Next Article in Journal
Evolution of Waste Cooking Oil Collection in an Area with Long-Standing Waste Management Problems
Next Article in Special Issue
Plasma Technology and Its Relevance in Waste Air and Waste Gas Treatment
Previous Article in Journal
Organizational Governance, Social Bonds and Information Security Policy Compliance: A Perspective towards Oil and Gas Employees
Previous Article in Special Issue
Evaluation of Parallel-Series Configurations of Two-Phase Partitioning Biotrickling Filtration and Biotrickling Filtration for Treating Styrene Gas-Phase Emissions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biological Waste Air and Waste Gas Treatment: Overview, Challenges, Operational Efficiency, and Current Trends

by
Daniel Dobslaw
1,* and
Oliver Ortlinghaus
2
1
Institute of Sanitary Engineering, Water Quality and Solid Waste Management, University of Stuttgart, Bandtäle 2, 70569 Stuttgart, Germany
2
Liutec Labor für Innovative Umwelttechnik Ing.—GmbH, Lindberghweg 179, 48155 Münster, Germany
*
Author to whom correspondence should be addressed.
Sustainability 2020, 12(20), 8577; https://doi.org/10.3390/su12208577
Submission received: 1 September 2020 / Revised: 12 October 2020 / Accepted: 14 October 2020 / Published: 16 October 2020
(This article belongs to the Special Issue Sustainable Waste Air and Biogas Treatment)

Abstract

:
International contracts to restrict emissions of climate-relevant gases, and thus global warming, also require a critical reconsideration of technologies for treating municipal, commercial, industrial, and agricultural waste gas emissions. A change from energy- and resource-intensive technologies, such as thermal post-combustion and adsorption, as well to low-emission technologies with high energy and resource efficiency, becomes mandatory. Biological processes already meet these requirements, but show restrictions in case of treatment of complex volatile organic compound (VOC) mixtures and space demand. Innovative approaches combining advanced oxidation and biofiltration processes seem to be a solution. In this review, biological processes, both as stand-alone technology and in combination with advanced oxidation processes, were critically evaluated in regard to technical, economical, and climate policy aspects, as well as present limitations and corresponding solutions to overcome these restrictions.

1. Introduction

Studies of the International Energy Agency (IEA) and British Petroleum showed that by 2040 the global gross domestic product is expected to double, while greenhouse gas emissions will rise by around 30% [1,2]. However, even emissions at the current level are already leading to unquestionable environmental changes and global warming. According to a study by the German Advisory Council on Global Change (WBGU) [3], a reduction of CO2 equivalents of industrial countries from approximately 12 t CO2,eq∙a−1∙inhabitant−1 to less than 1 t CO2,eq∙a−1∙inhabitant−1 by the year 2050 is strictly required to achieve the international objective of limiting global warming to 2 °C, equivalent to a reduction by more than 90%. In order to achieve this goal, it is inevitable that agricultural, municipal and industrial waste gas applications have to be addressed through innovative energy and resource-efficient technologies. Biological processes already meet these requirements but show restrictions in case of treatment of complex volatile organic compound (VOC) mixtures and space demand.
In the following sections, biological waste gas treatment processes are firstly presented in an overview and then with regard to technical, economic, and ecological aspects. Based on the current market situation for these technologies and the emerging markets of emission treatment of sewage sludge drying plants, the challenges for the application of this technology are also presented. This review is concluded by presentation of current trends in the innovative combination of biological processes with advanced-oxidation technologies to overcome the listed restrictions. This process combinations show high energy and resource efficiencies and are therefore sustainable process solutions. The application of these sustainable process combinations will be necessary to meet the requirements in reduction of greenhouse gas emissions, to overcome current restrictions of biological waste gas processes, and to increase the applicability of bio-based waste gas treatment.

2. Overview and Application

Besides competing waste gas and waste air purification techniques, like adsorption, physical or chemical scrubbing, membrane processes, condensation, thermal oxidative combustion processes, UV oxidation, non-thermal plasma (NTP), or thermal plasma, biological processes represent an important technology established at large-scale in treatment of volatile organic compounds (VOCs) and odors in agricultural, municipal, and industrial applications and are used either as stand-alone technology or in combination with the mentioned non-biological approaches. They can be subdivided into the three basic processes of bioscrubber, biofilter (BF), and biotrickling filter (BTF), in which the respective application, in particular, is defined by the polarity and, thus, the water solubility of the waste gas components to be treated [4,5,6,7]. Their process structure, function, applicability and technology-specific pros and cons are as follows.

2.1. Bioscrubber

The bioscrubber is a wet, two-stage system consisting of a column for physical absorption of waste gas compounds and a regenerator for biological regeneration of the loaded liquid phase [6,8]. Within the absorber, the waste gas is commonly conducted in counterflow to the dispersed liquid phase in order to enhance the number of separation stages and, thus, the purification efficiency of the column [6]. The purification efficiency can be further increased in a first step by installation of package materials to enhance the degree of turbulence and, thus, the contact time of both phases. The loaded phase is purified by biotransformation and mineralization processes of the suspended microorganisms in the aerated regenerator and subsequently dispersed in the absorber via nozzles in a closed loop operation, consisting of regeneration and loading again during contact with the waste gas [8].
The loading capacity and loading kinetics of the liquid phase depend on substance-specific properties (water solubility, Henry coefficient, pollutant concentration) [6], as well as process-specific parameters (residual load, spray density, exhaust air temperature, salinity, spray pattern, droplet size, degree of turbulence, contact time, number of separation stages, concentration of co-components, etc.). The requirements of the absorber stage (low pollutant load) conflict with the requirements of the regenerator for a high transformation kinetics (high pollutant load), so that in real operation a self-regulating operating point is usually established. Due to the wet operation of the system, bioscrubbers prove to be well suited for the treatment of polar components [4,6,7]. Since the distribution equilibrium of polar compounds is on the side of the aqueous phase, and thus the tendency of desorption is rather low, the bio-scrubber can be used for comparatively high pollutant concentrations of up to 2 g C∙m−3 [7]. Due to the high swamp volume and the wet mode of operation, it also allows a shaving of pollutant peaks, as well as a simple dosage of alkaline, solutions of nutrients, or even nitrification inhibitors for neutralization and bacterial growth [6,7,9,10]. On the other hand, the bioscrubber is not suitable for the treatment of waste gas with a high proportion of lipophilic components due to its design as a wet system. Since the 1970s, approaches with solubilizers mainly based on silicone oils were successfully introduced, creating a smooth transition to the biotrickling filter, but failed due to the high persistence of these compounds and the resulting disposal problems [11,12,13,14,15,16]. Current advances favor phthalic acid esters or polyethylene glycols on a technical scale. However, the very high biodegradability requires, on the one hand, a semi-continuous addition of these solubilizers into the aqueous phase and, on the other hand, an increased elutriation rate with higher wastewater volumes due to the strongly increasing biomass growth in the system. The process also requires a comparatively complex control and regulation technology and the expensive provision of compressed air (approximately 3 kW∙kg C−1) [5,7,10,17]. The most common reason for an insufficient cleaning performance of an established bioscrubber is an insufficient oxygen input [4], caused either by a too high biomass concentration or by aging effects at the aeration elements (clogging of ceramic aerators, pore expansion or cracks in membrane aerators). As a consequence, increased foam formation due to cell lysis, a drop in pH combined with acidic odor emissions, and migration of the suspended microbial consortium onto the absorber’s packing often occur (see Figure 1).

2.2. Biofilter

In terms of polarity, water regime, control, and regulation technology, as well as investment and operating costs, the biofilter is the counterpart of the bioscrubber [5,7]. The biofilter is a biologically active fixed bed where waste gas compounds are firstly absorbed and then mineralized by a fixed community of bacteria and fungi. As filter bed, inexpensive organic packing materials, such as compost, bark materials, forest residues, like root wood, peat, heather, coconut fibers, sawdust, or wood chips, inorganic materials, such as lava or marble quarry, or mixtures of these materials are mostly used [5,9]. Torn burl wood or a filter mat is often drawn in as the lowest layer to prevent blocking of the gas distributor. Due to the dry operation mode and to delay compaction of the package material, irrigation of the package material is strictly limited and waste gas humidification is mainly performed by a pre-scrubber in case of unsaturated waste gas conditions. Additionally, the mass transfer of mainly lipophilic waste gas compounds via the liquid phase/biofilm into microbial cells is considerably increased. Thus, biofilters are mainly used for applications dominated by lipophilic waste gas compounds in the concentration range up to 500 mg C∙m−3 and for odor abatement [5]. Here, the adsorptive properties of the organic materials have a positive effect on the purification performance, as the components present in the trace range accumulate and reach concentration levels which allow induction of microbial degradation pathways in terms of energy.
The inhomogeneous packing material and related problems are the main cons of the biofilters, where inhomogeneity leads to an inhomogeneous gas distribution and results in inactive volume fractions of the packing material of about 40–60% [6]. The active volume fraction can be further reduced by the formation of additional flow channels and partial drying or wetting of the system. Hence, performance levels of biofilters are generally lower than those of bioscrubbers or biotrickling filters and tend to decline at long-term conditions [5,7]. This process-related lower degradation performance is usually countered by larger-volume plants. Particularly during the cold season, the filter tends to be more susceptible to wetting and has a considerable amount of highly contaminated condensate.
As wooden packing material may act as a carbon source in microbial deficiency situations, the microbial attack in combination with acidification leads to a sprouting of calcium carbonate fixed in the wood (approximately 23–48% in the ash) [18]. Hence, the wood becomes soft, decomposes more easily, and tends to intensified compaction and clogging by the fine fractions produced. The lifetime of wooden materials is approximately 7 years under mesophilic conditions but drops to less than 2 years under thermophilic conditions [7].
A significant problem of the biofilter is pollutant-induced acidification [4,7]. Usually, acidification occurs to a greater extent with alcohols, aldehydes, organic acids and esters, and to a lesser extent with ketones, ethers, aromatics, and unsubstituted aliphatics, since the volatile components of the exhaust air are biotransformed into organic acids as storage substances [4]. Acid formation also occurs during the biodegradation of organosulfur compounds, such as methanethiol, dimethyl sulfide, and dimethyl disulfide, which are mineralized either via H2S or dimethyl sulfoxide to sulfuric acid as the final product [19]. Due to the already acidic filter material, as well as a pollutant-induced acidification of the filter material, the microbial flora in biofilters has a significantly higher fungal content, which can lead to typical fungal odor.

2.3. Biotrickling Filter

The biotrickling filter takes an intermediate position between the bio-scrubber and the biofilter, consisting of an absorber column and a rudimentary swamp. Besides an increase in turbulence and contact times for enhanced VOC mass transfer, the package material, mostly made of plastics, serves as growth surface for the microbial biofilm. Here, biotransformation of contaminants dominantly takes place [6]. The commonly not aerated swamp phase serves as a reservoir for the circulation pump but may be slightly aerated to exclude potential odor problems.
In industrial applications, irrigation can be continuous or intermittent and the flow rate can be both adapted to hydrophilic or lipophilic characteristics of the waste gas compounds and the wetting behavior of the package [20,21]. In case of agricultural or municipal applications, which are often characterized by NH3 emissions, continuous spraying is strictly necessary to minimize N2O emissions, since otherwise an average of 26% of the NH3-N load is released as N2O-N into the clean gas [22,23]. Since the flow rate is still only approximately 1–10% of that of the bioscrubber, this procedure is called semi-wet system, but it can still be used for dosage of nutrients or alkaline for neutralization, as well as pre-humidification of the waste gas [5,9]. Likewise, a temporary increase of the irrigation density with simultaneous elutriation allows a limited discharge of excess biomass, potentially toxic intermediates or excess salts to avoid salination [9]. In case of continuous irrigation, a countercurrent operation must be considered.
The gas velocity in the column is usually 0.1 m∙s−1 for industrial applications and is thus higher by a factor of 3–6 than for biofilters, which means that biotrickling filters can be built up more compactly. Biotrickling filters in agricultural applications operate at even higher gas velocities of 0.3–1.4 m∙s−1 [24]. However, the more compact design leads to higher specific loads, which may lead to clogging [5,9,25]. For example, Kan and Deshusses [26], Wang et al. [27] and Ryu et al. [28] describe the use of biotrickling filters for treatment of highly loaded toluene-containing waste gas with volume-specific loads of up to 4000 g toluene∙m−3∙h−1. The maximum elimination rates achieved here were 408, 480, and 3700 g toluene∙m−3∙h−1, respectively. Assuming a yield of 0.2 and appearance of first clogging phenomena, like inhomogeneous flow distributions and partial drying/wetting from approximately 60 kg∙m−3 onwards, these elimination rates result in a calculated operating time of only 80–735 h excluding increasing effects of conservative metabolism. In the case of Ryu et al. [28], a real operating time of 145 days (corresponds to a yield of only 0.01) was achieved at extremely short contact times of 3.5 s. Datta and Philip [29] described a rotating disc reactor for treatment of a mixture of methyl ethyl ketone, methyl isobutyl ketone, ethylbenzene, o-xylene, and toluene and achieved almost complete conversion with a maximum total elimination performance of 569 g C∙m−3∙h−1. However, specific loads or elimination rates in the range of 10–100 g C∙m−3∙h−1 should be aimed for in order to ensure stable long-term operation [30].
Further disadvantages of biotrickling filters are the low buffer capacity against pollutant peaks and low moisture content of the plastic package materials due to the lack of adsorption capacity of the filter material. The use of plastic carriers also causes problems during time intervals without pollutant loads, as no metabolizable carbon source is available. Artificial addition of pollutants either into the swamp phase or into the conventional exhaust air route is a feasible solution.

