overloading of a biofilter unit[92]
Methanol, toluene, trichloroethylene (TCE)0.076 × 1.3; 2.72CeliteFungi; from previously working BTFpH = 4
EBRT: 120 s
Inlet loading: TCE (10–40 g m−3 h−1), methanol (22–237 g m−3 h−1), toluene (17–101 g m−3 h−1)
The elevated inlet loading of a primary substrate (i.e. carbon source like methanol) inhibited TCE removal; biofilters with lower ratios of methanol and toluene to TCE resulted in better removal of TCE, as well as better responses to increased TCE loads[93]
Trichloro-
ethylene (TCE), methanol
0.076 × 1.3; 2.72CeliteFungi; from previously working BTFpH = 4
EBRT: 120 s
Inlet concentrations: TCE (28–80 ppmv), methanol (103–711 ppmv)
Higher concentrations of methanol in the inlet stream allowed for the higher removal of TCE, but only for methanol concentrations below the inhibition level; TCE was more bioavailable for higher methanol concentrations; preferential degradation of hydrophilic over hydrophobic substrate was observed; elimination efficiency of methanol was not affected by variations in TCE concentration[94]
Benzene, toluene, ethylbenzene, xylene0.084 × 0.45; 2Kaldnes ringsActivated sludge from denitrification-nitrification section from Valladolid WWTPEBRT: 30 min
Inlet loading of xylene, toluene and ethylebenze was 1.4 g m−3 h−1 for each and 1.5 g m−3 h−1 for benzene
Oxygen-free conditions of biodegradation were applied; REs above 90% were noted for toluene and ethylbenzene (relative ease of biodegradation compared to benzene and xylene); removal of xylene was limited due to mass transfer limitations; benzene removal was poor due to toxic intermediate products of biodegradation which were overcome by the application of UV to trickling liquid; the coupling of BTF with a UV gas pre-treatment didn’t increase the removal of BTEX in the system; low similarity of inoculum and developed microbial community composition was detected[95]
Cyclohexane, methyl acetate0.11 × 1; 4.75Volcanic rock and ceramsiteActivated sludge from the secondary sedimentation tank from Beijing Gaobeidian WWTPEBRT: 88.3 s
Inlet concentrations of cyclohexane and methyl acetate are up to 250 and 800 mg m−3
REs of over 90% for both investigated VOCs[96]
Toluene, formaldehyde, benzo-a-pyrene0.079 × 1.05 VermiculiteFusarium solani B1 (fungi) and Rhodococcus erythropolis DSM 43066EBRT: 31 s
Inlet loadings: benzo-a-pyrene (373 g m−3 h−1), toluene (33.5 g m−3 h−1), formaldehyde (34.8 g m−3 h−1)
Elimination capacities for benzo-a-pyrene, toluene, and formaldehyde reached 215, 31, and 22.5 g m-3 h-1, respectively, for conditions indicated in reference conditions given to the left; the biofilter was divided into three sections, and the first stage was preferentially colonized mainly by fungi, while the other two were colonized with bacteria[39]
Chloroform, dichlorobromo-methane, ethanol0.076 × 1.3; 2.72CeliteFrom previously working BTFEBRT: 5 min
Inlet concentrations of trihalomethanes (total) 14 ppmv, inlet concentration of ethanol 25–200 ppm
The application of a co-metabolite (ethanol) resulted in a similar increase in the elimination of trihalomethanes as the use of a surfactant; natural surfactant was found to be more efficient in aiding the BTF performance than the synthetic one[97]
Toluene, ethyl benzene, p-xylene, m-xylene, o-xylene0.1 × 0.38; 3Ceramic particlesMixed acclimated microbial consortia from petroleum polluted soil EBRT: 60 s
Inlet concentrations of each VOC from 50 to 160 mg m−3
Continuous and discontinuous BTF feeding strategies were compared; continuous feeding resulted in the higher removal of VOCs than the discontinuous mode, while switching from discontinuous to continuous resulted in a great increase in process efficiency [98]
n-Butanol, cyclohexane0.08 × 0.68; 2.5Peat, perlite, ceramic Raschig ringsNo additional inoculation (only microbes naturally occurring in peat material)EBRT: 46–60 s
Inlet loading of n-butanol (19.8–99 g m−3 h−1) and cyclohexane (45–180 g m−3 h−1)
Presence of butanol increased the removal efficiency of cyclohexane, starvation episodes slightly decreased biofiltration performance, and the longer the starvation period, the longer time that was needed for the recovery of removal efficiency[99]
Ethanol, hexane0.08 × 0.68; 2.5Peat, perlite, ceramic Raschig ringsNo additional inoculation (only microbes naturally occurring in peat material)EBRT: 60 s
Inlet loading of ethanol (18.45–38.5 g m−3 h−1) and hexane (25–140 g m−3 h−1)
The presence of ethanol increased the removal efficiency of hexane; ethanol starvation caused only a slight decrease in removal efficiency of hexane; a lower than previously reported volume by volume ratio of ethanol ensured the efficient biofiltration of hexane[100]
Cyclohexane, ethanol0.08 × 0.68; 2.5Polyurethane foamCandida albicans, Candida subhashiiEBRT: 60 s
Inlet loadings of VOCs in the range from 36 to 90 g m−3 h−1
Feeding a BTF with a mixture of ethanol and cyclohexane resulted in the higher removal of cyclohexane compared to when ethanol was added to the gas stream polluted with cyclohexane only; negligible concentrations of VOCs were detected in the liquid phase in steady-state conditions[101]
n-Hexane, dichloromethaneNo informationBamboo charcoal based polyurethane foamActivated sludge from the pharmaceutical factory WWTPEBRT: 20–150 s
Inlet concentrations of n-hexane (100 mg m−3) and dichloromethane (150 mg m−3)
The applied packing material offered large and uniform pores, thus improving the mass transfer from gas to liquid phase; further, the modification of polyurethane foam with charcoal increased its porosity, resulting in greater surface area available for biofilm development; the packing offered higher resistance to transient conditions[42]
m-Xylene, toluene0.0885 × 0.7; 2.4Diatomaceous earth pelletsAcclimated activated sludgeEBRT: 25–60 s
Inlet concentrations of m-xylene (250–1500 mg m−3) and toluene (250 mg m−3)
The mechanisms of synergistic removal of m-xylene in the presence of toluene were discussed in terms of toxicity effects, co-metabolism, and suitable concentration proportions (1:2 for toluene to m-xylene)[49]
Toluene, hexane, α-pinene, trichloroethylene (TCE)0.1 × 0.34;Polyurethane foamCandida subhashiiEBRT: 30 s
Inlet concentrations of VOCs in the range from 200 to 450 mg m−3; removal efficiency in the range 20–45%
A biodegradation pattern for investigated VOCs was found: toluene > n-hexane > α-pinene > TCE; this sequence was discussed in terms of Hansen solubility parameters; biotrickling filter configuration proved to be more efficient in biodegradation of VOCs than the conventional biofilter [102]

4. Modeling of Biofiltration and Biotrickling Filtration Processes

The use of mathematical models can help to explore the phenomena occurring during various processes. The modeling of a bioprocess is a representation of the biological, chemical, and physical processes occurring in the bioreactor, and aims at the selection and optimization of several process parameters that can affect the process performance. Mathematical modeling helps in understanding the process, as well as aids in the reactor design and scaling up [69,103]. In fact, modeling is a means of translating concepts and observations into mathematical equations [104,105,106]. Process modeling should indicate or help to identify the expected course of the process as a result of changes introduced among the system variables.
The approaches towards the process modeling of conventional biofiltration and biotrickling filtration are different; however, they share several common features. According to the development of biological methods for air treatment, the modeling of biofiltration in conventional biofilters was initiated earlier than for biotrickling filters. The main differences in the modeling approaches result from differences in the process, reactor design, and operation. Conventional biofilters usually use organic media, and no regular flow of a liquid through a packing occurs. Additionally, a treated gas is pre-humidified prior to entering the biofilter. In biotrickling filters, inert materials are used as packing elements and a continuous or intermittent flow of liquid occurs through the packing. However, in practice, there are a mix of above-mentioned features in both biofilters and biotrickling filters. For example, biofilters require additional watering of the packing, and newly developed packing media are being applied [56,107]. Thus, various similar phenomena take place in both the BF and BTF, which allows for similar descriptions when using mathematical modeling [108].
Both the BF and BTF involve a complex system of various aspects of chemical, physical, and biological processes. Thus, plenty of elements must be taken into consideration when proposing a mathematical model of these processes, including mass balances for treated compounds, oxygen, and products of biodegradation (CO2, biomass, other intermediate products) with respect to different phases within a bioreactor i.e., the gas phase, the solid phase with biofilm, and the liquid phase [104]. Additionally, in order to set up some simplifications (model assumptions) and develop a reliable model, various problems and phenomena must be taken into account. Among others, these include oxygen limitation, estimation of available interfacial area, adsorption on the solid phase, biodegradation aspects with kinetic or mass transfer limitations, uniformity wetting of both packing and biofilms, and interactions between target compounds when VOCs mixtures are treated.
The modeling of conventional biofiltration is well described in the literature. However, the modeling of biotrickling filtration is distinctly different than that of conventional biofiltration. One of the available reviews of mathematical models for biotrickling filtration was published by Deshusses and Shareefdeen [104]. The authors revised the models for both conventional biofiltration and biotrickling filtration, and stressed the differences and complications regarding the modeling of biotrickling filtration. The authors state that, besides the fact that BTF includes one more phase to be modeled (i.e., liquid phase) than BF, some aspects of modeling the BTF may be even easier than for the BF. Shareefdeen and Singh [109] point out that the main differences that make the modeling of BTFs more complicated include the existence of the liquid phase, the increased complexity of the process, and the various possible mass transfer pathways (direct transfer of the target pollutant to the biofilm or indirect to the biofilm via liquid phase). On the other hand, due to the utilization of inert materials, such as packing elements, it is possible to determine the active surface area of the packing more easily and precisely. It is also possible to determine other parameters, such as the concentration in the liquid phase or in the biofilm, which are hardly measured for biological packing in conventional biofilters. Another advantage is related to the possibility of more in-depth biofilm analysis, due to its macroscopic character (biofilm is usually clearly visible).
However, this paper deals with biotrickling filtration. Due to the historical perspective, similarities, and initial bases for the development of a mathematical description of the biofiltration processes, an overview of conventional biofiltration modeling is also provided.

