You are on page 1of 15

Biotechnology Advances 36 (2018) 1079–1093

Contents lists available at ScienceDirect

Biotechnology Advances
journal homepage: www.elsevier.com/locate/biotechadv

Research review paper

Biofiltration of volatile organic compounds using fungi and its conceptual T


and mathematical modeling

Alberto Vergara-Fernándeza, , Sergio Revahb, Patricio Moreno-Casasa, Felipe Scotta
a
Green Technology Research Group, Facultad de Ingeniería y Ciencias Aplicadas, Universidad de los Andes, Chile
b
Departamento de Procesos y Tecnología, Universidad Autónoma Metropolitana-Cuajimalpa, Mexico

A R T I C LE I N FO A B S T R A C T

Keywords: Volatile organic compounds (VOCs) are ubiquitous contaminants that can be found both in outdoor and indoor
Biofiltration air, posing risks to human health and the ecosystems. The treatment of air contaminated with VOCs in low
VOC concentrations can be effectively performed using biofiltration, especially when VOCs are hydrophilic. However,
Fungi the performance of biofilters inoculated with bacteria has been found to be low with sparsely water soluble
Bacteria
molecules when compared to biofilters where fungi develop. Using conceptual and mathematical models, this
Modeling
review presents an overview of the physical, chemical and biological mechanisms that explain the differences in
Computational fluid dynamics
the performance of fungal and bacterial biofilters. Moreover, future research needs are proposed, with an em-
phasis on integrated models describing the biological and chemical reactions with the mass transfer using high-
resolution descriptions of the packing material.

1. Introduction adults (Elliott et al., 2006), induce inflammatory reaction of airways,


immunological abnormalities and neurological symptoms (Win-Shwe
Atmospheric pollution has become one of the major causes of pre- et al., 2013) and an increased carcinogenic risk for formaldehyde,
mature death in developed and developing countries. Using data from benzene, acetaldehyde and naphthalene (Sarigiannis et al., 2011). The
the Global Burden of Diseases Study 2015, Cohen et al. (2017) reported changes and mechanisms behind some of these responses to VOCs ex-
that fine particulate material (PM2.5) alone was responsible for 4.2 posure have been recently studied in mice (Wang et al., 2012) and lung
million deaths in 2015, compared to 3.5 million in 1990. Nearly 60% of cells (Gostner et al., 2016). However, the lack of understanding of the
these deaths took place in east and south Asia. While the role of other mechanisms involved, did not preclude establishing links between
contaminants is well understood, such as ozone causing an additional VOCs exposure and health issues. In this regard, a study lasting 12 years
0.25 million deaths, the impact of volatile organic compounds (VOCs) and including nearly 60,000 Toronto residents, found that the exposure
has not been fully quantified, despite having a role in ozone formation to ambient levels of benzene, n-hexane and total hydrocarbons were
and that often they are associated to PM and polycyclic aromatic hy- associated with an increased risk of cancer mortality (Villeneuve et al.,
drocarbons (PAHs) (Kundu and Stone, 2014). VOCs comprise a group of 2013).
organic chemicals with a high vapor pressure at room temperature and Depending on the airflow and their concentration, VOCs can be
are commonly present in indoor and outdoor air (Khan and Ghoshal, abated using purely physico- chemical technologies such as gas mem-
2000). VOCs are emitted from industrial sources, such as petroleum brane separation, condensation and adsorption (typically used for high
refineries and chemical plants (Hoyt and Raun, 2015), commercial VOCs concentrations and low to medium air flows), incineration and
activities, such as gasoline refilling (Harley et al., 2006); and at the catalytic oxidation (for high concentration of VOCs and large air flows)
residential level, wood and fossil fuels burning for heating and cooking and biotechnological processes for dilute VOCs streams. The reviews by
(Evtyugina et al., 2014) and products used indoor like aerosols, paints Revah and Morgan-Sagastume (2005) and Delhoménie and Heitz
and cleaners (Bernstein et al., 2008). VOCs such as the low-boiling (2005) represent comprehensive assessments of the applicability of the
hydrocarbons (pentane, hexane, heptane, etc.), alcohols, aldehydes, different technologies which can be complemented with the work of
halogenated hydrocarbons, ethers and PAHs represents 7% to 10% of Estrada et al. (2012) to gain a broad perspective of the applicability and
all atmospheric pollutants emitted (Delhoménie and Heitz, 2005). costs of the most commonly applied VOCs abatement technologies.
These compounds have been shown to reduce pulmonary function in Biofiltration has become a viable economic and technical


Corresponding author at: Mons. Álvaro del Portillo 12455, Las Condes, Santiago 7620001, Chile.
E-mail address: aovergara@miuandes.cl (A. Vergara-Fernández).

https://doi.org/10.1016/j.biotechadv.2018.03.008
Received 9 January 2018; Received in revised form 9 March 2018; Accepted 14 March 2018
Available online 17 March 2018
0734-9750/ © 2018 Elsevier Inc. All rights reserved.
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

alternative for the treatment of airstreams with low concentrations of on whether using an approach tailored to each spatial and temporal
volatile organic and inorganic compounds (such as H2S and NH3) in scale or using an aggregated (or lumped) model. Despite the approach
waste gas streams. It is based on the capacity of an active microbial used, models must produce macroscopic outputs, often the only mea-
population, usually growing on a solid support in a biofilm, to trans- sured quantities in an experimental system, to compare model predic-
form the pollutants into less harming molecules, namely carbon di- tions, such as elimination capacity for contaminants and pressure drops,
oxide, sulfates, nitrates, etc. (Revah and Morgan-Sagastume, 2005) in a with the operation of the system being modeled. This section aims at
process that shares similarities with the microbial degradation of pol- introducing the general principles involved in the phenomena of mass
lutants in soils (Ren et al., 2018). and energy transport, as well as the biomass growth and biodegradation
The inherent heterogeneity and complexity of biofilters make the kinetics. Each aspect will be the subject of the forthcoming sections,
analysis and mathematical modeling of these systems a challenging dedicated to reviewing the application of these principles to the mod-
task. In such models, not only mass transport, fluid flow through porous eling of fungal biofilters. Before introducing the physical, chemical and
beds and biochemical reactions are entangled; but also, the properties biological mechanisms involved in biofiltration of air, a precision re-
of the biofilm, such as the thickness of the active biomass layer, and of garding biofilters and biotrickling filters needs to be made. While bio-
the solid inert support play a role, for example in determining the area filters have a static water phase associated to the biofilm and the sup-
of the biofilm (Spigno and Tronci, 2015; Vergara-Fernández et al., port; in biotrickling filters water is constantly supplied producing a thin
2008). water film flowing over the packing and the biofilm. However, in the
As the deployment of industrial biofilters has consistently grown laboratory and industrial practice, biofilters are intermittently supplied
since the late 80’s, so has been the increasing complexity of the con- with water, or a nutrient solution, to ensure appropriate moisture
ceptual and mathematical models developed for VOCs abatement in contents at regular time intervals, some are washed for as much as
fixed bed biofilters. In this regard, reviews in the area are devoted to the 20 min in each hour (Devinny and Ramesh, 2005). On the other hand,
conceptual analysis of biofiltration systems (Devinny and Ramesh, because unsaturated gravitational flow in porous media is difficult to
2005), to the use of fungi in biofilters (Kennes and Veiga, 2004), the control, biotrickling filters are likely to include patches that are not
mass transfer aspects in biofilters (Kraakman et al., 2011) and on how properly covered with water.
to reduce the mass transfer limitations (Cheng et al., 2016; Ferdowsi For highly soluble organic compounds, such as alcohols, the parti-
et al., 2017). However, a revision of the action of fungi in the treatment tion coefficient favors a higher concentration in the water phase, which
of volatile organic compounds and its modeling, including detailed promotes the biodegradation of the contaminant by increasing its
descriptions of the phenomena occurring in the reactor (mass, heat and concentration in the environment surrounding the biofilm (Cohen,
momentum transport) and of the microbial growth and biodegradation 2001). On the other hand, in fungal biofilters, if a layer of flowing water
kinetics, is missing. would be created over the solid support, then it would decrease the
In this review, results from biofiltration experiments are expressed surface area created by the aerial hypha, thus impairing the transfer of
in terms of the empty bed residence time (EBRT, s), VOCs inlet load (L, VOCs, especially hydrophobic ones, to the biomass phase (Sugai-
g m−3
reactor h
−1
), biofilters elimination capacity (EC, g m−3
reactor h
−1
) and Guérios et al., 2015).
removal efficiency (RE, %) according to: Fig. 1 shows a conceptual model of a biofilter. Contaminants are
Vr transported into the biofilter by the air at rates that justify assuming
EBRT = that the flow is laminar. However, dispersion occurs due to the tortu-
Q (1)
osity of the porous packing. As the air flows through the packing,
Q contaminants are transferred from the gas phase to the liquid phase
EC = (CG0 − CG (out ) )
Vr (2) embedded in the biofilm (for bacterial biofilms), or to the fungal hypha
and to the water embedded in the non-aerial portion of the biomass in
Q fungal biofilters. The contaminants diffuse or are transported to the
L= (CG0)
Vr (3) biofilm and then degraded by the microbial catabolic activity. As most
(CG0 − CG (out ) ) biofilters are used to treat polluted air and thus are based on the activity
RE = 100 of aerobic microorganisms, the availability of oxygen is also an im-
CG0 (4)
portant issue as well as the retro-diffusion of metabolic products such as
where CG0 and CG(out) are inlet and outlet VOCs concentration, re- CO2. Relevant variables affecting the mass transfer of contaminants and
spectively, Q is airflow and Vr is the reactor volume. oxygen in the reactors include the flow regimes, pressure drop, void
This review is organized as follows: the individual biological and fraction of the packed-bed, biofilm wetting; and for trickle-bed bio-
physicochemical phenomena involved in fungal biofiltration are re- filters, the flow of liquid and liquid hold-up (Lobo et al., 1999) and
viewed and compared with the bacterial case; (Section 2) next, the therefore are often considered in models of biofilter operation.
modeling efforts to integrate these aspects, understanding of the ex- In biofilters, the mass transfer parameters associated with the fluid
perimental information and optimization of biofilters design are re- mechanics of the system, e.g. the diffusion coefficient of the con-
viewed in Section 3. Finally, the characteristics and performance, in taminant in air, have been traditionally estimated using the correlation
terms of elimination capacity, are compared and analyzed for fungal proposed by Fuller et al. (1966), based on the molecular diffusivity of
and bacterial biofilters treating several VOCs. the species. Similarly, the diffusion coefficients of the nutrients from the
The remaining of this paper is devoted to reviewing the current liquid media in the biofilm have been estimated using de Nernst-Haskell
understanding of the physical, chemical and biological phenomena in- equation (Vergara-Fernández et al., 2008) and the axial dispersion
volved in fungal biofiltration and its modeling compared to bacterial coefficient, a measure of the deviation from plug-flow regime in a re-
biofiltration. actor, has been approximated using the correlation proposed by
Edwards and Richardson (1968). However, the aerial mycelia and the
2. Physicochemical and biological phenomena in a fungal biofilter colonization of the void space in the biofilter, when filamentous fungi
and its mathematical modeling are used, might lead to discrepancies between the predicted and ob-
served values of these coefficients.
Biofilters are multiphasic reactors that respond to diverse spatial
and temporal scales. From molecular diffusion and reaction to micro- 2.1. Mass transfer of species in the gas phase and reaction in the biofilm
bial growth and gas or liquid channeling, modeling of phenomena in-
volved in the operation of biofilters presents challenges to the modeler The temporal variation of the concentration of a compound; either a