3. Technical, Economic, and Climate Policy Evaluation of Waste Air Treatment Techniques

3.1. Technical Aspects

Biological waste air processes are particularly used in agricultural, municipal, and industrial waste air applications, especially for issues arising from husbandries, animal meal and sewage sludge drying, composting, landfilling, biogas treatment, and production processes in chemical, plastics, paper, and wood industries. Waste air from agricultural and municipal applications is typically characterized by large-volume flows (approximately 10,000–100,000 m³∙h−1) at variable concentrations of VOCs (approximately 10–300 mg C∙m−3), ammonia (up to 500 mg NH3∙m−3), nitrous oxide (5–120 mg N2O∙m−3), odorous compounds (over 20,000 OU∙m−3), germs (103–105 CFU∙m−3), and possibly methane (10–150 mg C∙m−3) and, to a lesser extent, H2S. Industrial production processes are also characterized by high volume flows (approximately 10,000–500,000 m³∙h−1) with enhanced concentrations of VOCs (approximately 100–500 mg C∙m−3, in some cases up to 2000 mg C∙m−3) [7,31]. The typical spectrum of waste gas compounds ranges from polar compounds, such as alcohols and ketones (e.g., in the printing and paint industry), mixtures of polar and lipophilic components to situations with solely lipophilic compounds, like styrene and toluene derivatives (e.g., in trickle resins, film, and plastic production). Depending on the production processes, additional inorganic compounds, like NH3, N2O, or H2S, may occur at concentrations above 1 g∙m−3 (e.g., at pig bristle hydrolysis). Furthermore, waste air flows are characterized by highly variable dust loads, humidity (0–100% of relative humidity), and ranges in temperature (ambient level up to approximately 60 °C). These operational parameters typically exclude treatment processes, such as condensation, membrane processes, thermal plasmas, or oxidative catalytic processes, from application. Alternative processes, like absorption, thermal oxidative processes, non-thermal plasma, UV-oxidation, and adsorption, are not suitable as stand-alone single-stage systems under these circumstances, as well. A summary of the suitability and limits of the respective processes is given in Table 1.
As shown in Table 1, physisorption is no suitable technique in presence of poorly water-soluble components and permanent gases due to its limitation to hydrophilic components. Although chemisorption permits improved elimination of H2S, odors, germs, and individual aromatic components, particularly when oxidizing agents are added, lipophilic components with a low tendency to oxidize also prove unsuitable for this process. The suitability of non-thermal plasma, as well as UV-oxidation, highly depends on substance properties. An efficient treatment of the waste gas with low specific input energy (SIE) values requires an increased reactivity of the waste gas components (e.g., aldehydes) or the presence of an unsaturated bond (e.g., alkenes, styrene or similar). Although adsorption is, meanwhile, also suitable for the treatment of polar components due to the development of impregnated carbons, the specific loadability is still many times lower than for lipophilic components. Especially in dusty exhaust air streams with high relative humidity, the process requires preconditioning or has high operating costs. However, thermal oxidative processes have proven to be suitable and therefore strongly dominate this field of industrial applications.
Similar to the non-biological scrubbers, the bioscrubber is also restricted to polar waste gas compounds [6,7]. Due to low irrigation density, however, both biofilter and biotrickling filter are eligible for treatment of a wide range of polarities and are therefore suitable for various applications [5]. In particular, the biotrickling filter benefits from the variable water management, which allows the discharge of acid reaction products, like H2SO4, HCl, or HF, or the addition of alkalines for neutralization [5,9,32].

3.2. Economic Aspects

From an economic point of view, biological processes are clearly preferable to post-combustion processes, as their investment costs reach only 20–50% of the costs of a post-combustion plant. As incinerators operating at non-adiabatic conditions require both primary and electrical energy, they have high operating costs. In contrast, the operating costs of biological treatment processes are mainly defined by the electrical energy demands of the compressor and, if necessary, the circulation pump [5,7,8,9]. Operational costs for dosage of alkalines and nutrients are normally below 1000 €·a−1 for treatment plants of a capacity up to 20,000 m³·h−1. An overview of investment and operating costs of biofilters, biotrickling filters, thermal oxidation, and adsorption, as well as non-thermal plasmas as comparative processes, presented in literature, is compiled in Table 2. The calculations of this study base on electrical energy costs of 0.15 €·kWh−1 and natural gas costs of 0.416 €·m−3.
As shown, the investment costs for biotrickling filters are in the range of 2.50–18.78 € per cubic meter of hourly waste air capacity, depending on the size of the plant, while the investment costs for thermal oxidation are between 12.80–34.83 € per cubic meter of hourly waste air capacity [35,37,38] this study. The operating costs of biotrickling filters are in the range of 0.12–0.45 €·1000 m−3 and, thus, a factor of 12–17 below those of thermal oxidation processes (1.50–7.67 €·1000 m−3) [35,37,38,41]. Due to lower requirements in treatment of agricultural exhaust gases (target set to 70% in minimum for ammonia, dust, and odor), biofiltration processes in agricultural applications are operated at significantly higher filter volume loads (100–175 m³·m−3·h−1 in industry vs. 1000–11,000 m³·m−3·h−1 in agricultural applications), which allows a more compact design [24]. Accordingly, the specific operating costs are significantly reduced by a factor of 2. For clarification, the investment and operating costs of a biotrickling filter versus combustion process for an industrial styrene-containing exhaust air stream (20–25 °C; 50% r.H.; 150–500 mg styrene·m−3, 13,000 m³·h−1 gas flow) and of a biotrickling filter for a pig fattening site (1000 places, 91,000 m³·h−1 gas flow) are summarized in Table 3. In summary, it can be pointed out that biofiltration processes with optional pre-treatment processes are competitive technologies compared to conventional thermal oxidation processes at lower investment and operational costs. Furthermore, in combination with additional advanced oxidation processes, like NTP or UV-oxidation, bioavailability and biodegradability and, finally, the elimination efficiency may be enhanced, leading to an additional reduction in investment and operating costs [45] (see Section 6).

3.3. Ecological Aspects

From an ecological point of view, the application of biofiltration processes is also clearly preferable to combustion processes, since thermal oxidative processes are characterized by high greenhouse gas emissions (CO2, CH4 as slip gas). It is also known that N2O emissions are formed during thermal oxidative processes, which contribute to a considerable greenhouse potential. Assuming the German energy mix of 2019 (production of 0.401 kg CO2∙kWh−1) [46], secondary emissions of a thermal incineration process reach 58,477 kg CO2,eq∙a−1 for treatment of a waste air volume flow of 1000 m³∙h−1 at 8000 h per year. Since the energy demand of biofiltration processes is quantitatively defined by the energy requirements of the compressor and N2O emissions of converted ammonia-based fertilizers, emissions of CO2 equivalents for biological processes are approximately 3565–3939 kg CO2,eq∙a−1 and, thus, approximately 6–7% of the global warming potential of thermal incineration processes. However, this advantage might be completely negated in case of incorrectly operated plants for NH3 elimination, as secondary emissions of 1–11% of NO or 3–29% of N2O of the original ammonia load can be formed by bioconversion processes of NH3 [47,48,49,50,51]. Distinguished by source the emissions of CO2 equivalents of an incineration process and of a non-thermal plasma process with post plasma catalyst (PPC), as discussed in Section 7, are presented in Table 4. Compared to the thermal process, the NTP process reveals a lower emission load of 67.3% or 39,359 kg CO2,eq∙a−1 but is up to 15,598 kg∙a−1 above the biofilter. Since the balancing of Table 4 for NTP processes assumes a complete mineralization of the pollutant by a stand-alone NTP process, a process combination of NTP and biotrickling filter (BTF) appears as an ecologically advantageous alternative, since the pollutants are then hydrophilized by partial oxidation in the NTP stage and can thus be degraded by a smaller BTF stage (see Section 7).

4. Current Market Position of Biological Waste Air Treatment Processes

As shown before, biological waste air treatment processes prove to be particularly suitable for applications, like agricultural animal husbandry, drying of animal meal and sewage sludge, waste composting, or applications in similar industrial sectors, but they particularly compete with scrubber processes (agriculture) and thermal incineration or adsorption processes (industrial sectors). In agricultural application, both regular surveys of the German Federal Statistical Office on livestock [53] and a study by Hahne [54] allow a good estimation of the market share of biological waste air treatment processes in this market segment. In Germany, at the recording date of 1 March 2016, there were about 120.966, 40.267, 19.556, 9.808, 49.093, and 42.145 farms for cattle, pig, sheep, goats, poultry, and horses/donkeys, respectively. Emissions of these farms are characterized by low carbon emissions but considerable odor emissions, mainly caused by NH3, which is solely emitted to more than 500,000 t∙a−1 with further upward trend by stock farming.
Nevertheless, the installation of an exhaust air purification plant is not generally required and only demanded in specific individual cases of the approval procedure, for example when the distance to residential buildings is below minimum requirements [55]. Accordingly, the installation of waste air treatment systems is being carried out tentatively in Germany. In the period from 1997 to 2010, only 1014 waste air treatment plants were installed (pigs: 762, poultry: 170, others: 82) [54]; hence, only 0.4% of all farms were equipped with treatment technologies in 2011. Single-stage biofilters were installed in 26.4% of the cases, single stage biotrickling filters in 24.0%, and multi-stage biological systems in 43.2%. Only in 6.4% of the cases non-biological systems, such as chemical scrubbers, were implemented. However, a considerable expansion of installations may be expected after the entry into force of the federal filter decrees of North Rhine-Westphalia (2013), Lower Saxony (2013), Schleswig-Holstein (2014), Thuringia (2016), and Brandenburg (2017), since, in the case of new installations, as well as considerable constructive changes to existing pig farms, the installation of waste air treatment plants is generally required [56]. If necessary, installation can also be imposed on smaller farms, but state-specific installation obligation varies between 1000 (Thuringia) and 10,000 (Brandenburg) fattening places. Nevertheless, the number of waste air treatment systems has thus risen sharply since 2013, but no current market figures are available. Since cattle farming causes approximately 51% of agricultural ammonia emissions, the expansion of these decrees to this application is expected in the next few years.
Biological waste air processes are widely used in the environmental and waste management industry, but also in some other fields of industrial and manufactural application (see Table 5) [57,58]. Further applications are known from automotive industry, cement industry, energy production, printing industry, the re-food sector, and other industrial applications. In all market sectors, they are mainly used for odor abatement, although isolated applications for the treatment of high carbon loads are also known.
The share of industrial waste air applications with VOC contents of several 100 mg C∙m−3 and potentially odorous compounds at temperatures less than 50 °C can be estimated at 10–15% of the total number of applications where waste air treatment is required, but only about 1% of these applications are assumed to be treated by biological waste air treatment processes. In this respect, there is a growth potential of 1400% for biological processes, which might be further increased by marketing of intelligent and economically interesting combined processes as described in Section 7 [59].

