You are on page 1of 41

EVALUATION OF MBT BIOFILTERS DESIGN CRITERIA AND EFFICIENCY IN AIR QUALITY MANAGEMENT.

Abstract With increasingly stringent environmental regulations governing plant emissions of volatile organic compounds (VOCs), there is a growing need for robust procedures for the design of optimal VOCs recovery processes or systems that rely on growth of a biofilm layer on an inert organic support such as compost or peat (biofilters). One of the more cost-effective and environmentally efficientknow-hows incorporated in VOCs recovery procedures is biofiltration.Due to the efficient elimination of many of the drawbacks of conventional physical and chemical techniques, bio-oxidation techniques have been applied more frequently to control air-borne pollutant emissions recently in the UK and Europe at large. The quest for a renewable energy-from-waste technology to efficiently manage waste resources has prompted the need for the development of mechanical biological treatment (MBT) and anaerobic digestionfacilities which leavethe environment in a perilous state at the annals of volatile organic contaminants emission. As a result, for the reason of co-existence of people and industry, it is required at local and regional echelons that these contaminants be reduced to threshold levels. To that end, proliferation of rigid legislation to control the emissions of air pollutants has also been observed as a driving force in the search for a more economical and environmentally friendly solution for the waste gas emissions. Studies show that the use of biofilters in VOCs oxidation is not out of place. The wide range employment of biofilters in the treatment of biological waste gas streams has been recognized as a panacea to the management of the air quality at the vicinity of AD and MBT plants. However, though biofiltration is firmly established technology for the control of emissions of volatile organic compounds VOCs,the design criteria, performance and efficiency of MBT biofilters in the control of these gas emissions is still not clearly understood by many. In particular, data on contaminants and their removal require development to inform operation and regulation. This work uses a case study methodology describing the dynamic physical and biological processes occurring in a packed peat biofilter at Global Renewables Limited
1

and wood chips biofilter at TEG Environmental Limited to evaluate the relationship between the design criteria, performance and the efficiency of contaminants removal. It compares the relationship directly with conventional thermal oxidation process at Orchid Environmental Limited to opt for an optimum design criterion efficient for VOCs removal.

Introduction Recently, the realization of the energy-from-waste technology which has called for the establishment of anaerobic digestion (AD) and mechanical biological treatment (MBT) facilities for mixed waste streams.This appears to be changing the landscape of waste treatment options in the U.K. as many local authorities are seeking environmentally compatible methods for the diversion of residual municipal solid waste (MSW) from landfill. In addition, the development of stringent European Union (EU) landfill Directive has prompted the need for the establishment of more energy-from-waste facilities as auniversal remedy to the phenomenon of greenhouse gas emission from landfill sites and the resulting effect of global warming. However, the option of AD and MBT technology has its attendant consequences. It is that the gas emitted from the plants is believed to be composed of hazardous and volatile organic contaminants which pose numerous problems to communities at the vicinity of these plants if operations are not coordinated. To this end, the nuisance of the offending odour (Smet and Van Langenhove, 1998 p273) has become one of the biggest challenges to the management of these plants as the Environment Agency in the UK as well as the statutory instruments strictly limit and require that the odour be removed to forestall any environmental and health concerns. The scope of this research covers the control of VOCs emission from AD processes by the method of biotechnology. The environment Act 1995(b) is a legislative instrument in the UK which restricts the concentration of air pollutants or volatile organic contaminants above a particular threshold limit. This serves as a statutory duty on the treatment of waste gas category where the possibility of these waste gas emissions would be high. These pollutants are subject to severe environmental constrains because of their negative impact on health (Moretti and Mukhopadhyay 1983). On this note, biofilters are seen to be the retrofits of
2

the existing chemical and physical techniques to completely rid the gas of the pollutants (Devinny et al., 1999). In response to this phenomenon, Leson and Winer in 1991 reported that, the removal of volatile organic compounds (VOCs) from contaminated airstreams has become a major air pollution concern. In principle, the development of this concept of pollutants removal culminated in what is scientifically termed as biofiltration (Zarook et al., 1997). Malhautier et al (2005, pp16-17) in a reviewed paper stated that the mechanism of biofiltration revolves around two principal phenomena: the transfer of contaminants from the gas phase to the support medium and then the bioconversion of pollutants to biomass, carbon dioxide and water. The history of biofiltration dates back as far as 1950s and has been used in the North America, Europe, the Netherlands, Japan and New Zealand for treating high volume of airstreams containing low concentrations of a range of pollutants (Devinny et al., 1999). Biodegradation of pollutants is effectual in the biofilm by microorganisms known as bacteria. Malhautier et al (2005, p17) discussed that the biofilm contains different microorganisms supported by a packing material. Quigley et al (2004,p320) discussed that biofiltration is an air pollution control technology that uses microorganisms to oxidize volatile organic compounds (VOCs) and oxidizable inorganic gases (Ammonia gas, Nitrogen dioxide and Sulphur dioxide) into carbon dioxide, water and biomass. Malhautier et al (2003, p145) added that the removal of the pollutants is achieved by biofilters. The biofilter is the resting ground for the oxidizing bacteria which oxidize a variety of inorganic and organic compounds into innocuous gases (Janni et al, 1998). Moreover, in order to better control and optimise biofiltration for a full scale implementation (Bonnin et al 1990) cited in Malhautie r et al (2003 p146), it is pertinent to research into the design criteria and performance of the biofilters composed of different packing material or biomass in order to evaluate the option which effectively removes the contaminants to the lowest concentration. The term biofilter is a reactor for the conversion of contaminated air into harmless products of carbon dioxide, water and mineral salts. In the reactor, the contaminated air is passed through a bed of porous moist medium, and the contaminants are adsorbed to the medium surface where they are degraded by the microorganisms in the medium (Shareefdeen and Singh, 2005). Due to end-of-life cycle impact of products or services on the environment, this work will also delve into the life cycle impact assessment of biofilters to examine the

environmental impact when they lose their usefulness. As per the efficiency of contaminants removal, the paper compares three biofilters of different media in a case study from TEG environmental operations limited, Orchid environmental limited and Global renewables operations, Lancashire. All located in the North West province of England, UK. The use of biofilters in the past decades in industrial applications has seen significant progress compared to traditional chemical and physical gas treatment technologies, due to their efficiency and competitive cost (Prado et al., 2009, p2515). However, in most cases, full-scale biofilters design is still based on pilot-scale performed either on-site with the actual air stream or in laboratories with a synthetic air stream (Deshusses and Johnson 2000, p461). The reason being that data on the elimination of contaminants are either not available or were obtained under different conditions and are thus not authentic for full-scale biofilter design. Following from this, though a large body of experimental data has been available, the basic knowledge necessary for biofilter design based on hypothetical concepts or mathematical models is still implicit and missing. Hence, there is the need for the development of standard biofilter design criteria and fundamental data for pollutant removal (Zhu et al., 2004). This will help designing biofilters and understanding comprehensively the factors that influence the elimination of pollutants in biofilters. This research work was intended to address the paradigm shift in the philosophy of the design, performance and control of the contaminants removal and the end-of-life impact of the biofilter technology. The objectives of this research are: y to determine a suitable design criteria and media for VOCs removal. What are the operating parameters? y to compare and obtain an authentic data for the VOCs removal efficiency of the biofilters under study and y to compare the environmental impacts of the biofilters base on their life cycles assessment with the conventional pollutant removal by thermal oxidation technique employed by Orchid Environmental Limited. In all, the research hypothesis is that there is an optimum design and media for an economic and environmentally acceptable biofilter.
4