1080
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Fig. 1. Scheme of a conceptual model of a fungal biofilter showing the different scales involved.

pollutant, oxygen or any other chemical species, in a differential vo- ∂CG, i ∂2CG, i
= Db, i +r
lume of gas can be obtained by assuming that there is no radial dis- ∂t ∂x 2 (7)
persion, that a species enters the control volume by advection (bulk
transport in the air) but its transport is delayed by axial dispersion where, Cb,i = Cb,i(t, xb) is the concentration of contaminant i at a given
(accounted for using an effective dispersion coefficient, DDz) leaving the time and thickness in the biofilm (x). The reaction rate (r) accounts for
control volume at a rate governed by the specific mass flux of compo- the catabolic reaction that allows the microorganisms embedded in the
nent i from the gas phase to the interface (NG,i), the specific area of the biofilm to oxidize the contaminant.
gas-liquid or gas-biofilm interface (a, interface surface per unit volume In fungal biofilters (Fig. 2.a), a contaminant is transported directly
of biofilter bed), and the void fraction occupied by the gas (ε). These from the gas phase to the aerial hyphae (Vergara-Fernández et al.,
purely physical transport mechanisms are regarded as the mass balance 2016). This situation can be represented by a simplified form assuming
for component i in the gas phase (Eq. 5). a negligible concentration gradient within the hyphae and an overall
mass transfer coefficient (KG):
∂CG, i ∂CG, i ∂2CG, i a
= vg − DDz − NG, i ∂CG, i CG, i
∂t ∂z ∂z 2 ϵ (5) = K G a ⎛⎜ − Cb, i⎞⎟ + r
∂t m
⎝ i, G , b ⎠ (8)
where, CG,i = CG,i(t,z) represents the concentration of contaminant i in
the gas phase at a given position in the axial direction (z) of the reactor where, mi,G,b represents the partition coefficient of contaminant i be-
at a certain time (t) and υg is the gas velocity. Considering that com- tween the gas phase and the biomass.
ponent i is being transferred from the bulk of the gas phase to a stagnant As it can be seen, the driving force for the transfer of the con-
layer of gas, the specific mass flux (NG,i) can be calculated using the taminant to the biofilm is determined by the difference between the
Fick's law: equilibrium concentration at the gas-biofilm interface, where the par-
tition coefficient plays a key role, and the concentration of the pollutant
Cb, i in the biofilm, controlled by the catabolic activity of the microbial
NG, i = Db, i ∂ population.
∂x x=0 (6)
The diffusivities of low molecular weight contaminants in the gas
Where, Db,i is the diffusion coefficient in the biofilm, Cb,i is the phase are in the range of 1⋅10−5 m2 s−1 (Tang et al., 2015) while in
concentration of contaminant i in the biofilm, and coordinate x de- water is close to 1⋅10−9 m2 s−1 (Harms and Bosma, 1997). The diffu-
scribes the thickness of the stagnant layer of gas and x = 0 denotes the sivity within biofilms can be expected to be lower than in pure water. If
interface between the stagnant layer of gas and its bulk (Fig. 1). the biofilm is conceptualized as a cluster of cells surrounded by water
The next transfer step delivers the contaminant from the stagnant and exopolymers, then, transfer of substances within the biofilm can be
layer of gas to a layer of water in biotrickling filters, or to the biofilm in regarded as a combination of advection within voids and water chan-
biofilters. In the latter, and when bacteria are the oxidizing micro- nels permeating the biofilm and diffusion within cell clusters and the
organisms, the transport across a layer of water can be omitted if the exopolymeric matrix (Guimerà et al., 2016). Thereby, since a biofilm
biofilm is assumed to be a pseudo-homogeneous phase composed of cannot be regarded as a homogeneous system, the measured diffusion
water and microorganisms (Fig. 2.b). The mass balance of any com- coefficient is an effective diffusion coefficient (De).
ponent, considering reaction and diffusion within the biofilm according In fungal biofilters the aerial mycelium is in direct contact with the
to Fick's Law, can be written as: gas phase while the non-aerial mycelium lies within the watery biofilm.
The formation of the aerial hyphae requires decreasing water surface

1081
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Fig. 2. Conceptual model of an aerial biofilm (a) and a monolayer biofilm (b) (Vergara-Fernández et al., 2016).

Fig. 3. Model for the formation of aerial hyphae aided by hydrophobins. The model was adapted from Wösten et al. (1999). A very similar model was proposed to explain the aerial
growth of the bacteria Streptomyces coelicolor (Claessen, 2003).

tension from 72 mJ m−2 to values as low as 30 mJ m−2 (Wösten et al., 2009). These proteins, containing 100 or more amino acids (Schor
1999). To overcome the barrier imposed by the surface tension of et al., 2016; Vigueras et al., 2008), also form a hydrophobic coating on
water, the submerged hyphal cells secrete hydrophobins, see Fig. 3. the aerial hyphae, fruiting bodies and spores, allowing hyphae to co-
Hydrophobins are thus surfactant proteins (with molecular weight of lonize hydrophobic materials (Wösten et al., 1994), possibly playing a
≈10 kDa), capable to self-assembly at hydrophilic-hydrophobic inter- role in preventing desiccation (Linder et al., 2005) and facilitating the
faces and which protect the fungal aerial structures (Vigueras et al., dispersion of spores into the air (Bayry et al., 2012). Hydrophobins

1082
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

promote the solubilization of hydrophobic VOCs in biofilters by acting supported by the fact that when F. solani is grown in glucose, or other
as a coating of the aerial mycelium. The solubilizing of hydrophobic soluble carbon sources such as 1-hexanol and glycerol, the partition
VOCs can be quantified using partition coefficients. coefficients are higher than those obtained when grown in n-hexane (in
In general, models reported in the available literature are based on the gas phase).
mass transfer principles, with the main objective of determining the In another approach, Salehahmadi et al. (2012), in their mathe-
elimination capacity of fungal biofilters. In order to do this, the gas matical model for fungal biofiltration, applied the method described by
phase and biofilm transport equations are mainly considered, as well as Mackay (Mackay, 2001), in order to determine the n-hexane/biomass
their relation to diffusion conditions of the contaminants and nutrients, (K) partition coefficient. The model is based on that the organic content
dispersion (described in the previous section), partition coefficient, and of the biomass is responsible for the hydrophobic VOCs absorption,
mass transfer coefficient, among others (Kraakman et al., 2011). which in turn can be related to the octane/air (Koa) (Eq. 12) coefficient.
1
2.1.1. Partition coefficient = 4.1·10−4 [K oa yρb ]
Kbiomass (12)
In general, fungal biofilters are modeled considering a contaminant
partition coefficient in the gas phase/biomass as an air/water Henry's where, ρb is the biofilm density and y represents the organic content.
constant. Systems were this assumption was made include α-pinene
degradation using an Ophiostoma specie (Jin et al., 2006), xylene and 2.1.2. Axial dispersion
mixed bacteria and fungi (Li and Liu, 2006), hexane and Fusarium solani Axial dispersion is the deviation from the ideal plug-flow regime in
(Arriaga and Revah, 2009; Hernández-Meléndez et al., 2008; Vergara- a pipe, column or reactor. It can be caused by multiple factors, in-
Fernández et al., 2006), toluene and a fungi consortium (Dorado et al., cluding the packing geometry and size, a non-uniform distribution of
2008). Nevertheless, the partition is modified when fungal biomass is the feed stream, channeling or short-circuiting creating dead zones, and
present in the aqueous biofilm. This situation was observed by Vergara- specifically for fungal biofilters, the growth of aerial hyphae occupying
Fernández et al. (2006), by determining that n-hexane solubility can the void space (Prenafeta-Boldú et al., 2008).
increase up to 200 times with respect to the solubility of water in The correlation proposed by Edwards and Richardson (1968) was
presence of filamentous fungi. used by Vergara-Fernández et al. (2008) in the estimation of the axial
The partition coefficient (Kbiomass) of fungal biomass may be ex- dispersion coefficient in the phenomenological modeling of a fungal
pressed as proposed by Vergara-Fernández et al. (2006) and Davison biofilter treating n-hexane. Helfferich (Helfferich, 1985) proposed an
et al. (Davison and Barton, 2000): empirical correlation for the axial dispersion coefficient in a packed-bed
as a function of the bed porosity (ε), gas velocity (ug), packing diameter
Cheadspace (dp) and the specie diffusion coefficient (DA) (Eq. 13). Although this
KBiomass =
Cbiomass (9) expression does not account for the presence of bacterial or fungal
biofilms, it was nevertheless used by Salehahmadi et al. (2012) in
where, Cheadspace and Cbiomass are the concentration of the contaminant
fungal biofilters.
in the headspace and biomass, respectively.
The partition coefficient of the dried biomass can be expressed as a D = (0.45 + 0.55ϵ) DA + 0.5dP ug /ϵ (13)
function of its composition and the individual partition coefficients of
the lipid-free biomass and the lipid component by Eq. (10). Similarly, In a different approach, Prenafeta-Boldú et al. (2008) showed that
the partition coefficient for the biomass on a wet basis can be calculated the gas residence time distribution (RTD) can be determined using an
by Eq. (11) that includes the water fraction and the partition coefficient air dispersion model. The model was experimentally validated in a
in water (which correspond to the Henry Coefficient). fungal biofilter inoculated with Cladophialophora sp. using methane as
tracer-gas. The dispersion study was carried out in two laboratory scale
lipid − free
1 xlipid xbiomass biofilters packed with granular perlite and polyurethane foam. The
= +
dry
Kbiomass Klipid lipid − free
Kbiomass (10) recorded pulse RTD curves were modeled as being equivalent to a series
of CSTR (Eq.15). Two additional CSTR compartments of constant vo-
1
dry
xbiomass x water lume were added to account for the void spaces present at the top and
wet
= dry
+ at the bottom of the packing. Then, the following equations represent
Kbiomass Kbiomass Kwater (11)
the proposed CSTR system for the estimation of the RTD:
where, x represents the mass fractions of water, lipids or biomass; and K dM0 Q
represents the partition coefficient in biomass, water and lipids. = − M0
dt Vin (14)
Vergara-Fernández et al. (2011b; 2006), obtained biomass partition
coefficient values with Fusarium solani in the range of 0.033–0.04 for n- dMi n
= (Mi − 1 − Mi ), i = 1, 2, …, n with Mi − 1 = 0 if i = 1
hexane and 0.015–0.032 for n-pentane, respectively, which are around dt τ (15)
three orders of magnitude lower as compared to water. Furthermore,
dMn + 1 Q
Vergara-Fernandez et al. (2011b), determined the n-pentane/dry-bio- = (Mn − Mn + 1)
dt Vout (16)
mass partition coefficient in microcosms for the same fungus when
grown in four packing materials (compost, peat, perlite and vermicu- where Mi is the contaminant gas concentration in tank i and τ is the
lite) at different temperatures (15 °C, 25 °C and 35 °C). Results indicated mean residence time of the tracer gas equivalent to that of the flowing
that the n-pentane/wet-biomass partition coefficients in organic air, the effective bed void volume can be calculated as:
packing material were 160-fold lower (0.21 ± 0.09) than those in
Ve = τ ·Q (17)
water (33.2 ± 9.4), while for inorganic packing material they were
700-fold lower (0.05 ± 0.04). The authors attributed this result to the Spigno and De Marco Faveri (2005) analyzed the RTD in a biofilter
fact that, when grown over an inorganic packing material, the growth model of a fungal biofilter inoculated with Aspergillus niger for n-hexane
of F. solani is supported only by n-pentane in the gas phase and not from abatement. They found a small axial dispersion coefficient of
the carbonaceous material present in the organic packing material. 1.22·10−4 m2 s−1.
Thus, an increased hydrophobicity of the surface of the mycelium,
caused by the presence of hydrophobins, would result in a more effi- 2.1.3. Diffusion coefficient in the biofilm
cient mass transfer of n-pentane from the gas phase. The role of hy- The main parameters affecting the diffusion of a contaminant in the
drophobins in the biofiltration of slightly soluble contaminants is fungal biofilm are the temperature and the biomass density, where the