5. Biofiltration in Sludge Drying Processes

In recent years the drying of municipal and industrial sludges can be observed as a new market segment for biological exhaust air treatment. This development is based on the following three factors [60,61,62]:
  • according to German legislation sewage sludge may no longer be applied to agricultural fields;
  • in the course of the pending German coal exit, sewage sludge is increasingly seen as a substitute for lignite; and
  • sewage sludge ashes are rich in phosphates, which is why sewage sludge is becoming increasingly important as a source of phosphorus, despite the high costs of recovery. However, recovery requires a thermal pretreatment.
In the drying process, solar-supported low-temperature drying (up to approximately 50 °C) in greenhouse-like halls, as well as high-temperature processes (80 °C to approximately 300 °C) in the form of belt dryers, plate dryers, fluidized bed dryers, drum dryers, and paddle dryers, have been established as technical processes. Biofilters with one or more preconditioning stages have become established for the waste gas treatment at most drying locations. Due to the strongly varying drying temperatures, this results in strongly varying treatment times in the dryer (minutes in the high-temperature dryers, up to seven days in low-temperature dryers) [60], as well as high variations in both waste gas temperature and pollutant concentrations [63,64,65]. For selection of the right waste gas treatment technology, preliminary tests are therefore necessary to characterize the actual emissions. Typical plant configurations for low-temperature drying include a simple plant combination of acidic scrubber for ammonia elimination and a biofilter for odor elimination at volume loads of 250 m3∙m−3∙h−1, while high-temperature processes require more complex treatment systems consisting of at least two upstream scrubbers and a thermophilic biofilter with a volume load of less than 100 m3∙m−3∙h−1.
It should be pointed out that these plant configurations are suitable for compliance with the limit values of the German Technical Instructions on Air Quality Control (TA Luft). As an exception, methane emissions cannot be effectively treated by these processes (see Section 6.7).

6. Marketing Obstacles and Challenges

Despite considerable economic and ecological advantages, as well as a significant underrepresentation of biological processes in waste air treatment applications according to Section 4, biological waste air treatment processes still represent a niche solution due to the following technical limitations, which must be overcome to establish these processes successfully.

6.1. Limitation in Efficiency

Well-adjusted biofiltration processes may achieve stable efficiencies of about 85% for VOCs and above 95% for odor abatement. Further increases in efficiency are associated with an exponential increase in maintenance.

6.2. Limitation in VOC Crude Gas Levels

According to the German Technical Instructions on Air Quality Control clean gas levels of VOCs are restricted to 50 mg C∙m−3 in total. If an efficiency of 85% for treatment of VOC emissions according to 5.1. is considered, biological processes are firstly restricted to crude gas concentration of 333 mg C∙m−3 in maximum.

6.3. Limitation in Space Requirement

Higher crude gas levels can be successfully treated by increasing the empty bed residence time (EBRT), thus reducing volume specific VOC loads. However, the lower the specific loads, the higher the space requirements of the biofiltration processes are [5,7]. Based on process specific gas velocities (3–5 m∙s−1 for non-biological processes ex. adsorption; approximately 1.5 m∙s−1 for bio-scrubber; 0.3–1.0 m∙s−1 for biotrickling filter; less than 0.3 m∙s−1 for biofilter and adsorption), the required plant volume of biological processes is higher by a factor of 6–10 than the plant volume of non-biological processes.

6.4. Clogging

Biofiltration processes normally operate at volume-specific loads of 125 m³∙m−3∙h−1 for VOC elimination and 175 m³∙m−3∙h−1 for odor abatement, equivalent to VOC elimination efficiencies up to 100 g C∙m−3∙h−1, respectively. By reducing the space requirements, and hence the plant volume, the volume-specific loads increase. During biodegradation VOCs are converted both to mineralization products, like CO2, H2O, NH3, HCl, or similar, and especially energy, which is partly used in biomass formation. An increase in the volume-specific load considerably accelerates the total biomass production. If approaches to get rid of excessive biomass are not successful, the risk of biomass-induced clogging of the package increases considerably and operational time of the plant decreases dramatically [66,67]. Such approaches might be adequate operation parameters of the package materials (material, size, structure), physical and mechanical approaches (changes in irrigation or flow pattern, flushing, fluidization, high pressure nozzles, periodical or continuous extraction of the packing material and subsequent cleaning, submerged system, rotating contactors, squeezing techniques, or temperature shifts), chemical methods (shifts in pH or salinity, further addition of chemicals), advanced oxidation processes (NTP, UV oxidation, ozonation), biological approaches (shifts in microbial communities, addition of predators), or limitations (nutrients, starvation periods) [25]. With the exception of starvation and nutrient limitation, only a few studies have dealt with the validation of these methods for clogging prevention in pilot-scaled or industrial-scaled waste air treatment systems at gas flows of 100 m³∙h−1 or higher [25,68,69,70]

6.5. Limitation in Transformation Kinetics

A wide range of xenobiotic chemicals is used in industrial and commercial processes. By increasing the xenobiotic character of the chemicals, the biological transformation kinetics and, thus, the volume specific elimination efficiency decreases [71]. Therefore, larger plant sizes and, thus, more expensive installations are required for successful treatment of corresponding waste gas flows [5,71]. The low transformation kinetics of xenobiotic compounds can be based either on a low affinity of the relevant enzymes to their new substrates or on low induction by these new substrates during transcription of the associated gene sequences. The same applies to lipophilic waste gas compounds, as well as components with very high vapor pressures which, due to the limited water solubility, are only present at low concentrations in the aqueous phase or biofilm, and therefore show low bioavailability. Additionally, biodegradation follows a 1st order reaction kinetics. However, the objective of both compact and cost-effective biological waste gas treatment systems seems realistic in case of processes combining non-biological and biological approaches, in which the upstream non-biological stage enhances hydrophilicity and, thus, bioavailability of waste gas compounds by partial oxidation reactions (see Section 7 for details).

6.6. Grade of Persistence

The xenobiotic character of chemicals increases by higher ratio of anthropogenic substituents and binding types that do not occur naturally in the environment. Examples are halogen substituents, nitro groups or nitrile bonds. Since the exposure of microorganisms to these substituents and bonds has taken place for a short time from an evolutionary point of view, microorganisms generally have not adapted/optimized corresponding degradation enzymes nor established new types of highly efficient pathways. The absence of suitable enzymes may lead to failure of the application.
On the other hand, this postulate was contradicted by various biodegradation studies of xenobiotic compounds. As examples, biodegradation of 2-halobenzoates [72], 3-fluorobenzoates [73], 2,4-dichlorophenoxyacetate and 2-chloro-4-methylphenoxyacetate [74,75], fluorene and other PAHs [76,77], aryl, alkyl, biaryl and cyclically condensed biaryl ethers [78], fluorobenzene [79,80], 2-chlorotoluene [32,81], and triclosan [82,83,84] were shown successfully.

6.7. Limitation in Degradation of VOC Mixtures

With increasing complexity, number and variation of chemical substituents, the complexity of regulatory mechanisms both at gene and protein levels in biodegradation of waste gas compounds also increases. VOCs that belong to the same chemical group or lead to the same degradation pathway can usually be degraded microbially without any loss of enzymatic performance when presented as a mixture (e.g., alkyl aromatics, like toluene, styrene, and xylene isomers, can be degraded simultaneously by the same enzyme set) [85,86]. A comprehensive compilation of test series for the simultaneous degradation of aromatics can be found in [85]. However, simultaneous biodegradation of VOCs belonging to different chemical groups is more difficult, i.e., treatment of 1,2-dichlorobenzene in mixture with toluene and o-xylene [85], alcohols and ethers [87], methoxypropyl acetate and xylene isomers [88], or tert-butanol and acetone [89], as shown by different studies of our research group. Moreover, this effect was even observed in case of closely related groups of substances, like benzene in mixture with toluene, styrene, and xylene isomers [86]. While a simultaneous degradation of toluene, styrene, and xylene isomers at degradation rates corresponding to their specific water solubility occurred, benzene was not degraded due to a different initial ring oxidizing enzyme system.
Even though a natural compound, it should be mentioned that methane reveals poor biodegradability due to its high vapor pressure, low solubility in water, and high chemical stability. Especially in concentrations less than 100 ppm, and therefore relevant for the majority of large-scale applications, methane conversion was observed at very high EBRTs of 72–96 h, and performance further declined in presence of VOCs [90,91,92,93].
Hence, besides the affiliation to individual substance groups, the water solubility of components plays a decisive role in microbial biodegradation, as water solubility and, thus, bioavailability decreases with increasing lipophilicity. Accordingly, multiple studies pointed out the preferential degradation of hydrophilic VOCs in mixtures of both hydrophilic and lipophilic compounds. As an example, enhanced input loads of n-hexane and benzene in a ternary mixture of methanol, n-hexane, and benzene had no significant effect on methanol removal efficiency (>95%), while the removal efficiency of n-hexane was severely affected by changes in methanol loads (drop from 75% to 32%) [94,95]. Comparable results were observed for mixtures of BTEX and n-hexane or benzene and n-hexane, where the removal efficiency of n-hexane as a very lipophilic compound decreased in presence of these aromatic compounds [96,97].
Therefore, simultaneous degradation of VOCs representing different chemical groups is often not possible and sufficient degradation may not be achieved. However, this problem mainly occurs in bioscrubber systems with homogeneous distribution of microbial consortia and pollutants, while this problem is of less relevance in biofiltration processes due to different established microbial consortia along the height of the packing material, i.e., hydrophilic and therefore highly bioavailable compounds are commonly mineralized near to the crude gas inlet of the packing, while lipophilic compounds are degraded downflow.

6.8. Limitation in Gas Temperature

The vast majority of microbial primary degraders grow under mesophilic conditions, i.e., approximately 10–40 °C. The transformation optimum is usually around 30 °C. Additionally, thermotolerant or thermophilic microorganisms were also applied in treatment processes for waste air temperatures of 50 °C or higher, like odor emissions of a cocoa factory [31], emissions of sludge drying processes [98], wood processing industry [31,99,100], or treatment of VOC mixtures, like benzene, n-hexane, and toluene [101]. Even though the process adapted microorganisms exhibit higher transformation kinetics due to the higher temperature according to the van´t-Hoff equation, biomass yield decreases simultaneously due to increasing maintenance metabolism [102]. The start-up process of industrial plants is thus considerably longer and more difficult. In particular, stable operation of such thermophilic plants is fundamentally more challenging, since they require constant temperatures above 45 °C, which commonly cannot be guaranteed during maintenance cycles. Besides recurring changes between mesophilic and thermophilic conditions, increases in vapor pressures and, thus, lower bioavailability, as well as toxicity, are the main reasons for operational problems. Tests in our research group showed that the biodegradability of formaldehyde as a VOC compound in wood processing industry emissions particularly depends on the waste gas temperature and therefore exhibits high toxicity at thermophilic conditions [103]. In contrast, degradability of methanol and α-pinene as concomitant substances was successfully shown in literature [99,100]. However, significant formaldehyde emissions were exemplarily observed in hot waste gas streams of biogas plants, combined heat and power plants, wood and pulp industry, plastic coating, and ceramics industry [104].

7. Combination of (Non-)Biological Processes

With increasing complexity of waste gas situations multi-stage processes and process combinations are required for successful treatment. For example, three-stage process combinations of acid scrubber, alkaline scrubber and biofilter (rarely a condenser as alternative) represent the state of the art in waste air treatment of sewage sludge drying plants. The waste gas at temperatures of 20–150 °C (depending on the drying process) is firstly cooled adiabatically and cleaned of ammonia, as well as dust particles, in the acidic stage. Separation of H2S, mercaptans, and organic acids takes places in the downstream alkaline scrubber, in which cleaning efficiency can be considerably improved by addition of oxidizing agents. The residual VOCs in the waste gas are treated by a biofilter as third stage.