Literature Review Air pollution The evolution of industrial advancement has of late seen the ecosystem of man being hinged at a very perilous state due the exploitation of the resources of his own niche (the earth)resulting in what is being described as environmental nuisance. To this end, man by his own activities sometimes becomes concerned about the medical conditions of the air purity infiltrated by emissions from the industry that put environmental air into an unacceptable respiratory state. An inference from this indicates that the air is being pollutedby odorous gases of volatile organic compounds (VOCs) or inorganic compounds which subsequently require remediation to forestall public health concerns. On this note, de Nevers (1995) defined air pollution as the presence of undesirable material in air, in quantities large enough to produce harmful effects. The consequence of these undesirable materials pose a lot of problems among other things that include human health,change of global environment conditions as well as aesthetic offenses in the form of brown or hazy air or odorous smells (de Nevers, 1995). Methods of Odor and VOC Control As per the scope of this research which searches the best biofilter design to control waste gas odor, the contaminants under consideration are the volatile organic compounds group. Volatile organic compounds (VOCs) are liquids or solids that contain organic carbon (carbon bonded to carbon, hydrogen ,nitrogen, or sulphur, but not carbonate carbon as in CaCO 3 nor carbide carbon as in CaC2 or CO or CO 2), which vaporize at significant rates (de Nevers, 1995). The treatment of off-gases has been practiced over the past decades and is primarily based on non-biological methods such as condensation, activated carbon adsorption, absorption/scrubbing and incineration (Asadi et al., 2009). In the condensation process, cooling and compression condense contaminant vapours from air. This process is economical for higher boiling point compounds and more
5

concentrated airstreams. In the adsorption process, pollutants are adsorbed onto adsorbents such as activated carbon. This process is effective when the concentration in the airstreams is low. Regeneration of the adsorbents is done using steam or hot air. However, recovery of compounds is costly and spent adsorbents such as solid waste need to be landfilled or incinerated which leads to not any other problem other than environmental concern. In absorption or scrubbing, pollutants from the air are absorbed into a scrubbing solution such as water or solvents. Often, chemical costs are high and liquid waste needs further treatment. Nozzle maintenance, complex chemical feed and control systems and high operating costs are a few problems associated with this method (Devinny et al., 1999). The incineration technology has been widely used due to its high efficiency. To increase the efficiency of incineration and reduce fuel requirement for combustion, several forms of this technology, such as recuperative, regenerative and catalytic oxidation, are practiced. Regardless, incineration is an expensive method due to high energy requirements and it is not economical if concentration levels are low and large airflow volumes need treatment. This process also produces the highest amount of greenhouse gases as terms of carbon dioxide or nitrous gases emission. Biological methods are effective and economical for biodegradable odorants and VOC contaminants. Air-phase bioreactors used in industries for odor and VOC removal include biofilters, biotrickling filters and biosrubbers. In general, highly soluble and low molecular weight VOCs such as methanol, ethanol, aldehydes, acetates, ketones and some aromatic hydrocarbons and inorganic compounds such as hydrogen sulphide and ammonia are easily biodegradable in these bioreactors. Low molecular weight aliphatic hydrocarbons such as methane, pentane and some chlorinated compounds are difficult to biodegrade (Devinny et al., 1999). Unique bioreactors for VOC and odor control include rotating drum biofilters, horizontal flow biofilters, foamed emulsion bioreactors, short contact time biotrickling filters (Gabriel et al., 2002), higher plantbased biofilters (Guilbault, 2002) and microwave concentrator/biofilter integrated systems (Webster et al., 2002).

Biological treatment Biological treatment is a cost-effective and economical method for the removal of low concentrations of contaminants in large quatities (Ardjamand et al. 2005). The contaminants are sequestered for the waste gas to an aqueous phase where microbial reaction occurs. Through oxidative and occasionally reduction reactions, the contaminants are converted to carbon dioxide, water vapour, and organic biomass. These air pollutants may be either organic or inorganic vapors and are used as energy and sometimes as a carbon source for maintenance and growth by the microorganism populations. In general, the microbes used for biological treatment are organisms that are naturally occurring. These microbial populations may be dominated by one particular microbial species or also engaged in many of the same ecological relationships such as predation or parasitism. Such relationship is a precondition for balance within the system. In addition, particular contaminants of interest must be biodegradable and non-toxic for biological air treatment to be successful (Devinny et al., 1999). It is interesting to note that, the most successful removal in gas-phase bioreactors occurs for low molecular weight and highly soluble organic compounds with simple bond structures (Iranpour et al., 2005). Compounds with complex bond structures generally require more energy to be degraded, and this energy is not always available to the microbes. Hence, little or no biodegradation of these types of compounds occurs. Instead, microorganisms degrade those compounds that are readily available and easier to degrade. Organic compounds such as alcohols, aldehydes, ketones and some simple aromatics demonstrate complete biodegradability. Some compounds that show moderate to slow degradation include phenols, chlorinated hydrocarbons, polyaromatic hydrocarbons and highly halogenated hydrocarbons. Inorganic compounds such as hydrogen sulphide and ammonia are also biodegraded well. Interestingly, Devinny et al. (1999) noted certain anthropogenic compounds may not biodegrade at all because microorganisms do not possess the requisite enzymes to break the bond structure of the compound effectively (see Table 3). Compounds that are biologically treatable can come from a wide array of sources.

Table 3 Biodegradability of various contaminants in a biofilter Contaminants Aliphatic hydrocarbons Methane Propane Butane Pentane Isopentane Hexane Cyclohexane Acetylene Aromatic hydrocarbons Benzene Phenol Toluene Xylene Styrene Ethylbenzene chlorinated b hydrocarbons Carbon tetrachloride Chloroform Dichloromethane Bromodichloromethane 1,1,1-Trichloroethane 1,1-Dichloroethane Tetrachloroethane Trichloroethane 1,2-Dichloroethane 1,1-Dichloroethane Vinyl chloride 1,2-Dichlorobenzene Chlorotoluene 1 1 3 ? ? ? 1a 1a ? ? 1 ? 1 2 3 3 2 2 3 1 ? ? 1 1 2 1 1 Biodegradability

Nitrogen-containing carbon compounds Amines Aniline Nitrile Acrylonitrile Pyridine Sulphur-containing compounds Carbon disulphide Dimethyl sulphide Dimethyl disulfide Methyl mercaptan Thiocyanates 2 2 2 1 1
b

3 3 1 ? 1 carbon

Oxygenated carbon compounds alcohols Methanol Ethanol Butanol 2-Butanol 1-Propanol Aldehydes Formaldehyde Acetaldehyde Carbonic acids (esters) Butyric acid Vinyl acetate Ethyl acetate Butyl acetate Isobutyl acetate Ethers Diethyl ether 3 3 3 3 2 3 3 3 1 1 3 3 3 3 3 3

Dioxane Methyl tert-butyl ether Tetrahydrofuran Ketones Acetone Methyl ethyl ketone Methyl isobutyl ketone Inorganic b compounds Ammonia Hydrogen sulphide Nitrogen oxide
a

1 1 3

3 3 3

3 3 1

Indicates that cometabolism or anaerobic treatment has been identified within a

biofilter.
b

Indicates that a change in filter bed may occur with treatment of these compounds.

This change may negatively affect performance. Note: 1 = some biodegradability; 2 = moderate biodegradability; 3 = good biodegradability; ?= unknown. Source: Biofiltration for air pollution control (Devinny et al., 1999).

Biofilters Biofilters are reactors in which a humid polluted air stream is passed through a porous packed bed on which a mixed culture of pollutant-degrading organisms is naturally immobilized (Deshusses, 1997). Outlined in figure 1 below, the removal of the gaseous pollutant in a biofilter is the result of a complex combination of different physicochemical and biological phenomena. The microorganisms grow in the biofilm on the surface of a medium or are suspended in the water phase surrounding the medium particles (figure 2). The filter-bed medium consists of comparatively inert substances such as compost or peat which ensure large surface attachment areas and additional nutrients supply (Quigley et al., 2004). As the air passes through the bed, the contaminants in the air phase sorb into the
10

Figure 2

Internal mechanism of a biofilter. Contaminant air (CG) passes through the

filter bed medium (compost, soil, peat, etc.) with oxygen (O2) and sorbs into a microbial biofilm/liquid phase attached to the filter medium. Microbes convert the contaminants to carbon dioxide (CO2) and water (Devinny et al., 1999) cited in Ardjmand et al. (2005).

biofilm and onto the filter medium, where they are biodegraded (figure 3). Biofilters are not filtration units as strictly defined (Devinny et al., 1999). Instead, they are systems that use a combination of basic processes: absorption, degradation, adsorption, and desorption of gas phase contaminants. Biofilters usually incorporate water irrigation (Zhu et al., 2004), to control moisture content and add nutrients. Basically, the gas stream is humidified before entering the biofilter reactor (Ardjmand et al., 2005). However, if humidification proves inadequate then,direct irrigation of the bed may be needed (Lu et al., 2002).