1083
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

diffusion coefficient for biofilms is found to be slightly lower than the Above, Vr is the reactor total volume, VE is the total volume of the
one for water. However, this parameter should be considered only for support, dhyphae is the hyphae diameter obtained from morphological
the cases in which the biomass attached to the support is modeled (not observations and Lhyphae,Total is the total length of the hyphae which can
the aerial biomass), since it is customary to consider that the con- be calculated as shown in Eq. (26).
taminant reacts instantaneously once it is in contact with the surface of An important aspect to be considered when simulating the pressure
the aerial hyphae (Vergara-Fernández et al., 2016, 2008). drop in a fungal biofilter at the initial stages of operation, is the amount
The diffusion coefficient for fungal biofilters, might be approxi- of static liquid present in the bed. This behavior can generate a de-
mated by applying the empirical correlation of Wilke and Chang viation of up to 11% from experimental measurements when it is not
(1955). This method was used by Spigno and De Marco Faveri (2005) accounted for in the mathematical model (Vergara-Fernández et al.,
by assuming the value of the diffusion coefficient to be 40% of its value 2008). The maximum pressure drop obtained in the latter study goes up
in water, with a resulting estimated diffusion coefficient of 0.74 to about 45 mmH2O m−1 bed. On the other hand, Zamir et al. (2011a,
×10−9 m2 s−1. Instead, Arriaga and Revah (2009) applied the ex- 2011b) obtained a pressure drop approximately four times smaller
pression for impermeable sphere-packed porous media by Neale and (11 mmH2O m−1 bed), than those obtained by Vergara-Fernández et al.
Nader (1973) in a mathematical model of hexane degradation in fungal (2008) using a gas flow three times lower and half the residence time in
biofilters, obtaining the following effective diffusivity for hexane (Eq. the toluene elimination in a fungal biofilter inoculated with the fungus
18): Phanerochaete chrysosporium.
Def 2ϵ
= 2.3. Drying, metabolic activity and heat generation
DHwater 3−ϵ (18)

where, Def is the effective diffusivity in the biofilm, ε is the bed void Packing and biofilm moisture content, metabolic activity and heat
fraction, and DHwater is the diffusivity in water. The bed void fraction generation are intimately intertwined. Moisture balance is affected by
can be obtained according to: thermodynamic phenomena, the equilibrium between the content of
water in the biofilm and the moisture in the gas stream, the water
Vwater
ϵ= transfer rate from the solid support to the biofilm (in biofilters).
Vtotal (19) Humidification of the contaminated air stream entering the biofilter is a
where, Vwater is the volume occupied by the water in a total volume common practice to avoid biofilm drying, incomplete humidification of
(Vtotal) packed with a support material. The authors reported a Def/ the air stream (relative humidity < 100%) evaporates moisture from
DHwater of 0.35 for bacterial biofilms, and 0.19 for fungal biofilms, with the solid support and biofilm at a rate that depends on inlet tempera-
effective diffusivity in fungi of 4.1×10−7 m2 h−1. ture, initial relative humidity and empty bed residence time. The heat
liberated by the oxidation of gaseous pollutants, especially those with
high upper heating values such as alcohols, alkanes and alkenes, in-
2.2. Pressure drop
creases the evaporation rate. van Lith et al. (1997) estimated that a
biofilter operated with inlet contaminant concentration of 0.5 g m−3
The pressure drop in fungal biofilters is one of the most important
and volumetric load of 50 g m−3 h−1 loses 30 L of water per cubic
operational parameters to be considered. The increase in the pressure
meter of packing in a period of 3 to 6 days. This drying mechanism is
drop can be caused by several factors including entrapment of particles
especially relevant near the biofilter influent where the highest VOCs
transported in the air entering to the reactor, bacterial growth and by
concentrations are found. In this regard, Gostomski et al. (1997) pro-
the aerial growth of filamentous fungi, diminishing the void fraction,
vide experimental results showing that the degradation caused by heat
and therefore reducing the permeability of the porous media of the
generation and water evaporation started in a narrow zone at the inlet
bioreactor. Since air entering to biofilters is commonly prefiltered and
of the bed, however, as the loss of water inhibited the microbial activity
humidified, the load of particles that can cause cloging of the reactor is
of this area, the degradation region progressed through the bed.
reduced, along with the concentration of highly water soluble VOCs.
Insufficient moisture prevents the formation of either bacterial or
Estrada et al. (2013) found pressure drop increments for fungal (Pae-
fungal wet biofilms, that can support microbial growth and respiration,
cilomyces variotti) and bacterial biofilters of 91 to 912 Pa m−1
bed and 91 to
promotes the contraction of the support and the consequent medium
372 respectively, for the treatment of a VOC mixture, showing how
cracking reducing retention times (Swanson and Loehr, 1997). On the
fungi can introduce significant resistance to the moving flow compared
other hand, excessive moisture reduces mass transfer rates of hydro-
to bacteria. Similar values, between 98 and 350 Pa m−1 bed, were obtained
phobic substances, reduces biofilm surface area by clogging available
Prenafeta-Boldú et al. (2008) in a fungal biofilter inoculated with Cla-
pore space, increases pressure drop (Ferdowsi et al., 2017; van Lith
dophialophora sp. strain CBS 110553 for the removal of toluene using
et al., 1997) and creates anaerobic zones that promote odor formation
perlite granules as packing material.
and low degradation rates (Ferdowsi et al., 2017; Swanson and Loehr,
The pressure drop evolution over time in a fungal biofilter was
1997).
evaluated by Vergara-Fernández et al. (2008) by applying the Darcy
The moisture content of the biofilter support, both for fungal and
equation (Eq. 20).
bacterial biofilters, is one of the main parameters affecting reactor
ΔP 72ω (1 − ϵ)2 ⎞ ⎛ μf ⎞ performance. The control of the moisture content requires under-
⎛ ⎞ = ⎛⎜
⎝H ⎠ ⎝ ϵ 2dp2
⎟⎜ ⎟ vg standing the dynamics of drying of the biologically active support due
⎠ ⎝ ρg g ⎠ (20)
to the air temperature changes at the inlet, the air relative humidity,
where, ΔP is the pressure drop, H is the biofilters height, ϖ is the tor- and the metabolic heat produced by the oxidation of VOCs. Morales
tuosity factor, ε is the bed porosity in the biofilters, dp is the particle et al. (2003) developed a mathematical model to study the drying effect
diameter, μf dynamic viscosity of the fluid, ρg is the gas density, g is the on a biofilter inoculated with five different types of bacteria and two
acceleration of gravity, and vg is the gas superficial velocity. types of yeast used for the elimination of toluene. The model considers
Where the void fraction can be evaluated as a function of fungal the effect of temperature, water content, and contaminant concentra-
growth by application of Eq. (21): tion in the biological reaction. The latter model allows the prediction of
the biofilter performance and water evaporation from the support

ϵ(t ) =
Vr − ⎡VE +
⎣ ( )d
π
4
2
hyphae Lhyphae, Total (t ) ⎤

media because of the metabolic heat generation and the variation of the
inlet air current relative humidity. The water vapor in the gaseous
Vr (21)
phase (air absolute humidity) can be estimated from Eq. (22):

1084
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

∂XH F ∂XH −RV ρC 2008; Spigno and De Marco Faveri, 2005). In contrast, studies carried
= +
∂t ρa ηg ∂Z ηg (22) out by Vergara-Fernández et al. (2016, 2008) show that the aerial
growth of filamentous fungi plays an important role in the VOCs bio-
where XH is the absolute humidity of air, F is the air mass flow, ηg is the degradation. This is mostly due to the increase of transport area, which
ratio of gas-phase volume to elementary representative volume, ρa is the can reach values up to 107 m2 m−3, in comparison to 102 m2 m−3 for
real density of air, ρc is the apparent density of the packing material, non-aerial biofilm (Arriaga and Revah, 2009).
and RV is the water evaporation rate based on dry support. Vergara-Fernández et al. (2016) developed a simple mathematical
When analyzing the energy balance in the biofilter, the convective model with the objective of describing the contribution of aerial and
heat must be considered due to its importance in: the biofilter drying, non-aerial biodegradation of n-pentane using Fusarium solani. They
the water evaporation and the reaction heat, as stated in Eq. (23): found that, although the aerial mycelium represents only 23% of the
∂T biomass on a dry weight basis, is responsible for 71.6% of n-pentane
β = Qconv + Qevap − Qrx degradation. This result highlights the key role of the aerial hyphae on
∂t (23)
hydrophobic VOCs degradation. In this study, together with mass bal-
where, β is a parameter associated to the accumulation of heat and it ances in the gas phase and the bioreactor biofilm, the mass transfer
represents the average heat capacity of the medium, Qconv represents surface area (av) was determined, according to Eq. (25), where the
the convective volumetric heat, Qevap is the evaporative volumetric biofilm area (aa) can be determined as reported by Arriaga and Revah
heat, and Qrx depicts the volumetric heat associated to VOCs degrada- (2009), considering a total hyphal length, Lhyphae (Eq. 25).
tion and T is the operation temperature. The combustion heat from the
reaction can be estimated from each VOCs combustion enthalpy. The a v = am + aa ·α (25)
CO2 production rate because of the VOC mineralization can be used as a where, α denotes the fraction of the aerial biomass. Hence, the super-
correction factor of the combustion enthalpy. ficial area available for VOCs transfer can be defined as:
Water content can affect also other important aspects of fungal
Vbiomass
biofilters such as spore production, emission and retention. The effect of Lhyphae =
dh yphae 2
biofilter water content on spores generation was studied by Vergara-
Fernández et al. (2012b) and Saucedo-Lucero et al. (2014) operating
π ( 2 ) (26)