4.1. Modeling of Conventional Biofiltration

In accordance to the historical development and innovation in biological treatment of polluted gases, modeling of such processes started from the modeling of conventional biofiltration. The development of biofiltration models was directly connected to results from the progress of knowledge and understanding of complex phenomena occurring during the biofiltration process. The first mathematical model of biofiltration was proposed by Ottengraf and van den Oever [110]. This model served as the basis for many models of biofiltration, and the following assumptions are undertaken in this model:
  • target compounds (i.e., compounds removed from air in biofiltration system) are transported within the biofilm layer via diffusion, described by an effective diffusion coefficient;
  • the biofilm thickness is much smaller that the diameter of a bed particle;
  • the biodegradation of target compounds in the biofilter can be described by the Michaelis–Menten equation;
  • the flow of the gas phase is a plug flow;
  • mass transfer resistance at the gas side in the gas-biofilm interface is negligibly small.
Deshusses and Hamer [111] provided summary formulae for the Ottengraf and van den Oever model while considering three operational conditions that depend on the biodegradation kinetics. These operating situations are valid for the bottom of an expanded clay sphere-packed biofilter with a downward flow of gas phase:
C G C G , i n = exp ( h K 1 m i u G )
C G C G , i n = 1 ( h K 0 u G C G , i n )
C G C G , i n = [ 1 h u G · K 0 D e a 2 m i C G , i n δ ] 2
where a (m2 m−3) is an interfacial area per unit volume, CG (mg m−3) is a component concentration in the gas phase, De (m2 s−1) is an effective diffusion coefficient, h (m) is the height of the bed, in refers to the biofilter inlet, K0 (g m−3 s−1) and K1 (s−1) are zero- or first-order reaction rate constants, respectively, mi is a distribution coefficient of a component i, and uG (m s−1) is the superficial gas velocity.
Shareefdeen and Baltzis [112] stated that the assumptions from steady-state conditions cannot be directly applied to transient conditions, which is true both for conventional as well as biotrickling filters.
Based on the available literature, in 2002 Świsłowski [113] concluded that the biofiltration process, together with possibilities to predict the process performance, was valid for steady-state conditions only. Świsłowski postulated that much less data were available on the unsteady-state conditions, e.g., when the biofilter is undergoing process start-up and stabilization. An empirical relationship was proposed for the determination of the liquid–solid mass transfer of target compounds in the liquid phase of the biofilm.
Dupasquier et al. [114] investigated the effects of inlet gas mixture composition and possible co-metabolic mechanisms for the removal of methyl tert-butyl ether and pentane, and proposed a model describing the substrate / co-substrate biodegradation of VOC mixtures in a biofilter.
Deshuses and Shareefdeen [104] proposed a general division of biofiltration models into steady-state models and transient models. They examined the Ottengraf and van den Oever model with further modifications within the group of steady-state models. These models assume 1st or 0th order biodegradation that can be described by Michaelis–Menten kinetics. Additionally, two possible cases can be identified, depending on factors such as microbial consortia in the biofilter or treated compounds. The other group of models is the transient models. Shareefdeen and Baltzis [112] proposed a model taking into account the adsorption phenomena. They concluded that models describing steady-state conditions are not applicable to transient conditions, e.g., during the biofilter start-up or variations in process conditions, which happen frequently in practical applications. This is because of two problems: the adsorption and absorption of VOCs at the packing and in the liquid phase, as well as the problem of non-uniform biomass distributions, especially during the development of a biofilm during the start-up period.
Hodge and Devinny [115] developed a model describing basic transport and biological processes for a biofilter treating ethanol vapors. The model included a transfer of ethanol between air and solid/water phases, the biodegradation of ethanol, the resulting CO2 production, and the pH changes due to the accumulation of CO2. These pH changes are related to acidification, and are caused by the uptake of CO2 and its partial oxidation to acetic acid. A model incorporated axial dispersion and excluded the plug flow assumption, while no large-scale turbulence within a biofilter, homogeneity of filter material composition, uniform biomass distribution, first-order biodegradation kinetics, or stoichiometric production of CO2 from ethanol and oxygen were assumed. The following relationship for ethanol concentration in air phase was proposed:
C G = C G 0 e x p ( b 1 k m x Z )
where CG0 (mg cm−3) is the initial ethanol (i.e., substrate) concentration in the gas phase, b1 (h−1) is a first-order biodegradation rate constant, km is the ratio of a substrate mass in solids/water phase to its mass in the gas phase, x (cm) is a coordinate position in the biofilter, and Z (cm h−1) is the axial interstitial velocity in the gas phase.
In 1995, a new approach to biofilter modeling was proposed by Deshusses et al. [116]. This approach concentrated on dynamic conditions during the biofiltration process, and the model considered a biofilter as a bioreactor comprised of several finite sections, for each of which transient mass balances can be determined. Competitive inhibition kinetics were incorporated in the model and described by the following expression:
R s   j , n , w = V m j S j , n , w K m j ( 1 + I n , w K i ) + S j , n , w
where Rsj (kg m−3 s−1) is a degradation rate constant, Vm (kg m−3 s−1) is a maximum degradation rate, Sj (kg m−3) is the concentration of competitive substrate (i.e., biodegradation inhibitor), Kmj (kg m−3) is the Michaelis–Menten constant of component j, Ki (kg m−3) is the inhibition constant of I on the removal rate of j; and subscripts n and w denote biofilm/sorption volume subdivisions and biofilter layer subdivisions, respectively.
In 1997, Shareefdeen proposed a general transient biofilter model for single and multiple VOC removal from air, which included a so-called interactive kinetics assumption, multi-component adsorption phenomena, axial dispersion effects, and oxygen limitation. Amanullah et al. [117] proposed a model taking into account convection and dispersion phenomena in the gas phase, partial coverage of the solid support by the biofilm, interphase mass transfer between the gas phase and the aqueous biofilm with an equilibrium partition at the interface before diffusion, direct adsorption to the available uncovered solid adsorbent media, transfer between the biofilm and the solid support, and the biological reactions in both the biofilm and the adsorbent. The authors pointed out that the model can be successfully applied to model conventional biofiltration for a wider range of process parameters than previously proposed models. Additionally, it was postulated that a considerable role in the elimination of VOCs from the gas phase may be attributed to adsorption of support elements not covered by biofilm. The adsorbed compounds may then undergo biodegradation, or at least such adsorption may help in handling the fluctuating inlet loads.
The problems of biomass overgrowth and biofilter clogging were investigated and modeled by Ozis et al. [118]. This model of biofilm growth was based on the combination of percolation theory and a cellular automaton model.
Further development of biofiltration modeling includes increasingly broader use of software for the modeling and control of industrial-scale biofilters [104,119], and neural networks are being used for the development of models with dynamic conditions.
A model combining dynamic conditions during biofiltration with biomass growth and resulting pressure drops was proposed by Xi and co-workers [120]. Vaiškūnaitė and Zagorskis [121] developed a model for the treatment of VOC mixtures. Recently, the influences of substrate degradation and inhibition mechanisms were investigated and modeled [122].
A brief overview of models proposed for the description and prediction of conventional biofiltration processes, including the most important assumptions and simplifications, is given in Table 4.

4.2. Modeling of Biotrickling Filtration

Due to the complexity of various physical, chemical and biological processes that constitute biotrickling filtration, its mathematical modeling is challenging; however, such modeling may reveal important information about the biotrickling filtration mechanisms and process optimization strategies [103].
The beginning of biotrickling filtration modeling is related to works by Hartmans, Dick, and co-workers regarding the steady-state removal of dichlorobenzene [133,134]. Mpanias and Baltzis proposed a model taking into account kinetic and oxygen limitations during the biodegradation of VOCs [135]. Okkerse and co-workers proposed a dynamic model incorporating the possible acidification of a trickling liquid, as well as biomass accumulation and distribution along the biofilter [136]. Dynamic models, including the mass transfer limitations of hydrophobic VOCs in aqueous systems and non-uniform biofilm wetting, were then developed [137,138]. Further, models, including the substrate competition and the cross-inhibition of substrates during biotrickling filtration were developed [139,140]. The effects of possible absorption of VOCs by the liquid phase were included in the model proposed by Hernandez et al. [141]. The modeling of biotrickling filtration for mixtures of hydrophilic volatile organic compounds, including the effects of intermittent trickling patterns, were proposed by San-Valero and co-workers [103]. Recently, Kalantar et al. [142] proposed a dynamic model for the biotrickling filtration of mixed hydrophilic and hydrophobic VOCs. Brief overviews of state-of-the-art biotrickling filtration models for the time period before 2018 were provided by Sharvelle et al. [143], Ahmed et al. [144], San-Valero et al. [103], and Gąszczak et al. [145]. A few review descriptions on biotrickling filtration models are also available in other papers referred to in this work.
The modeling of the biotrickling filtration process, which is similar to conventional biofiltration, requires knowledge on the courses of single sub-processes, such as mass transfer of the target compound, oxygen limitation from the gas to liquid/biofilm phases, biodegradation kinetics, and hydrodynamic behavior of the bioreactor. San-Valero et al. [146] determined the mass transfer coefficients for isopropanol and oxygen for biotrickling filters packed with random or structured media, and proposed the use of a global mass transfer coefficient to describe the mass transfer phenomena, which included a simple methodology to aid in the mathematical modeling of the biotrickling filtration of hydrophilic VOCs.