11

To that end, the overall effectiveness of a biofilter is largely governed by the properties and characteristics of the support medium, which include porosity, degree of compaction, water retention capabilities and the ability to host microbial populations (Quigley et al., 2004). Critical biofilter operational and performance parameters include the microbial inoculum, medium pH, temperature, and the medium moisture and nutrients as in Table 2 below. Table 2 Typical Biofilter Operating Conditions for waste Air Treatment Parameter Biofilter layer height Biofilter Area Waste air flow Biofilter surface loading Biofilter volumetric loading Bed void volume Mean effective gas residence time Pressure drop per meter of bed height Inlet pollutant and /or odor concentration Operating temperature Inlet air relative humidity Water content of the support material pH of the support material Typical removal efficiencies Typical value 1-1.5 1-3000m 2 50-300,000m 3h-1 5-500m 3m-2h-1 5-500m 3m-3h-1 50% 15-60s 0.2-1.0 cm water gauge (max. 10 cm) 0.01-5 g m-3, 500-50,000 OU m -3 15-30 C >98% 60% by mass pH 6-8 60-100%

Source: Deshusses, M.A., Biodegradation of Mixture of Ketone Vapours in Biofilters for the Treatment of Waste Air, Ph.D. thesis, Swiss Federal Institute of Technology, Zurich, 1994, cited in Devinny et al. (1999). The contaminant vapours and oxygen are transported in humid air by forced convection. Interphase mass transfer occurs and, provided that the biofilter bed particles are small, interfacial equilibrium is achieved so that gas-phase resistance can be neglected, in the biofilm, simultaneous diffusion and biodegradation of the pollutants occurs as a result of growing or resting microorganisms. The sorption of all chemical species involved is possible either after diffusion through the biof ilm (Deshusses, 1997)

12

or directly from the gas-solid contact occurring on plain portions of particles (Shareefdeen and Baltzis, 1994). Furthermore, the probability of metabolite accumulating on the biofilm is certain as bioconversion progresses (Devinny and Hodge, 1995). The metabolites experience the same simultaneous diffusion, bioconversion or sorption process. The toxic or acidic metabolites are neutralised by the addition of limestone or other pH regulators mostly mixed with the medium support before the packing of the biofilter (Deshusses, 1997). In addition, the fact that formation of metabolites is seldomly perceived indicates that they are degraded faster than the primary volatile compounds either by auxiliary strains in direct contact with the primary contaminant converters. The evolved carbon dioxide from the oxidation process diffuses back and is subsequently transferred to the gas phase (Deshusses, 1997).

Biofilter Media Media selection for a biofilter is based on the ability to support bacterial growth (Quigley et al., 2004). However, the criteria vary depending on the reactor types. The performance of a biofilter for odor or VOC control depends on the nature of the carrier or support media where the adhesion of microorganisms takes place (Devinny et al., 1999), resulting in the development of a biofilm due to contaminant degradation. Large surface area, pressure drop, cost, chemical reactivity and void space are important factors considered in selecting a biofilter medium. For biofilters, in addition to biological properties, media must provide good absorption capacity, adsorption property, pH buffering capacity, good pore structure and very low compaction over time (Leson and Winer 1991). Although the selection of biofilter media should be based on all these parameters, frequently only media with good biodegradation properties such as peat, compost, soil are selected, without giving consideration to structural, mass transfer and adsorption properties (Devinny et al., 1999). Some of these issues have been resolved with the introduction of vendor -supplied manufactured media that are produced with definitive specifications. For example, Biorem,s manufactured medium consists of hydrophilic mineral cores coated with hydrophobic sorption material to give a high specific surface area. For an optimum biofilter, the media must be coated with nutrient13

rich organic material for microbe hosting and suitable binders to provide product strength, buffer capacity and stability (Devinny et al., 1999). Through research and experience,biofilter vendors have now put emphasis on media development and selection for more efficient odor and VOC removal performance that meets compliance. In subsequent headings, medium selection criteria, engineering specifications and limitations are discussed in detail. MICROBIOLOGY OF BIOFILTERS Several groups of microorganisms, primarily bacterial species, are responsible for the degradation of the air pollutants in bioreactors. Naturally occurring biofilter media such as peat and compost contain about 1 billion microorganisms per gram, capable of degrading odor and VOC contaminants present in the air. Activated sludge suspensions from sewage treatment plants serve as inoculum for many compounds (Ottengraf and Diks, 1990), but poorly biodegradable compounds such as chlorinated hydrocarbons

and aromatics require inoculation with specially cultivated organisms (Ottengraf et al., 1986). In manufactured media supplied by some vendors, inoculum is added during the media-production stage. During treatment, the introduction of odorants or VOC contaminants into a bioreactor shifts the distribution of the existing microbial populations toward a strain that can metabolize the target odor or VOC pollutant s. Compared to pure cultures, the use of microbial consortia or mixed cultures is common in large-scale applications for odor or VOC contaminants. Wastewater sludge pr a filter bed containing a diverse natural community are the most frequently used, non-defined inoculums for the start-up of bioreactors. However, some studies with single strain or defined consortium inoculation at start-up have reported high contaminant removal efficiencies (Cox et al., 1997). Biofilters using natural organic carriers are expected to contain a wide range of organisms including bacteria, fungi, yeasts, algae and protozoa as compared to synthetic or inert carriers. However, biofilter inoculation may speed up the start-up period.

14

In operation, pressure drop and clogging are well-known phenomenon in biofilters, due to the nature of the carrier, particle size and shape, moisture content, superficial gas velocity and excessive microbial growth (Devinny et al., 1999). Clogging due to extensive biomass formation is more common with inert material where a nutrient solution is fed at regular intervals. Limiting the supply of some nutrients such as phosphate and potassium ion or nitrogen is a possible means to reduce biomass growth (Wubker and Friedrich 1996). Packing Media and biofilte r types In recent years, biofiltration has been widely employed in odor removal in the waste related treatment industries. Industrial sectors including rendering, food processing, flavour manufacturers, and composting facilities are selecting biofilter systems for odor and VOC removal in their facilities. In biofilters, as contaminated air is passed through a bed of media, the contaminants and oxygen are first transferred to the biofilms formed on the surface of the media particles and then metabolized by bacteria. In order to sustain microbial growth on the media particles, moisture is provided by saturating the process air before it enters the biofilter unit. The moisture is also provided by intermittent, occasional spray irrigation of the media. The media within a biofilter are normally composed of material such as peat, wood bark, soil, compost, coated ceramic particles, synthetically manufactured media, or a combination of these products. If properly designed, biofilters can provide complete removal of the odor and VOC contaminants present in waste air. Chapter 6 covers the historical development, fundamental theory, process mechanisms, and important parameters of the biofiltration process. The major difference between biotrickling filters and biofilters is the presence of continuous water flow in the reactor. The water phase carries nutrients for the microorganisms and is usually neutralized before recirculation, for pH-control purposes. Microbial oxidation takes place in the water phase as well as in the immobilized biofilms attached on the media particles. Microorganisms in the biofilms degrade absorbed contaminants into harmless and odourless products. Excessive biomass growth and clogging are major problems encountered in biotrickling filters. Biotricklingbiofilters are more complex to construct and operate than classical