fungal biofilters for the elimination of n-pentane and n-hexane, re- where, Lhyphae is the hyphae length, Vbiomass is the biomass volume and
spectively. The number of colony forming units (CFU m−3 air ) in the PDA dhyphae is the hyphae diameter.
Petri dish according to Eq. (24), was used for the determination of the The total fungal interfacial area per reactor volume, aa, is:
spore concentration (SC):
dhyphae Lhyphae π
CFU aa =
SC = Vr (27)
Q·tm (24)
where, Vr is the reactor volume.
Where, tm is the elapsed time for spore collection and Q the filtered gas A similar approach was applied by Spigno and De Marco Faveri
flow rate. (2005), to establish the amount of aerial biofilm in the mass balance of
The results show that a low water content in the biofilter promote the gaseous phase of a fungal biofilter. This was carried by sensibility
sporulation, while a higher water content decreases sporulation and analysis of the variation of the contaminants concentration throughout
favors spore retention. Vergara-Fernandez et al. (2012), concluded that the biofilter for values of α between 0.25 and 1.0.
the periodic addition of mineral medium in biofilters based on in- The specific transport area obtained from several simulations reach
organic materials promotes rapid recovery of the biodegradation per- values around 2×105 m2 m−3, considering hyphae diameters between
formance and reduces spore emission. However, an excess of irrigation 2.1 μm and 2.9 μm (Arriaga and Revah, 2005; Vergara-Fernández et al.,
can increase the pressure drop across the biofilter bed and hinder hy- 2008). These values are between 50 and 130 times larger than those
drophobic VOCs transfer from the gas phase, with a subsequent de- obtained for bacterial biofilters (Iliuta and Larachi, 2004; Zhu et al.,
crease in EC. Interestingly, the increase in VOCs inlet load in a non-N- 2004).
limited scenario triggered spore emission and increased the EC, with
average spore counts of 1.8 ×104 CFU m−3 −3
air and ECmax of 110 g mreactor 2.4.2. Kinetics
−1
h . Finally, lower empty bed residence times (EBRTs) can support an Biofilter performance is ultimately determined by the catabolic ac-
increase in the emission of spores and a deterioration in the EC. Similar tivity of the microbial biomass which, in turn, depends both on the
results, between 2.4×103 to 9.0×104 CFU m−3, were reported by microbial mass present and the intrinsic activity of this biomass. The
Saucedo-Lucero et al. (2014) in the hexane abatement in a fungal amount of biomass results from growth and decay rates and the in-
biofilter inoculated with a fungal consortium, obtaining a EC of trinsic activity on the identity of the microbial population (genotype),
35 g m−3 h−1. how it is expressed (phenotype) and the reactor conditions (pH, tem-
perature, water content, presence of inhibitors, etc.). Substrate uptake
2.4. Biomass growth and biodegradation kinetics is utilized for biomass growth when conditions are appropriate using
the chemical energy generated by the catabolic reactions and producing
Though the main problem in biofilters relates to the description of CO2.
the mass and momentum transfer phenomena between the gas, the li- Modeling fungi growth decay has received less interest due to the
quid and the biofilm; the knowledge of the intrinsic microbial growth complex processes involved in the growth of filamentous fungi. The
and biodegradation rates is a key issue in the design and optimization of nature in which fungi grow, makes it troublesome to define cellular
this type of bioreactors. Coupling both phenomena can be done con- death when compared to unicellular organisms. Szewczyk and Myszka
sidering a micro approximation in order to determine the intrinsic (1994) applied an Arrhenius type of equation to represent specific
biodegradation kinetics (Iranmanesh et al., 2015). growth rate as a function of temperature, where there is a term de-
scribing growth and another describing cellular death. Smits et al.
2.4.1. Fungal biofilm (1999) did not develop an explicit model for cellular death, yet they
Fungal biofilm modeling in biofilters has been scarcely studied. In proposed an equation describing specific respiration activity of the
general it is considered as a flat biofilm with constant thickness microbial mass decay over time.
(Fig. 2b), within the range of several μm to > 100 μm (Dorado et al., In the other hand, hyphae elongation of the first hyphae, during the

1085
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

germination of one spore, has been represented by many authors as where, V is the reactor volume; Q is the gas flow; Cgi and Cgo are the
Monod type kinetics (Aynsley et al., 1990; Carlsen et al., 2000; inlet and outlet contaminant concentration, respectively; KS/X corre-
Larralde-Corona et al., 1997). This representation has been applied to sponds to the saturation constant, rmax is the maximum rate of biode-
determine time and branching frequency. Vergara-Fernandez et al. gradation and CLn is the concentration of a contaminant in the liquid.
(Vergara-Fernandez et al., 2011), considered the principles proposed in In the phenomenological model simulation of a fungal biofilter done
the Larralde-Corona et al. (1997) and López-Isunza et al. (1997) by Vergara-Fernández et al. (2008), it was found that there is a certain
models, to develop an individual hyphae growth model taking into degree of energetic decoupling throughout the start-up period. This
account the primary hypha and same diameter branching, with a cy- situation was verified by Vergara-Fernandez et al. (2014) in a following
lindrical shape and constant density. The total biomass and the total study. In the literature is possible to find several mathematical models
hyphae length were determined with the initially added spores, con- applied to biofiltration systems inoculated with bacteria and fila-
sidering that each spore germinates forming a sole primary hypha with mentous fungi, which have used the yield parameter as a constant
many branches. (Arriaga and Revah, 2009; Dorado et al., 2008; Spigno and De Marco
As important as knowing the rate of biomass growth, is to establish Faveri, 2005; Vergara-Fernández et al., 2008). The latter occurs be-
biodegradation rate of the contaminants involved in the biofiltration. cause the models applied only consider the yield parameter as con-
This is because the kinetic study of the degradation of the contaminants sumption of substrate (VOCs) for the biomass growth, and therefore
is contemplated as one of the main steps in the characterization of the neglecting cellular maintenance, product formation, and heat loss,
biofilter operation. It is important to consider that normally VOCs among others. Nevertheless, the presence of filamentous fungi increases
biodegradation studies for biofiltration are carried out as batch culture the solubility of hydrophobic VOCs, being able to generate a substrate
(microcosms), where the adsorption phenomena in the support, the excess with respect to the present nutrients, causing decoupling be-
pore diffusion and the gas dispersion can be considered negligible tween anabolism and catabolism leading to energy spilling. Vergara-
(Delhoménie et al., 2008; Iranmanesh et al., 2015; Vergara-Fernández Fernandez et al. (2014) suggested a model that allows to establish the
et al., 2006). The expression used for fungal microcosms with gaseous effect of the ratio of initial substrate concentration to biomass in the
substrate for the determination of the relation between the COV con- yield and the energetic decoupling coefficient (Eqs. 31 and 32).
sumption rate and the specific growth rate is shown in Eq. (28):
1 1 1 Cbi0/ X0
= +
dCG 1 dX Yobs (Yobs )max (YW )min Cbi0
=− + mX X0
+ KS/X (31)
dt YX / S dt (28)

where, CG is the limiting contaminant concentration, t is time, Yx/s is (Yobs )max − Yobs
Eu =
the yield coefficient and m is the maintenance coefficient, and X is the (Yobs )max (32)
fungal biomass. Vergara-Fernandez et al. (2011b) apply the same re- where Yobs is the observed cell yield, (Yobs)max denotes the observed
lation to determining the biodegradation of n-pentane, howbeit ig- growth yield of substrate-limited culture, (YW)min is the minimum en-
noring the maintenance coefficient. The maintenance coefficient can be ergy spilling-related cell yield, KS/X corresponds to the saturation con-
expressed as: stant, C i b0 depicts the initial VOCs concentration in wet biomass, ×0 is
[dCG / dt ]m initial biomass concentration, and Eu is the energy uncoupling coeffi-
m= cient.
X (29)
Results show that Yobs in the gaseous phase is inversely proportional
Vergara-Fernández et al. (2006) established that, under the micro- to the C i b0/ X0 ratio, and that over 60% of VOCs was consumed due to
cosms conditions, in the n-hexane fungal biodegradation, the main- the loss of energy, and a strong dissociation of catabolism and anabo-
tenance was approximately 2% of the total removal. This confirms the lism takes place for high C i b0/ X0 ratios.
importance of the presence of nutrients in biofilters to sustain high
uptake rates. 2.4.2.1. Temperature effect in the fungal biodegradation kinetics. As
The kinetic equations employed in biofiltration studies are nu- aforementioned, temperature effect in fungal growth is one of the
merous. From these the most commonly used for the elimination of main environmental outcomes considered in fungal biofiltration,
hydrophobic VOCs in fungal biofiltration systems are Haldane type however it is rarely considered in models due to its complexity.
models of inhibition by substrate, and the Monod type models. Different heat and mass transfer models acknowledge this effect over
Furthermore, a series of other expressions have been used to model the the increment of the partition coefficient (lower solubility) and bed
kinetics of biodegradation in fungal biofilters, which are summarized in drying (Salehahmadi et al., 2012). On the other hand, some models
Table 1. have incorporated the effect of temperature in the kinetics of
Kinetic studies for the biodegradation of gaseous compounds in contaminants biodegradation, as reported by Jin et al. (2007) and
biofiltration systems can be investigated from micro-kinetics (as shown Salehahmadi et al. (2012), assuming an Arrhenius type equation and an
above) or macro-kinetics. Nonetheless, the micro-kinetics methodology optimal biodegradation temperature, where an increment in
cannot be extended, in some cases, to systems in gaseous phase (Zamir temperatures below the optimal temperature increases the
et al., 2011a, 2011b). The latter happens because the biodegradation biodegradation while decreasing the reaction rate (Eq. 33).
and the transport phenomena involved in a liquid media system are not
comparable to those in solid media with biofilm. The macro-kinetic Ea
r (T ) = r (Topt ) exp ⎡ (T − Topt ) ⎤
method, for the determination of different kinetic parameters in fungal ⎢ RTTopt ⎥
⎣ ⎦ (33)
biofilters, has been applied by numerous authors (Mathur et al., 2006;
Spigno and De Marco Faveri, 2005; Zamir et al., 2011a, 2011b). The where, r(T) is the biological reaction rate, T denotes operation
different kinetic constants are found solving and linearizing the equa- temperature, Topt represents the optimal temperature, R depicts the
tion of change of concentration in the gaseous phase in the biofilter and ideal gas constant and Ea represents the activation energy of biological
the applied kinetic model. For example, for the case of Monod type reaction.
kinetics Eq. (30) is obtained (Spigno and De Marco Faveri, 2005; Zamir The ratio Ea/(RTTopt) in Eq. (33) can be assumed as constant for the
et al., 2011a, 2011b). operation range in biological processes (Jin et al., 2007). These authors
observed, during the biofiltration of α-pinene in a fungal bioreactor, an
V /Q KS/X 1 1
= + increased in removal efficiency when the operating temperature was
Cgi − Cg 0 rmax CLn rmax (30) increased from 20 to 30 °C, from 60% to 100% respectively, for an inlet

1086
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Table 1
Kinematic models used in VOCs biodegradation of fungal biofilters.