7.1. Innovative Combinations of Advanced Oxidation and Biological Processes

In addition to this conventional-looking process combination of two scrubbers and a biofilter, research recently focused on innovative process combinations of biological treatment and advanced oxidation technologies, like non-thermal plasma or UV oxidation. Depending on the sequence of the processes either a partial oxidation of waste gas compounds or a combined finishing and disinfection of the waste gas can be achieved. In the former case, VOCs are partially oxidized by the upstream advanced oxidation process; thus, biological availability of the waste gas compounds is achieved by increasing the water solubility of the intermediates formed [105]. In the latter case, waste gas compounds are mainly mineralized by the biological process and the downstream advanced oxidation process is used as finishing and disinfection step to enable recirculation of the treated air. UV process combinations have recently been used in treatment of chlorobenzene [106,107], dichloromethane [108], ethylbenzene [109], n-hexane [110], PCE [111], α-pinene [112,113,114], styrene [39,115], TCE [111], toluene [116,117], and o-xylene [118] as single substances or for mixtures of toluene and o-xylene [119], a mixture of ethyl acetate, toluene, ethylbenzene, xylene, ethyltoluene, and trimethylbenzene [120], and in a mixture of toluene, ethylbenzene, and xylene isomers [121]. NTP-based combinations were applied in treatment of chlorobenzene [122,123], 1,2-DCE [124], dimethyl disulfide [125], and limonene [126] as single substances and in a couple of different VOC mixtures [34,43,45,105,124,127,128,129,130,131,132]. Additionally, first tests with magnetic fields as activation step were performed [133,134]. Details of these studies are listed in Table 6 (magnetic fields), Table 7 (UV), and Table 8 (NTP). Data of Table 6, Table 7 and Table 8 shows that in principle all three process combinations (biological process and magnetic field, UV, or NTP) reveal a higher removal efficiency compared to the sole processes. This achieved efficiency is also above the cumulative efficiencies of the individual processes, i.e., the waste gas components are first partially oxidized in the advanced oxidation process and can then be degraded much more efficiently in the downstream biological stage due to the increased hydrophilicity and, thus, bioavailability of the intermediates. Comparing advanced oxidation processes, the use of these processes leads to an increase in efficiency of 3–15% (magnetic field), 20–50% (UV oxidation), and 25–80% (NTP) compared to the efficiency of the biological process alone. The total achievable efficiencies are the highest in the combination of NTP and a subsequent biological stage. Finally, it should be mentioned that a combination of non-thermal plasma and physical scrubbers for elimination of lipophilic VOCs was also presented [135].
The use of these process combinations appears to be of high interest especially with respect to biologically persistent compounds, since the partial oxidation of the persistent pollutants can significantly increase the biodegradability. For example, in the case of freon R-22, dehalogenated reaction products were identified after exposure to non-thermal plasma, which tend to exhibit higher biodegradability than the original substance [136].
In addition to the requirements of a multi-stage process due to complex waste gas situations, process combinations of UV/NTP with biofiltration processes are often used to reduce both investment costs and operating costs of the overall waste air treatment concept. Compared to a concept based purely on UV oxidation or NTP processes, the process combination has considerably lower electricity costs, since waste gas contaminants are only partially oxidized and intermediates formed can be further treated by biofiltration processes without significant operating costs. Compared to a sole biological process, the treatment of partially oxidized intermediates allows a significantly higher bioavailability and, thus, transformation kinetics compared to the original waste gas contaminants. Hence, the bioprocess can be built significantly smaller and thus more cost-efficient [45]. Furthermore, more stable reactive compounds of the pre-treatment, like ozone, affect the microbial community in the biofiltration process [108,122,125,137]. Thus, a continuous presence of the stressor ozone leads to a lower biomass yield, lower pressure losses, and, thus, a lower tendency to clogging [114,119,122,138].

7.2. Combination of Biological Processes

Besides combinations of non-biological and biological processes, multi-stage process combinations of biological processes are also used. These combinations are commonly used in case of VOC mixtures of strongly divergent polarity with high pollutants loads, so that in case of a single-stage biotrickling filter or biofilter long-term stability seems to be questionable due to expected clogging. As an example, a combination of BTF and BF was used for treatment of H2S, dimethyl sulfide and dimethyl disulfide [139]. The water-soluble hydrogen sulfide was completely mineralized in the BTF, while the degradation of the two water-insoluble organosulfur compounds took place in the biofilter. Gerl et al. [88] examined biodegradation of a mixture of polar methoxypropyl acetate and lipophilic xylene isomers at a 1:1 ratio of the ester and the aromatics by a combination of bioscrubber and biotrickling filter in comparison to a biofilter as reference system. While the performance of the biofilter was approximately 20% with increasing pressure losses as secondary problem, the combined process revealed long-term stability in operation at elimination efficiencies or more than 70% at the same EBRT.

8. Conclusions

International studies expect a further increase in greenhouse gas emissions over the next 20 years despite increasing efforts to improve energy efficiency in technical processes. In order to meet the international goal of limiting global warming to 2 °C, industrialized countries will have to reduce emissions of CO2 equivalents by more than 90%. That way, mankind simply can no longer afford to generate immense CO2 emissions by waste gas treatment technologies. Development and application of sustainable, energy and resource efficient processes for waste air treatment is obligatory. By a large extent, existing biological waste air treatment techniques meet these requirements. According to the limitations mentioned above, these processes reveal moderate performance levels in case of simultaneous degradation of VOC mixtures and, therefore, have high space requirements. These problems and corresponding marketing obstacles, as well, can be counteracted by the innovative approach of combining advanced oxidation and biofiltration processes, so that compact, energy-efficient, and low-emissions technologies are available for agricultural, municipal, and industrial waste gas applications. Although biological processes, UV oxidation, and NTP processes, as stand-alone processes, have been intensively described, combinations of these processes have rarely been discussed. Experiences in the semi- and large-scale technical scale are even more limited. Future research will therefore have to deal with the process stability (NTP process), the stable transformation efficiency (UV process), the influence of interfering components (especially dusts and aerosols in water vapor saturated emission situations; NTP and UV processes), and the in-depth effect of reactive oxidative components on the performance and growth of the microbial consortium of the subsequent biological stage.