4.3. Problems and Challenges for Modeling of Biotrickling Filtration

The following problems and challenges can be identified when modeling biotrickling filters:
  • non-uniform wetting of the packing and biofilm
  • oxygen limitation
  • biokinetic or mass transfer and diffusion limitations, depending on the substrate type
  • stripping of pollutants at the top of the biotrickling filter (in the vicinity of a trickling liquid sprayer)
  • existence of active and inactive biomass zones
  • determination or estimation of the gas-biofilm/liquid interface area
  • the need to include gas–liquid / gas–liquid–biofilm mass transfer as well as adsorption of VOCs in the solid phases; however, problems with adsorption may be only relevant for rare package materials, as most materials are made of plastics with low adsorption capacity
  • interactions between target compounds.
An important aspect of biotrickling filtration modeling is related to the description of biodegradation kinetics. Kinetics is one of the most important parameters affecting the process rate and performance, as well as expected model fitting to experimental data. In order to model the performance of a biotrickling filter, or any other biofilter type, it is necessary to assume a rate equation describing the consumption rate of the substrate by microorganisms [147]. This is an important step of model development. Mpanias and Baltzis proposed a model suitable for any VOC for which the biodegradation kinetics are known [135]. The following most important assumptions were undertaken in this model and are frequently valid for many other models: steady-state conditions are assumed; VOC and oxygen are transferred from the gas to liquid phase, and then from liquid to biofilm; biodegradation of the VOC takes place in the biofilm layer; the biodegradation rate is limited only by the concentrations of VOC and oxygen; the biodegradation takes place in the biolayer (bioactive layer) of the biofilm; anaerobic degradation of the VOC is not considered; concentrations of VOC and oxygen in gas/liquid phases follow Henry’s law; and the concentrations of VOC and oxygen in the biofilm are the same as in the liquid phase.
Assumptions undertaken in models for conventional biofilters and biotrickling filters are similar to some extent, e.g., kinetics and biodegradation description. The clearest difference in the approach to modeling of these processes is due to mass transfer from gas to biofilm phases. Usually, in biotrickling filters, it is assumed that the VOC is firstly transferred to the liquid phase, and then to the biofilm, which may result in, peak concentrations of the VOCs that can biodegrade, depending on the trickling mode (continuous or intermittent) [103].
The approach and means of BTF modeling have been developing, as indicated by selected milestones reported in Figure 1, as seen in the perspective of the overview of process modeling, which is given in detail in Table 5. As with conventional biofiltration, steady-state models were initially proposed, followed by the dynamic models, which were more suitable for the transient process conditions. A common feature during the development of each mathematical model of biotrickling filtration, especially for dynamic models, included the definition of mass balances for gas, biofilm, and possible liquid phases. These were the initial steps during the formulation of phenomenological models [69]. Model development also requires accepting the description of diffusion phenomena and biodegradation kinetics, as well as defining limitations (e.g., oxygen limitation). All this data are reflected in various model parameters, that either need to be experimentally determined for a specific system, or may be calculated or accepted based on the available literature. It must be stressed that several parameters may be compound-specific or biofilm/microbial-type specific, and thus require individual revision or calculation. Another important step of model development includes model calibration and validation by using either new or available experimental data. Upon the development of new models and the inclusion of more assumptions and limitations, the models became more and more complex, and seldom resulted in a single equation allowing for one result, such as the determination of VOC concentrations in treated gas leaving the biofilter. Due to the fact that the majority of models revised in Table 5 are phenomenological models, together with their complexity and usually individual character, specific model equations regarding computational values, such as mass balances, diffusion coefficients, or formulae for model functions were intentionally omitted in this review. In the case of interest in the specific model, the reader is kindly asked to refer to the original source of the literature reference.
The progress in BTF modeling has mainly included new assumptions to make the models more closely replicate practical/industrial applications and problems. However, among plenty of models, similar approaches are common, including those regarding the gas phase flow pattern, distinguishing the contacting phases and interfaces, and biodegradation kinetics descriptions. The last aspect, i.e., biodegradation kinetics, seems to be of special importance, due to the biological character of biotrickling filtration; thus, recent models include multiple substrate approaches or induction/inhibition effects [148,149]. Currently, the development of biotrickling filtration and process modeling is bieng observed in the field of biogas upgrading [150,151,152]. This reflects a close and reasonable relationship between scientific studies and the current need for the management and utilization of natural and renewable resources, for which biotrickling filtration can be of special importance.
Table 5. Overview of selected mathematical models for biotrickling filtration.
Table 5. Overview of selected mathematical models for biotrickling filtration.
Target Compound (s)Biofilter ParametersModel DescriptionReference
DichlorobenzeneVertical Perspex pipe (din = 0.29 m;
H = 1 m) packed with polypropylene elements (Filterpack) inoculated with adopted dichloromethane degrading microorganisms
Simulation of the biotrickling filter as a series of perfectly stirred interconnected reactors; computer software was used for calculating dichloromethane conversion using immobilized biocatalysts; the modeling gave surprisingly good results when compared to experimental data, however no mass transfer resistance between liquid and biofilm was included [134]
DichloromethaneCeramic saddles Steady-state model with the assumption of zeroth order kinetics; negligible resistance for the mass transfer of VOCs from gas to liquid phase; the model allowed for determining concentration profiles along co- and counter-current biofilter[133,153]
Theoretical modelNo experimental verification; modeling studyThe model incorporates simplified equations describing the gas flow and biodegradation; the model indicated two aspects of biotrickling filtration: the inclusion of both stripping in counter-current operation, and the influence of absorption/biodegradation in the liquid phase[154]
TolueneStainless-steel biotrickling filter
packed with Celite (pelletized diatomaceous earth) inoculated with activated sludge
A modification of a conventional model for synthetic-media biotrickling filters using one substrate, uniform biomass, and one type of microbial population; uniform pollutant concentrations in the gas phase are assumed; only gaseous and biofilm phases are considered [155]
m-ChlorobenzeneGlass column (din = 0.152 m;
H = 0.74 m) packed with Intalox ceramic saddles inoculated with a stable microbial consortium
The model incorporates oxygen limitation, non-uniform wetting of biofilm, and step-wise mass transfer of pollutants from the gas to liquid to biofilm phase; reactions occur only in the effective biolayer; no anaerobic degradation of VOCs is assumed; constant density of the biofilm is assumed[135]
Pentane, izobuteneA biotrickling filter (din = 0.0508 m;
H = 0.5 m) packed with a structured polyethylene packing inoculated with microbial consortium taken from previous studies
Steady-state model assuming a biofilm covered with a liquid phase; a two-step approach of modeling, including dependence on mass transfer and then biodegradation kinetics is proposed; results of the modeling suggest both kinetic and mass transfer limitations during the biotrickling filtration of pentane and izobutene[156]
Acetone, tolueneStainless-steel bioreactor packed with pelletize diatomaceous earth particles (H = 1.14 m) Conceptual model incorporating active and inactive regions of biomass as well as the simultaneous growth and detachment of biofilm; modeling results indicated high mass transfer resistance for the liquid phase[157]
Diethyl etherGlass BTF divided into 7 sections, din = 0.076 m and total height of packing H = 0.61 m; packing material: Celite (porous ceramic particles); inoculation with enriched microbial culture from activated sludge The model is a development of [157]; Glass BTF is divided into 7 sections, din = 0.076 m, and total height of packing H = 0.61 m; packing material: Celite (porous ceramic particles); assumes inoculation with enriched microbial culture from activated sludge[137]
DichloromethaneGlass fiber enforced plastic column, din = 0.396 m and H = 2.7 m; packed with a corrugated sheet 60° PVC cross-flow cooling packing, inoculated with a biomass suspension of Hyphomicrobium sp. GJ21A dynamic model assuming the development of active (i.e. dichloromethane-degrading microorganisms) as well as inactive biomass; dicholormethane undergoes mass transfer from the gas to liquid phase to biofilm phase; a model for the biodegradation of acidifying VOCs; the model predicts process performance and related gasphase and liquid VOC concentration profiles, as well as the effects of initial VOC concentration increases in a treated gas, like bed clogging[136]
Carbon disulfideAcrylic tube of din = 0.29 m and height of packing H = 1.2 m; packed with Pastdec 12060 structured PVC packing; inoculation with Thiobacilii consortia
Modeling of mass transfer from the gas to liquid to biofilm phase; a single criterion was proposed for identifying the rate-limiting step of biotrickling filtration; a model for describing substrate gas-phase and liquid concentrations and their effects on mass transfer; low substrate concentration in the liquid phase illustrates mass transfer limitations, while higher concentrations suggest kinetic limitation in the biofilm; two-parameter steady-state, one-species, one-directional and heterogeneous model; the model predicts substrate concentration profiles for co- and counter-current modes[158]
Diethyl etherA four-section BTF packed to the total height H = 0.122 m with Celite (pelletized diatomaceous earth), inoculated with ethyl ether acclimated and enriched aerobic culture, originating from activated sludgeConceptual model incorporating the reduction of the gas–liquid interface area and the increase in liquid velocity occurring during biomass growth and biofilm development; the paper presents very clear assumptions for the model; a dynamic model for biotrickling filtration was developed, with a focus on nutrient addition and the removal of clogging overgrowing biomass; the model included two effects of backwashing, i.e. biomass thickness reduction and related changes in the reactor specific surface area; the model allows for qualitative description of the biotrickling filtration process[159]
Chlorobenzene, ortho-chlorobenzeneGlass BTF column of din = 0.152 m and H = 0.79 m packed with Intalox ceramic saddles, inoculated with microbial community from previous studiesThe model is a development of [135] and includes cross-inhibition effects during the biofiltration of a mixture of chlorobenzene (with kinetic competition for biodegradation)[160]
Hydrogen sulfidePilot scale biofilter—insulated vertical fiberglass cylinder with
din = 0.608 m and H = 1.8 m, packed with lava rock particles, inoculum from activated sludge
Conceptual model incorporating detailed mass transfer phenomena, biofilm formation, and biomass decay; a dynamic model for biofiltration of real waste air streams from wastewater treatment plants[161]
Differential biotrickling filter filled with a open-pore polyurethane foam cube (side wall of 0.04 m) with already developed biofilm from previous processesConceptual model for predicting mass transfer phenomena; the model considers the wetting of packing, as well as gas/liquid/biofilm mass transfer issues; uniform biofilm thickness, non-uniform biofilm wetting, plug flow for gas phase; no reaction in the liquid is assumed; the model proved that the biofiltration of H2S may be highly limited by the external mass transfer [138]
Mono-chlorobenzenePerspex tube (din = 0.05 m; H = 1 m); coal particles on acrylic plastic mesh; inoculation with activated sludgeThe model utilizes Monod kinetics for the description of biodegradation; biodegradation is evaluated solely as an aerobic process; the plug flow of gas phase and negligible axial dispersion are assumed, the biofilter is regarded isothermic; the model predicts the concentration profile of the VOC in the gas phase, in the biofilm, and in the trickling liquid retained on solid particles; the model follows first-order kinetics[162]
Dimethyl sulfide and methanolPMMA biofilter (din = 0.101 m; H = 0.33 m); Packed with Nova inert i.e. porous silica packing, inoculated with activated sludgeModel assumptions: plug flow of gas phase, no mass transfer resistance, rectangular geometry for biofilm, methanol and DMS are the only growth substrates for microbes, microbial species degrading both substrates are the same; the model incorporates competitive and activation functions of methanol on DMS biofiltration and sets the basis for the description and modeling of VOC mixtures, and successfully predicted the feeding strategies for enhanced biofiltration performance[139]
StyreneGlass biofilter (din = 0.005 m, H = 0.41 m) packed with coconut coir inoculated with Pseudomonas putidaModel assumptions include: negligible internal and external mass transfer resistance, plug flow bioreactor, uniform distribution of microorganisms throughout the biofilter packing, equilibrium at gas–liquid/liquid–solid interfaces; a deterministic model based on the kinetic parameters is proposed for describing biofiltration behavior for styrene abatement[163]
CO2 as a model compound for absorption studiesTypical packing for biofilters and biotrickling filters, including: porous ceramic rings, Pall rings, polyurethane foam, Lava rock, compost mixtureThe paper presents an approach for determining and setting the reference values of mass transfer coefficients for various packing materials used in biofiltration/biotrickling filtration; the approach and obtained values are useful for modeling purposes for a broad range of biofilter configurations; this is an important contribution to further studies of biofiltration modeling[164,165]
Ammonia and hydrogen sulfideAcrylic tube (din = 0.1 m; H = 0.61 m); polypropylene, polystyrene or PVC packing elements The model combines the models described in [153,154]; the model describes co- and counter -current biotrickling filtration for the simultaneous treatment of gas and liquid contaminants; the model assumptions include a plug flow, no oxygen or nutrient limitation, constant biofilm density, substrate concentration at the liquid-biofilm interface is the same as in the bulk liquid, and adsorption, as well as direct transfer of substrate to the biofilm, are insignificant[143]
MethaneFeasibility study—no single experimental verification of the modelModeling of biotrickling filtration using methanotrophic bacteria; a general, theoretical model for the estimation of methane removal is proposed; the assumptions of model are taken from [155][166]
Ethyl acetate, toluene (separately) and their mixtureBiofilter with intermittently supplied liquid solution; biofilter
(din = 0.136 m, H = 0.95 m) packed with fibrous peat inoculated with pre-acclimated activated sludge
A dynamic model with the following assumptions: plug flow of gas phase, mass transfer is limited by the diffusion resistance in the liquid phase, adsorption phenomena are negligible, biofilm develops only on the surface of packing elements, inhibition in biodegradation is accounted for; cross-inhibition of ethyl acetate on the toluene removal is confirmed[140]
EthyleneTheoretical approachModeling with the use of genetic algorithms; inverse modeling is used to estimate values of biofiltration parameters[167]
TRS (total reduced sulfur)Theoretical approachSensitivity analysis on the mathematical model for biotrickling filtration of TRS revealed that biofilm properties and kinetics had the most important influence on the model behavior [168]
HexanePVC column (din = 0.083 m, H = 0.53 m); abiotic conditions were tested for determining mass transfer from gas to liquid phaseModeling resulting in an isomorphous equation describing the maximum fraction of VOC that can be transferred from the gas to liquid phase within the comparison of three types of bioreactors; the biotrickling filter working as a two-liquid phase system exhibited the highest mass transfer performance[141]
Hydrogen sulfidePilot scale biotrickling filter (din = 0.3 m, H = 1.1 m) packed with polyurethane foam cubes inoculated with WWTP activated sludgeA model describing biomass accumulation (the production of solids) related to the pressure head loss increase due to the occurrence of solids by sulfur transformations; the growth rate of biomass, measurements of pressure drops, biomass increase, effluent COD, and total solids were successfully modeled[169]
Nitric oxideA plexiglass cylinder (din = 0.08 m, H = 0.3 m) packed with porous ceramic particles inoculated with Chelatococcus daeguensis TAD 1The model incorporates gaseous chemical oxidation for NO removal during biotrickling filtration, which is novel in biotrickling filtration modeling; the assumptions include plug glow of a gas phase, planar geometry of the biofilm, NO as a single rate-limiting substrate, no excessive biomass and uniform distribution of the biofilm; Monod-type kinetics and mass transfer in the gas–biofilm interface are exploited[170]
Ethyl acetatePlexiglass column (din = 0.062 m, H = 1 m) packed with walnut shell, inoculated with Pseudomonas putida PTCC 1694Experimental model based on zeroth order kinetics with diffusion limitation, assuming gas plug flow, constant biofilm thickness, and diffusion following Fick’s law; paper presents optimal results of various model results for the experimental data[171]
Benzene, toluene, ethylbenzene, o-xylene (BTEX)Perspex pipe (din = 0.14 m, H = 0.6 m) packed with corn-cobs, inoculated with Bacillus sphaericusCDR (convection–diffusion reactor) model was used to generate VOC concentration profiles along the biofilter [172]
Hydrogen sulfideA theoretical modelA dynamic model as an extension of a differential reactor model by Kim and Deshusses [138]; the model includes mass balances in gas, liquid, wetted, and non-wetted biofilm; biotrickling filtration performance is mostly affected by H2S concentration and gas flow rate[144]
IsopropanolMethacrylate column (din = 0.144 m,
H = 1.2 m); two packing materials were compared: random and structured polymeric media
A model aimed at describing mass transfer properties in biotrickling filters for a hydrophilic compound, including the mass transfer of oxygen, which is a limiting factor in such conditions; the results showed that global mass transfer coefficients describes the mass transfer phenomena in biotrickling filters with the highest accuracy[146]
Hydrogen sulfideDifferential biotrickling filter A dynamic model for describing the biotrickling filtration of high loads of H2S; assumptions of the model include: biofilm is completely covered with a liquid, adsorption phenomena are negligible, biofilm is uniformly distributed over packing elements, biomass contains active and non-active parts; gas–liquid mass transfer is described by the gas–liquid mass transfer coefficient, the biodegradation kinetics of hydrogen sulfide follow the Haldane equation; the model reveals a high dependence of H2S biofiltration on oxygen mass transfer[173]
Hydrophilic VOCs: ethanol, ethyl acetate, 1-ethoxy-2-propanolTwo systems were investigated: a laboratory biotrickling filter (din = 0.144 m, H = 1 m) and an industrial biotrickling filter (packing volume: 49 m3) filled with polypropylene rings inoculated with activated sludgeA dynamic model for biotrickling filtration with intermittent trickling patterns; the liquid phase is considered mobile during spraying episodes and considered stagnant for non-spraying periods; it makes similar assumptions to [173], except for: biodegradation kinetics follow a Monod equation, the presence of biomass in the biofilter increases the resistance for mass transfer from the gas to liquid phase, oxygen limitation is assumed; the model predicts performance, outlet emission peaks, and decreases in concentrations of outlet gases after trickling stops[103]
MethanePVC column (din = 0.153 m, H = 1.2 m) packed with polyethylene rings inoculated with methanotrophic bacteria, i.e., Methylomicrobium album (ATCC 33003) and Methylocystis sp. (ATCC 49242)A dynamic mathematical model based on mass transfer and kinetic parameters; the model assumptions include: plug flow of gas phase with no axial dispersion, gas–biofilm equilibrium follows Henry’s law, diffusion is described by Fick’s law, biofilm develops only on the outer surface of packing elements, the surface properties of the biofilm are the same as those for water, uniform distribution and no accumulation of biomass, no adsorption of the pollutant[174]
StyreneLaboratory (din = 0.144 m, H = 1.23 m) and industrial biofilters (volume of packing: 0.6 m3), packed with polypropylene rings and structured packing, respectivelyA three-phase dynamic model accounting for the convection, mass transfer, diffusion, and biodegradation phenomena of a biotrickling filtration process; the model is based on mass balances of the above-listed phenomena; the model assumes gas phase plug flow, a negligible share of adsorption and axial dispersion pertaining to biofiltration performance, that biodegradation takes place only within the biofilm; a numerical solution to the model by Matlab software; a limiting factor for the biotrickling filtration of hydrophobic compounds is their water solubility [69]
StyrenePilot-scale stainless-steel biotrickling filter (din = 1.084 m, H = 3.51 m) packed with polypropylene Ralu rings inoculated with microorganisms (E-93486 strain)Simple one substrate model was proposed; the paper presents verification of a two-substrate and one-substrate model for styrene removal from air; oxygen is not a limiting factor for the biotreatment of styrene-polluted air; the simple one-substrate model gave the lowest values of mean relative error, and can be used for predicting the performance of styrene biotrickling filtration[145]
Hydrogen sulfideA full-scale SULPHUS system was applied using a structured packing media (din = 3m, H = 5 m)The paper revises and restates selected models for biodegradation kinetics in biofilms, as well as proposes novel assumptions to be included in the biofilter models; a broad set of data is fit to the existing models, and the most appropriate models for various ranges of H2S inlet concentrations are proposed[147]
Biogas upgradingPVC column (din = 0.03 m, H = 1 m) packed with open pore polyurethane foam inoculated with hydrogenotrophic methanogens (Methanospirillum hungatei and additional biomass suspension)The developed conceptual model includes the transport of hydrogen from the gas to wetted and non-wetted biofilm fractions; the model assumptions include: the uniform spread and thickness of biofilm, biofilm thickness remains constant and cannot be not fully wetted, dynamic variations in biofilm and adsorption on packing material are neglected, diffusion in the biofilm is described by Fick’s law, axial dispersion is considered (e.g., due to high substrate concentration), the most sensitive parameters of the model are the gas flow rate, specific surface area, biofilm thickness, and maximum reaction rate; the model can be used for the optimization of biotrickling filters for biogas upgrading,[172]
Toluene and methanolGlass column (din = 0.065 m, H = 0.8 m) packed with a mixture of pumice grains and HDPE pall rings inoculated with a consortium of microorganisms, including Pandoraea pnomenusa DSM 16536 and Ralstonia eutrohpa PTCC1615Dynamic model based on mass transfer through gas, liquid, and biofilm phases is developed; assumptions include neglected physical adsorption, plug flow for gas phase, no axial dispersion, planar biofilm geometry, neglected methanol stripping from liquid, and no oxygen limitation; the model was calibrated and validated; the model indicated the fractions of a biofilm / biotrickling filter that are active in the degradation of either toluene or methanol; sensitivity analyses were performed; model can predict the dynamics of VOC biotrickling filtration, together with the determination of kinetic constants[142]
Hydrogen sulfide (biogas upgrading)A laboratory-scale biotrickling filter (din = 0.0714 m, H = 0.7 m) packed with polypropylene Pall rings, inoculated with aerobic sludgeLiquid–gas mass transfer correlations are provided; a dynamic model for predicting physio-chemical and biological process during biogas desulfurization using a biotrickling filter; the model is a three-phase model that can describe biotrickling filter operation for variable H2S inlet loading[173]
COD—Chemical Oxygen Demand.