15

biofilters. For chlorinated VOCs and compounds that produce intermediate acidic byproducts, however, biotrickling filters are very effective (Devinny et al., 1999). While biofilters and biotrickling employ immobilized organisms, bioscrubbers utilize dispersed (suspended) cultures. Bioscrubbers consist of two units: the usual scrubber in which VOCs and odorous compounds are transported from the air to a water phase and a liquid-phase bioreactor where the water exiting the scrubber is subject to biological treatment in a liquid phase. The bioreactor, which contains suspended cultures, requires sufficient oxygen through aeration to maintain a high level of biodegradation. However, due to volatilization from bioreactor and scrubber operations, emissions may not be completely eliminated. Thus, a secondary polishing unit may be required in many cases. As in biotrickling filters, the water phase allows the addition of nutrients to the system. Bioscrubbers can handle varying inlet concentrations and flow rates, or shock-loading conditions more easily. Chapter 8 covers the basic principles, variations of designs, bioscrubber technology. Membrane bioreactor technology is becoming popular for odor and VOC removal. Membrane bioreactors are based on the utilization of microporous hydrophobic membranes that are made up of materials such as polypropylene and polyethylene. The membranes provide high gas permeability, while not allowing water transport across the membranes. Through diffusion, gases in the membranes are transferred into the surrounding liquid phase that contains essential nutrients and microorganisms capable of degrading contaminants. The arrangement is very much like a shell and tube heat exchanger where, in this case, the tubes are made of very small-diameter porous fibers. The membranes are manufactured to give a very high specific surface area. Ergas and McGrath (1997) reported several applications involving the use of membrane bioreactors for the removal of contaminants from airstreams. BIOFILTRATION REVIEW Biofiltration has become in the past few decades the most feasible and cost-effective technology for the management of environmental air quality (Prado et al., 2009). Van Groenestijn and Kraakman (2005) in a review paper stated that, air pollution has become an increasing environmental and health concern. To this end, Biofiltration is becoming more popular because of improving reliability of biofilters, and in addition, it

16

is a green technology that uses no chemicals thereby creating no issues of potentially hazardous media disposal (Easter et al., 2005). Thus, its significant integration in waste treatment facilities such as mechanical biological treatment (MBT) and for that matter anaerobic digestion (AD) plants which are capable of generating effluent gas composed of odorous pollutants such as hydrogen sulphide, volatile organic compounds (VOCs) and pathogens contamination. In view of this, waste air treatment at these facilities usually focuses on odour nuisance complains from neighbouring communities (Iranpour et al., 2005, p254). Due to alarming and terrific growing concerns about the toxicity and carcinogenicity of these VOCs, the development of biofiltration technology to a more consistent and effective standard is persistently gaining authorization. This is reflected in the hazardous air pollution regulation on VOC emissions in the UK- Air Quality standards which limit VOCs emissions. In the wake of environmental nuisance caused by these waste management facilities , one of the more recent practices to reduce the odour emissions is by the use of biofilters (Nicolai, n.d,p952). Biofiltration is not a new technology as such. However, it is an adaptation of natural atmospherecleaning processes. The biofilter consists of simple structured packed bed completely surrounded by microscopic bacteria (Khan and Ghoshal, 2000, p534). The microorganisms convert the gaseous contaminants to carbon dioxide, water vapour and organic biomass. Biofilters use porous solid medium to support microorganisms and allow access to the contaminants in the gas flow. The contaminated influent air passes over the filter media (Khan and Ghoshal, 2000) and is engulfed by a biofilm. The contaminants are adsorbed to the biofilm from the gas and are further oxidized by the microbes. Deshusses (1997) pointed out that, the packing medium which supports the biofilm functions as a reservoir for water, pollutants and nutrients by adsorption on its matrix and absorption in pore water. The biofilm is where the microorganisms breakdown the contaminants. The media of biofilters include peat , soil, compost ,wood chips, straw or a combination of two or more (Nicolais and Janni, 2001), cited in (Nicolais and Lefers ,n.d, p952). The pore structures, specific area, flow resistance and retention ability in the biofilters can be very different, depending on the support material used for biomass growth (Khan and Ghoshal, 2000, p534). Natural filter media like soil, compost or bark cakes up easily

17

during operation, but materials such as plastic, ceramics and metal have , tremendously, larger bed porosity for a priori design of the bed pore space (Devinny et al., 2001). In the past years tremendous attention was given to biofiltration as waste gas purification technology due the enormous advantages it has compared to traditional methods (Khan and Ghoshal, 2000, p534). Biofiltration is a microscopic ecological technology by which contaminated air is passed through a porous packed bed that supports the micro-organism to oxidize volatile organic compounds into innocuous products (Janni et al., 1998). The extent of adsorption is a function of the chemical characteristics of the specific pollutant such as water solubility, Henry s constant and molecular weight (Devinny et al., 1999). The microorganisms convert the gaseous contaminants to carbon dioxide, water vapour, inorganic products and organic biomass (Liu et al., 1994) Biofilters are classified according to the configuration (open or closed) and flow sequence (up-flow, down-flow or horizontal flow). For the fact that it is the microorganisms in the biofilter that provide the actual work of breaking down odorous compounds, it is imperative the transformations and interaction of these

microorganisms is understood. Our knowledge of microorganisms ecosystem is fragmentary and poor (Devinny, 1999) and for that matter the research on the biofilter design performance and efficiency with respect to changes in microbial community and pollutant gasses in the inlet and outlet is essential. Several factors militate against the design, performance and efficiency of biofilters. Among the lot include clogging or channelling, biofilter residence time, pressure drop, biofilter moisture, biofilter media depth, pH, temperature, biofilter media degradation, direction of airflow and biofilter cost (Janni et al., 1998 p19). The design parameters of biofilters include bed thickness, media type, empty bed contact time, moisture control and media bed orientation. For a minimized establishment cost and maximized effectiveness, the design parameters must be well set up since they influence the operating factors greatly and as such the efficiency of the biofilters. In addition, consistent operational data on contaminants and removal efficiencies is essential for the operation of biofilters; this is required to meet emissions legislation. This research will also focus on the life span and operation of biofilters. To undertake

18

this exposition, three case studies were established at different waste management companies to obtain the requisite information and data for the writ e up. The companies are the MBT plant at Global Renewables, Lancashire, and Composting plant at TEG Environmental Limited and as a direct comparison with the biofilters use by Orchid Environmental Limited (mechanical treatment thermal Steam process. Each

organisation uses different biofilter type to neutralise volatile contaminants from waste air streams from MBT and AD plants. DESCRIBING BIOFILTER PERFORMANCE In general, a competent description of intrinsic biofilter performance should take account of the inlet gas concentration, gas flow rate, volume of the biofilter system, removal efficiency and the elimination capacity (Devinny et al., 1999).For easy understanding of the performance andmechanism of activities in biofilters, it is imperative to clearly outline common terms that are pertinent to field of biotechnology. Empty bed Residence Time The term empty bed residence time (also empty bed contact time or empty bed detention time ) relates the flow rate to the size of the biofilter. It is defined as the empty bed filter volume divede by the air flow rate (Devinny et al., 1999). That is EBRT ! Vf Q

Where EBRT =empty bed residence time (seconds, minutes); Vf = filter bed volume (m3, ft 3, etc); and Q =air flow rate (m3 h-1, scfm, etc) The empty bed residenc time overestimates the actual treatment time. The medium occupies a substantial fraction of the biofilter, reducing the volume within which the air flows and shortening the contact time. Even so, it is a commonly used parameter because it is calculated. The true residence time, which is the actual time a parcel of air will remain in the biofilter, is defined as the total filter bed volume multiplied by the bed porosity of the filter medium, divided by the air flow rate: Vf v Q

19

Where

=true residence time (seconds, minutes);

= porosity = volume of void

space/volume of filter material. In literature, the terms empty bed residence time (EBRT) and true residence time ( ) are both commonly used. The difference between these two terms is the porosity factor ( ) and can be quite substantial. The effects of the empty bed residence time or true residence time on the performance of a biofilter are parallel. Generally, as either EBRT or increases either by reducing the

volumetric flow rate or by increasing the volume of medium, the system performance will improve. For many particular biofilter sites, the flow rate is fixed and a function of the contaminant producing process. Hence, reactor volume is often the only variable that can be increased. Increasing the porosity or size of the filter bed can increase this volume. However, biofilters with larger volumes and longer gas residence times are more expensive.Typical vapour residence times for commercial and industrial applications range from 25 seconds for the treatment of odor and low VOC concentrations to over a minute for high concentrations of VOCs (Leson and Winer, 1991).

Surface (or Volumetric) and Mass loading Rate Surface (or volumetric) and mass loading rate are terms used to define the amount of air or contaminant that is being treated. Both terms are normalized, allowing for comparision between reactors of different sizes. Surface loading rate is defined as the volume of gas per unit area of filter material per unit time (in metric units as m3 of gas per m2 of bed suface per hour). Similarly, the volumetric loading rate is defined as the volume of gas per unit volume of filter material per unit time ( metric units as m3 of gas per m3 of filter material per hour). surface loading ! Where A= filter area (m2, ft 2). Q A

20

Volumetric loading !