Type Model Contaminant Type of fungi Reference

Haldane μ = μmax
S n-hexane Fusarium solani (Vergara-Fernández et al., 2006)
S2
S + KS +
KI
Monod μ = μmax
S Toluene n-pentane Bacterial/fungal consortium Fusarium solani (Dorado et al., 2008)
S + KS
(Vergara-Fernández et al., 2016)
Monod-Haldane n-hexane; nitrogen source Fusarium solani (Vergara-Fernández et al., 2008)
⎡ ⎤
μ = μmax ⎢
S ⎥ ⎡ SN ⎤
⎢ S2 ⎥ ⎣ SN + KN ⎦
⎢ S + K S +
⎣ KI ⎥

Webb μ = μmax
S (1 + S / K ) n-hexane Fungal consortium (Iranmanesh et al., 2015)
S2
S + KS +
KI
Yano μ = μmax
S n-hexane Fungal consortium (Iranmanesh et al., 2015)
S2
S + KS +
KI (1 + S / K )
Teissier n-hexane Fungal consortium (Iranmanesh et al., 2015)

μ = μmax ⎡exp

( ) − exp ( ) ⎤⎦

S
KI

S
KS

Aiba S n-hexane Fungal consortium (Iranmanesh et al., 2015)


S ⎛⎜exp ⎛− ⎞ ⎞⎟
⎜ ⎟

⎝ ⎝ KI ⎠ ⎠
μ = μmax
S + KS
Andrews μ = μmax
S n-hexane Aspergillus niger (Spigno and De Marco Faveri, 2005)
S⎞
(S + K S ) ⎛1 +
⎜ ⎟

⎝ KI ⎠
Exponential μ = μmax
S n-hexane Fungal consortium (Iranmanesh et al., 2015)
S
(S + K S ) exp ⎜⎛− ⎟⎞
⎝ KI ⎠

concentration of 150 ppm. However, an opposite trend was observed at (2008) found that the toluene elimination capacity in a biofilter in-
temperatures between 30 and 40 °C. A similar effect was observed by oculated with species of Pseudomonas decreases as the inlet load of
Vergara-Fernández et al. (2012b) in the biofiltration of n-pentane with hydrogen sulfide increases. On the other hand, Woertz et al. (2001)
the fungus Fusarium solani, obtaining an increase in the ECmax from 45 found that toluene was consumed as a source of carbon and reducing
to 65 g m−3 h−1 when the temperature increased from 15 °C to 25 °C, power in a fungal biofilter for the oxidation of nitric oxide, with a
however a decrease to 60 g m−3 h−1 was obtained when the tempera- maximum removal efficiency of 93%.
ture was increased to 35 °C.

3. Modeling of VOCs elimination in fungal biofilters


2.4.2.2. Biodegradation kinetics of a mixture of contaminants. The
presence of more than a sole contaminant in different emission
The following paragraphs illustrate the evolution in the models
sources is common, hence the study of fungal biofiltration under the
describing fungal biofilters, explicitly stating the physicochemical and
aforementioned condition is mandatory. Iranmanesh et al. (2015)
biological phenomena considering in the models. Fig. 4 displays this
studied the effect of the presence of toluene in the biodegradation of
information in a graphical fashion. Among the first models for fixed bed
n-hexane using an undefined fungi consortium. The results were fitted
biofilters, Ottengraf and Van Den Oever (1983) considered steady-state
into a Haldane equation for biodegradation of both substrates (Eq. 34).
operation, axial plug flow, diffusion of the pollutant in the biofilm and
μmax ,1 CG,1 μmax ,2 CG,2 zero or first-order kinetics. Shareefdeen and Baltzis (1994) developed
μ= +
CG2 ,1 CG2 ,2 an unsteady-state model of a biofilter treating toluene vapors in air
KX / S,1 + CG,1 + KI ,1
+ I2,1 CG,2 KX / S,2 + CG,2 + KI ,2
+ I1,2 CG,1
considering mass balances in the bacterial biofilm, gas phase and solid
(34) support. The reactions performed by the bacteria in the biofilm were
mathematically described using Monod kinetics.
where μ is the specific rate of growth, μmax is the specific rate of
Deshusses et al. (1995) later proposed a model for the degradation
maximum growth, CG is the contaminant concentration, KX/S is the
of methyl ethyl ketone (MEK) and methyl isobutyl ketone (MIBK) in a
constant of affinity contaminant, KI is the inhibition constant and Ii,j
biofilter. The microbial kinetics consider a Monod-type expression in-
indicates the degree of interaction of substrate i affecting the
corporating competitive inhibition towards the MEK and MIBK mixture.
biodegradation of substrate j (high values indicate higher degree of
A different approach was adopted by Hodge and Devinny (1995), who
inhibition). On the other hand, Vergara-Fernández et al. (2008) used a
propose a model describing transfer between the air and the biotic
coupled kinetic model of Haldane for the contaminant and a Monod
solids/water phases, biological substrate degradation, CO2 production
model for the use of nutrients (nitrogen source) (Eq. 35), to establish
and accumulation and pH changes resulting from CO2 accumulation. A
the rate of growth of the filamentous fungus Fusarium solani in the n-
mathematical model describing the biofiltration of a mixture of hy-
hexane biofiltration.
drophilic and hydrophobic compounds was developed by Mohseni and
CG ⎤ ⎡ CN Allen (2000). Based on the work of Ottengraf and Van Den Oever
μ = μmax ⎡ ⎤
⎢ Cg + KX / S + C 2/ KI ⎥ ⎢ CN + KN ⎥ (1983) for a single COV, the steady-state model considers the biofilm as
⎣ G ⎦⎣ ⎦ (35)
an organic matrix and uses Monod kinetics with inhibition.
where, CN is the nutrient concentration and KN is the nutrient affinity Although mass transfer and biodegradation kinetics are at the core
constant. of most biofilter models, other aspects of the biofilter operation have
Finally, the kinetics of microbial growth and pollutants metaboli- been considered. Morales et al. (2003) included in their model the
zation can be influenced by inorganic gases, such as ammonia, hy- drying and loss of activity of the humid solid support (peat) induced by
drogen sulfide or nitrogen oxides, with effects over VOCs degradation the metabolic heat generated by the degradation of the gaseous sub-
ranging from beneficial to detrimental. For example, Galera et al. strate. Iliuta and Larachi (2004) addressed the problem of extensive

1087
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Fig. 4. Summary of models to describe the biofiltration of hydrophobic VOCs using fungi (except for the model of Ottengraf and Van Den Oever, 1983) and their assumptions.

biomass growth and clogging considering a unidirectional dynamic of intra and inter-particle phenomena occurring in the biofilter. As the
flow model based on the volume-average mass, momentum and species aerial growth in filamentous microorganisms develops, it occupies the
balance equations coupled with conventional diffusion/reaction equa- empty void between the solid support particles in the biofilter.
tions describing apparent kinetics in the biofilm. Filamentous growth in fungal biofilters favors the transport from the
Several models have been reported describing fungal biofilter in gas to the biotic phase by both the increase in the exchange surface, due
contrast to those models described in the previous paragraphs, where to hyphal growth, and by increased solubility of the hydrophobic VOCs
no distinction is made as which are the agents (bacteria, yeast or fungi) due to the hydrophobic nature of the fungal surface (Vergara-Fernández
of the biological activity. Spigno et al. (2003) and Spigno and De Marco et al., 2006).
Faveri (2005) proposed a model for the elimination of n-hexane by The diffusive transport of the contaminant across the biofilm cannot
Aspergillus niger based on a steady-state description considering axial be analyzed using a model based in an aerial biofilm. In such model,
dispersion in the gas phase. However, the model treats the biofilm using there is no homogeneous film where diffusion can occur, thereby it is
the same assumptions involved in microbial consortiums or bacterial assumed that the contaminant is in direct contact with the hypha of the
biofilters modeling, considering a homogeneous and constant fungal filamentous fungi (Vergara-Fernandez et al., 2012).
biofilm where the reactions take place in the so called “effective bio- Vergara-Fernández et al. (2008) developed a mathematical model
layer”, where the pollutants diffuse. Both models were developed using considering the physical and biological phenomena in a fungal biofilter.
a steady state assumption. The biofilter is mathematically described and the main physical, kinetic
An improvement including actual partition coefficient in the fungi, data and morphological parameters of aerial hyphae were obtained by
biofilm thickness, superficial area and effective diffusivity was pre- independent experiments for model verification. The model proposed
sented by Arriaga and Revah (2009). The model is based on the work by describes the increase in the transport area by the growth of the fila-
Ottengraf and Van Den Oever (1983) modified by considering that a mentous cylindrical mycelia and its relation with n-hexane elimination
fraction of the biomass is active in degrading the substrate, a fraction in quasi-stationary state in a biofilter. In order to mathematically de-
controlled by the diffusion rate within the biofilm. Another modifica- scribe the system, four processes were considered: (1) mass transfer of
tion from the original model is that zero-order kinetic was assumed. VOCs in the bulk gas, (2) mass transfer of VOCs into the gas layer
Dorado et al. (2008) developed a dynamic model describing the around the mycelium considering the experimentally obtained gas-
elimination of toluene in a bacterial-fungal biofilter, the model was biomass partition coefficient, (3) mass transfer and uptake of the ni-
calibrated and validated against experimental data. The model is based trogen source through the elongating mycelia, (4) and the kinetic of
on mass balances, including convection, absorption, diffusion and mycelial growth. The model describing fungal growth includes Monod-
biodegradation. The biofilter was initially inoculated with a bacterial Haldane kinetic and hyphal elongation and ramification. A maximum
consortium, and after 60 days of operation, the biofilter was colonized EC of 250 g m−3 h−1 was predicted in simulations and experimentally
by fungi as confirmed by a change in the reactor pH and microscopy. validated.
In general, previous models were developed assuming the biofilm as Recently, Vergara-Fernández et al. (2016) developed a mathema-
a pseudo-homogeneous phase. However, in the case of aerial biofilms, tical model predicting the influence of the fraction of aerial biomass
such as those generated by the hyphae of filamentous fungi, those and its partition coefficient on the biodegradation of n-pentane by the
pseudo-homogeneous phase models do not provide all the information filamentous fungi Fusarium solani in a fixed-bed continuous biofilter,