5. Concluding Remarks and Future Perspectives

Recent research results with a focus on process modeling for the treatment of air polluted with VOCs in biotrickling filters was presented and discussed. Additionally, due to the historical perspective on the development of biotrickling filtration and its modeling, a brief overview of models for conventional biofiltration was also presented. In the cases of both biological treatment methods mentioned above, modeling started with the description of process performance under steady-state conditions. However, due to dynamic changes in both waste gas flow rates and pollutant concentrations frequently faced in real-life applications, dynamic models for transient conditions were developed. Along with the development of biofiltration modeling, model assumptions and objectives were adjusted to the possible real/industrial conditions, including variable concentrations and compositions of a gas phase component, intermittent spraying patterns, non-uniform growth and distribution of biomass, and non-uniform wetting of a packing/biofilm. These developments resulted in the formulation of various mathematical expressions, which were usually complex and problem-specific, that can aid process design and optimization. However, besides considerable progress in the description of biotrickling filtration mechanisms, the challenges listed in 2005 by Devinny and Ramesh [108] still remain. These include a need for further studies on the effects of substrate interactions when treating VOC mixtures, the effects of substrate types on biodegradation in the biotrickling filter, and a need to predict and model biofilm shape and physico-chemical properties. Nevertheless, no universal model of biofiltration, both for conventional and biotrickling filtration, is available.
Current research on biotrickling filtration concentrates on the means of enhancing process performance, especially with respect to the treatment of hydrophobic volatile organic compounds. As indicated by San-Valero et al. [69], industrial biotrickling filters can present different performance behaviors from laboratory-scale systems, which are mainly due to fluctuating conditions with respect to gas flow rate, shut down periods, and variable inlet concentrations. This statement underlines the special need for models suitable for dynamic conditions during waste air biofiltration. Hopefully, the overview of recent research shows the increasing number of papers on studies of real waste gas streams or model mixtures containing multiple VOC mixtures that mimic real streams. The application of biotrickling filtration towards biogas upgrading is also rising in importance. This direction of studies and process development seems to be of special importance, and especially consider the current economic and energetic crisis in Europe and in the world.
The future of investigations on biotrickling filtration should still be concentrated on the scaling up process, and offer more industrial applications that produce more results of practical meaning. This is especially important from the perspective of process modeling, which need more general and simplified models for direct industrial applications. In light of sustainability requirements and circular economic interests, combining BTF processes with other processes, including those treating waste with waste, seems to be an unavoidable direction of development. New perspectives are open in terms of managing the by-products of biofiltration with the possibility for energy production, e.g., using microbial fuel cells and the utilization of waste streams from other processes, such as biofilter packing materials or green chemical additives to aid the performance of biofilters. More experimental data or pilot-scale tests are also needed for models that become valuable for practical industrial applications.

Funding

This research received no external funding.