Author Contributions

Conceptualization and Methodology, D.D. and O.O.; calculations, D.D.; investigation, D.D. and O.O.; writing—original draft preparation, review and editing, D.D. and O.O.; visualization, D.D.; project administration, D.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We would like to thank Isabelle Franzen-Reuter (FH Münster) and Martina Kiel (University of Stuttgart) for their critical technical and lingual editing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. International Energy Agency IEA. World Energy Outlook 2017; International Energy Agency: Paris, France, 2017. [Google Scholar]
  2. British Petroleum PLC. BP Energy Outlook-2018 ed.; British Petroleum PLC: London, UK, 2018. [Google Scholar]
  3. WBGU, Wissenschaftlicher Beirat der Bundesregierung Globale Umweltveränderung. Kassensturz für den Weltklimavertrag-der Budgetansatz, Sondergutachten; WBGU: Berlin, Germany, 2009; ISBN 978-3-936191-26-4. [Google Scholar]
  4. Ravi, R.; Philip, L.; Swaminathan, T. Comparison of biological reactors (biofilter, biotrickling filter and modified RBC) for treating dichloromethane vapors. J. Chem. Technol. Biotechnol. 2010, 85, 634–639. [Google Scholar] [CrossRef]
  5. 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]
  6. Singh, A.; Shareefdeen, Z.; Ward, O.P. Bioscrubber Technology. In Biotechnology for Odor and Air Pollution Control; Shareefdeen, Z., Singh, A., Eds.; Springer: Berlin/Heidelberg, 2005; pp. 169–193. ISBN 3-540-23312-1. [Google Scholar]
  7. Kolekar, S.; Patil, A.; Shingare, S.P.; Mandake, M.B. Biofilter system for air pollution control: A review. Int. Res. J. Eng. Technol. 2019, 6, 2053–2058. [Google Scholar]
  8. Humeau, P.; Pré, P.; Le Cloirec, P. Optimization of Bioscrubber Performances: Experimental and Modeling Approaches. J. Environ. Eng. 2004, 130, 314–321. [Google Scholar] [CrossRef]
  9. 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]
  10. Van Groenestijn, J.W. Bioscrubbers. In Bioreactors for Waste Gas Treatment; Alloway, B.J., Trevors, J.T., Kennes, C., Veiga, M.C., Eds.; Springer: Dordrecht, The Netherlands, 2001; pp. 133–162. ISBN 978-90-481-5772-3. [Google Scholar]
  11. Zamir, S.M.; Babatabar, S.; Shojaosadati, S.A. Styrene vapor biodegradation in single- and two-liquid phase biotrickling filters using Ralstonia eutropha. Chem. Eng. J. 2015, 268, 21–27. [Google Scholar] [CrossRef]
  12. Song, T.; Yang, C.; Zeng, G.; Yu, G.; Xu, C. Effect of surfactant on styrene removal from waste gas streams in biotrickling filters. J. Chem. Technol. Biotechnol. 2012, 87, 785–790. [Google Scholar] [CrossRef]
  13. San-Valero, P.; Gabaldón, C.; Penya-Roja, J.M.; Quijano, G. Enhanced styrene removal in a two-phase partitioning bioreactor operated as a biotrickling filter: Towards full-scale applications. Chem. Eng. J. 2017, 309, 588–595. [Google Scholar] [CrossRef] [Green Version]
  14. San-Valero, P.; Álvarez-Hornos, J.; Ferrero, P.; Penya-Roja, J.M.; Marzal, P.; Gabaldón, C. Evaluation of Parallel-Series Configurations of Two-Phase Partitioning Biotrickling Filtration and Biotrickling Filtration for Treating Styrene Gas-Phase Emissions. Sustainability 2020, 12, 6740. [Google Scholar] [CrossRef]
  15. Rezaei, M.; Moussavi, G.; Naddafi, K.; Johnson, M.S. Enhanced biodegradation of styrene vapors in the biotrickling filter inoculated with biosurfactant-generating bacteria under H2O2 stimulation. Sci. Total Environ. 2020, 704, 135325. [Google Scholar] [CrossRef]
  16. Rene, E.R.; Montes, M.; Veiga, M.C.; Kennes, C. Styrene removal from polluted air in one and two-liquid phase biotrickling filter: Steady and transient-state performance and pressure drop control. Bioresour. Technol. 2011, 102, 6791–6800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. LFKW University of Stuttgart. Betriebsdaten LFKW Jahresbericht 2017. Available online: http://www.iswa.uni-stuttgart.de/lfkw/betriebsdaten/2017Uebersicht.html (accessed on 14 October 2020).
  18. Misra, M.K.; Ragland, K.W.; Baker, A.J. Wood ash composition as a function of furnace temperature. Biomass Bioenergy 1993, 4, 103–116. [Google Scholar] [CrossRef]
  19. Schafer, H.; Myronova, N.; Boden, R. Microbial degradation of dimethylsulphide and related C1-sulphur compounds: Organisms and pathways controlling fluxes of sulphur in the biosphere. J. Exp. Bot. 2010, 61, 315–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. 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]
  21. San-Valero, P.; Penya-Roja, J.M.; Sempere, F.; Gabaldón, C. Biotrickling filtration of isopropanol under intermittent loading conditions. Bioprocess. Biosyst. Eng. 2013, 36, 975–984. [Google Scholar] [CrossRef] [Green Version]
  22. Melse, R.W.; Mosquera, J. Nitrous oxide (N2O) emissions from biotrickling filters used for ammonia removal at livestock facilities. Water Sci. Technol. 2014, 69, 994–1003. [Google Scholar] [CrossRef]
  23. Sapkota, A.; Haghverdi, A.; Avila, C.C.E.; Ying, S.C. Irrigation and Greenhouse Gas Emissions: A Review of Field-Based Studies. Soil Syst. 2020, 4, 20. [Google Scholar] [CrossRef]
  24. KTBL. Abluftreinigung in der Schweinehaltung-Verfahren, Einsatzbereiche, Leistungen und Kosten: Kuratorium für Technik und Bauwesen in der Landwirtschaft e.V. 2010. 1–6. Available online: https://docplayer.org/19734418-Abluftreinigung-in-der-schweinehaltung-verfahren-einsatzbereiche-leistungen-und-kosten.html (accessed on 14 October 2020).
  25. 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]
  26. Kan, E.; Deshusses, M.A. Development of foamed emulsion bioreactor for air pollution control. Biotechnol. Bioeng. 2003, 84, 240–244. [Google Scholar] [CrossRef]
  27. Wang, L.; He, S.; Xu, J.; Li, J.; Mao, Z. Process Performance of a Biotrickling Filter Using a Flow-Directional-Switching Method. Clean Soil Air Water 2013, 41, 522–527. [Google Scholar] [CrossRef]
  28. Ryu, H.W.; Kim, S.J.; Cho, K.S.; Lee, T.H. Toluene degradation in a polyurethane biofilter at high loading. Biotechnol. Bioproc. E 2008, 13, 360–365. [Google Scholar] [CrossRef]
  29. Datta, A.; Philip, L. Performance of a rotating biological contactor treating VOC emissions from paint industry. Chem. Eng. J. 2014, 251, 269–284. [Google Scholar] [CrossRef]
  30. VDI. Guideline No. 3478-2: Biological Waste Gas Purification-Biological Trickle Bed-Reactors, 2008th ed.; Verein Deutscher Ingenieure VDI, Ed.; Beuth Verlag GmbH: Berlin, Germany, 2008. [Google Scholar]
  31. van Groenestijn, J.W.; Kraakman, N.J.R. Recent developments in biological waste gas purification in Europe. Chem. Eng. J. 2005, 113, 85–91. [Google Scholar] [CrossRef]
  32. Dobslaw, D.; Engesser, K.-H. Biodegradation of gaseous emissions of 2-chlorotoluene by strains of Rhodococcus sp. in polyurethane foam packed biotrickling filters. Sci. Total Environ. 2018, 639, 1491–1500. [Google Scholar] [CrossRef]
  33. Eitner, D. Biofilter in der Abluftreinigung: Biomassen-Planung-Kosten-Einsatzmöglichkeiten. Bwk Das Energ. Fachmag. 1989, 41, L24–L29. [Google Scholar]
  34. Jager, J.; Bockreis, A.; Steinberg, I.; Rohde, C.; Laser, U. Einsatz Eines Alternativen Abluftreinigungsverfahrens Kombiniert mit dem Bestehenden Biologischen Abluftreinigungsverfahren in Einer Mechanisch-Biologischen Restabfallbehandlungsanlage. BWPLUS-Förderkennzeichen BWD 21013, TU Darmstadt; PUDI Document Server of the State Institute for the Environment Baden-Württemberg, 2004; Available online: https://pudi.lubw.de (accessed on 14 October 2020).
  35. Loy, J. Biotrickling Technologie-Geruchsreduzierung und Lösungsmittelbeseitigung aus Abluft mit Biotrickling Technologie. 2006. Available online: http://www.loy.tc/documents/biotrickling-presentation_dt.ppt.pdf (accessed on 14 October 2020).
  36. Estrada, J.M.; Kraakman, N.J.R.B.; Lebrero, R.; Muñoz, R. A sensitivity analysis of process design parameters, commodity prices and robustness on the economics of odour abatement technologies. Biotechnol. Adv. 2012, 30, 1354–1363. [Google Scholar] [CrossRef]
  37. Prado, O.J.; Gabriel, D.; Lafuente, J. Economical assessment of the design, construction and operation of open-bed biofilters for waste gas treatment. J. Environ. Manag. 2009, 90, 2515–2523. [Google Scholar] [CrossRef]
  38. Estrada, J.M.; Kraakman, N.J.R.B.; Muñoz, R.; Lebrero, R. A comparative analysis of odour treatment technologies in wastewater treatment plants. Environ. Sci. Technol. 2011, 45, 1100–1106. [Google Scholar] [CrossRef]
  39. Álvarez-Hornos, F.J.; Martínez-Soria, V.; Marzal, P.; Izquierdo, M.; Gabaldón, C. Performance and feasibility of biotrickling filtration in the control of styrene industrial air emissions. Int. Biodeterior. Biodegrad. 2017, 119, 329–335. [Google Scholar] [CrossRef] [Green Version]
  40. Deshusses, M.A.; Cox, H.H.J. Biotrickling Filters for Air Pollution Control. In Encyclopedia of Environmental Microbiology; Bitton, G., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2003; ISBN 0471263397. [Google Scholar]
  41. Gabriel, D.; Deshusses, M.A. Technical and economical analysis of the conversion of a full-scale scrubber to a biotrickling filter for odor control. Water Sci. Technol. 2004, 50, 309–318. [Google Scholar] [CrossRef]
  42. Dobslaw, D. Herausforderungen in der Biologischen und Nicht-Biologischen Abluftreinigung; Universität Stuttgart: Stuttgart, Germany, 2020. [Google Scholar] [CrossRef]
  43. Rafflenbeul, R. Nichtthermische Plasmaanlagen (NTP) zur Luftreinhaltung in der Abfallwirtschaft. Müll Abfall 1998. [Google Scholar] [CrossRef]
  44. Wang, C.; Xi, J.; Hu, H.; Kang, I. Effects of design parameters on performance and cost analysis of combined ultraviolet-biofilter systems treating gaseous chlorobenzene based on mathematical modeling. Front. Environ. Sci. Eng. 2012, 6, 588–594. [Google Scholar] [CrossRef]
  45. Dobslaw, D.; Schulz, A.; Helbich, S.; Dobslaw, C.; Engesser, K.-H. VOC removal and odor abatement by a low-cost plasma enhanced biotrickling filter process. J. Environ. Chem. Eng. 2017, 5, 5501–5511. [Google Scholar] [CrossRef]
  46. Umweltbundesamt. Entwicklung der Spezifischen Kohlendioxid-Emissionen des Deutschen Strommix in den Jahren 1990–2019; Umweltbundesamt: Dessau-Roßlau, Germany, 2020; ISSN 1862-4359. [Google Scholar]
  47. Clemens, J.; Cuhls, C. Greenhouse gas emissions from mechanical and biological waste treatment of municipal waste. Environ. Technol. 2003, 24, 745–754. [Google Scholar] [CrossRef]
  48. Schachinger, I. Untersuchungen an Hochleistungs-Feinkorn-Beton.38; DAfSTb-Forschungskolloquium: München, Germany, 2000; pp. 55–66. [Google Scholar]
  49. Schuch, R. Stickstoffeliminierung aus Betriebsabwässern der Metallverarbeitung. Ph.D. Thesis, Universität Karlsruhe, Karlsruhe, Germany, 2000. [Google Scholar]
  50. VDI. Guideline No. 3475 Emission Control-Biological Waste Treatment Facilities-Anaerobic Digestion and Post Treatmen, 2015th ed.; Verein Deutscher Ingenieure VDI, Ed.; Beuth Verlag GmbH: Berlin, Germany, 2015. [Google Scholar]