Q Vf

The mass loading rate (either surface of volumetric) is the mass of the contaminant entering the biofilter per unit area or volume of filter material per unit time, often expressed as grams per m2 or m3 of filter material per hour. Because flow remains constant through a filter bed, the mass loading along the length of the bed will decline as contaminant is removed. However, generally an overall mass loading rate for a system is defined as Mass loading (surface) ! Where C Gi = inlet concentration (g m -3). ass loading (volumetri c) ! Q v C Gi Vf Q v C Gi A

Removal efficiency and elimination capacity Removal efficiency and elimination capacity are used to describe the performance of a biofilter. Removal efficiency (RE) is the fraction of the contaminant removed by the biofilter, expressed as a percentage: C  C Go Removal efficiency ! Gi Vf v 100

Where C Gi = inlet concentration (ppmv, g m -3); CGo = outlet concentration (ppmv, g m -3). Elimination capacity (EC) is the mass of contaminant degraded per unit volume of filter material per unit time. Typical units for elimination capacity are grams of pollutant per m3 of filter material per hour. An overall elimination capacity is generally defined: Eliminatio n capacity !

C Gi  C Go v Q
Vf

Elimination capacity = Volumetric mass loading RE

21

Removal efficiency is an incomplete descriptor of biofilter performance because it varies with contaminant concentration, airflow, and biofilter size and only reflects the specific conditions under which it is measured. One of the most appropriate descriptors of pollutants removal in a biofilter is probably the Henry law coefficient, because it links gas phase and biofilm concentrations (Johnson and Deshusses, 1997).The elimination capacity allows for direct comparison of the results of two different biofilter systems because the volume and flow are normalized by definition; however, elimination capacity is also a function of input concentrations.

E.C Maximum Slope= 1 Elimination R.E. =100% Capacity

Critical E.C. where Load equals E.C.at Highest R.E

Load Figure 3.Typical elimination capacity vs. load curve. Elimination capacity is always equal to or less than the load. The ratio between elimination capacity and load is the removal efficiency of the system.

Effluent concentration (or percent removed) is still commonly used as the goal of regulatory compliance. Elimination capacity can only be equal to or less than the mass loading rate. Under low load conditions, the elimination capacity essentially equals the load, and the system is calculated to be at 100% removal efficiency (see figure 3 above). By increasing the load on a system, a point will reached where the overall mass loading rate will exceed the
22

overall elimination capacity, generating removal efficiencies less than 100%. This point is typically called the critical load or critical elimination capacity. The decline in overall removal efficiency may be explained differently depending on which parameter is increased to increase the overall mass loading rate. If the flow rate is increased or the volume decreased, the residence time is reduced, and the contaminant may not have sufficient time to diffuse into the biofilm and be readily oxidized. Conversely, if the concentration is increased and the flow rate and volume remain the same, the biofilm may not be able to absorb the increase in concentration, with some of the contaminant simply passing through the system untreated. As the loading rate continues to increase, a maximum overall elimination capacity (ECmax) will eventually be reached. This maximum overall elimination capacity is independent of contaminant concentration and residence time within a reasonable range of operating conditions. Elimination capacities for conventional biofilters treating common pollutants typically range from 10 to 300g m -3h-1. MECHANISM OF BIOFILTRATION The ultimate operating mechanisms are complex and thus must be controlled to ensure success in the degradation of contaminants efficiently. The biofilter contains a porous medium whose surface is covered with water and microorganisms. Treatment begins with the transfer of the contaminants from the air stream to the water phase. The dissolved contaminant is moved by diffusion and by advection in the air. The contaminant may form complexes with organic compounds in the water. It may adsorb to the exopolysaccharides released by the biofilm-forming cells or to the cells themselves. Ultimately, biotransformation converts the contaminant to biomass, metabolic by-products, or carbon dioxide and water. Chlorine or sulphur present in the contaminant will disappear as chloride and sulphate. The biodegradation is carried by a complex ecosystem of degraders, competitors, predators that are at least partially organized into a biofilm. Gas transfer In the first step of the mechanisms, gas transfer, the movement of the contaminants from the air to the water phase occurs according to physical laws. At equilibrium, a partition between the air and water develops according to Henry s Law in which the
23

concentrations in the water varies proportionally to those in the air. This means an increase in the concentration of the contaminants in the air phase will witness a corresponding increase in the water phase. The transportation of VOCs from the gas phase to the aqueous phase is a key process (Zhu et al., 2004). Basically, Henry s constants are almost all well below 1, even for hydrophobic compounds. Constants for hydrophilic compounds are very low (Devinny et al., 1999) and this makes them easily biodegraded due to the fact that the contaminants get retarded and tend to dissolve at faster rate in the aqueous phase leading to their removal in the biofilter.Biofilters thus are very suitable to remove VOCs with moderate to low Henry s constants (Zhu et al., 2004). Dehusses and Johnson (2000) studied biofilter elimination capacity and discovered that the biodegradation of VOCs in biofiters is influenced greatly by their availability or by the Henry s constants. Previous studies also suggested that the mass transfer of gas-phase substances into a biofilm may not be limited by the aqueous phase since transport of gas-phase compounds into the biofilm can occur directly through non-wetted areas (Zhu et al., 2001). In effect, the partition and retardation phenomenon is affected by the adsorption of contaminants by the biomass and the support medium. The ratio between the mass of material in the air and the mass in all the other phases, measured within a given volume of biofilter determines the retardation effect (Devinny et al., 1999). At equilibrium, concentrations of the contaminants decline substantially as the air moves from the regions near the inlet to regions near the outlet and decreases drastically deep within the biofilter than at the air-water interface ( Devinny et al., 1999, p24). Transfer Rate In addition to this mechanism is the transfer rate of the contaminants. The economy of biofilters is dependent on how quick air can be passed through. Pollutants mass transfer from air to water is of prime importance in the construction of any industrial biofilter. On the other hand there must be a hydrostatic phenomenon in the biofilter where a laminar flow of the water is being observed (Devinny et al., 1999). Added to this, downward movement may occur if condensation or irrigation is heavy; however, operations are always in that direction to minimise it to avoid leachate production.
24

Under these conditions, diffusion is predominant and contaminants move place to palce in the water creating opportunity for efficient biodegradation. This notwithstanding, the transfer of contaminants from the air to the water and solids in a biofilter is a fundamental step in treatment and is sometimes loosely referred to as adsorption or dissolution. Contaminant molecules may be simply dissolved in the water, but they may also be absorbed on the surface of the medium, taken up by living cells, adsorbed on the surface of biofilm organic matter, absorbed within organic matter in the biofilm or medium, or collected at the surface of the water. For highly soluble contaminants such as ethanol, the dissolved form may be dorminant, and the volume of the water phase will have considerable influence on the amount transferred from the air (Hodge and Devinny, 1997). For more hydrophobic contaminants, the major reservoir may be material adsorbed on the surface of the medium and absorbed within the organic matter. Despite the fact that adsorption phenomenon in biofilters is poorly understood, it remains very imperative to biofilter operation. The ratio of the total amount of contaminant in the water-and solids phase to the amount in the air determines the residence time of the contaminant in the biofilter. Contaminants at the surface of the medium or inlarge pores may be available for biodegradation, while those in pores too small for microorganisms may not be. Adsporption and desorption fro different media or from different parts of a single medium will occur at different rates. The general goal of biofilter design in this respect is to achieve the maximum possible concentration of contaminant in the forms that are available for biodegradation. The best way to this is not always evident. Models have been formulated to give an explicit explanation of this phenomenon. One of such is the Freundlich model. In Freundlich model, it is assumed that sites for adsorption are not limited, and the amount of contaminant held depends on the concentration in the water (Devinny et al., 1999). The Freudlich model presumes that the adsorption capacity is unlimited in range over which the relationship is applied: increasing the liquid phase concentration will always increase the amount of contaminant adsorbed. However, according to Langmuir, adsorption occurs at specific sites and that each site can be occupied by only one contaminant molecule (Devinny et
25