Data Availability Statement

Data are available in a publicly accessible repository as well as in a publicly accessible repository that does not issue DOIs. All data referred to in this paper are accessible according to the list of references.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Wu, H.; Yin, Z.; Quan, Y.; Fang, Y.; Yin, C. Removal of methyl acrylate by ceramic-packed biotrickling filter and their response to bacterial community. Bioresour. Technol. 2016, 209, 237–245. [Google Scholar] [CrossRef] [PubMed]
  2. Anand, S.S.; Philip, B.K.; Mehendale, H.M. Volatile Organic Compounds. Encycl. Toxicol. Third Ed. 2014, 967–970. [Google Scholar] [CrossRef] [Green Version]
  3. Sbai, S.E.; Mejjad, N.; Norelyaqine, A.; Bentayeb, F. Air quality change during the COVID-19 pandemic lockdown over the Auvergne-Rhône-Alpes region, France. Air Qual. Atmos. Health 2021, 14, 617–628. [Google Scholar] [CrossRef] [PubMed]
  4. Conticini, E.; Frediani, B.; Caro, D. Can atmospheric pollution be considered a co-factor in extremely high level of SARS-CoV-2 lethality in Northern Italy? Environ. Pollut. 2020, 261, 114465. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, Y.; Deng, J.; Xie, S.; Wang, Z.; Dai, H. Catalytic removal of volatile organic compounds using ordered porous transition metal oxide and supported noble metal catalysts. Cuihua Xuebao/Chin. J. Catal. 2016, 37, 1193–1205. [Google Scholar] [CrossRef]
  6. Li, Y.; Chang, H.; Yan, H.; Tian, S.; Jessop, P.G. Reversible Absorption of Volatile Organic Compounds by Switchable-Hydrophilicity Solvents: A Case Study of Toluene with N, N-Dimethylcyclohexylamine. ACS Omega 2021, 6, 253–264. [Google Scholar] [CrossRef]
  7. Gil, R.R.; Ruiz, B.; Lozano, M.S.; Martín, M.J.; Fuente, E. VOCs removal by adsorption onto activated carbons from biocollagenic wastes of vegetable tanning. Chem. Eng. J. 2014, 245, 80–88. [Google Scholar] [CrossRef]
  8. Gupta, V.K.; Verma, N. Removal of volatile organic compounds by cryogenic condensation followed by adsorption. Chem. Eng. Sci. 2002, 57, 2679–2696. [Google Scholar] [CrossRef]
  9. Song, M.; Kim, K.; Cho, C.; Kim, D. Reduction of Volatile Organic Compounds (VOCs) Emissions from Laundry Dry-Cleaning by an Integrated Treatment Process of Condensation and Adsorption. Processes 2021, 9, 1658. [Google Scholar] [CrossRef]
  10. Poddar, T.K.; Majumdar, S.; Sirkar, K.K. Removal of VOCs from air by membrane-based absorption and stripping. J. Memb. Sci. 1996, 120, 221–237. [Google Scholar] [CrossRef]
  11. Bai, B.; Qiao, Q.; Li, J.; Hao, J. Progress in research on catalysts for catalytic oxidation of formaldehyde. Cuihua Xuebao/Chin. J. Catal. 2016, 37, 102–122. [Google Scholar] [CrossRef]
  12. Dobslaw, D.; Engesser, K.H.; Störk, H.; Gerl, T. Low-cost process for emission abatement of biogas internal combustion engines. J. Clean. Prod. 2019, 227, 1079–1092. [Google Scholar] [CrossRef]
  13. Wang, C.; Xi, J.Y.; Hu, H.Y.; Yao, Y. Effects of UV pretreatment on microbial community structure and metabolic characteristics in a subsequent biofilter treating gaseous chlorobenzene. Bioresour. Technol. 2009, 100, 5581–5587. [Google Scholar] [CrossRef] [PubMed]
  14. Sivachandiran, L.; Thevenet, F.; Rousseau, A. Isopropanol removal using MnXOY packed bed non-thermal plasma reactor: Comparison between continuous treatment and sequential sorption/regeneration. Chem. Eng. J. 2015, 270, 327–335. [Google Scholar] [CrossRef]
  15. Liang, Z.; Wang, J.; Zhang, Y.; Han, C.; Ma, S.; Chen, J.; Li, G.; An, T. Removal of volatile organic compounds (VOCs) emitted from a textile dyeing wastewater treatment plant and the attenuation of respiratory health risks using a pilot-scale biofilter. J. Clean. Prod. 2020, 253, 120019. [Google Scholar] [CrossRef]
  16. Rybarczyk, P.; Szulczyński, B.; Gębicki, J.; Hupka, J. Treatment of malodorous air in biotrickling filters: A review. Biochem. Eng. J. 2019, 141, 146–162. [Google Scholar] [CrossRef]
  17. Dobslaw, D.; Ortlinghaus, O. Biological Waste Air and Waste Gas Treatment: Overview, Challenges, Operational Efficiency, and Current Trends. Sustain. 2020, 12, 8577. [Google Scholar] [CrossRef]
  18. Barbusiński, K.; Urbaniec, K.; Kasperczyk, D.; Thomas, M. Biofilters versus bioscrubbers and biotrickling filters: State-of-the-art biological air treatment. In From Biofiltration to Promising Options in Gaseous Fluxes Biotreatment: Recent Developments, New Trends, Advances, and Opportunities; Elsevier: Amsterdam, The Netherlands, 2020; pp. 29–51. ISBN 9780128190647. [Google Scholar]
  19. Runye, Z.; Kennes, C.; Zhuowei, C.; Lichao, L.; Jianming, Y.; Jianmeng, C. Styrene removal in a biotrickling filter and a combined UV–biotrickling filter: Steady- and transient-state performance and microbial analysis. Chem. Eng. J. 2015, 275, 168–178. [Google Scholar] [CrossRef]
  20. Helbich, S.; Dobslaw, D.; Schulz, A.; Engesser, K.-H. Styrene and Bioaerosol Removal from Waste Air with a Combined Biotrickling Filter and DBD–Plasma System. Sustainability 2020, 12, 9240. [Google Scholar] [CrossRef]
  21. Cheng, Z.W.; Zhang, L.L.; Chen, J.M.; Yu, J.M.; Gao, Z.L.; Jiang, Y.F. Treatment of gaseous alpha-pinene by a combined system containing photo oxidation and aerobic biotrickling filtration. J. Hazard. Mater. 2011, 192, 1650–1658. [Google Scholar] [CrossRef]
  22. Zeng, P.; Li, J.; Liao, D.; Tu, X.; Xu, M.; Sun, G. Performance of a combined system of biotrickling filter and photocatalytic reactor in treating waste gases from a paint-manufacturing plant. Environ. Technol. (United Kingd.) 2016, 37, 237–244. [Google Scholar] [CrossRef]
  23. 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]
  24. Rene, E.R.; Špačková, R.; Veiga, M.C.; Kennes, C. Biofiltration of mixtures of gas-phase styrene and acetone with the fungus Sporothrix variecibatus. J. Hazard. Mater. 2010, 184, 204–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Cox, H.H.J.; Deshusses, M.A. Biotrickling Filters; Springer: Dordrecht, The Netherlands, 2001; pp. 99–131. [Google Scholar]
  26. Deshusses, M.A.; Gabriel, D. Bio trickling filter technology. In Biotechnology for Odor and Air Pollution Control; Springer: Berlin/Heidelberg, Germany, 2005; pp. 147–168. ISBN 3540233121. [Google Scholar]
  27. Sheoran, K.; Siwal, S.S.; Kapoor, D.; Singh, N.; Saini, A.K.; Alsanie, W.F.; Thakur, V.K. Air Pollutants Removal Using Biofiltration Technique: A Challenge at the Frontiers of Sustainable Environment. ACS Eng. Au 2022, 2, 378–396. [Google Scholar] [CrossRef] [PubMed]
  28. Lamprea Pineda, P.A.; Demeestere, K.; Toledo, M.; Van Langenhove, H.; Walgraeve, C. Enhanced removal of hydrophobic volatile organic compounds in biofilters and biotrickling filters: A review on the use of surfactants and the addition of hydrophilic compounds. Chemosphere 2021, 279, 130757. [Google Scholar] [CrossRef]
  29. Meena, M.; Sonigra, P.; Yadav, G. Biological-based methods for the removal of volatile organic compounds (VOCs) and heavy metals. Environ. Sci. Pollut. Res. 2020, 28, 2485–2508. [Google Scholar] [CrossRef] [PubMed]
  30. Schiavon, M.; Ragazzi, M.; Rada, E.C.; Torretta, V. Air pollution control through biotrickling filters: A review considering operational aspects and expected performance. Crit. Rev. Biotechnol. 2016, 36, 1143–1155. [Google Scholar] [CrossRef] [PubMed]
  31. Pachaiappan, R.; Cornejo-Ponce, L.; Rajendran, R.; Manavalan, K.; Femilaa Rajan, V.; Awad, F. A review on biofiltration techniques: Recent advancements in the removal of volatile organic compounds and heavy metals in the treatment of polluted water. Bioengineered 2022, 13, 8432–8477. [Google Scholar] [CrossRef]
  32. Lee, S.-H.; Kurade, M.B.; Jeon, B.-H.; Kim, J.; Zheng, Y.; Salama, E.-S. Water condition in biotrickling filtration for the efficient removal of gaseous contaminants. Crit. Rev. Biotechnol. 2021, 41, 1279–1296. [Google Scholar] [CrossRef]
  33. Kraakman, N.J.R.; Rocha-Rios, J.; Van Loosdrecht, M.C.M. Review of mass transfer aspects for biological gas treatment. Appl. Microbiol. Biotechnol. 2011, 91, 873–886. [Google Scholar] [CrossRef]
  34. Yang, C.; Qian, H.; Li, X.; Cheng, Y.; He, H.; Zeng, G.; Xi, J. Simultaneous Removal of Multicomponent VOCs in Biofilters. Trends Biotechnol. 2018, 36, 673–685. [Google Scholar] [CrossRef]
  35. Gomez, M.; De Oca, M.; Danila, V.; Zagorskis, A.; Januševičius, T. Effects of Water Content and Irrigation of Packing Materials on the Performance of Biofilters and Biotrickling Filters: A Review. Processes 2022, 10, 1304. [Google Scholar] [CrossRef]
  36. Marycz, M.; Brillowska-Dąbrowska, A.; Muñoz, R.; Gębicki, J. A state of the art review on the use of fungi in biofiltration to remove volatile hydrophobic pollutants. Rev. Environ. Sci. Biotechnol. 2022, 21, 225–246. [Google Scholar] [CrossRef]
  37. Gabriel, D.; Deshusses, M.A. Retrofitting existing chemical scrubbers to biotrickling filters for H2S emission control. Proc. Natl. Acad. Sci. USA 2003, 100, 6308–6312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ferdowsi, M.; Avalos Ramirez, A.; Jones, J.P.; Heitz, M. Elimination of mass transfer and kinetic limited organic pollutants in biofilters: A review. Int. Biodeterior. Biodegrad. 2017, 119, 336–348. [Google Scholar] [CrossRef]
  39. Vergara-Fernández, A.; Yánez, D.; Morales, P.; Scott, F.; Aroca, G.; Diaz-Robles, L.; Moreno-Casas, P. Biofiltration of benzo[A]pyrene, toluene and formaldehyde in air by a consortium of Rhodococcus erythropolis and Fusarium solani: Effect of inlet loads, gas flow and temperature. Chem. Eng. J. 2018, 332, 702–710. [Google Scholar] [CrossRef]
  40. Guo, J.; Gao, Q. Enhancement of ethylbenzene removal from contaminated gas and corresponding mechanisms in biotrickling filters by a biosurfactant from piggery wastewater. J. Environ. Manage. 2021, 277, 111411. [Google Scholar] [CrossRef]
  41. Zhanga, Y.; Denga, W.; Qina, Y.; Yanga, Z.; Liua, J.; Lia, J. Research on Simultaneous Removal of Cyclohexane and Methyl Acetate in Biotrickling Filters. In Proceedings of the 2nd International Conference of Recent Trends in Environmental Science and Engineering (RTESE’18), Niagara Falls, ON, Canada, 10–12 June 2018; Available online: https://avestia.com/RTESE2018_Proceedings/files/paper/RTESE_107.pdf (accessed on 31 October 2021).
  42. Zhao, M.; Hu, L.; Dai, L.; Wang, Z.; He, J.; Wang, Z.; Chen, J.; Hrynsphan, D.; Tatsiana, S. Bamboo charcoal powder-based polyurethane as packing material in biotrickling filter for simultaneous removal of n-hexane and dichloromethane. Bioresour. Technol. 2022, 345, 126427. [Google Scholar] [CrossRef]
  43. Sempere, F.; Gabaldón, C.; Martínez-Soria, V.; Marzal, P.; Penya-roja, J.M.; Javier Álvarez-Hornos, F. Performance evaluation of a biotrickling filter treating a mixture of oxygenated VOCs during intermittent loading. Chemosphere 2008, 73, 1533–1539. [Google Scholar] [CrossRef] [PubMed]
  44. Jiménez, L.; Arriaga, S.; Muñoz, R.; Aizpuru, A. Effect of extended and daily short-term starvation/shut-down events on the performance of a biofilter treating toluene vapors. J. Environ. Manage. 2017, 203, 68–75. [Google Scholar] [CrossRef] [PubMed]
  45. 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] [PubMed]
  46. Dobslaw, D.; Woiski, C.; Winkler, F.; Engesser, K.H.; Dobslaw, C. Prevention of clogging in a polyurethane foam packed biotrickling filter treating emissions of 2-butoxyethanol. J. Clean. Prod. 2018, 200, 609–621. [Google Scholar] [CrossRef]
  47. Baltzis, B.C.; Wojdyla, S.M.; Zarook, S.M. Modeling Biofiltration of VOC Mixtures under Steady-State Conditions. J. Environ. Eng. 1997, 123, 599–605. [Google Scholar] [CrossRef]
  48. Yao, X.; Shi, Y.; Wang, K.; Wang, C.; He, L.; Li, C.; Yao, Z. Highly efficient degradation of hydrogen sulfide, styrene, and m-xylene in a bio-trickling filter. Sci. Total Environ. 2022, 808, 152130. [Google Scholar] [CrossRef]
  49. Dou, X.; Liu, J.; Qi, H.; Li, P.; Lu, S.; Li, J. Synergistic removal of m-xylene and its corresponding mechanism in a biotrickling filter. Process Biochem. 2022, 118, 404–412. [Google Scholar] [CrossRef]
  50. Zehraoui, A.; Hassan, A.A.; Sorial, G.A. Effect of methanol on the biofiltration of n-hexane. J. Hazard. Mater. 2012, 219–220, 176–182. [Google Scholar] [CrossRef] [PubMed]
  51. Cheng, Y.; Li, X.; Liu, H.; Yang, C.; Wu, S.; Du, C.; Nie, L.; Zhong, Y. Effect of presence of hydrophilic volatile organic compounds on removal of hydrophobic n-hexane in biotrickling filters. Chemosphere 2020, 252, 126490. [Google Scholar] [CrossRef] [PubMed]
  52. Qian, H.; Cheng, Y.; Yang, C.; Wu, S.; Zeng, G.; Xi, J. Performance and biofilm characteristics of biotrickling filters for ethylbenzene removal in the presence of saponins. Environ. Sci. Pollut. Res. 2017 2530 2017, 25, 30021–30030. [Google Scholar] [CrossRef] [PubMed]
  53. Karre, A.; Jones, K.; Boswell, J.; Paca, J. Evaluation of VOC emissions control and opacity removal using a biological sequential treatment system for forest products applications. J. Chem. Technol. Biotechnol. 2012, 87, 797–805. [Google Scholar] [CrossRef]
  54. Jiade, W.; Jianmeng, C. Removal of dichloromethane from waste gases with a bio-contact oxidation reactor. Chem. Eng. J. 2006, 123, 103–107. [Google Scholar] [CrossRef]
  55. Dobslaw, D.; Schöller, J.; Krivak, D.; Helbich, S.; Engesser, K.H. Performance of different biological waste air purification processes in treatment of a waste gas mix containing tert-butyl alcohol and acetone: A comparative study. Chem. Eng. J. 2019, 355, 572–585. [Google Scholar] [CrossRef]
  56. Zhang, Y.; Liu, J.; Xing, H.; Li, J. Performance and fungal diversity of bio-trickling filters packed with composite media of polydimethylsiloxane and foam ceramics for hydrophobic VOC removal. Chemosphere 2020, 256, 127093. [Google Scholar] [CrossRef] [PubMed]