  51. VDLUFA. Kreislaufwirtschaft mit der Landwirtschaft-quo Vadis; VDLUFA: Speyer, Germany, 2005. [Google Scholar]
  52. Grömping, M.; Kollbach, J.S. Stickstoffrückbelastung. Stand. der Technik 1996/97; Zukünftige Entwicklungen; TK-Verl. Thomé-Kozmiensky: Neuruppin, Germany, 1996; ISBN 3-924511-85-3. [Google Scholar]
  53. German Federal Statistical Office. Land-und Forstwirtschaft, Fischerei-Viehbestand. Fachserie 3, Reihe 4.1.; Destatis: Wiesbaden, Germany, 2017.
  54. Hahne, J. Entwicklung der Abluftreinigung in der Tierhaltung in Deutschland. Landtechnik 2011, 66, 289–293. [Google Scholar] [CrossRef]
  55. Hahne, J. Abluftreinigung-ein Weg zu einer emissionsarmen Tierhaltung. Forschungsreport des Thünen-Instituts 1/2012. Available online: https://www.thuenen.de/media/ti-themenfelder/Klima_und_Luft/Abluftreinigung_in_Tierstaellen/FoReport_1-2012_Abluft.pdf (accessed on 14 October 2020).
  56. Ministry for the Environment, Energy and Nature Conservation of Thuringia. Immissionsschutz rechtliche Anforderungen an Tierhaltungsanlagen und Einführung der Geruchs-Immissionsrichtlinie (GIRL): Sog. Filtererlass. Available online: https://umwelt.thueringen.de/fileadmin/001_TMUEN/Unsere_Themen/Boden_Wasser_Luft_Laerm/Immissionsschutz/filtererlass_thuringen_2016.pdf (accessed on 14 October 2020).
  57. Braun, R.; Holubar, P.; Pias, C. Biologische Abluftreinigung in Österreich. Studie im Auftrag des Bundesministeriums für Umwelt, Jugend und Familie, Sektion II.; Bundesministerium für Umwelt, Jugend und Familie: Wien, Austria, 1994. [Google Scholar]
  58. Margesin, R.; Schneider, M.; Schinner, F. Praxis der biotechnologischen Abluftreinigung; Margesin, R., Schneider, M., Schinner, F., Eds.; Springer: Berlin/Heidelberg, Germany, 1996; ISBN 978-3-540-59335-5. [Google Scholar]
  59. Shareefdeen, Z.; Singh, A. Biotechnology for Odor and Air Pollution Control; Springer: Berlin, Germany, 2005; ISBN 978-3-540-23312-1. [Google Scholar]
  60. Bennamoun, L. Solar drying of wastewater sludge: A review. Renew. Sustain. Energy Rev. 2012, 16, 1061–1073. [Google Scholar] [CrossRef]
  61. Heindl, A.; Gröbl, T.; Görlich, M.; Graf, M. Thermische Behandlung von Klärschlamm. Teil 1: Rechtliche Vorschriften, Eigenschaftren und Trocknung von Klärschlamm. Korresp. Abwasser Abfall 2020, 67, 286–294. [Google Scholar]
  62. Heindl, A.; Gröbl, T.; Görlich, M.; Graf, M. Thermische Behandlung von Klärschlamm. Teil 4: Emissionen und Dauerbetrieb von Wirbelschichtverbrennungsanlagen. Korresp. Abwasser Abfall 2020, 67, 526–534. [Google Scholar]
  63. Bauerfeld, K.; Dellbrügge, R.; Dichtl, N.; Großer, A.; Paris, S. Technology transfer-oriented research and development in the wastewater sector-validation at industrial-scale plants (EXPOVAL)-Subgroup 6: Solar sewage sludge drying-First results from investigations with a pilot plant. EXPOVAL 2014, 10, 370–380. [Google Scholar]
  64. Maurer, C.; Müller, J. Ammonia (NH3) emissions during drying of untreated and dewatered biogas digestate in a hybrid waste-heat/solar dryer. Eng. Life Sci. 2012, 12, 321–326. [Google Scholar] [CrossRef]
  65. Awiszus, S.; Meissner, K.; Reyer, S.; Müller, J. Ammonia and methane emissions during drying of dewatered biogas digestate in a two-belt conveyor dryer. Bioresour. Technol. 2018, 247, 419–425. [Google Scholar] [CrossRef] [PubMed]
  66. Delhoménie, M.-C.; Bibeau, L.; Gendron, J.; Brzezinski, R.; Heitz, M. A study of clogging in a biofilter treating toluene vapors. Chem. Eng. J. 2003, 94, 211–222. [Google Scholar] [CrossRef]
  67. Jin, Y.; Veiga, M.C.; Kennes, C. Development of a novel monolith-bioreactor for the treatment of VOC-polluted air. Environ. Technol. 2006, 27, 1271–1277. [Google Scholar] [CrossRef]
  68. Webster, T.S.; Cox, H.H.J.; Deshusses, M.A. Resolving operational and performance problems encountered in the use of a pilot/full-scale biotrickling filter reactor. Environ. Prog. 1999, 18, 162–172. [Google Scholar] [CrossRef]
  69. Zhao, L.; Huang, S.; Wei, Z. A demonstration of biofiltration for VOC removal in petrochemical industries. Environ. Sci. Process. Impacts 2014, 16, 1001–1007. [Google Scholar] [CrossRef]
  70. Lim, J.S.; Hwang, J.W.; Choi, C.Y.; Kim, B.H.; Park, S.H. A Pilot-Scale Rotating Drum Biotrickling Filter for Removing Gaseous Styrene. KEM 2005, 277–279, 517–522. [Google Scholar] [CrossRef]
  71. Iranpour, R.; Cox, H.H.J.; Deshusses, M.A.; Schroeder, E.D. Literature review of air pollution control biofilters and biotrickling filters for odor and volatile organic compound removal. Environ. Prog. 2005, 24, 254–267. [Google Scholar] [CrossRef]
  72. Engesser, K.H.; Schulte, P. Degradation of 2-bromo-, 2-chloro- and 2-fluorobenzoate by Pseudomonas putida CLB 250. Fems Microbiol. Lett. 1989, 60, 143–147. [Google Scholar] [CrossRef]
  73. Engesser, K.H.; Auling, G.; Busse, J.; Knackmuss, H.-J. 3-Fluorobenzoate enriched bacterial strain FLB 300 degrades benzoate and all three isomeric monofluorobenzoates. Arch. Microbiol. 1990, 153, 193–199. [Google Scholar] [CrossRef] [Green Version]
  74. Pieper, D.H.; Stadler-Fritzsche, K.; Engesser, K.H.; Knackmuss, H.J. Metabolism of 2-chloro-4-methylphenoxyacetate by Alcaligenes eutrophus JMP 134. Arch. Microbiol. 1993, 160, 169–178. [Google Scholar] [CrossRef] [PubMed]
  75. Pieper, D.H.; Reineke, W.; Engesser, K.-H.; Knackmuss, H.J. Metabolism of 2,4-dichlorophenoxyacetic acid, 4-chloro-2-methylphenoxyacetic acid and 2-methylphenoxyacetic acid by Alcaligenes eutrophus JMP 134. Arch. Microbiol. 1988, 150, 95–102. [Google Scholar] [CrossRef] [Green Version]
  76. Trenz, S.P.; Engesser, K.H.; Fischer, P.; Knackmuss, H.J. Degradation of fluorene by Brevibacterium sp. strain DPO 1361: A novel C-C bond cleavage mechanism via 1,10-dihydro-1,10-dihydroxyfluoren-9-one. J. Bacteriol. 1994, 176, 789–795. [Google Scholar] [CrossRef] [Green Version]
  77. Ghosal, D.; Ghosh, S.; Dutta, T.K.; Ahn, Y. Current State of Knowledge in Microbial Degradation of Polycyclic Aromatic Hydrocarbons (PAHs): A Review. Front. Microbiol. 2016, 7, 1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Engesser, K.H. Kritische Reaktionen beim bakteriellen Abbau Aromatischer Xenobiotika. Habilitationsschrift, Universität Stuttgart, Stuttgart, Germany, 1992. [Google Scholar]
  79. Carvalho, M.F.; Ferreira, M.I.M.; Moreira, I.S.; Castro, P.M.L.; Janssen, D.B. Degradation of fluorobenzene by Rhizobiales strain F11 via ortho cleavage of 4-fluorocatechol and catechol. Appl. Environ. Microbiol. 2006, 72, 7413–7417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Strunk, N.; Engesser, K.-H. Degradation of fluorobenzene and its central metabolites 3-fluorocatechol and 2-fluoromuconate by Burkholderia fungorum FLU100. Appl. Microbiol. Biotechnol. 2013, 97, 5605–5614. [Google Scholar] [CrossRef] [PubMed]
  81. Dobslaw, D.; Engesser, K.-H. Degradation of 2-chlorotoluene by Rhodococcus sp. OCT 10. Appl. Microbiol. Biotechnol. 2012, 93, 2205–2214. [Google Scholar] [CrossRef] [PubMed]
  82. Chen, X.; Nielsen, J.L.; Furgal, K.; Liu, Y.; Lolas, I.B.; Bester, K. Biodegradation of triclosan and formation of methyl-triclosan in activated sludge under aerobic conditions. Chemosphere 2011, 84, 452–456. [Google Scholar] [CrossRef] [Green Version]
  83. Lee, D.G.; Zhao, F.; Rezenom, Y.H.; Russell, D.H.; Chu, K.-H. Biodegradation of triclosan by a wastewater microorganism. Water Res. 2012, 46, 4226–4234. [Google Scholar] [CrossRef]
  84. Wang, S.; Yin, Y.; Wang, J. Microbial degradation of triclosan by a novel strain of Dyella sp. Appl. Microbiol. Biotechnol. 2018, 102, 1997–2006. [Google Scholar] [CrossRef]
  85. Dobslaw, D. Der Bakterielle Abbau von Halogen-und Methylsubstituierten Aromatengemischen und Dessen Technische Anwendung in der Biologischen Abluftreinigung; Oldenbourg-Industrieverl: München, Germany, 2009; ISBN 978-3-8356-3192-2. [Google Scholar]
  86. Liao, D.; Li, E.; Li, J.; Zeng, P.; Feng, R.; Xu, M.; Sun, G. Removal of benzene, toluene, xylene and styrene by biotrickling filters and identification of their interactions. PLoS ONE 2018, 13, e0189927. [Google Scholar] [CrossRef] [PubMed]
  87. Dobslaw, D.; Dobslaw, C.; Fütterer, N.; Engesser, K.-H. Biologische Abluftreinigung von Druckereiablüften: Optimierung eines Biowäschers. Chem. Ing. Tech. 2007, 79, 1679–1685. [Google Scholar] [CrossRef]
  88. Gerl, T.; Engesser, K.-H.; Fischer, K.; Dobslaw, D. Biologische Abluftreinigung einer Lackierabluft im Kombinationsverfahren. Chem. Ing. Tech. 2016, 88, 1145–1150. [Google Scholar] [CrossRef]
  89. 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]
  90. Albanna, M.; Warith, M.; Fernandes, L. Kinetics of biological methane oxidation in the presence of non-methane organic compounds in landfill bio-covers. Waste Manag. 2010, 30, 219–227. [Google Scholar] [CrossRef]
  91. Heipieper, H.J.; de Bont, J.A.M. Methane oxidation by Dutch grassland and peat soil microflora. Chemosphere 1997, 35, 3025–3037. [Google Scholar] [CrossRef]
  92. Wilde, C.; Dobslaw, D.; Engesser, K.H. Applicability and Limits of Biological Waste Air Treatment Techniques. In Proceedings of the TISD2010—3rd Technology and Innovation for Sustainable Development International Conference, Nong Khai, Thailand, 4–6 March 2010. [Google Scholar]
  93. 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]
  94. 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]
  95. Zehraoui, A.; Sorial, G.A. Treatment of Dynamic Mixture of n -Hexane, Benzene, and Methanol and Fungi Community Characterization in an Integrated Scheme of Cyclic Adsorption/Desorption Beds and Trickle Bed Air Biofilter. Air Soil Water Res. 2015, 8, ASWR.S23688. [Google Scholar] [CrossRef] [Green Version]
  96. Amin, M.M.; Rahimi, A.; Bina, B.; Nourmoradi, H.; Hassanvand, M.S.; Mohammadi-Moghadam, F.; Norouzi, S.; Heidari, M. Biodegradation of n-hexane as single pollutant and in a mixture with BTEX in a scoria/compost-based biofilter. Process. Saf. Environ. Prot. 2017, 107, 508–517. [Google Scholar] [CrossRef]
  97. Hassan, A.A.; Sorial, G.A. Biofiltration of n-hexane in the presence of benzene vapors. J. Chem. Technol. Biotechnol. 2010, 85, 371–377. [Google Scholar] [CrossRef]
  98. Yang, K.; Li, L.; Ding, W.; Liu, J.; Xue, S. A full-scale thermophilic biofilter in the treatment of sludge drying exhaust: Performance, microbial characteristics and bioaerosol emission. J. Chem. Technol. Biotechnol. 2018, 93, 2216–2225. [Google Scholar] [CrossRef]
  99. Dhamwichukorn, S.; Kleinheinz, G.T.; Bagley, S.T. Themophilic biofiltration of methanol and alpha-pinene. J. Ind. Microbiol. Biotechnol. 2001, 26, 127–133. [Google Scholar] [CrossRef]
  100. Kong, Z.; Farhana, L.; Fulthorpe, R.R.; Allen, D.G. Treatment of volatile organic compounds in a biotrickling filter under thermophilic conditions. Environ. Sci. Technol. 2001, 35, 4347–4352. [Google Scholar] [CrossRef] [PubMed]