al., 1999). This implies that if the concentration in the liquid phase is small, most of the adsorption sites are not occupied. However, when the concentration in the liquid phase is very high, essentially all the adsorption sites will be occupied and the amount adsorbed will be a constant independent of concentration. Thus, it is pertinent to note that adsorption capacity of the medium in the biofilter is a function of the concentration of contaminant in the air in several cases. Increasing air concentrations will cause more contaminant to adsorb, and a drop to lower concentration will cause it to be released. Hence, the biofilter serves as a buffer for concentration fluctuations (Devinny et al., 1999). Contaminant Biodegradation The biofilm The key element in destroying contaminants is the biofilm. This is a mass of organisms growing on the surface of the solid medium and carrying out the metabolic activities which transform the contaminant to harmless products. Biofilter efficiency is highly dependent on the activity of the microorganisms. The structure of biofilms is poorly known. The biofilm is formed as a result of biomass accumulation at the interface between the solid medium and the gas. Resident microorganisms in the biofilm protrude above the water into the air (Hugler et al., 1996) where they are able to attack and metabolize contaminants. The organisms depend on the carbon for growth and as the process goes on, some the microorganisms are in turn preyed upon resulting in a complex system. This complex microbial ecosystem is what most of the time is termed as the biofilm significant to growth of the microorganisms. Kinetics The kinetics of contaminant degradation are often modelled using the Michaelis-Menten equation for enzyme mediated reactions (Devinny et al., 1999) which assumes that the number of microorganisms present in the biofilm is not changing. However, in many cases the biomass will be developing as the contaminant is consumed. Thus since contaminant degradation is the result of microbial activity, the kinetics of contaminant degradation are closely related to the kinetics of microorganisms growth. Growth is commonly presumed to be proportional to the size of the microbial population. This depends on the concentration of contaminant available for use by the microorganisms.
26

Besides, Alexander (1999) noted that kinetics in bioiflters may depart from the normal performance rule in the sense that the following could be observed as the process proceeds: diffusional barriers may exist, the substrate may be sorbed on the medium, the degrading microorganisms may be aided or hindered by the presence of other compounds, inorganic nutrients or oxygen may be limiting rather than the substrate, several species with different degradation constants may be active, or predators may limit the numbers of bacteria. In addition, cells may be aggregated, affecting substrate transport processes and acclimation of the microbial population may be necessary. Furthermore, it is noted that degradation kinetics are further complicated because at low concetrations the contaminant may not diffuse into the biofilm rapidly enough to penetrate the full length of the biofilm. Under such conditions less biomass will be utilized and less substrate consumed as the concentrations decline. Thus degradation becomes limited by the diffusion phenomenon rather than by biological activity. The degrading processes that exist in the biofilter comprise aerobic and anaerobic processes. In the biofilter, aerobic processes predominate due to the flow of large amounts of through them. Contaminants are normally present in low concentrations so that even their complete oxidation dos not deplete the oxygen available in the air. According to Devinny et al. (1999), aerobic degradation is often desired because it is generally faster and less likely to produce objectionable by-products. Anaerobic conditions may occur at the bottom of biofilms or within remote pores in the support medium where the supply of oxygen is diffusion limited. This occurs if the degradation rates are high and oxygen consumption within the biofilm exceeds the rate at which it can diffuse inward through the water layer. In poorly operated biofilter, where there is too much water or the support medium is decomposing and compacting, the air flow may be channelled. This creates anaerobic regions within the biofilter due to the fact that no air passes through them. To this end, Devinny et al. (1999), indicated that the prevailing anaerobic conditions may generate sulphides, mercaptans, ammonia, short-chain carboxylic acids, or other odorous or toxic products which are mostly the major cause of biofilter replacement. It is as the result of these processes occurring in the absence of oxygen, it is presumed that they were done by microorganisms in anaerobic regions beneath the aerobic biofilm.
27

The products generated in the degradation process include carbon dioxide, water, or sulphate and nitrate. Alternatively, the complex compounds might undergo several different transformations in several microbial species before mineralization. Additionally, the oxidation process also generates heat which is sufficient to raise the temperature from 20 C to 38 C (Devinny et al., 1999). It is thus possible to use the heat generation as a measure of degradative activity; however, several factors must be taken into account in order to calculate a heat balance. DESIGN CRITERIA FOR BIOFILTERS Biofilter medium selection criteria for optimum perfomance The component to be considered in designing a biofilter is the medium and its selection (Quigley et al., 2004). The selection of the media for biofilter applications has evolved over time and continues to evolve. Materials such as soil, peat bark, wood chips, compost, heather and inert additives such as perlite and plastics have all been used successfully as media for biofilters. The most common media are soil, bark and compost. A good biofilter medium should (Quigley et al., 2002): Support a large diverse microbial population Have a low pressure drop Provide pH buffering capabilities Produce clear drainage water (leachate) Have the ability to retain microbes Have high bearing strength Be physically stable The design of the medium involves a number of discretionary decisions, that include the medium, particle sizing, cross-sectional depth, surface loading rate per square metre of the medium, porosity and desired service life. These parameters are dependent on the odorous air-stream characteristics, including contaminants of concern and loading rates (Easter et al., 2005).

28

In addition, although the selection of biofilter media should be based on all these parameters, only biofilter media such as peat, compost, soil and chicken manure which have excellent biodegradation properties are selected without giving consideration to structural, mass transfer and adsorption properties (Shareefdeen and Singh, 2005).

Controlling factors of biofilter operation As part of the design criteria to ensure optimum conditions and efficient operation of biofilters, a variety of parameters must be satisfied in the design (Quigley et al., 2004 p320). Added to this Devinny et al. (1999) noted that these factors a must be carefully selected to produce a single principal goal which is to provide suitable environment to sustain the microorganisms responsible for the biofiltration process. Among these parameters include the temperature, size of the biofilter, moisture content, pH, airflow distribution, leachate control and drainage and the packing material selection (Quigley et al., 2004). However, Devinny et al (1999, p51) noted that the three most significant factors for an efficient biofilter are medium moisture content, pH and bed temperature. Furthermore, the other factors are also important as they influence medium lifetime or removal performance to a lesser extent than do these three factors (Devinny et al., 1999). The principal means of treatment in the biofilters is the action of the pollutant oxidizing bacteria which implies that the controlling parameters in a biofilter are meant to control the activity of the process culture. Moisture Content Moisture control is one of the most important tools in the maintenance of biofiltration medium. Medium that is too dry does not support for the microbes to ensure efficient contaminants removal. On the hand, from Quigley et al (2004), extremely wet medium becomes too dense and compact, resulting in reduced porosity and high pressure drop. High back pressure results in reduced air flow (Quigley et al., 2004 p321). In view of this phenomenon, it is a pre-requisite in the design of biofilters that a prehumidification system is established to moisten the odorous air flow to nearly 100% relative humidity (Lu et al., 2002). Janni et al (1998) outlined that the overall effect of the reduced porosity and high back pressure results in an unconstructive impact on the microorganisms and reduced efficiency VOCs or odour removal. The effectiveness of the
29

biofilter depends entirely on the health and effectiveness of the biofilm at capturing and decomposing the pollutants and odours. Prehumidifying the inlet air stream is recommended to maintain the requisite moisture in the filter medium. Alternatives for prehumidifying the air include the use of spray nozzles in the biofilter inlet air duct, spray chambers ahead of the biofilter using water to prevent biofilter desiccation (Janni et al., 1998 p11). The activity of the microorganisms is a function of water content (Devinny et al., 1999). The presence of moisture in the biofilm is imperative and thus affects the transfer of contaminants from the air and the physical properties of the medium. With regard to biofilter affinity for water, Devinny et al. (1999) postulated that most biofilters being solids have some sort of dipole moment to attract the polar water molecules. However, some packing materials such as peat are hydrophobic and are not easily moistened. The observation being that most packing materials when allowed to dry lose their hydrophilic property and become hydrophobic (Devinny et al., 1999 p52). This means rewetting of dry out biofilters will require some time to assume their water retention potential. The mechanism is that as a dry porous medium encounters water vapour, it develops the tendency to draw water molecules out of the air that contain the dissolved organic contaminants. To this end, as the water molecules are drawn to the surfaces of the biofilter medium, the adsorption of the contaminants is initiated in the biofilm leading their breakdown by the microorganisms housed in the film.