1088
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

obtaining an EC close to 680 g m−3 h−1. the ratio of the advection to diffusion rates, can be effective in ana-
lyzing the interplay of mass transfer and biodegradation kinetics.
3.1. Artificial neural networks models in fungal biofilters Different authors have used this method to identify whether the op-
eration of the reactor was mass transfer or kinetically limited (Vergara-
As discussed previously, several phenomenological mathematical Fernández et al., 2008; Xi et al., 2015).
models have been developed to describe the performance of fungal In air biopurifications systems, such as biofilters and biotrickling
biofilter systems. The evolution of these models has led to cumbersome filters, the pollutant degrading microorganisms are attached to a
sets of non-lineal equations of momentum, heat and mass transfer. packing media. The packing media can be conformed by different types
Furthermore, these models require the experimental determination of a of porous materials such as compost, mineral grains and artificial foams
series of parameters, such as kinetic constants, constants of pollutant (Kim et al., 2007; Yang et al., 2011; Znad et al., 2007). The con-
diffusion in the biofilm, oxygen consumption and carbon dioxide pro- taminated air flows through this media which translate into direct
duction, heat and mass transfer coefficients, among others. A solution contact of air with the attached biomass (biofilters), or diffusion of the
to the aforementioned set of equations, lays in the use of a non-linear contaminant in a flowing liquid in which the biomass is immersed
modeling method that focuses on the simulation and prediction of the (biotrickling filters). There, the biomass degrades the contaminant
biofilter performance, based exclusively on available data (Zamir et al., which is converted to innocuous products and more biomass (Moe and
2011a, 2011b). One of these alternatives, suggested in a recent pub- Irvine, 2001). The effectivity of the latter process depends, among
lication, is to use artificial neural networks to model biological pro- others, on how much of the contaminant is bioavailable for the de-
cesses. Chairez et al. (2009), proposed a model using continuous neural grading microorganisms (Cheng et al., 2016), either from the bulk gas
networks with the aim to design and predict the behavior and elim- to the biofilm, or from the gas phase to liquid phase and from there to
ination capacity of a fungal biofilter (Paecilomyces variotti) for toluene the biofilm.
abatement, as an alternative to the complexity and simplification as- The porous media accommodates in a biofilter generating void
sumptions of phenomenological models. This model concludes that the spaces or pores, as well as porous channels that are irregular in shape
application of neural networks in this type of problems reduces the cost (Devinny and Ramesh, 2005). Porosity and porous channel configura-
of the biofilter performance due to the reduction in the number of re- tion plays a major role on residence time, pressure drop and pre-
quired sensors. ferential flow, and therefore affecting the overall performance of the
Additionally, Spigno and Tronci (2015) developed a model for the biofilter. And this is only the tip of the iceberg, since the growth of
elimination of hexane in a fungal biofilter inoculated with Aspergillus biofilm over time may generally lead to changes in all previous char-
niger, based on hybrid models, obtained by the connection of the mass acteristics (Devinny and Ramesh, 2005). Furthermore, depending on
balance equation with a feedforward neural network, used as a black- the pores size, for the case of biotrickling filters, the flow may be
box model of the reaction and transport phenomena. A similar ap- subjected to capillarity effects.
proach for the development of an experimental model for the analysis From the previous description, it seems relevant to study in depth
of a compost biofilter treating n-hexane was developed by Zamir et al. the dynamics of the air flowing through the porous media selected for
(2011a, 2011b), over a wide range of inlet loads (between 8 and the biopurification treatment system, since it might affect the actual
598 g m−3 h−1). The neural model showed the ability to predict the treatment. In other words, how does the selected porous bed will en-
performance of biofilter under intermittent operating conditions. The hance pressure drop, boost preferential flow increasing death zones and
results suggested that neural network can be considered as a suitable therefore reducing bioavailability in the biofilter in average, and finally
alternative to conventional knowledge-based models; however, the decreasing residence time. In this regard, computational fluid dynamic
sudden changes in the value of removal efficiency have an impact on studies (computational simulations) coupled with a realistic re-
the neural networks training/generalization pattern. presentation of the media may be of great help. This can be done by
discretizing the porous volume within the biofilter in computational
3.2. Modeling for design and operational optimization cells, and solving the flow of air passing through it, by means of the
Navier-Stokes equation, with the corresponding boundary conditions.
Optimization of the operation of a biofiltration system requires Through this type of studies, it might be possible to understand at the
knowledge of the rate-limiting steps controlling the efficiency of the microscale level the complex physical phenomena involved in the in-
system. Depending on system's setup and operational parameters, bio- teraction between the fluid and the porous structure. Validation can be
filters can be operated under either mass transfer or kinetically limited achieved by comparison of pressure drop and flow velocity with ex-
conditions. Several strategies can be used to determine the rate-limiting perimental studies. Since flow velocities are usually very low, and be-
step. In biotrickling filters, bioscrubbers and bubble tank reactors a cause of the practical limitations in measuring flow velocity at the pore,
straightforward strategy is to increase the biomass concentration in the a residence time distribution (RTD) study can be carried out, and
liquid phase. If a change in the amount of pollutant removed per reactor compared to the RTD simulation results. This is in a way, and indirect
volume is achieved when biomass concentration is increased, then the measurement of the velocity distribution in the biofilter. Once vali-
system is kinetically limited (Kraakman et al., 2011). However, this dated, important properties of the porous structure can be assessed,
procedure is not feasible in a biofilter. Barton et al. (1999) and such as permeability (in the flow and transverse directions), pre-
Iranmanesh et al. (2015) suggested a technique to distinguish between ferential flow (by defining a threshold velocity), etc. Moreover, for the
kinetically o mass transfer limitation by changing the operating tem- case of trickling filters a two-phase (liquid and air) computational si-
perature of the system. An important increase in removal efficiency as mulation may deliver a better understanding of the combination of flow
temperature increases measured at different gas velocities is indicative conditions that enhance contaminant diffusion into the liquid phase.
of a kinetically-limited system as mass transfer parameters such as so- Taking into account that surface tension might be an important vari-
lubility (increasing with temperature), diffusion (increasing with tem- able, a capillarity force should be added to the Navier-Stokes equations.
perature) and Henry's constant (decreasing with temperature) are less Thus, again, a detailed and realistic description of the porous structure
sensitive to temperature compared to biodegradation parameters. is needed.
Mathematical modeling of biofilters can be used to ascertain mass As previously mentioned, a highly resolved description of the
transfer and kinetic limitations. Particularly, sensitivity analysis of di- porous media is very important to find reliable information at the pore
mensionless numbers such as the Damköhler number, the ratio of the scale. That can be achieved by digitalizing the actual porous media to
reaction rate to the mass transfer rate; the Thiele number, defined as the be used in the biofilter. This is possible nowadays by, for example,
ratio of the reaction rate to the diffusion rate, and the Peclet number, computational tomography. Figs. 5 are good examples of the level of

1089
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Fig. 5. Left: Clipped computational tomography image of a


biofilter column with vermiculite as porous media. The
vermiculite has an average diameter above 4 mm. The
column is comprised by a PVC pipe with flanges at the top
and bottom. The porous media is supported at the bottom
by a fixed plastic screen. The pipe has an internal diameter
of 7.5 cm, the pipe with flanges has 30 cm in height. Right:
Computational tomography image of 16 by 16 by 8 mm in
length, height and depth, respectively, of a polyurethane
foam.

detail that can be obtained for a grain (vermiculite) porous structure concentration of the contaminant and, (iii) for hydrophilic pollutants in
and a synthetic foam (polyurethane), respectively. Once validation is the gas phase (panel D in Fig. 6), its performance is inferior to bacterial
achieved, different studies of interest can be carried out in order to biofilters.
better understand several important characteristics and properties of As discussed in Section 3, these observations have been accounted
the effects of the porous media in the final performance of the bior- for, in both the conceptual and mathematical modeling of fungal bio-
eactor. filters, through the use of partition coefficients of the contaminant in
the biomass, which can be considered as an attempt to capture the
surface properties of the aerial hyphae (such as the presence of hy-
4. Comparison of fungal and bacterial biofilters and its drophobins), and by including descriptions of the increased surface area
performance available for the transport of contaminants to the fungal biomass.
Finally, applications of fungal biofiltration under development are
In this section, a comparison of experimental results from biofilters represented by the abatement of extremely low solubility contaminants,
operated with bacteria or fungi treating several VOCs is performed. such as methane (Lebrero et al., 2016), and large and complex mole-
Instead of using the rather common approach of revisiting each pub- cules such as polycyclic aromatic hydrocarbons (Vergara-Fernández
lished report and describing the results (for example the recent review et al., 2018).
of Cheng et al. (2016)), we adopt a comparative approach as follows.
Four categories of VOCs were distinguished: BTEX (benzene, toluene,
ethylbenzene and xylene) and styrene as slightly hydrophobic VOCs, 5. Conclusion
with partition coefficients between 1.66 (xylene) to 4.68 (styrene),
hydrophobic VOCs (α-pinene with a partition coefficient of 8.33), al- Biological technologies for the abatement of pollutants in diluted air
kanes as very hydrophobic VOCs (partition coefficients > 40) and streams offer economic advantages over physicochemical methods as
water-soluble VOCs. The literature was surveyed to collect experi- shown by the industrial application of bacterial biofiltration in the last
mental results of reactors treating these VOCs by either bacterial or decades. However, when the organic pollutants to be treated are hy-
fungal biofiltration. When possible, the inlet concentration of the con- drophobic, the performance, in terms of elimination capacity and inlet
taminant, the EBRT, L and EC were collected at 100% relative efficiency load, of bacterial biofilters is typically lower than the ones attained in
(also called critical load) and at the maximum recorded EC. These va- fungal biofilters.
lues were used to construct Fig. 6 A to D, where it can be seen that the As many factors, including the mass transfer of VOCs, the microbial
fungal biofilters outperform its bacterial counterpart in the treatment of kinetics, the structure of the packing media and the fluid flow in the
hydrophobic VOCs. On the other hand, the data is scarce on the use of reactor, can have an impact on the performance of a biofilter, mathe-
fungal biofilter for the abatement of hydrophilic compounds, and that matical models are often developed to disentangle the effects of each
the available information shows no distinct advantages for the fungal variable in the overall performance of the biofilter. For fungal reactors,
based biofilters over bacterial ones. models have evolved from those in which the same considerations as in
The high performance of fungal biofilters for the treatment of hy- bacterial biofiltration were used, to models specially designed to ac-
drophobic VOCs over bacterial biofilters has been previously reported count for traits, such as the development of hyphae, that are unique to
in literature for individual compounds (Estévez et al., 2005; Estrada fungi.
et al., 2013; Jorio et al., 2009; Kennes and Veiga, 2004; Vergara- Further research is still needed for the optimization of the packing
Fernández et al., 2011). However, to the best of our knowledge, Fig. 6 to improve the flow of the gas within the biofilter, minimizing death
represents the first systematic comparison reported to date comparing zones and short-circuiting, thus reducing the size of a biofilter for a
multiple results grouped by contaminant hydrophobicity for bacterial given task. This can be accomplished by a combination of highly re-
and fungal biofilters. From its analysis, it can be concluded that bio- solved descriptions of the porous media with computational fluid dy-
filters where fungi are used as the biocatalyst: (i) attained high elim- namics. Coupling such models with mathematical descriptions of the
ination capacities for hydrophobic compounds compared to its bacterial biodegradation kinetics and mass transfer could provide new insights in
counterparts at higher inlet loads, keeping a high relative efficiency fungal biofilters optimization using conventional packing (vermiculite
(panels A to C in Fig. 6), (ii) can be operated at higher inlet and perlite), but could also open the possibility of designing new

1090
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Fig. 6. Elimination capacity and load for fungal (circle) and bacterial (square symbols) biofilters. Panel A: biofilters treating benzene (B), toluene (T), styrene (S) and xylene (X). Panel B:
biofilters treating α-pinene. Panel C: biofilters treating n-pentane (C5), n-hexane (C6) and n-heptane (C7). Panel D: biofilters treating methanol (M), ethanol (EtOH), formaldehyde (F) and
methyl-propyl-ketone (MPK). Colored symbols indicate the inlet contaminant concentration according to the color bar at the bar at the right hand side of each plot, while the size of the
marker is proportional to the EBRT. A complete list of the references and data used to generate this figure is presented as supplementary material.

structured and unstructured packing. 18, 87–92. http://dx.doi.org/10.1002/ep.670180212.