  101. Hu, Q.-Y.; Wang, C.; Huang, K.-X. Biofiltration performance and characteristics of high-temperature gaseous benzene, hexane and toluene. Chem. Eng. J. 2015, 279, 689–695. [Google Scholar] [CrossRef]
  102. Mohammad, B.T.; Veiga, M.C.; Kennes, C. Mesophilic and thermophilic biotreatment of BTEX-polluted air in reactors. Biotechnol. Bioeng. 2007, 97, 1423–1438. [Google Scholar] [CrossRef]
  103. Neugart, E. Charakterisierung und Betrieb einer Methanol abbauenden, thermophilen Mischkultur. Bachelor’s Thesis, Universität Stuttgart, Stuttgart, Germany, 2015. [Google Scholar]
  104. 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]
  105. Schiavon, M.; Schiorlin, M.; Torretta, V.; Brandenburg, R.; Ragazzi, M. Non-thermal plasma assisting the biofiltration of volatile organic compounds. J. Clean. Prod. 2017, 148, 498–508. [Google Scholar] [CrossRef]
  106. Wang, C.; Xi, J.Y.; Hu, H.Y.; Yao, Y. Advantages of combined UV photodegradation and biofiltration processes to treat gaseous chlorobenzene. J. Hazard Mater. 2009, 171, 1120–1125. [Google Scholar] [CrossRef]
  107. Wang, C.; Xi, J.-Y.; Hu, H.-Y. Reduction of toxic products and bioaerosol emission of a combined ultraviolet-biofilter process for chlorobenzene treatment. J. Air Waste Manag. Assoc. 2009, 59, 405–410. [Google Scholar] [CrossRef]
  108. Jianming, Y.; Wei, L.; Zhuowei, C.; Yifeng, J.; Wenji, C.; Jianmeng, C. Dichloromethane removal and microbial variations in a combination of UV pretreatment and biotrickling filtration. J. Hazard Mater. 2014, 268, 14–22. [Google Scholar] [CrossRef] [PubMed]
  109. Hinojosa-Reyes, M.; Rodríguez-González, V.; Arriaga, S. Enhancing ethylbenzene vapors degradation in a hybrid system based on photocatalytic oxidation UV/TiO2-In and a biofiltration process. J. Hazard Mater. 2012, 209–210, 365–371. [Google Scholar] [CrossRef] [PubMed]
  110. Saucedo-Lucero, J.O.; Arriaga, S. Photocatalytic oxidation process used as a pretreatment to improve hexane vapors biofiltration. J. Chem. Technol. Biotechnol. 2015, 90, 907–914. [Google Scholar] [CrossRef]
  111. 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]
  112. Mohseni, M.; Prieto, L. Biofiltration of hydrophobic VOCs pretreated with UV photolysis and photocatalysis. IJETM 2008, 9, 47. [Google Scholar] [CrossRef]
  113. Koh, L.-H.; Kuhn, D.C.S.; Mohseni, M.; Allen, D.G. Utilizing ultraviolet photooxidation as a pre-treatment of volatile organic compounds upstream of a biological gas cleaning operation. J. Chem. Technol. Biotechnol. 2004, 79, 619–625. [Google Scholar] [CrossRef]
  114. Cheng, Z.-W.; Zhang, L.-L.; Xi, J.-Y.; Chen, J.-M.; Hu, H.-Y.; Jiang, Y.-F. Recovery of biological removal of gaseous alpha-pinene in long-term vapor-phase bioreactors by UV photodegradation. Chem. Eng. J. 2011. [Google Scholar] [CrossRef]
  115. Runye, Z.; Christian, K.; 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]
  116. Palau, J.; Penya-Roja, J.M.; Gabaldón, C.; Álvarez-Hornos, F.J.; Martínez-Soria, V. Effect of pre-treatments based on UV photocatalysis and photo-oxidation on toluene biofiltration performance. J. Chem. Technol. Biotechnol. 2012, 87, 65–72. [Google Scholar] [CrossRef]
  117. Wei, Z.; Sun, J.; Xie, Z.; Liang, M.; Chen, S. Removal of gaseous toluene by the combination of photocatalytic oxidation under complex light irradiation of UV and visible light and biological process. J. Hazard Mater. 2010, 177, 814–821. [Google Scholar] [CrossRef]
  118. Mohseni, M.; Zhao, J.L. Coupling ultraviolet photolysis and biofiltration for enhanced degradation of aromatic air pollutants. J. Chem. Technol. Biotechnol. 2006, 81, 146–151. [Google Scholar] [CrossRef]
  119. Moussavi, G.; Mohseni, M. Using UV pretreatment to enhance biofiltration of mixtures of aromatic VOCs. J. Hazard Mater. 2007, 144, 59–66. [Google Scholar] [CrossRef]
  120. 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]
  121. 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. 2016, 37, 237–244. [Google Scholar] [CrossRef]
  122. Jiang, L.; Li, H.; Chen, J.; Zhang, D.; Cao, S.; Ye, J. Combination of non-thermal plasma and biotrickling filter for chlorobenzene removal. J. Chem. Technol. Biotechnol. 2016, 91, 3079–3087. [Google Scholar] [CrossRef]
  123. Zhu, R.; Mao, Y.; Jiang, L.; Chen, J. Performance of chlorobenzene removal in a nonthermal plasma catalysis reactor and evaluation of its byproducts. Chem. Eng. J. 2015, 279, 463–471. [Google Scholar] [CrossRef]
  124. Jiang, L.; Li, S.; Cheng, Z.; Chen, J.; Nie, G. Treatment of 1,2-dichloroethane and n -hexane in a combined system of non-thermal plasma catalysis reactor coupled with a biotrickling filter. J. Chem. Technol. Biotechnol. 2018, 93, 127–137. [Google Scholar] [CrossRef]
  125. Wei, Z.S.; Li, H.Q.; He, J.C.; Ye, Q.H.; Huang, Q.R.; Luo, Y.W. Removal of dimethyl sulfide by the combination of non-thermal plasma and biological process. Bioresour. Technol. 2013, 146, 451–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Steinberg, I.; Rohde, C.; Bockreis, A.; Jager, J. Increase of the purification efficiency of biofilters by the use of a complementary ionisation step. Waste Manag. 2005, 25, 375–381. [Google Scholar] [CrossRef]
  127. Kim, H.; Han, B.; Hong, W.; Ryu, J.; Kim, Y. A New Combination System Using Biotrickling Filtration and Nonthermal Plasma for the Treatment of Volatile Organic Compounds. Environ. Eng. Sci. 2009, 26, 1289–1297. [Google Scholar] [CrossRef]
  128. Jiang, L.; Zhu, R.; Mao, Y.; Chen, J.; Zhang, L. Conversion characteristics and production evaluation of styrene/o-xylene mixtures removed by DBD pretreatment. Int. J. Environ. Res. Public Health 2015, 12, 1334–1350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Karatum, O.; Deshusses, M.A. A comparative study of dilute VOCs treatment in a non-thermal plasma reactor. Chem. Eng. J. 2016, 294, 308–315. [Google Scholar] [CrossRef]
  130. Schiavon, M.; Scapinello, M.; Tosi, P.; Ragazzi, M.; Torretta, V.; Rada, E.C. Potential of non-thermal plasmas for helping the biodegradation of volatile organic compounds (VOCs) released by waste management plants. J. Clean. Prod. 2015, 104, 211–219. [Google Scholar] [CrossRef]
  131. Mäule, J.; Reiser, M.; Kranert, M. Neue Techniken zur Abgasreinigung in der Abfallwirtschaft durch nichtthermische Plasmatechnologie: Verfahrenskombination aus Biofilter und NTP-Technologie. Abschlussbericht zum DBU-Forschungsvorhaben 17988; Universität Stuttgart: Stuttgart, Germany, 2005. [Google Scholar]
  132. Hołub, M.; Brandenburg, R.; Grosch, H.; Weinmann, S.; Hansel, B. Plasma Supported Odour Removal from Waste Air in Water Treatment Plants: An Industrial Case Study. Aerosol Air Qual. Res. 2014, 14, 697–707. [Google Scholar] [CrossRef] [Green Version]
  133. Quan, Y.; Wu, H.; Guo, C.; Han, Y.; Yin, C. Enhancement of TCE removal by a static magnetic field in a fungal biotrickling filter. Bioresour. Technol. 2018, 259, 365–372. [Google Scholar] [CrossRef]
  134. Quan, Y.; Wu, H.; Yin, Z.; Fang, Y.; Yin, C. Effect of static magnetic field on trichloroethylene removal in a biotrickling filter. Bioresour. Technol. 2017, 239, 7–16. [Google Scholar] [CrossRef]
  135. Brandenburg, R.; Kovačević, V.V.; Schmidt, M.; Basner, R.; Kettlitz, M.; Sretenović, G.B.; Obradović, B.M.; Kuraica, M.M.; Weltmann, K.D. Plasma-Based Pollutant Degradation in Gas Streams: Status, Examples and Outlook. Contrib. Plasma Phys. 2014, 54, 202–214. [Google Scholar] [CrossRef]
  136. Oda, T.; Yamaji, K.; Takahashi, T. Decomposition of Dilute Trichloroethylene by Nonthermal Plasma Processing—Gas Flow Rate, Catalyst, and Ozone Effect. IEEE Trans. Ind. Appl. 2004, 40, 430–436. [Google Scholar] [CrossRef]
  137. 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]
  138. Xi, J.; Saingam, P.; Gu, F.; Hu, H.-Y.; Zhao, X. Effect of continuous ozone injection on performance and biomass accumulation of biofilters treating gaseous toluene. Appl. Microbiol. Biotechnol. 2014, 98, 9437–9446. [Google Scholar] [CrossRef]
  139. Malhautier, L.; Soupramanien, A.; Bayle, S.; Rocher, J.; Fanlo, J.-L. Potentialities of coupling biological processes (biotrickler/biofilter) for the degradation of a mixture of sulphur compounds. Appl. Microbiol. Biotechnol. 2015, 99, 89–96. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Clogging of the package material by enhanced migration of the biomass out of the swamp phase onto the packing material. Source: self-developed.
Figure 1. Clogging of the package material by enhanced migration of the biomass out of the swamp phase onto the packing material. Source: self-developed.
Sustainability 12 08577 g001
Table 1. Suitability of waste air treatment processes based on nearly 70 in-house industrial studies. +: high suitability, +/−: moderate suitability, −: no/low suitability.
Table 1. Suitability of waste air treatment processes based on nearly 70 in-house industrial studies. +: high suitability, +/−: moderate suitability, −: no/low suitability.
ParameterPhy. ScrubberChem. ScrubberIncinerationNTPUV-OxidationAdsorptionBioscrubberBiotrickling FilterBiofilter
Aliphatics++/−+/−+++
Alcohols++++/−+/−++++/−
Aldehydes++++++/−+++
Ketones++++/−+/−++++
Organic acids+++++++++/−
Esters+++++++++/−
Ethers++/−+/−+++
Aromatics+/−++/−+/−+++
Sulphurorganics++/−+/−+/−+++/−
Haloorganics+/−+/−+/−+++/−
CH4++/−
NH3+++/−+++++
N2O+/−+/−
H2S+/−++/−+/−+/−+/−+++/−
Odor+/−++++++/−+/−+/−
Germs+++++++
Dust++++/−
Humidity+++/−+++
Table 2. Compilation of investment and operational costs of waste air treatment processes. Calculations of this study base on electrical energy costs of 0.15 €·kWh−1, natural gas costs of 0.416 €·m−3, and a minimum size of 10,000 m³∙h−1.
Table 2. Compilation of investment and operational costs of waste air treatment processes. Calculations of this study base on electrical energy costs of 0.15 €·kWh−1, natural gas costs of 0.416 €·m−3, and a minimum size of 10,000 m³∙h−1.
ProcessInvestment Costs (€·(m³/h)−1)Operational Costs (€·1000 m−3)Reference
Biofilter1.53–5.110.44–0.48[33]
0.90–1.500.70–1.00[34]
4.10–12.800.25–0.77[35]
2.80–4.000.25[36]
2.500.23[37]
-0.24[38]
Biotrickling filter13.00–15.000.71[39]
2.81–8.130.04–1.27 [40]
150.13–0.45[38]
-0.12[41]
2.500.20[37]
1.41–18.780.14–0.31This study
Adsorption2.50–5.003.00–3.50[42]
Chemical scrubber 0.70–1.90 (odor)
18.85–23.20 (VOC)
0.40 (odor)
0.50 (VOC)
[34]
4.10–25.600.50–2.55[35]
Thermal oxidation13.00–15.002.27[39]
12.8–25.61.50–7.67[35]
34.831.62This study
Non-thermal plasma5.0–8.00.30–0.84[42]
1.20–1.40 (odor)
10.00–12.00 (VOC)
0.30 (odor)
1.00 (VOC)
[34]
8.52–30.680.19–0.68[43]
UV13.652.62[44]
Table 3. Cost comparison between biotrickling filter (BTF) and a regenerative combustion (RPC) for treatment of an industrial styrene-containing air of 13,000 m³·h−1 and of a biotrickling filter for treatment of pig fattening waste gases of 91,000 m³·h−1, based on actual quotes. Calculations use electrical energy costs of 0.15 €·kWh−1 and natural gas costs of 0.416 €·m−3.