  57. López, M.E.; Rene, E.R.; Malhautier, L.; Rocher, J.; Bayle, S.; Veiga, M.C.; Kennes, C. One-stage biotrickling filter for the removal of a mixture of volatile pollutants from air: Performance and microbial community analysis. Bioresour. Technol. 2013, 138, 245–252. [Google Scholar] [CrossRef] [PubMed]
  58. Wu, C.H.; Shih, J.C.; Lin, C.W. Continuous production of power using microbial fuel cells with integrated biotrickling filter for ethyl acetate-contaminated air stream treatment. Int. J. Hydrog. Energy 2016, 41, 21945–21954. [Google Scholar] [CrossRef]
  59. Liu, S.H.; Fu, S.H.; Chen, C.Y.; Lin, C.W. Enhanced processing of exhaust gas and power generation by connecting mini-tubular microbial fuel cells in series with a biotrickling filter. Renew. Energy 2020, 156, 342–348. [Google Scholar] [CrossRef]
  60. Valdebenito-Rolack, E.; Díaz, R.; Marín, F.; Gómez, D.; Hansen, F. Markers for the Comparison of the Performances of Anoxic Biotrickling Filters in Biogas Desulphurisation: A Critical Review. Processes 2021, 9, 567. [Google Scholar] [CrossRef]
  61. Liu, S.H.; Tsai, S.L.; Lai, Y.R.; Lin, C.W.; Huang, Y.W. Improving the performance of biotrickling filter microbial fuel cells in treating exhaust gas by adjusting the oxygen content of the anode tank. Chemosphere 2021, 278, 130390. [Google Scholar] [CrossRef] [PubMed]
  62. Korth, B.; Pous, N.; Hönig, R.; Haus, P.; Corrêa, F.B.; Nunes da Rocha, U.; Puig, S.; Harnisch, F. Electrochemical and Microbial Dissection of Electrified Biotrickling Filters. Front. Microbiol. 2022, 13, 1512. [Google Scholar] [CrossRef] [PubMed]
  63. Szulczyński, B.; Rybarczyk, P.; Gospodarek, M.; Gębicki, J. Biotrickling filtration of n-butanol vapors: Process monitoring using electronic nose and artificial neural network. Mon. Fur Chem. 2019, 150, 1667–1673. [Google Scholar] [CrossRef] [Green Version]
  64. Mirmohammadi, M.; Sotoudeheian, S.; Bayat, R. Triethylamine removal using biotrickling filter (BTF): Effect of height and recirculation liquid rate on BTFs performance. Int. J. Environ. Sci. Technol. 2017, 14, 1615–1624. [Google Scholar] [CrossRef]
  65. Aguirre, A.; Bernal, P.; Maureira, D.; Ramos, N.; Vásquez, J.; Urrutia, H.; Gentina, J.C.; Aroca, G. Biofiltration of trimethylamine in biotrickling filter inoculated with Aminobacter aminovorans. Electron. J. Biotechnol. 2018, 33, 63–67. [Google Scholar] [CrossRef]
  66. Palanisamy, K.; Mezgebe, B.; Sorial, G.A.; Sahle-Demessie, E. Biofiltration of Chloroform in a Trickle Bed Air Biofilter Under Acidic Conditions. Water Air Soil Pollut. 2016, 227, 478. [Google Scholar] [CrossRef] [PubMed]
  67. Mezgebe, B.; Palanisamy, K.; Sorial, G.A.; Sahle-Demessie, E.; Hassan, A.A.; Lu, J. Comparative Study on the Performance of Anaerobic and Aerobic Biotrickling Filter for Removal of Chloroform. Environ. Eng. Sci. 2018, 35, 462–471. [Google Scholar] [CrossRef]
  68. San-Valero, P.; Dorado, A.D.; Martínez-Soria, V.; Gabaldón, C. Biotrickling filter modeling for styrene abatement. Part 1: Model development, calibration and validation on an industrial scale. Chemosphere 2018, 191, 1066–1074. [Google Scholar] [CrossRef] [PubMed]
  69. Chuang, F.-J.; Chou, M.-S.; Chang, H.-Y. Biotrickling filtration of airborne styrene: A comparison of filtration media. J. Air Waste Manag. Assoc. 2018, 68, 369–376. [Google Scholar] [CrossRef]
  70. Li, K.; Zhou, J.; Wang, L.; Mao, Z.; Xu, R. The styrene purification performance of biotrickling filter with toluene-styrene acclimatization under acidic conditions. J. Air Waste Manag. Assoc. 2019, 69, 944–955. [Google Scholar] [CrossRef] [PubMed]
  71. Zhang, Y.; Liu, J.; Qin, Y.; Yang, Z.; Cao, J.; Xing, Y.; Li, J. Performance and microbial community evolution of toluene degradation using a fungi-based bio-trickling filter. J. Hazard. Mater. 2019, 365, 642–649. [Google Scholar] [CrossRef]
  72. Zhang, Y.; Liu, J.; Li, J.; Yue, T. Effects of filler voidage on pressure drop and microbial community evolution in fungal bio-trickling filters. Chemosphere 2021, 273, 129710. [Google Scholar] [CrossRef]
  73. De Vela, R.J.L.; Gostomski, P.A. Design and Performance of a Toluene-Degrading Differential Biotrickling Filter as an Alternative Research Tool to Column Reactors. J. Environ. Eng. 2020, 147, 04020159. [Google Scholar] [CrossRef]
  74. Salamanca, D.; Dobslaw, D.; Engesser, K.-H. Removal of cyclohexane gaseous emissions using a biotrickling filter system. Chemosphere 2017, 176, 97–107. [Google Scholar] [CrossRef]
  75. Vergara-Fernández, A.; Hernández, S.; Revah, S. Elimination of hydrophobic volatile organic compounds in fungal biofilters: Reducing start-up time using different carbon sources. Biotechnol. Bioeng. 2011, 108, 758–765. [Google Scholar] [CrossRef] [PubMed]
  76. Botros, M.M.; Hassan, A.A.; Sorial, G.A.; Botros, M.M.; Hassan, A.A.; Sorial, G.A. Role of Fungal Biomass in N-Hexane Biofiltration. Adv. Microbiol. 2017, 7, 673–688. [Google Scholar] [CrossRef] [Green Version]
  77. Yu, G.; Wang, G.; Wang, S.; Yang, C.; Chen, H.; Zhu, Y.; Yu, L.; Li, J.; Kazemian, H. Performance promotion and its mechanism for n-hexane removal in a lab-scale biotrickling filter with reticular polyurethane sponge under intermittent spraying mode. Process Saf. Environ. Prot. 2021, 152, 654–662. [Google Scholar] [CrossRef]
  78. Den, W.; Huang, C.; Li, C.H. Effects of cross-substrate interaction on biotrickling filtration for the control of VOC emissions. Chemosphere 2004, 57, 697–709. [Google Scholar] [CrossRef] [PubMed]
  79. Den, W.; Ravindran, V.; Pirbazari, M. Photooxidation and biotrickling filtration for controlling industrial emissions of trichloroethylene and perchloroethylene. Chem. Eng. Sci. 2006, 61, 7909–7923. [Google Scholar] [CrossRef]
  80. Lakey, M.; Manager Victoria, G.; Pitt, M. 36th Annual Qld Water Industry Operations Workshop, Dual Phase Biotrickling Filter Treatment of H2S & VOC’S; 2011. In Proceedings of the 36th Annual Qld Water Industry Operations Workshop, Clive Berghofer Recreation Centre, USQ, Toowoomba, Australia, 31 May–2 June 2011; Available online: http://www.wioa.org.au/conference_papers/2011_qld/documents/Matthew_Lakey.pdf (accessed on 19 June 2022).
  81. Balasubramanian, P.; Philip, L.; Murty Bhallamudi, S. Biotrickling filtration of complex pharmaceutical VOC emissions along with chloroform. Bioresour. Technol. 2012, 114, 149–159. [Google Scholar] [CrossRef]
  82. Lebrero, R.; Rodríguez, E.; Estrada, J.M.; García-Encina, P.A.; Muñoz, R. Odor abatement in biotrickling filters: Effect of the EBRT on methyl mercaptan and hydrophobic VOCs removal. Bioresour. Technol. 2012, 109, 38–45. [Google Scholar] [CrossRef] [PubMed]
  83. He, Z.; Li, J.; Chen, J.; Chen, Z.; Li, G.; Sun, G.; An, T. Treatment of organic waste gas in a paint plant by combined technique of biotrickling filtration with photocatalytic oxidation. Chem. Eng. J. 2012, 200–202, 645–653. [Google Scholar] [CrossRef]
  84. Li, J.; Ye, G.; Sun, D.; An, T.; Sun, G.; Liang, S. Performance of a biotrickling filter in the removal of waste gases containing low concentrations of mixed VOCs from a paint and coating plant. Biodegradation 2012, 23, 177–187. [Google Scholar] [CrossRef] [PubMed]
  85. Lebrero, R.; Rodríguez, E.; Pérez, R.; García-Encina, P.A.; Muñoz, R. Abatement of odorant compounds in one- and two-phase biotrickling filters under steady and transient conditions. Appl. Microbiol. Biotechnol. 2013, 97, 4627–4638. [Google Scholar] [CrossRef] [PubMed]
  86. Zehraoui, A.; Hassan, A.A.; Sorial, G.A. Biological treatment of n-hexane and methanol in trickle bed air biofilters under acidic conditions. Biochem. Eng. J. 2013, 77, 129–135. [Google Scholar] [CrossRef]
  87. Zehraoui, A.; Wendell, D.; Sorial, G.A. Treatment of hydrophobic VOCs in trickling bed air biofilter: Emphasis on long-term effect of initial alternate use of hydrophilic VOCs and microbial species evolution. J. Air Waste Manage. Assoc. 2014, 64, 1102–1111. [Google Scholar] [CrossRef] [Green Version]
  88. Zehraoui, A.; Kapoor, V.; Wendell, D.; Sorial, G.A. Impact of alternate use of methanol on n-hexane biofiltration and microbial community structure diversity. Biochem. Eng. J. 2014, 85, 110–118. [Google Scholar] [CrossRef]
  89. Zehraoui, A.; Wendell, D.; Sorial, G.A. Biodegradation of a ternary mixture of hydrophobic and hydrophilic VOCs in trickle bed air biofilters. Water. Air. Soil Pollut. 2014, 225, 1–14. [Google Scholar] [CrossRef]
  90. Viotti, P.; Schiavon, M.; Gavasci, R.; Capodaglio, A.G. Remoção de benzeno e tolueno de um efluente de refinaria por absorção e filtração "biotrickling". Rev. Ambient. E Agua 2015, 10, 720–727. [Google Scholar] [CrossRef] [Green Version]
  91. Vanek, T.; Halecky, M.; Paca, J.; Zapotocky, L.; Gelbicova, T.; Vadkertiova, R.; Kozliak, E.; Jones, K. A two-stage combined trickle bed reactor/biofilter for treatment of styrene/acetone vapor mixtures. J. Environ. Sci. Health—Part A Toxic/Hazard. Subst. Environ. Eng. 2015, 50, 1148–1159. [Google Scholar] [CrossRef]
  92. Chheda, D.; Sorial, G.A. Effect of a Ternary Mixture of Volatile Organic Compounds on Degradation of TCE in Biotrickling Filter Systems. Water. Air. Soil Pollut. 2016, 227, 1–11. [Google Scholar] [CrossRef]
  93. Chheda, D.; Sorial, G.A. Evaluation of co-metabolic removal of trichloroethylene in a biotrickling filter under acidic conditions. J. Environ. Sci. (China) 2017, 57, 54–61. [Google Scholar] [CrossRef] [Green Version]
  94. Akmirza, I.; Pascual, C.; Carvajal, A.; Pérez, R.; Muñoz, R.; Lebrero, R. Anoxic biodegradation of BTEX in a biotrickling filter. Sci. Total Environ. 2017, 587–588, 457–465. [Google Scholar] [CrossRef]
  95. Mezgebe, B.; Sorial, G.; Wendell, D.; Sahle-Demessie, E. Effectiveness of biosurfactant for the removal of trihalomethanes by biotrickling filter. Eng. Rep. 2019, 1, e12031. [Google Scholar] [CrossRef]
  96. Feng, R.; Xu, M.; Li, J.; Huang, S.; Zhao, G.; Tu, X.; Sun, G.; Guo, J. Structure and predictive functional profiling of microbial communities in two biotrickling filters treated with continuous/discontinuous waste gases. AMB Express 2019, 9, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Rybarczyk, P.; Szulczyński, B.; Gospodarek, M.; Gębicki, J. Effects of n-butanol presence, inlet loading, empty bed residence time and starvation periods on the performance of a biotrickling filter removing cyclohexane vapors from air. Chem. Pap. 2019, 74, 1039–1047. [Google Scholar] [CrossRef] [Green Version]
  98. Rybarczyk, P.; Szulczyński, B.; Gębicki, J. Simultaneous Removal of Hexane and Ethanol from Air in a Biotrickling Filter—Process Performance and Monitoring Using Electronic Nose. Sustainability 2020, 12, 387. [Google Scholar] [CrossRef] [Green Version]
  99. Rybarczyk, P.; Marycz, M.; Szulczyński, B.; Brillowska-Dąbrowska, A.; Rybarczyk, A.; Gębicki, J. Removal of cyclohexane and ethanol from air in biotrickling filters inoculated with Candida albicans and Candida subhashii. Arch. Environ. Prot. 2021, 47, 26–34. [Google Scholar] [CrossRef]
  100. Marycz, M.; Rodríguez, Y.; Gębicki, J.; Muñoz, R. Systematic comparison of a biotrickling filter and a conventional filter for the removal of a mixture of hydrophobic VOCs by Candida subhashii. Chemosphere 2022, 306, 135608. [Google Scholar] [CrossRef]
  101. San-Valero, P.; Penya-Roja, J.M.; Javier Álvarez-Hornos, F.; Marzal, P.; Gabaldón, C. Dynamic Mathematical Modelling of the Removal of Hydrophilic VOCs by Biotrickling Filters. OPEN ACCESS Int. J. Environ. Res. Public Health 2015, 12, 746–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Deshusses, M.A.; Shareefdeen, Z. Modeling of Biofilters and Biotrickling Filters for Odor and VOC Control Applications. Biotechnol. Odor Air Pollut. Control 2005, 213–231. [Google Scholar] [CrossRef]
  103. Kucharska, K.; Hołowacz, I.; Konopacka-Łyskawa, D.; Rybarczyk, P.; Kamiński, M. Key issues in modeling and optimization of lignocellulosic biomass fermentative conversion to gaseous biofuels. Renew. Energy 2018, 129, 384–408. [Google Scholar] [CrossRef]
  104. Fedailaine, M.; Moussi, K.; Khitous, M.; Abada, S.; Saber, M.; Tirichine, N. Modeling of the Anaerobic Digestion of Organic Waste for Biogas Production. Procedia Comput. Sci. 2015, 52, 730–737. [Google Scholar] [CrossRef] [Green Version]
  105. Lebrero, R.; Rodríguez, E.; Collantes, M.; De Juan, C.; Norden, G.; Rosenbom, K.; Muñoz, R.; Maria, L.; Capelli, T.; Invernizzi, M. Comparative Performance Evaluation of Commercial Packing Materials for Malodorants Abatement in Biofiltration Comparative Performance Evaluation of Commercial Packing Materials for Malodorants Abatement in. Biofiltration. Appl. Sci 2021, 2021, 2966. [Google Scholar] [CrossRef]
  106. Devinny, J.S.; Ramesh, J. A phenomenological review of biofilter models. Chem. Eng. J. 2005, 113, 187–196. [Google Scholar] [CrossRef]
  107. Shareefdeen, Z.; Singh, A. Biotechnology for Odor and Air Pollution Control; Springer: Berlin/Heidelberg, Germany, 2005; ISBN 3540233121. [Google Scholar]
  108. Ottengraf, S.P.P.; Oever, A.H.C. Van Den Kinetics of organic compound removal from waste gases with a biological filter. Biotechnol. Bioeng. 1983, 25, 3089–3102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Deshusses, M.A.; Hamer, G. Bioprocess Engineering The removal of volatile ketone mixtures from air in biofilters. Bioprocess Eng. 1993, 9, 141–146. [Google Scholar] [CrossRef]
  110. Shareefdeen, Z.; Baltzis, B.C. Biofiltration of toluene vapor under steady-state and transient conditions: Theory and experimental results. Chem. Eng. Sci. 1994, 49, 4347–4360. [Google Scholar] [CrossRef]
  111. Świsłowski, M. Removal of organic substances vapors from air in a biofilter with natural packings (in Polish). Ph.D. Dissertation, Gdansk University of Technology, Gdansk, Poland, 2002. [Google Scholar]
  112. Dupasquier, D.; Revah, S.; Auria, R. Biofiltration of methyl tert-butyl ether vapors by cometabolism with pentane: Modeling and experimental approach. Environ. Sci. Technol. 2002, 36, 247–253. [Google Scholar] [CrossRef]