Temperature The function of a biofilter depends on the gas and medium temperature. Microorganisms operate efficiently at temperatures ranging from about 15 C to 30 C (Quigley et al., 2004). The higher the temperature, the higher the metabolic rate and the biodegradation rate up to a temperature of about 40C. However, solubility and adsorption rates decrease with increasing temperature. The heat required to sustain the bed temperature is in two parts: one from the odorous air and the rest from the

30

metabolic activity of the microorganisms. Thus in the design of biofilters it is very significant to control the temperature of the odorous gas before entering filter medium. Microbial activity and biofilter success are both strongly influenced by temperature (Devinny et al., 1999 p60). A microorganism is a tiny sack of chemicals and enzymes whose existence is a countless of biochemical reactions that run more rapidly as the temperature increases (Quigley et al., 2004). Utmost reaction rates roughly double when the temperature raises 10 C and microbial metabolic activity increases proportionally through the range in which the synchronization of the reactions can be maintained. Usually, a warmer reactor will treat contaminants more rapidly and efficiently. However, there should be maximum limits of temperature as high

temperature may make some reactions occur so rapidly that metabolic coordination is disturbed (Devinny et al., 1999). The reason being that each enzyme is made of proteins and has a temperature limit beyond which it is denatured and is no longer effective in aiding microbial activity. As the temperature rises, each microorganism approaches a limit where it can no longer be effective and metabolic activity drops off rapidly. Basically, the cell is destroyed by the heat.

31

Maximum Activity for Well Acclimated Biofilter Rate of Metabolic Activity

Psychrophile Mesophile Thermophile

Temperature Figure 2 Temperature effects on species of microorganisms and biofilter activity (Devinny et al., 1999 p61).

From figure 2 above, the analysis by Devinny et al (1999) suggests that the metabolism of the microbes deteriorates as the temperature declines reducing the rate at which biodegradation of the contaminants proceeds. Ultimately the microorganisms become basically inactive and again may die as necessary functions cease. While some microorganisms are active at temperatures near the freezing point of water, and many can survive freezing, none can be active while frozen. As shown in figure 2, the combination of these factors produces a temperature-activity curve for a given cell which typically rises with temperature to a limiting value and falls rapidly (Devinny et al., 1999). Every microbial species has its temperature range over which they can survive, reflecting the diversity of natural environments.

32

Oxygen Content For effective operation of biofilters, oxygen content may be an issue of limitation in the biofilm where biodegradation is executed. In high-performance biofilters, oxygen limitation may occur in the biofilm (Devinny et al., 1999 p68). The idea of oxygen content observed as a limitation(Zhu et al., 2004), does not essentially denote that oxygen is completely depleted in the biofilm, but refers to the general situation where the rate of biodegradation is affected by the concentration of oxygen. First and

foremost, the reality of oxygen limitation in an air biofilter might appear paradoxical due to the fact that air contains 21% oxygen. However, the reason for oxygen limitation is that the oxygen gas-liquid partition coefficient is 33.5 (Devinny et al., 1999) which implies that greater proportion of the oxygen is in the gas phase somewhat than dissolved. In this respect, Devinny et al., (1999) studied that the dissolved oxygen concentration in equilibrium with air at 25 C is approximately 8.1mg L-1, or 0.25mMol. On the basis of this phenomenon, Cussler (1997) stated that the diffusion coefficient of oxygen in water is about 2.1 10 -9 m2s-1and common VOC diffusion coefficients are on the order of 0.8 to 1.3 10 -9 m2s-1. Assuming identical diffusion coefficients for both oxygen and VOCs if the stoichiometric amount of oxygen required to degrade the treated VOC is greater than the interfacial oxygen concentration of 0.25mMol, the probability of oxygen being consumed in the biofilm prior to VOCs treatment is sufficiently significant. This implies that the effective biofilm growth, according to Williamson and McCarty (1976) is controlled by oxygen rather than VOC availability. Aside this, it is reasonable to conclude that maximum elimination capacity of VOCs in a biofilter is dependent on threshold concentrations of oxygen. To this end, Devinny and Hodge (1995) discuss that oxygen deprivation is undesirable since it can lead to partially oxidized by-products such as carboxylic acids or aldehydes which can cause nuisance odors. Table 1 shows some values for theoretical minimum VOC concentration that cause oxygen limitation. Another important phenomenon with regard to this is that oxygen limitation could occur in biofilters treating VOCs with low values of Henry s constant due to their more partitioning into water and biofilm (Zhu et al., 2004). VOCs with low values of Henry constant or partition coefficient implies that they are hydrophilic (Devinny et al., 1999).

33

The values show that treatment of hydrophilic compounds (Zhu et al., 2004) due to a more favourable partition in the water is most probably limited by oxygen at concentrations normally prevailing in the biofilters whereas treatment of hydrophobic compounds will perhaps not be subject to oxygen limitation (Devinny et al., 1999). Table 1 Estimated Threshold concentrations for Oxygen Limitation and Maximum

Elimination Capacity for the treatment of Various VOCs Maximum VOC Lowest Elimination Amount of concentration of capacity , oxygen needed contaminant in the assuming an for complete air to induce oxygen oxygen oxidation depletion in biofilm transfer rate (gm-3) of 200gm -3 h-1 3 0.0009 88 5 0.0245 110

Compound

Dimensionless Henry s coefficient

Ethanol 0.000257 Ethyl 0.0055 acetate Toluene 0.275 9 0.7112 64 Hexane 74.13 7.5 195 95 Calculated assuming complete aerobic oxidation; oxygen and VOC diffusion coefficient to be identical; a temperature of 20 to 25 C. Source: Devinny et al (1999) Biofiltration for air pollution control.

Kok (1991) estimated the maximum elimination capacity of biofilters for various VOCs based on the hypothetical oxygen transfer rate of 200g m-3 h-1 and found maximum elimination capacities to be well in the range of values of commonly observed values (Table 1). Oxygen limitation may influence biodegradation rates even if oxygen is not completely depleted in the biofilm according to double Michaelis-Menten reaction kinetic (Devinny et al., 1999).

34

To this end, Michaelis-Mentenexperimented that the reaction rate defines the biodegradation rate of the VOCs in which both substrate-VOC and the oxygen play a very significant role (Devinny et al., 1999). However, it is practicallydifficult to experimentally determine the equation constraints in biofilters due to the failure to measure or control conditions in the microenvironment within the biofilm despite the fact that it is easy to determine in bioreactors with suspended culture. Apparently, the lack of simple experimental methods to study the effects of oxygen concentration on contaminant removal kinetics poses a challenge. Hence, identification of meaningful measures to remediate the limitation but oxygen limitation in full scale-scale biofilters remains remain objects of further study (Zhu et al., 2004). Experimental controls have been performed on substantive oxygen outlines only in a restricted number of paradigms (Mirpuri et al., 1997) which revealed that oxygen could definitely become restrictive. However, Deshusses et al (1996) disclosed that increasing the oxygen content in the air has no stimulus in the elimination of the treated pollutants. Furthermore, another type of oxygen limitation may occur in biofilters. This is when part of the biofilter bed has much lower air permeability than the rest of the bed resulting from bed compaction, improper air distribution, or medium loading (Devinny et al., 1999). This leads to the formation of regions where only limited air exchange takes place. With time, undesirable anaerobic conditions develop in these regions. With regard to this, it is somewhat difficult for the biofilter operator to determine whether a portion of the bed is anaerobic. Emissions of metabolites or malodorous compounds and medium acidification are signs of anaerobic activity. If dead zones constitute a significant portion of the bed, reduced performance will be witnessed (Devinny et al., 1999). The ultimate proof that dead zones are present is given by residence time distribution testing, either by pulsing a tracer gas into the inlet or by performing smoke tests in open beds. Unfortunately, there is no easy remedy to the existence of anaerobic regions. If the nuisance odors are released or emission threshold limits are violated, the biofilter medium must be unloaded and carefully repacked (Deshusses and Johnson, 2000). In many instances, repacking the biofliter structure will provide an opportunity to inspect and maintain the biofliter structure, as well as possibly replacing the medium.