Bayry, J., Aimanianda, V., Guijarro, J.I., Sunde, M., Latg, J.P., 2012. Hydrophobins-un-
ique fungal proteins. PLoS Pathog. 8, 6–9. http://dx.doi.org/10.1371/journal.ppat.
Acknowledgments 1002700.
Bernstein, J.A., Alexis, N., Bacchus, H., Bernstein, I.L., Fritz, P., Horner, E., Li, N., Mason,
The present work has been sponsored by CONICYT – Chile (National S., Nel, A., Oullette, J., Reijula, K., Reponen, T., Seltzer, J., Smith, A., Tarlo, S.M.,
2008. The health effects of nonindustrial indoor air pollution. J. Allergy Clin.
Commission for Scientific and Technological Research) (Fondecyt Immunol. 121, 585–591. http://dx.doi.org/10.1016/j.jaci.2007.10.045.
1160220). Carlsen, M., Spohr, A.B., Nielsen, J., Villadsen, J., 2000. Morphology and physiology of an
α-amylase producing strain of Aspergillus oryzae during batch cultivations.
Biotechnol. Bioeng. 49, 266–276. http://dx.doi.org/10.1002/(SICI)1097-
Appendix A. Supplementary data 0290(19960205)49:3<266::AID-BIT4>3.0.CO;2-I.
Chairez, I., García-Peña, I., Cabrera, A., 2009. Dynamic numerical reconstruction of a
Supplementary data to this article can be found online at https:// fungal biofiltration system using differential neural network. J. Process Control 19,
1103–1110. http://dx.doi.org/10.1016/j.jprocont.2008.12.009.
doi.org/10.1016/j.biotechadv.2018.03.008.
Cheng, Y., He, H., Yang, C., Zeng, G., Li, X., Chen, H., Yu, G., 2016. Challenges and
solutions for biofiltration of hydrophobic volatile organic compounds. Biotechnol.
References Adv. 34, 1091–1102. http://dx.doi.org/10.1016/j.biotechadv.2016.06.007.
Claessen, D., 2003. A novel class of secreted hydrophobic proteins is involved in aerial
hyphae formation in Streptomyces coelicolor by forming amyloid-like fibrils. Genes
Arriaga, S., Revah, S., 2005. Removal of n-hexane by Fusarium solani with a gas-phase Dev. 17, 1714–1726. http://dx.doi.org/10.1101/gad.264303.
biofilter. J. Ind. Microbiol. Biotechnol. 32, 548–553. http://dx.doi.org/10.1007/ Cohen, Y., 2001. Biofiltration – the treatment of fluids by microorganisms immobilized
s10295-005-0247-9. into the filter bedding material: a review. Bioresour. Technol. 77, 257–274. http://
Arriaga, S., Revah, S., 2009. Mathematical modeling and simulation of hexane degrada- dx.doi.org/10.1016/S0960-8524(00)00074-2.
tion in fungal and bacterial biofilters: effective diffusivity and partition aspects. Can. Cohen, A.J., Brauer, M., Burnett, R., Anderson, H.R., Frostad, J., Estep, K., Balakrishnan,
J. Civ. Eng. 36, 1919–1925. http://dx.doi.org/10.1139/L09-090. K., Brunekreef, B., Dandona, L., Dandona, R., Feigin, V., Freedman, G., Hubbell, B.,
Aynsley, M., Ward, A.C., Wright, A.R., 1990. A mathematical model for the growth of Jobling, A., Kan, H., Knibbs, L., Liu, Y., Martin, R., Morawska, L., Pope, C.A., Shin, H.,
mycelial fungi in submerged culture. Biotechnol. Bioeng. 35, 820–830. http://dx.doi. Straif, K., Shaddick, G., Thomas, M., van Dingenen, R., van Donkelaar, A., Vos, T.,
org/10.1002/bit.260350810. Murray, C.J.L., Forouzanfar, M.H., 2017. Estimates and 25-year trends of the global
Barton, J.W., Davison, B.H., Klasson, K.T., Gable, C.C., 1999. Estimation of mass transfer burden of disease attributable to ambient air pollution: an analysis of data from the
and kinetics in operating trickle-bed bioreactors for removal of VOCS. Environ. Prog.