  113. Hodge, D.S.; Devinny, J.S. Modeling Removal of Air Contaminants by Biofiltration. J. Environ. Eng. 1995, 121, 21–32. [Google Scholar] [CrossRef]
  114. Deshusses, M.A.; Hamer, G.; Dunn, I.J. Behavior of Biofilters for Waste Air Biotreatment. 1. Dynamic Model Development. Environ. Sci. Technol 1995, 29, 1048–1058. [Google Scholar] [CrossRef]
  115. Amanullah, M.; Farooq, S.; Viswanathan, S. Modeling and Simulation of a Biofilter. Ind. Eng. Chem. Res. 1999, 38, 2765–2774. [Google Scholar] [CrossRef]
  116. Ozis, F.; Bina, A.; Devinny, J.S. Biofilm Growth-Percolation Models and Channeling in Biofilter Clogging. J. Air Waste Manage. Assoc. 2007, 57, 882–892. [Google Scholar] [CrossRef] [Green Version]
  117. Optimization of Biofiltration for Odor Control: Model Calibration, Validation, and Applications on JSTOR. Available online: https://www.jstor.org/stable/25045568 (accessed on 18 August 2022).
  118. Xi, J.; Kang, I.; Hu, H.; Zhang, X. A biofilter model for simultaneous simulation of toluene removal and bed pressure drop under varied inlet loadings. Front. Environ. Sci. Eng. 2014 93 2014, 9, 554–562. [Google Scholar] [CrossRef]
  119. Vaiškūnaitė, R.; Zagorskis, A. Modelling of a biofiltration process of volatile organic compound mixtures in a biofilter. Biotechnol. Biotechnol. Equip. 2016, 30, 1217–1222. [Google Scholar] [CrossRef] [Green Version]
  120. Süß, M.; Visscher, A. De Experimental and numerical study of steady state stability in a toluene biodegrading biofilter. Sci. Rep. 2022, 12, 12510. [Google Scholar] [CrossRef]
  121. Shareefdeen, Z.; Baltzis, B.C.; Oh, Y.-S.; Bartha, R. Biofiltration of methanol vapor. Biotechnol. Bioeng. 1993, 41, 512–524. [Google Scholar] [CrossRef] [PubMed]
  122. Shareefdeen, Z.; Baltzis, B.C. Biological Removal of Hydrophobic Solvent Vapors from Airstreams. Adv. Bioprocess Eng. 1994, 397–404. [Google Scholar] [CrossRef]
  123. Deshusses, M.A.; Hamer, G.; Dunn, I.J. Behavior of Biofilters for Waste Air Biotreatment 2. Experimental Evaluation of a Dynamic Model. Environ. Sci. Technol 1995, 29, 1059–1068. [Google Scholar] [CrossRef] [PubMed]
  124. Zarook, S.M.; Shaikh, A.A.; Ansar, Z. Development, experimental validation and dynamic analysis of a general transient biofilter model. Chem. Eng. Sci. 1997, 52, 759–773. [Google Scholar] [CrossRef]
  125. Mohseni, M.; Allen, D.G. Biofiltration of mixtures of hydrophilic and hydrophobic volatile organic compounds. Chem. Eng. Sci. 2000, 55, 1545–1558. [Google Scholar] [CrossRef]
  126. Raghuvanshi, S.; Babu, B.V. Modeling and Simulation of Biofilters Operated in Periodic Mode. Available online: https://www.researchgate.net/publication/252695298_Modeling_and_Simulation_of_Biofilters_Operated_in_Periodic_Mode (accessed on 19 June 2022).
  127. Rene, E.R.; Veiga, M.C.; Kennes, C. Experimental and neural model analysis of styrene removal from polluted air in a biofilter. J. Chem. Technol. Biotechnol. 2009, 84, 941–948. [Google Scholar] [CrossRef] [Green Version]
  128. Ranjbar, S.; Ghaemi, A. Mathematical Modeling for Volatile Organic Compounds Removal in a Biofilter: Model Validation and Sensitivity Analysis. Iran. J. Chem. Eng. 2013, 10, 77. [Google Scholar]
  129. Meena, V.; Rajendran, L.; Kumar, S.; Jansi Rani, P.G. Mathematical modeling of gas phase and biofilm phase biofilter performance. Egypt. J. Basic Appl. Sci. 2016, 3, 94–105. [Google Scholar] [CrossRef] [Green Version]
  130. Salih, M.; Shareefdeen, Z.; Khouri, S. A novel approach to the solution of a steady state biofilter model. Environ. Eng. Res. 2020, 25, 779–787. [Google Scholar] [CrossRef]
  131. Diks, R.M.M.; Ottengraf, S.P.P. Verification studies of a simplified model for the removal of dichloromethane from waste gases using a biological trickling filter. Bioprocess Eng. 1991, 6, 93–99. [Google Scholar] [CrossRef] [Green Version]
  132. Hartmans, S.; Tramper, J. Dichloromethane removal from waste gases with a trickle-bed bioreactor. Bioprocess Eng. 1991, 6, 83–92. [Google Scholar] [CrossRef]
  133. Mpanias, C.J.; Baltzis, B.C. An Experimental and Modeling Study on the Removal of Mono-Chlorobenzene Vapor in Biotrickling Filters. Biotechnol Bioeng 1998, 59, 328–343. [Google Scholar] [CrossRef]
  134. Okkerse, W.J.H.; Ottengraf, S.P.P.; Osinga-Kuipers, B.; Okkerse, M. Biomass Accumulation and Clogging in Biotrickling Filters for Waste Gas Treatment. Evaluation of a Dynamic Model Using Dichloromethane as a Model Pollutant. 1999, 63, 418–430. [Google Scholar] [CrossRef]
  135. Zhu, X.; Alonso, C.; Suidan, M.T.; Cao, H.; Kim, B.J.; Kim, B.R. The effect of liquid phase on VOC removal in trickle-bed biofilters. Water Sci. Technol. 1998, 38, 315–322. [Google Scholar] [CrossRef]
  136. Kim, S.; Deshusses, M.A. Development and experimental validation of a conceptual model for biotrickling filtration of H2S. Environ. Prog. 2003, 22, 119–128. [Google Scholar] [CrossRef]
  137. Zhang, Y.; Liss, S.N.; Allen, D.G. Modeling the Biofiltration of Dimethyl Sulfide in the Presence of Methanol in Inorganic Biofilters at Steady State. Biotechnol. Prog. 2008, 24, 845–851. [Google Scholar] [CrossRef] [PubMed]
  138. Álvarez-Hornos, F.J.; Gabaldón, C.; Martínez-Soria, V.; Marzal, P.; Penya-roja, J.M. Mathematical modeling of the biofiltration of ethyl acetate and toluene and their mixture. Biochem. Eng. J. 2009, 43, 169–177. [Google Scholar] [CrossRef]
  139. Hernández, M.; Quijano, G.; Muñoz, R.; Bordel, S. Modeling of VOC mass transfer in two-liquid phase stirred tank, biotrickling filter and airlift reactors. Chem. Eng. J. 2011, 172, 961–969. [Google Scholar] [CrossRef]
  140. Kalantar, M.; Zamir, S.M.; Ferdowsi, M.; Shojaosadati, S.A. Removal of toluene in a biotrickling filter in the presence of methanol vapors: Experimental study, mathematical modeling, and kinetic parameters optimization. J. Environ. Chem. Eng. 2021, 9, 104617. [Google Scholar] [CrossRef]
  141. Sharvelle, S.; Arabi, M.; McLamore, E.; Banks, M.K. Model Development for Biotrickling Filter Treatment of Graywater Simulant and Waste Gas. I. J. Environ. Eng. 2008, 134, 813–825. [Google Scholar] [CrossRef]
  142. Ahmed, W.; Shareefdeen, Z.M.; Jabbar, N.A. Dynamic modeling and analysis of biotrickling filters in continuous operation for H2S removal. Clean Technol. Environ. Policy 2014, 16, 1757–1765. [Google Scholar] [CrossRef]
  143. Gąszczak, A.; Bartelmus, G.; Burghardt, A.; Rotkegel, A.; Sarzyński, R. Experiments and modelling of a biotrickling filter (BTF) for removal of styrene from airstreams. J. Chem. Technol. Biotechnol. 2018, 93, 2659–2670. [Google Scholar] [CrossRef]
  144. San-Valero, P.; Penya-Roja, J.M.; Álvarez-Hornos, F.J.; Gabaldón, C. Modelling mass transfer properties in a biotrickling filter for the removal of isopropanol. Chem. Eng. Sci. 2014, 108, 47–56. [Google Scholar] [CrossRef] [Green Version]
  145. Beigi, B.H.M.; Thorpe, R.B.; Ouki, S.; Winter, P.; Waalkens, A. Hydrogen sulphide and VOC removal in biotrickling filters: Comparison of data from a full-scale, low-emission unit with kinetic models. Chem. Eng. Sci. 2019, 208, 115033. [Google Scholar] [CrossRef]
  146. Dallinger, A.; Duldhardt, I.; Kabisch, J.; Schlüter, R.; Schauer, F. Biotransformation of cyclohexane and related alicyclic hydrocarbons by Candida maltosa and Trichosporon species. Int. Biodeterior. Biodegrad. 2016, 107, 132–139. [Google Scholar] [CrossRef]
  147. Dewidar, A.A.; Sorial, G.A. Effect of surfactin on removal of semi-volatile organic compound: Emphasis on enhanced biofiltration performance. Environ. Res. 2021, 193, 110532. [Google Scholar] [CrossRef] [PubMed]
  148. López, L.R.; Bezerra, T.; Mora, M.; Lafuente, J.; Gabriel, D. Influence of trickling liquid velocity and flow pattern in the improvement of oxygen transport in aerobic biotrickling filters for biogas desulfurization. J. Chem. Technol. Biotechnol. 2016, 91, 1031–1039. [Google Scholar] [CrossRef]
  149. López, L.R.; Dorado, A.D.; Mora, M.; Gamisans, X.; Lafuente, J.; Gabriel, D. Modeling an aerobic biotrickling filter for biogas desulfurization through a multi-step oxidation mechanism. Chem. Eng. J. 2016, 294, 447–457. [Google Scholar] [CrossRef] [Green Version]
  150. Taha, A.; Patón, M.; Rodríguez, J. Model-based design and operation of biotrickling filters for foul air H2S removal at wastewater networks. J. Environ. Chem. Eng. 2022, 10, 107372. [Google Scholar] [CrossRef]
  151. Diks, R.M.M.; Ottengraf, S.P.P. Bioproeess Engineering Verification studies of a simplified model for the removal of dichloromethane from waste gases using a biological trickling filter (Part II). Bioprocess Eng. 1991, 6, 131–140. [Google Scholar] [CrossRef]
  152. Ockeloen, H.F.; Overcamp, T.J.; Grady, C.P.L., Jr. Engineering Model for Fixed-Film Bioscrubbers. J. Environ. Eng. 1996, 122, 191–197. [Google Scholar] [CrossRef]
  153. Alonso, C.; Suidan, M.T.; Sorial, G.A.; Smith, F.L.; Biswas, P.; Smith, P.J.; Brenner, R.C. Gas Treatment in Trickle-Bed Biofilters: Biomass, How Much Is Enough? Biotechnol Bioeng 1997, 54, 583–594. [Google Scholar] [CrossRef]
  154. Barton, J.W.; Hartz, S.M.; Klasson, K.T.; Davison, B.H. Microbial Removal of Alkanes From Dilute Gaseous Waste Streams: Mathematical Modeling of Advanced Bioreactor Systems. 93è98 J. Chem. T Echnol. Biotechnol 1998, 72, 93–98. [Google Scholar] [CrossRef]
  155. Alonso, C.; Suidan, M.T.; Kim, B.R.; Kim, B.J. Dynamic mathematical model for the diodegradation of VOCs in a biofilter: Biomass accumulation study. Environ. Sci. Technol. 1998, 32, 3118–3123. [Google Scholar] [CrossRef]
  156. Lobo, R.; Revah, S.; Viveros-García, T. An Analysis of a Trickle-Bed Bioreactor: Carbon Disulfide Removal. Biotechnol. Bioeng. 1999, 63, 98–109. [Google Scholar] [CrossRef]
  157. Alonso, C.; Zhu, X.; Suidan, M.T.; Kim, B.R.; Kim, B.J. Mathematical Model of Biofiltration of VOCs: Effect of Nitrate Concentration and Backwashing. J. Environ. Eng. 2001, 127, 655–664. [Google Scholar] [CrossRef]
  158. Baltzis, B.C.; Mpanias, C.J.; Bhattacharya, S. Modeling the removal of VOC mixtures in biotrickling filters—PubMed. Biotechnol. Bioeng. 2001, 72, 389–401. [Google Scholar] [CrossRef] [PubMed]
  159. Martin, R.W.; Li, H.; Mihelcic, J.R.; Crittenden, J.C.; Lueking, D.R.; Hatch, C.R.; Ball, P. Optimization of biofiltration for odor control: Model calibration, validation, and applications. Water Environ. Res. 2002, 74, 17–27. [Google Scholar] [CrossRef] [PubMed]
  160. Mathur, A.K.; Sundaramurthy, J.; Balomajumder, C. Kinetics of the removal of mono-chlorobenzene vapour from waste gases using a trickle bed air biofilter. J. Hazard. Mater. 2006, 137, 1560–1568. [Google Scholar] [CrossRef] [PubMed]
  161. Das, C.; Chowdhury, R.; Bhattacharya, P. Three phase biofilter model for the removal of styrene through the microbial route. Int. J. Chem. React. Eng. 2008, 6, 1–9. [Google Scholar] [CrossRef]
  162. Kim, S.; Deshusses, M.A. Determination of mass transfer coefficients for packing materials used in biofilters and biotrickling filters for air pollution control. 1. Experimental results. Chem. Eng. Sci. 2008, 63, 841–855. [Google Scholar] [CrossRef]
  163. Kim, S.; Deshusses, M.A. Determination of mass transfer coefficients for packing materials used in biofilters and biotrickling filters for air pollution control—2: Development of mass transfer coefficients correlations. Chem. Eng. Sci. 2008, 63, 856–861. [Google Scholar] [CrossRef]
  164. Yoon, S.; Carey, J.N.; Semrau, J.D. Feasibility of atmospheric methane removal using methanotrophic biotrickling filters. Appl. Microbiol. Biotechnol. 2009, 83, 949–956. [Google Scholar] [CrossRef] [PubMed]
  165. Lee, S.H.; Heber, A.J. Ethylene removal using biotrickling filters: Part II. Parameter estimation and mathematical simulation. Chem. Eng. J. 2010, 158, 89–99. [Google Scholar] [CrossRef]
  166. Silva, J.; Aroca, G. Sensitivity analysis of a model that describes the biofiltration of TRS compounds. Procedia—Soc. Behav. Sci. 2010, 2, 7741–7742. [Google Scholar] [CrossRef] [Green Version]
  167. Mannucci, A.; Munz, G.; Mori, G.; Lubello, C. Biomass accumulation modelling in a highly loaded biotrickling filter for hydrogen sulphide removal. Chemosphere 2012, 88, 712–717. [Google Scholar] [CrossRef] [PubMed]
  168. Liang, W.; Huang, S.; Yang, Y.; Jiang, R. Experimental and modeling study on nitric oxide removal in a biotrickling filter using Chelatococcus daeguensis under thermophilic condition. Bioresour. Technol. 2012, 125, 82–87. [Google Scholar] [CrossRef] [PubMed]
  169. Zare, H.; Najafpour, G.; Rahimnejad, M.; Tardast, A.; Gilani, S. Biofiltration of ethyl acetate by Pseudomonas putida immobilized on walnut shell. Bioresour. Technol. 2012, 123, 419–423. [Google Scholar] [CrossRef]
  170. Rahul; Mathur, A.K.; Balomajumder, C. Biological treatment and modeling aspect of BTEX abatement process in a biofilter. Bioresour. Technol. 2013, 142, 9–17. [Google Scholar] [CrossRef] [PubMed]
  171. López, L.R.; Dorado, A.D.; Mora, M.; Prades, L.; Gamisans, X.; Lafuente, J.; Gabriel, D. Modelling biotrickling filters to minimize elemental sulfur accumulation during biogas desulfurization under aerobic conditions. In Proceedings of the 7th European Meeting on Chemical Industry and Environment, EMChIE 2015, Tarragona, Spain, 10-12 June 2015. [Google Scholar]
  172. Cáceres, M.; Dorado, A.D.; Gentina, J.C.; Aroca, G. Oxidation of methane in biotrickling filters inoculated with methanotrophic bacteria. Environ. Sci. Pollut. Res. 2016, 24, 25702–25712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Dupnock, T.L.; Deshusses, M.A. Development and validation of a comprehensive model for biotrickling filters upgrading biogas. Chem. Eng. J. 2021, 407, 126614. [Google Scholar] [CrossRef]