35

It is difficult to define the situations specifically in which oxygen limitation will occur. However, oxygen limitation is more likely to occur in the case of high concentrations of easily degradable hydrophilic compounds (Zhu et al., 2004) and in systems where thick biofilims exist. From an operator view, there is no proven technique to remediate to oxygen limitation. By varying the oxygen content of the influent gas in biofilters treating acetone and propionaldehyde (both hydrophilic), Kirchner et al. (1992) discovered that the diffusion of oxygen in the biofilm is rate- limiting. However, oxygen content limitation has been uncommonly observed in biofiltration studies and operation (Zhu et al., 2004). In this regard, it was thus discovered by Zhu et al. (2004) that a rise in oxygen content at the influent gas has no effect on the performance of the biofilter when treating Diethyl ether. Similarly, Devinny et al. (1999) found that the addition of oxygen to the inlet air is unlikely to be economically feasible, and diluting the inlet air may result in unexpected effects; however,it may be of substanceto make a bench-scale investigation and treatment cost evaluation. The most rational option is to repack the biofilter so that interfacial area and oxygen transfer are maximised. Medium pH and Alkalinity The effect of pH and temperature on biofilter performance are similar in a number of ways (Devinny et al., 1999 p62). At a particular pH range, each species of the degrading microorganisms is more effective and inhibited or denatured if conditions shift outside this range. This is analogous to the temperature effect on the activity of the microorganisms. Rapid changes in pH are destructive to most species (Lu et al., 2002). However, some species do well at high pH and some at low pH as species tolerance to moderate pH are probably ubiquitous (Devinny et al., 1999 p62). The optimum pH for biofilter design and operation is about 7. Alkaline mechanisms are habitats for microorganisms, as are acid bogs. Ideally, prior to the operation of a biofilter for contaminant removal, a pilot test should be undertaken at many values of pH over a wide range in long-term tests to determine the global optimum. However, this is uncommonly performed due to the cost involved. Some biofilter packing media such as compost or peat have pHs as low as 4 or 5 because of the organic acids. Drastic changes in the pH cansubsequently lead to the failure of the biofilter. In situations where the acid producers get ahead of the acid consumers, the pH may fall and cause the digester to fail (Devinny et al., 1999 p63).
36

Such episodes can be controlled by using a biofilter medium with a high buffer capacity. Buffer capacity is the ability of the medium to resist pH changes (Devinny et al., 1999 p63). Nutrients In the face of effective biofilter performance, Zhu et al. (2004) note that biofilterperformance depends not only on the availability of VOCs but also on oxygen as well as nutrients and substrate biodegradability. The microorganisms in biofilters consume contaminants for the energy and carbon they provide. On the contrary, they also need mineral nutrients such as nitrogen, phosphorous, potassium, sulphur, calcium, magnesium, sodium iron and many others. Others may require vitamins that they cannot themselves synthesize. Efficient biofilter operation requires that the needed nutrients be provided in the form and quantities that will support vigorous microbial activity. Compost has the important advantage that the nutrients are present in the medium. Compost is made from plant tissues, sewage sludge, or other once-living tissues. As these undergo degradation, just those elements and compounds are released during degradation in the approximate proportions suitable for cell growth. However, it is possible that the rate of degradation and therefore the rate at which soluble nutrients are generated can be too slow. Gibbins and Loehr (1998) found that nutrients released rates were limiting in a compost-perlite biofilter treating a high load of toluene. When the biofilter is manufactured, it is possible to add slow-release fertilizers in order to replenishthe nutrients as they are lost to leaching or biotransformation (Devinny et al., 1999). It is thus necessary to monitor process in order to ensure that replacement rates are as high as loss rates. Contaminant load and surface load The mass of contaminant entering the biofilter per unit time and per unit volume is the contaminant load. It has major effects on biofilter performance. Removal efficiency may be excellent at low contaminant loads but poor when loads are high. To this effect, medium acidification and for that matter system failure is far more attainable when

37

loads are high and the acid-generating reactions in the degradation may run ahead of the acid-consuming ones. Biofilters which are operating at low loads may reach approximate steady state with respect to biomass and nutrients (figure 4). The microbial ecosystem is starved for substrate and predator species tend to consume the biomass as it is produced. Biofilters treating very high loads will grow biomass rapidly if nutrients are available. Th is however, leads to excessive pressure drop across the bed and means for removing the biomass should be instituted forestall poor performance. The ultimate way to maintain biofilter efficiency is to replace the filter medium or in the case of biotrickling filters, according to Sorial et al. (1997) to backwash with water.

Low-load Biofilter

Contaminant input

Steady State Biomass Nutrient cycle

Output CO 2 and H2O

High-load Biofilter
Contaminant input Growing Biomass Biomass Nutrient Additions Nutrient cycle Wasting (With nutrients) Output CO 2 and H2O

Figure 4

Comparison of biofilter operating regime (Devinny et al., 1999 p66).

Air flow Direction


38

To ensure uniform odorous gas loading within the biofilter medium bed, uniform air flow distribution and direction is very important to maintain (Quigley et al., 2004). In a start, most applications of enclosed biofilters, downward air flow has proven superior to up-flow. However, there are a number of success stories of biofilters operating in an up-flow mode, meaning that down-flow direction is not a prerequisite design condition. The advantage of down-flow direction is that it improves moisture control (Devinny et al., 1999). If drying of the biofilter medium should occur, it will usually start from the inlet side of the biofilter as the result of either unsaturated inlet air or production of metabolic heat concentrated in the inlet side. In the case of downward flow, moisture can be efficiently controlled by addition al water supply provided as irrigation on top of the bed, where it is most needed. In the up -flow mode, drying occurs preferentially at the bottom where it is difficult to provide additional moisture (Lu et al., 2002). Another reason to prefer the down-flow mode is that it allows better drainage, particularly at the bottom of the bed. In the case of upward air flow, great care should be given to drainage and to air distribution systems. There are, however, several cases in which up-flow is advantageous. During treatment of odors from a treatment plant, sulphuric acid is generated by sulphide-oxidizing organisms. This lowers the pH (Lu et al., 2002), particularly near the air inlet where the biological activity is concentrated. Similar phenomena are perceived for the treatment of chlorinated compounds where chloride formation will cause the pH to drop. Excessive pH drop is detrimental to pollutant removal (Devinny et al., 1999). In such a case, upward flow might be preferred, enabling trickling of water or a pH buffer from the top of the biofilter or, even better, from a lower irrigation system. The acidic end products, sulphate, chloride, etc are then easily washed out without leaching through the entire bed, as would occur if sulphate were concentrated at the top of the bed indown-flow systems. Further the usual problems associated with moisture control in up-flow biofilters can be reduced by the installation of a lower irrigation system (Lu et al., 2002).

39

Dust and grease Biofilters are designed on the assumption that the material received can be completely converted to gaseous or soluble products which are carried out of the biofilter. They are not designed to handle dust. In the short term, they will efficiently collect dust, because the air flow is divided into fine convoluted streams with a large surface-to-volume ratio. As the air passes through, prospects for dust particles to contact the walls of the pores are many. In a wet medium, dust particlesthat contact will likely stick (Allen and Van Til, 2005). If large amounts of dust are present, the biofilter will soon clog, and replacement of the medium will be required (Devinny et al., 1999). Unfortunately, there have been problems with a few large-scale biofilters treating dusty air from the pressboard industry, because the accumulation was not anticipated. Allen and van Til (2005) discovered that the solution to dust removal is to install wet scrubbers.Togna et al. (1997) found clogging in a biofilter designed for treated wood products waste and decided it had resulted from the accumulation of resinous lignin degradation products and benzoic acid. In this case, the design criteria of the biofilter would have to incorporate an adequate pretreatment system to remove grease and dust (devinny et al., 1999). The installation of such system probably will make biofilters uneconomical for some applications.

METHODOLOGY Description of Selected biofilters Biofilter start up and monitoring LCA of biofilters STATISTICAL ANALYSIS Conclusion Future studies

40

41

You might also like