Table 3. Cost comparison between biotrickling filter (BTF) and a regenerative combustion (RPC) for treatment of an industrial styrene-containing air of 13,000 m³·h−1 and of a biotrickling filter for treatment of pig fattening waste gases of 91,000 m³·h−1, based on actual quotes. Calculations use electrical energy costs of 0.15 €·kWh−1 and natural gas costs of 0.416 €·m−3.
BTF ResinRPC ResinBTF Pig Fattening
RPC (p = 262 kW, ∆T = 35 K)-250,000 €-
BTF151,000 €-49,000 €
Bypass during production shutdown-30,000 €-
Fresh air damper-2400 €-
Pre-filter-11,700 €-
Burn-out function-9500 €-
Dryer for compressed air-1900 €-
Compressor (at 6 bar)3500 € (1.3 L·s−1)5400 € (2 L·s−1)-
Ventilator4500 € (3500 Pa)21,500 € (for 100 °C, 5500 Pa)18,600 € (2100 Pa)
Chimney-20,500 €-
Assembly and commissioning 18,000 €35,000 €10,500 €
Transport and crane5000 €9500 €4000 €
Foundation30,000 €30,000 €20,000 €
Waste air pipes (25 m)2500 €2500 €3500 €
Natural gas connection -15,000 €-
Water connection (25 m, of which 10 m ice-free-proof)1750 €-3000 €
Waste water connection (10 m, ice-free-proof) 1200 €-2000 €
pH dosage5000 €-5000 €
Circulation pump800 €-500 €
Nutrient dosage1500 €--
Level control2500 €-2500 €
Control cabinet7000 €8000 €9500 €
Isolation and ice-free installations10,000 €--
Total investment244,250 €452,900 €128,100 €
Electricity demand ventilator130,000 kWh389,000 kWh648,000 kWh
Electricity demand pump11,200 kWh-9300 kWh
Electricity demand compressor3800 kWh5600 kWh-
Electricity demand pipe heating (10 W∙m−1)43 kWh-43 kWh
Peripheral electricity demand 550 kWh-300 kWh
Annual electricity demand (kWh)145,593 kWh394,600 kWh657,643 kWh
Annual electricity demand (€)21,840 €59,190 €98,650 €
Natural gas-94,200 €-
Chemicals1000 €-550 €
Parts of annual replacement1000 €8500 €250 €
Annual maintenance 1500 €5000 €700 €
Staff costs (on-site)6500 € 6500 €4000 €
Interest costs (2.5%)6200 € 11,500 €3200 €
Annual operational costs35,540 €184,890 €107,350 €
Specific investment costs (€·m−3 hourly air capacity)18.7834.831.41
Specific operational costs (€·1000 m−3)0.311.620.14
Table 4. Balancing of emissions of CO2, CH4, and N2O as greenhouse gases for combustion and non-thermal plasma (NTP) processes for treatment of a gas flow of 1000 m³∙h−1 for 8000 h∙a−1. Data based on CO2 emissions of the German energy mix in 2019 (0.401 kg CO2∙kWh−1). * analytical data of the authors; ** experience values for combustion. *** calculated value based on a concentration of 10 mg∙L−1 of ammonia as fertilizer in the swamp phase at pH 7.8 and ammonium/ammonia distribution of Kollbach [52].
Table 4. Balancing of emissions of CO2, CH4, and N2O as greenhouse gases for combustion and non-thermal plasma (NTP) processes for treatment of a gas flow of 1000 m³∙h−1 for 8000 h∙a−1. Data based on CO2 emissions of the German energy mix in 2019 (0.401 kg CO2∙kWh−1). * analytical data of the authors; ** experience values for combustion. *** calculated value based on a concentration of 10 mg∙L−1 of ammonia as fertilizer in the swamp phase at pH 7.8 and ammonium/ammonia distribution of Kollbach [52].
ParameterIncinerationNon-Thermal PlasmaBTF
QuantityIn CO2,eqQuantityIn CO2,eqQuantityIn CO2,eq
Natural gas12,729.5 m³∙a−125,003 kg∙a−10 m³∙a−10 kg∙a−10 m³∙a−10 kg∙a−1
Methane emissions24.0 kg∙a−1
(3 mg C∙m−3 **)
504 kg∙a−13.31 kg∙a−1
(0.58 ppm *)
69.6 kg∙a−10 kg∙a−10 kg∙a−1
Energy demand fan17.6 MWh∙a−1 (∆p = 5000 Pa) 7040 kg∙a−18.8 MWh∙a−1 (∆p = 2250 Pa) 3520 kg∙a−18.8 MWh∙a−1 (∆p = 2250 Pa) 3520 kg∙a−1
Energy demand compressor4 MWh∙a−1 (P = 0.5 kW)1600 kg∙a−10 kWh0 kg∙a−10 kWh0 kg∙a−1
Energy demand NTP0 kWh0 kg∙a−132 MWh∙a−1
(4 kWh∙1000 m−3)
12,800 kg∙a−10 MWh∙a−10 kg∙a−1
N2O formation78.5 kg∙a−1
(5 ppm **)
24,330 kg∙a−18.8 kg∙a−1
(0.57 ppm *)
2728 kg∙a−10.15–1.35 kg∙a−1
(0.32 ppm ***)
45–419 kg∙a−1
Sum 58,477 kg∙a−1 19,118 kg∙a−1 3565–3939 kg∙a−1
Reduction potential0 kg∙a−139,359 kg∙a−1 or 67.3%54,538–54,912 kg∙a−1 or 93.3–93.9%
Table 5. Industrial fields of application and shares of biological waste air treatment in Austria [57,58].
Table 5. Industrial fields of application and shares of biological waste air treatment in Austria [57,58].
SectorShare (%)
Municipal wastewater treatment48.0
Industrial wastewater treatment14.0
Food industry10.6–14.1
Composting 6.8
Chemical industry 3.4–8.4
Commercial4.6–6.5
Landfills4.6
Rendering plants4.6
Plastics industry1.9–3.4
Pulp and paper industry0.0–2.7
Electronics industry0.0–1.1
Ceramics industry0.0–1.1
Petroleum industry0.0–0.8
Wood processing industry0.0–0.8
Leather processing industry0.0–0.8
Table 6. Recent examples of volatile organic compound (VOC) removal combining static magnetic field and biofiltration.
Table 6. Recent examples of volatile organic compound (VOC) removal combining static magnetic field and biofiltration.
Efficiency (%)
CompoundConcentrationηMagnetic fieldηBioηCombinationMagnetic Flux Density (mT)Gas flow (m³∙h−1)EBRT (s)CombinationLiterature
TCE53.6–326.8 mg∙m−3-29.8–89.845.5–92.2600.048202.5MF + BTF[134]
TCE1000 mg∙m−3-39.452.9200.024405MF + BTF[133]
Table 7. Recent examples of VOC removal combining UV oxidation and biofiltration.
Table 7. Recent examples of VOC removal combining UV oxidation and biofiltration.
Efficiency (%)
CompoundConcentrationηUVηBioηCombinationSpecific Input Energy
(SIE, kWh∙1000 m−3)
Gas Flow (m³∙h−1)EBRT UV (s)EBRT Bio (s)CombinationLiterature
Chlorobenzene250–1500 mg∙m−3-5.4–81.831.6–10045–2300.1–0.328–8341–122UV + BF[106]
Chlorobenzene250 mg∙m−379–9030–4986–10076.70.327.640.8UV + BF[107]
Dichloromethane420 mg∙m−3-54.681.844.40.811060 (2)/70 (3)
60 (2)/70 (3)
36 (2)/46 (3)
UV + BTF[108]
650 mg∙m−3-36.385.4
410 mg∙m−3-463066.0
Ethylbenzene3000 mg∙m−336478322400.00795.360UV + BF[109]
n-Hexane33.3 mg∙m−321.550–5565–70-0.0136034.2UV + BF[110]
PCE32 ppm96.872.898.544.40.094120UV + BTF[111]
α-Pinene130 ppm20–5020–50100-0.3–0.39--UV + BF[112]
α-Pinene45 ppm73.3–94.3--9.7 W∙cm2−20.064–0.1511.5–3.5- (1)UV + BF[113]
95 ppm49.5–86.7--
145 ppm36.2–66.7
α-Pinene600–1500 mg∙m−3-25–7657–9818–400.9–2.01817/23/39UV + BTF[114]
α-Pinene1500–1600 mg∙m−315
15
70
71
80
84
800.451827/40/81UV + BTF
UV+BF
[114]
Styrene2.11–2.525 g∙m−315–1633–3457–5944.40.811030UV + BTF[115]
Styrene2–4 g∙m−3-27.2–61.864.8–87.344.40.811020/30/60 (2)
30/40/70 (3)
UV + BF vs. BF[115]
Styrene107 mg∙m−389393–981.85671031UV + BTF[39]
TCE307.5 ppm97.55094–10044.40.094120UV + BTF[111]
Toluene650 mg C∙m−35951–699231.61.142.7 45UV + BF[116]
1200 mg C∙m−37474–7984
Toluene210–500 mg∙m−353–58.292.394.1–96.7400.113.3 108UV + BF[117]
o-Xylene60–220 mg∙m−310–3550–6073–9941.7–86.00.186–0.3840.5/1.235/75UV + BF[118]
Toluene,
o-Xylene
44.7–121.4 ppm
44.7–121.4 ppm
29.8–132.7 ppm
16–36
10–14
9–12
42–77
22–27
32–34
90–100
92–99
99–100
0.38–0.560.5–1.545
30
30
UV + BF[119]
Ethyl acetate,2.4–24.894.576.599.00.32300010.87.2BTF + UV[120]
Toluene,5.4–63.975.088.598.0
Ethylbenzene38.4–150.781.672.894.7
Xylene21.3–23393.275.098.3
Ethyltoluene0–61.991.064.197.6
Trimethylbenzene0.6–32.588.871.497.6
VOC total46.8–485.986.874.196.6
Toluene,-72.7->99----UV + BTF[121]
Ethylbenzene,-77.2->99
m-Xylene,-71.9->99
o-Xylene,-74.8->99
p-Xylene-60.0->99
-: no data; (1): installation of biofiltration was suggested but not implemented; (2): BF or BTF combined with UV; (3): BF or BTF solely installed.
Table 8. Recent examples of VOC removal combining NTP and biofiltration.
Table 8. Recent examples of VOC removal combining NTP and biofiltration.
Efficiency (%)
CompoundConcentrationηNTPηBioηCombinationSpecific Input Energy
(SIE, kWh∙1000 m−3)
Gas Flow (m³∙h−1)EBRT NTP (s)EBRT Bio (s)CombinationLiterature
Chlorobenzene500 mg∙m−3
500 mg∙m−3
1000 mg∙m−3
1000 mg∙m−3
5868.797.69720.154.3260
60
60
30
NTP + BTF[122]
88.168.798.83194
-57.894.53194
4367.23194
Chlorobenzene750 mg∙m−373–100--972–32000.011–0.0335–15- (1)NTP + BTF[123]
1,2-DCE200 mg∙m−3
800 mg∙m−3
-
-
95
39
95–100
85
0.970.0533.24 90NTP + BTF[124]
Dimethyl disulfide76.216.1–24.55.8–15.987.7–96.3-0.110.610.3NTP + BTF[125]
Limonene0.5 ppm100---20--NTP + BF[126]
30 ppm30--
1,2-DCE,
n-Hexane
200 mg∙m−3
200 mg∙m−3
-
-
70
60
92–95
82–85
0.970.0533.2490NTP + BTF[124]
Toluene,116.5 ppm-86–9495.2-1.2--NTP + BTF[127]
p-Xylene,132.4 ppm-66–7095.8
Ethylene127.5 ppm 50–5481.3
Toluene,
n-Heptane,
95.6
49.4
-
-
60
23
>93
81–91
25.60.15660NTP + BF[105]
p-Xylene,60.8-87>93
Ethylbenzene,47.3-76>93
Benzene36.6-28>93
Toluene,95.6-6010035.60.15660NTP + BF[105]
n-Heptane,49.4-2389
p-Xylene,60.8-87100
Ethylbenzene,47.3-76100
Benzene36.6-28100
Styrene,
o-Xylene (1:1 ratio)
500
1000
2000
3000
3000
82.4
74.3
49.5
35.2
77.3
-
-
-
-
-
-
-
-
-
-
111.10.045
0.045
0.045
0.045
0.011
3.6
3.6
3.6
3.6
15
- (1)NTP + BTF[128]
MEK
Benzene
Toluene
3-Pentanone
MTBE
Ethylbenzene
n-Hexane
95–100 ppm50--97.20.3960.016- (2)NTP + BTF[129]
58--
74--
76--
80--
81--
90--
Terpenes
Chlorobenzene
Trichlormethane
Phenol
Odor
--->90-1000<1 BF + NTP[43]
Ethanol,233–60415–35--25.63∙10−40.98- (1)NTP + BF[130]
Ethyl acetate,216–7787–8--
Benzene,6–724–6--
Toluene,53–34414–20--
Octane65–35222–32--
Ethanol,233–60483–88--1253∙10−40.98- (1)NTP + BF[130]
Ethyl acetate,216–77849–90--
Benzene,6–7235–64--
Toluene,53–34468–91--
Octane65–35280–94
Ethanol,233–60484–95--2503∙10−40.98- (1)NTP + BF[130]
Ethyl acetate,216–77882–100--
Benzene,6–7265–92--
Toluene,53–34496–100--
Octane65–35297–100
n-Butanol,
2-Butoxyethanol,
Odor
4 mg C∙m−3
1 mg C∙m−3
700–3600 OU∙m−3
50–60
100
42.8–91.7
-
-
-
93–100
100
94.3–97.2
2.27500.310.6NTP + BTF[45]
Butanone,54 mg 40.71001002.7360-9BF + NTP[131]
Ethyl acetate,
Dimethyl disulfide,
Toluene,
Limonene
C∙m−3
15 mg C∙m−3
1 mg C∙m−3
12 mg C∙m−3
59 mg C∙m−3
53.3
100
41.7
89.8
100
100
71.4
33.3
100
100
83.3
93.2
Dimethyl disulfide,
Carbonyl sulfide,
Dimethyl sulfide,
Dimethyl trisulfide,
Odor
1.4 mg∙m−3
0.04 mg∙m−3
<0.2 mg∙m−3
<0.2 mg∙m−3
86,900–97,900 OU∙m−3
57–725280-16012-BF + NTP[132]
VOC mixture
Odor
80–160 mg∙m−3
9200–14,200 OU∙m−3
65,000 OU∙m−3
25
25
40
25
25
-
50
50
-
0.7
0.7
0.04
250
250
14,000
4.5
4.5
1.5
-
-
-
NTP + BF[34]
-: no data; (1): installation of biofiltration was suggested but not implemented; (2): BTF not installed due to high performance of NTP.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dobslaw, D.; Ortlinghaus, O. Biological Waste Air and Waste Gas Treatment: Overview, Challenges, Operational Efficiency, and Current Trends. Sustainability 2020, 12, 8577. https://doi.org/10.3390/su12208577

AMA Style

Dobslaw D, Ortlinghaus O. Biological Waste Air and Waste Gas Treatment: Overview, Challenges, Operational Efficiency, and Current Trends. Sustainability. 2020; 12(20):8577. https://doi.org/10.3390/su12208577

Chicago/Turabian Style

Dobslaw, Daniel, and Oliver Ortlinghaus. 2020. "Biological Waste Air and Waste Gas Treatment: Overview, Challenges, Operational Efficiency, and Current Trends" Sustainability 12, no. 20: 8577. https://doi.org/10.3390/su12208577

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