  174. López, L.R.; Mora, M.; Van der Heyden, C.; Baeza, J.A.; Volcke, E.; Gabriel, D. Model-Based Analysis of Feedback Control Strategies in Aerobic Biotrickling Filters for Biogas Desulfurization. Process. 2021, 9, 208. [Google Scholar] [CrossRef]
Figure 1. Milestones in the development of mathematical models for biotrickling filtration of air polluted with VOCs.
Figure 1. Milestones in the development of mathematical models for biotrickling filtration of air polluted with VOCs.
Processes 10 02531 g001
Table 1. Selected operating and process parameters and their effects of the performance of biotrickling filters treating air polluted with VOCs.
Table 1. Selected operating and process parameters and their effects of the performance of biotrickling filters treating air polluted with VOCs.
ParameterEffectsReferences
Liquid trickling velocity and frequencyThe mode (intermittent/continuous) and velocity of trickling affects the humidity of the packing (important for biomass growth), ensures a supply of minerals/microelements for microbial growth, rinses out biodegradation by-products and dead matter, and plays role in managing the pressure drop across the packing; increased trickling velocity may result in decreased mass transfer of the hydrophobic VOCs, but on the other hand may enhance the mass transfer of hydrophilic VOCs and result in more uniform water distribution throughout the biofilter bed[32,38]
pH, temperature and composition of trickling liquidTrickling liquid can buffer the pH in the optimal range for microbial growth (e.g., fungi prefer an acidic pH); the modification of trickling liquid composition using, e.g., surfactants (also biosurfactants produced in situ by microbial metabolism) or silicone oil, can enhance the removal of hydrophobic VOCs from air; the increase in process temperature may result in increased process efficiency, but also increased pressure drops across the packing layer[28,39,40]
Packing materialDepending on the porosity and surface area, packing materials define the available surface for biofilm development and growth, and consequently affect the interfacial area for mass transfer from the gas to biofilm/liquid phase; this parameter greatly affects the pressure drop in the biofilter (low porosity and high surface area may lead to fast clogging of the bed); Bio-based packing materials seem to be promising in the development of biofiltration methods for hydrophobic VOCs, e.g., those coming from the pharmaceutical industry.[30,32,41,42]
Feeding strategyContinuous, intermittent, or periodic shut-down of the polluted gas supply (e.g., for weekends); the mixing of gas streams from two sources with distinctly different waste gas compositions (e.g., hydrophilic and hydrophobic components)[43,44]
Biomass growth control strategyThe control of a pressure drop and avoidance of the bed clogging, e.g., by periodic switching of the co-/ counter-current gas-liquid flow patterns, or compressing and squeezing out of the packing (bed clogging results in the deterioration of biofilter performance, and sometimes the recovery is not possible); mechanical, chemical, or biological (e.g., predators) methods are used[45,46]
Table 4. Overview of selected mathematical models for conventional biofiltration.
Table 4. Overview of selected mathematical models for conventional biofiltration.
Target Compound(s)Process ParametersModel DescriptionReference
Mixture:
toluene, butylacetate, ethylacetate, butanol (Cin for each component ranged from 0.050 to 5.606 g m−3)
5-stage Quick fit glass columns
(din = 0.1524 m, H = 0.6 m)
Peat compost inoculated with toluene degrading organism
uG: 0.84 and 14.8 cm s−1
Steady-state theoretical model describing the elimination of the carbon sources in the filter bed; Monod equation is adapted for the microkinetics description, and model supports identification of either reaction or diffusion limitation
[110]
Methanol (Cin about 6.4 g m−3)Peat-perlite packed biofilter
(din = 0.05 m, H = 0.6 m)
uG = 6.4–12.7 m h−1
The model considers the consumption of two substrates i.e., methanol and oxygen; the diffusion and reaction of these two substrates in the biofilm at quasi-steady state conditions are considered with Andrews and Monod-type dependences on the concentrations of methanol and oxygen, respectively; the model assumes gas plug flow and uniform biofilm density along the biofilter; the model showed that the biofiltration process is mass-transfer limited by oxygen and kinetic-limited by methanol[123]
Toluene (IL range: 4.8–26.8 g m−3 h−1; Cin: 0.6–2.8 g m−3Plexiglass column
(din = 0.102 m, H = 0.686 m),
sterilized peat and perlite inoculated with toluene-degrading consortium
The model approaches the description of biofiltration at transient conditions; the model assumes that the biolayer may not be uniform at the surface of packing elements, and VOCs can be reversibly adsorbed on the packing elements not covered with biofilm; the biodegradation rate depends on the concentrations of both VOCs and oxygen, and can be determined based on batch suspended culture experiments; the air stream flow is a plug flow[112]
Benzene, toluenePlexiglass column (din = 0.1 m, H: 0.51 m for benzene and 0.69 for toluene); Contaminated soil-based microbial consortium inoculated on peat-perlite particlesThe model attempts to predict biofilter capabilities and aid during the biofilter design stage; the model describes and predicts with good accuracy steady-state biofiltration; results of the modeling and experiments revealed that benzene is harder to remove from air than toluene, and the biofiltration of hydrophobic compounds is less oxygen-limited than in the case of hydrophilic VOCs[124]
EthanolPolyvinyl chloride biofilter column (din = 0.076 m; H = 0.9 m); three packing media were compared: granular activated carbon, sintered diatomaceous earth, and compostDevelopment of a model by Shareefdeen and Baltzis using exclusion of a plug-flow assumption and the incorporation of axial dispersion in transient conditions; the model provided foundations for further biofilter developments, including the most important feature affecting the biofiltration performance i.e., packing media allowing for enhanced biodegradation, with the lowest fraction of area available for microbial growth, and which allows for high adsorption capacity for the treated compound[115]
Methyl ethyl ketone, methyl isobutyl ketone and their mixturePlexiglass column (din = 0.08 m,
H = 1 m); Bioton packing (mixture of compost and polystyrene spheres)
The proposed dynamic, diffusion reaction model describes both the transient and steady-state biofiltration performance, including kinetics with substrate competition[116,125]
Benzene, tolueneModel was verified using data from [124]A general transient biofiltration model incorporating mixing, oxygen limitation, and adsorption phenomena; the biofilm is assumed to not necessarily be uniformly formed over the bed particles; planar geometry of biolayer, reversible VOC adsorption and no biomass accumulation are assumed; [126]
Binary gas mixtures: benzene-toluene or ethanol-butanol3-sectional glass biofiltration column (din = 0.152 m, H = 0,915 m), packed with a mixture of peat moss and horticultural grade perliteSteady-state model for the biofiltration of binary VOC mixtures; the model incorporates potential kinetic interactions between the air pollutants (VOCs i.e., substrates), as well as the effects of oxygen limitation on biodegradation and biomass diversification in the biofilter packed bed; results suggest that the modeling of VOC mixture biofiltration is not straightforward when compared to modeling single compound biofiltration[47]
Model validation on selected literature dataAs = 0.457–1.9
Pe = 1–1000
m2,I = 0.002–20
A transient model in which dispersion phenomena, convection, adsorption, biomass diversification, oxygen limitation, diffusion, and biodegradation in both the biofilm and a solid support are considered.[117]
Methanol, α-pinenePlexiglass, 4-segment biofilter (din = 0.28 m, Htotal = 1.2 m); compost and wood chips as packing materialsA model approaching the description of biofiltration of a mixture of hydrophobic and hydrophilic VOCs; the steady-state model, including Monod kinetics with inhibition, showed that the prediction of biofiltration performance for α-pienene as a representative of hydrophobic VOCs can be achieved by using an air/biofilm partition coefficient instead of a typical air/water partition coefficient, due to lipophilic characteristics of the biofilm[127]
Hydrogen sulfidePilot-scale biofilter made of a fiberglass cylinder: din = 0.608 m; H = 1.2 m The BiofilterTM two-phase model was used; homogeneity of biofilm was assumed; the model has been incorporated into a software[119]
Butanol, methyl ethyl ketoneLiterature data were used for model validationThe model uses a linear driving force approach, assumptions include axially dispersed plug flow for the gas phase, pollutant adsorption phenomena, and no oxygen limitation, as well as first-order kinetics of biodegradation[128]
Hydrogen sulfideIndustrial-scale biofilter (volume of packing: 16.12 m3; H = 1.524 m); Biosorbens (inorganic media) packing material Use of biofilter models in real full-scale applications using a dedicated software[104]
EthanolAcrylic plastic biofilter column (din = 0.07 m, H = 0.25 m), sand and lava rock inoculated with raw sludge acclimated to ethanol degradationThe model combines a percolation theory and a cellular automaton model for the description of biofilm growth; it was found that, during ethanol biofiltration, oxygen concentration was a limiting factor in the upper parts of the biofilter, while ethanol concentration was limiting in the bottom regions (the supply of treated air from top of the biofilter); the model allows for predicting decreases in biofiltration performance with respect to increasing pressure drops across the packing[118]
StyrenePerlite inoculated with a mixed microbial culture from petrochemical refinery sludgeA back propagation neural network algorithm was used to model and predict biofiltration performance based on the inlet styrene gas-phase concentration and gas flow rates[129]
Methyl propyl ketone, toluene, p-xylene, n-butyl acetateModel was validated on literature experimental dataA dynamic model for predicting biofiltration performance; assumptions include gas phase plug flow, utilization of Henry’s law to describe gas-biofilm interface equilibrium, Fick’s law to describe diffusion, planar biofilm geometry, and adsorption phenomena[130]
TolueneA biofilter column: din = 0.12 m; H = 0.4 m; packing material: wood chips and propylene spheresBiofiltration model incorporating biofilter performance as a function of biomass growth and pressure drop for varied toluene loadings; toluene biodegradation follows Monod kinetics, biofilm is treated as a flat surface; in contrast to previously described models, this model employs microbial growth kinetics, including inert biomass formation, and the model has been verified for various flow rates and concentrations, which mimic the real variations in gas effluents[120]
Butyl acetate, butanol, xyleneA biofilter with dimensions: 0.5 × 0.48 × 2 m (length: width: height) packed with activated pine barkThe proposed mathematical model aims at predicting the biofiltration performance of VOC mixtures depending on the mixture composition and ratio of investigated VOCs; surface response methodology and finite difference approximation were employed to deliver model solutions; relatively simple equations were proposed to determine the elimination capacities of each studied VOC[121]
Methanol, α-pineneModel validation is based on the results obtained by Mohseni and Allen [127]The model gives approximate analytical expression of the concentration profiles of studied VOCs in air/biofilm phases using the Adomian decomposition method; non-linear differential equations have been solved analytically in this biofiltration model[131]
Methanol, α-pineneModel validated on experimental data from literatureAn approach of a biofilter model based on expressing the particular solution of governing equations using an integral containing Green’s function; the proposed model is a development of a model proposed by Mohseni and Allen [127][132]
ToluenePolycarbonate biofilter column (din = 0.1 m, H = 0.63 m) packed with sterilized compost and wood chips, inoculated with Nocardia sp.A hypothesis was formulated that a biofilter can progress from a single steady-state to a multi steady-state, driven by the growth of biofilm thickness and biofilm die-off during biofiltration; toluene removal was decreased with biofiltration time for increasing inlet toluene concentrations, which was proven to be related with the shift from a diffusion-limited biofilm high-activity state to a low-activity state; this can be a result of substrate degradation and consequent inhibition following Haldane kinetics and pollutant diffusion within the biofilm[122]
uG—superficial gas velocity; As—biofilm surface area per unit volume of particle; Pe—Peclet number (ratio of convective to diffusive mass transfer rates; product of Reynolds and Schmidt numbers); m2,i—distribution coefficient of substance i in an air/solid media system.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rybarczyk, P. Removal of Volatile Organic Compounds (VOCs) from Air: Focus on Biotrickling Filtration and Process Modeling. Processes 2022, 10, 2531. https://doi.org/10.3390/pr10122531

AMA Style

Rybarczyk P. Removal of Volatile Organic Compounds (VOCs) from Air: Focus on Biotrickling Filtration and Process Modeling. Processes. 2022; 10(12):2531. https://doi.org/10.3390/pr10122531

Chicago/Turabian Style

Rybarczyk, Piotr. 2022. "Removal of Volatile Organic Compounds (VOCs) from Air: Focus on Biotrickling Filtration and Process Modeling" Processes 10, no. 12: 2531. https://doi.org/10.3390/pr10122531

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