1091
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Global Burden of Diseases Study 2015. Lancet 389, 1907–1918. http://dx.doi.org/10. temperature and toluene presence. CLEAN - Soil, Air, Water 43, 104–111. http://dx.
1016/S0140-6736(17)30505-6. doi.org/10.1002/clen.201200318.
Davison, B.H., Barton, J.W., Klasson, K.T., Francisco, A.B., 2000. Influence of high bio- Jin, Y., Veiga, M., Kennes, C., 2006. Performance optimization of the fungal biode-
mass concentrations on alkane solubilities. Biotechnol. Bioeng. 68, 279–284. http:// gradation of α-pinene in gas-phase biofilter. Process Biochem. 41, 1722–1728.
dx.doi.org/10.1002/(SICI)1097-0290(20000505)68:3<279::AID-BIT6>3.0.CO;2-P. http://dx.doi.org/10.1016/j.procbio.2006.03.020.
Delhoménie, M.-C., Heitz, M., 2005. Biofiltration of air: a review. Crit. Rev. Biotechnol. Jin, Y., Guo, L., Veiga, M.C., Kennes, C., 2007. Fungal biofiltration of α-pinene: effects of
25, 53–72. http://dx.doi.org/10.1080/07388550590935814. temperature, relative humidity, and transient loads. Biotechnol. Bioeng. 96, 433–443.
Delhoménie, M.-C., Nikiema, J., Bibeau, L., Heitz, M., 2008. A new method to determine http://dx.doi.org/10.1002/bit.21123.
the microbial kinetic parameters in biological air filters. Chem. Eng. Sci. 63, Jorio, H., Jin, Y., Elmrini, H., Nikiema, J., Brzezinski, R., Heitz, M., 2009. Treatment of
4126–4134. http://dx.doi.org/10.1016/j.ces.2008.05.020. VOCs in biofilters inoculated with fungi and microbial consortium. Environ. Technol.
Deshusses, M.A., Hamer, G., Dunn, I.J., 1995. Behavior of Biofilters for waste air 30, 477–485. http://dx.doi.org/10.1080/09593330902778849.
Biotreatment. 1. Dynamic model development. Environ. Sci. Technol. 29, 1048–1058. Kennes, C., Veiga, M.C.M.C., 2004. Fungal biocatalysts in the biofiltration of VOC-pol-
http://dx.doi.org/10.1021/es00004a027. luted air. J. Biotechnol. 113, 305–319. http://dx.doi.org/10.1016/j.jbiotec.2004.04.
Devinny, J.S., Ramesh, J., 2005. A phenomenological review of biofilter models. Chem. 037.
Eng. J. 113, 187–196. http://dx.doi.org/10.1016/j.cej.2005.03.005. Khan, F.I., Kr Ghoshal, A., 2000. Removal of volatile organic compounds from polluted
Dorado, A.D., Baquerizo, G., Maestre, J.P., Gamisans, X., Gabriel, D., Lafuente, J., 2008. air. J. Loss Prev. Process Ind. 13, 527–545. http://dx.doi.org/10.1016/S0950-
Modeling of a bacterial and fungal biofilter applied to toluene abatement: kinetic 4230(00)00007-3.
parameters estimation and model validation. Chem. Eng. J. 140, 52–61. http://dx. Kim, D., Cai, Z., Sorial, G.A., Shin, H., Knaebel, K., 2007. Integrated treatment scheme of a
doi.org/10.1016/j.cej.2007.09.004. biofilter preceded by a two-bed cyclic adsorption unit treating dynamic toluene
Edwards, M.F., Richardson, J.F., 1968. Gas dispersion in packed beds. Chem. Eng. Sci. 23, loading. Chem. Eng. J. 130, 45–52. http://dx.doi.org/10.1016/j.cej.2006.10.033.
109–123. http://dx.doi.org/10.1016/0009-2509(68)87056-3. Kraakman, N.J.R., Rocha-Rios, J., Van Loosdrecht, M.C.M., 2011. Review of mass transfer
Elliott, L., Longnecker, M.P., Kissling, G.E., London, S.J., 2006. Volatile organic com- aspects for biological gas treatment. Appl. Microbiol. Biotechnol. 91, 873–886.
pounds and pulmonary function in the third National Health and nutrition ex- http://dx.doi.org/10.1007/s00253-011-3365-5.
amination survey, 1988–1994. Environ. Health Perspect. 114, 1210–1214. http://dx. Kundu, S., Stone, E.A., 2014. Composition and sources of fine particulate matter across
doi.org/10.1289/ehp.9019. urban and rural sites in the Midwestern United States. Environ. Sci. Process. Impacts
Estévez, E., Veiga, M.C., Kennes, C., 2005. Biodegradation of toluene by the new fungal 16, 1360–1370. http://dx.doi.org/10.1039/C3EM00719G.
isolates Paecilomyces variotii and Exophiala oligosperma. J. Ind. Microbiol. Larralde-Corona, C.P., López-Isunza, F., Viniegra-González, G., 1997. Morphometric
Biotechnol. 32, 33–37. http://dx.doi.org/10.1007/s10295-004-0203-0. evaluation of the specific growth rate ofAspergillus Niger grown in agar plates at high
Estrada, J.M., Kraakman, N.J.R., Lebrero, R., Muñoz, R., 2012. A sensitivity analysis of glucose levels. Biotechnol. Bioeng. 56, 287–294. http://dx.doi.org/10.1002/(SICI)
process design parameters, commodity prices and robustness on the economics of 1097-0290(19971105)56:3<287::AID-BIT6>3.0.CO;2-F.
odour abatement technologies. Biotechnol. Adv. 30, 1354–1363. http://dx.doi.org/ Lebrero, R., López, J.C., Lehtinen, I., Pérez, R., Quijano, G., Muñoz, R., 2016. Exploring
10.1016/j.biotechadv.2012.02.010. the potential of fungi for methane abatement: performance evaluation of a fungal-
Estrada, J.M., Hernández, S., Muñoz, R., Revah, S., 2013. A comparative study of fungal bacterial biofilter. Chemosphere 144, 97–106. http://dx.doi.org/10.1016/j.
and bacterial biofiltration treating a VOC mixture. J. Hazard. Mater. 250–251, chemosphere.2015.08.017.
190–197. http://dx.doi.org/10.1016/j.jhazmat.2013.01.064. Li, L., Liu, J.X., 2006. Removal of xylene from off-gas using a bioreactor containing
Evtyugina, M., Alves, C., Calvo, A., Nunes, T., Tarelho, L., Duarte, M., Prozil, S.O., bacteria and fungi. Int. Biodeterior. Biodegrad. 58, 60–64. http://dx.doi.org/10.
Evtuguin, D.V., Pio, C., 2014. VOC emissions from residential combustion of Southern 1016/j.ibiod.2006.07.002.
and mid-European woods. Atmos. Environ. 83, 90–98. http://dx.doi.org/10.1016/j. Linder, M.B., Szilvay, G.R., Nakari-Setälä, T., Penttilä, M.E., 2005. Hydrophobins: the
atmosenv.2013.10.050. protein-amphiphiles of filamentous fungi. FEMS Microbiol. Rev. 29, 877–896. http://
Ferdowsi, M., Avalos Ramirez, A., Jones, J.P., Heitz, M., 2017. Elimination of mass dx.doi.org/10.1016/j.femsre.2005.01.004.
transfer and kinetic limited organic pollutants in biofilters: a review. Int. Biodeterior. Lobo, R., Revah, S., Viveros-García, T., 1999. An analysis of a trickle-bed bioreactor:
Biodegrad. 119, 336–348. http://dx.doi.org/10.1016/j.ibiod.2016.10.015. carbon disulfide removal. Biotechnol. Bioeng. 63, 98–109. http://dx.doi.org/10.
Fuller, E.N., Schettler, P.D., Giddings, J.C., 1966. New method for prediction of binary 1002/(SICI)1097-0290(19990405)63:1<98::AID-BIT10>3.0.CO;2-8.
gas-phase diffusion coefficients. Ind. Eng. Chem. 58, 18–27. http://dx.doi.org/10. López-Isunza, F., Larralde-Corona, C.P., Viniegra-González, G., 1997. Mass transfer and
1021/ie50677a007. growth kinetics in filamentous fungi. Chem. Eng. Sci. 52, 2629–2639. http://dx.doi.
Galera, M.M., Cho, E., Tuuguu, E., Park, S.-J., Lee, C., Chung, W.-J., 2008. Effects of org/10.1016/S0009-2509(97)00078-X.
pollutant concentration ratio on the simultaneous removal of NH3, H2S and toluene Mackay, D., 2001. Multimedia Environmental Models: The Fugacity Approach, Second
gases using rock wool-compost biofilter. J. Hazard. Mater. 152, 624–631. http://dx. Ed. Lewis Publishers - CRC Press.
doi.org/10.1016/j.jhazmat.2007.07.025. Mathur, A.K., Sundaramurthy, J., Balomajumder, C., 2006. Kinetics of the removal of
Gostner, J.M., Zeisler, J., Alam, M.T., Gruber, P., Fuchs, D., Becker, K., Neubert, K., mono-chlorobenzene vapour from waste gases using a trickle bed air biofilter. J.
Kleinhappl, M., Martini, S., Überall, F., 2016. Cellular reactions to long-term volatile Hazard. Mater. 137, 1560–1568. http://dx.doi.org/10.1016/j.jhazmat.2006.04.042.
organic compound (VOC) exposures. Sci. Rep. 6, 37842. http://dx.doi.org/10.1038/ Moe, W.M., Irvine, R.L., 2001. Effect of nitrogen limitation on performance of toluene
srep37842. degrading Biofilters. Water Res. 35, 1407–1414. http://dx.doi.org/10.1016/S0043-
Gostomski, P.A., Sisson, J.B., Cherry, R.S., 1997. Water content dynamics in Biofiltration: 1354(00)00417-6.
the role of humidity and microbial heat generation. J. Air Waste Manage. Assoc. 47, Mohseni, M., Allen, D.G., 2000. Biofiltration of mixtures of hydrophilic and hydrophobic
936–944. http://dx.doi.org/10.1080/10473289.1997.10463952. volatile organic compounds. Chem. Eng. Sci. 55, 1545–1558. http://dx.doi.org/10.
Guimerà, Xavier, Dorado, Antonio David, Bonsfills, Anna, Gabriel, Gemma, David 1016/S0009-2509(99)00420-0.
Gabriel, X.G., 2016. Dynamic characterization of external and internal mass transport Morales, M., Hernández, S., Cornabé, T., Revah, S., Auria, R., 2003. Effect of drying on
in heterotrophic biofilms from microsensors measurements. Water Res. 102, Biofilter performance: modeling and experimental approach. Environ. Sci. Technol.
551–560. http://dx.doi.org/10.1016/j.watres.2016.07.009. 37, 985–992. http://dx.doi.org/10.1021/es025970w.
Harley, R.A., Hooper, D.S., Kean, A.J., Kirchstetter, T.W., Hesson, J.M., Balberan, N.T., Neale, G.H., Nader, W.K., 1973. Prediction of transport processes within porous media:
Stevenson, E.D., Kendall, G.R., 2006. Effects of reformulated gasoline and motor diffusive flow processes within an homogeneous swarm of spherical particles. AICHE
vehicle fleet turnover on emissions and ambient concentrations of benzene. Environ. J. 19, 112–119. http://dx.doi.org/10.1002/aic.690190116.
Sci. Technol. 40, 5084–5088. http://dx.doi.org/10.1021/es0604820. Ottengraf, S.P.P., Van Den Oever, A.H.C., 1983. Kinetics of organic compound removal
Harms, H., Bosma, T.N.P., 1997. Mass transfer limitation of microbial growth and pol- from waste gases with a biological filter. Biotechnol. Bioeng. 25, 3089–3102. http://
lutant degradation. J. Ind. Microbiol. Biotechnol. 18, 97–105. http://dx.doi.org/10. dx.doi.org/10.1002/bit.260251222.
1038/sj.jim.2900259. Prenafeta-Boldú, F.X., Illa, J., van Groenestijn, J.W., Flotats, X., 2008. Influence of syn-
Helfferich, F.G., 1985. Principles of adsorption & adsorption processes, by D. M. Ruthven, thetic packing materials on the gas dispersion and biodegradation kinetics in fungal
John Wiley & sons, 1984, xxiv + 433 pp. AICHE J. 31, 523–524. http://dx.doi.org/ air biofilters. Appl. Microbiol. Biotechnol. 79, 319–327. http://dx.doi.org/10.1007/
10.1002/aic.690310335. s00253-008-1433-2.
Hernández-Meléndez, O., Bárzana, E., Arriaga, S., Hernández-Luna, M., Revah, S., 2008. Ren, X., Zeng, G., Tang, L., Wang, J., Wan, J., Liu, Y., Yu, J., Yi, H., Ye, S., Deng, R., 2018.
Fungal removal of gaseous hexane in biofilters packed with poly(ethylene carbonate) Sorption, transport and biodegradation – an insight into bioavailability of persistent
pine sawdust or peat composites. Biotechnol. Bioeng. 100, 864–871. http://dx.doi. organic pollutants in soil. Sci. Total Environ. 610–611, 1154–1163. http://dx.doi.
org/10.1002/bit.21825. org/10.1016/j.scitotenv.2017.08.089.
Hodge, D.S., Devinny, J.S., 1995. Modeling removal of air contaminants by Biofiltration. Revah, S., Morgan-Sagastume, J., 2005. Methods for odor and VOC control. In:
J. Environ. Eng. 121, 21–32. http://dx.doi.org/10.1061/(ASCE)0733-9372(1995) Shareefdeen, Z., Singh, A. (Eds.), Biotechnology for Odour and Air Pollution. Springer
121:1(21). Berlin Heidelberg, pp. 29–64.
Hoyt, D., Raun, L.H., 2015. Measured and estimated benzene and volatile organic carbon Salehahmadi, R., Halladj, R., Zamir, S.M., 2012. Unsteady-state mathematical modeling
(VOC) emissions at a major U.S. refinery/chemical plant: comparison and prior- of a fungal Biofilter treating hexane vapor at different operating temperatures. Ind.
itizatio. J. Air Waste Manage. Assoc. 65, 1020–1031. http://dx.doi.org/10.1080/ Eng. Chem. Res. 51, 2388–2396. http://dx.doi.org/10.1021/ie2014718.
10962247.2015.1058304. Sarigiannis, D.A., Karakitsios, S.P., Gotti, A., Liakos, I.L., Katsoyiannis, A., 2011. Exposure
Iliuta, I., Larachi, F., 2004. Transient biofilter aerodynamics and clogging for VOC de- to major volatile organic compounds and carbonyls in European indoor environments
gradation. Chem. Eng. Sci. 59, 3293–3302. http://dx.doi.org/10.1016/j.ces.2004.05. and associated health risk. Environ. Int. 37, 743–765. http://dx.doi.org/10.1016/j.
004. envint.2011.01.005.
Iranmanesh, E., Halladj, R., Zamir, S.M., 2015. Microkinetic analysis of n -hexane bio- Saucedo-Lucero, J.O., Quijano, G., Arriaga, S., Muñoz, R., 2014. Hexane abatement and
degradation by an isolated fungal consortium from a Biofilter: influence of spore emission control in a fungal biofilter-photoreactor hybrid unit. J. Hazard.

1092
A. Vergara-Fernández et al. Biotechnology Advances 36 (2018) 1079–1093

Mater. 276, 287–294. http://dx.doi.org/10.1016/j.jhazmat.2014.05.040. 439–444.


Schor, M., Reid, J.L., MacPhee, C.E., Stanley-Wall, N.R., 2016. The diverse structures and Vergara-Fernandez, A., San Martín-Davison, J., Díaz-Robles, L.A., Soto-Sanchez, O.,
functions of surfactant proteins. Trends Biochem. Sci. 41, 610–620. http://dx.doi. Martin-Davison, J., Diaz-Robles, L., Soto-Sanchez, O., 2014. Effect of initial sub-
org/10.1016/j.tibs.2016.04.009. strate/inoculum ratio on cell yield in the removal of hydrophobic VOCs in fungal
Shareefdeen, Z., Baltzis, B.C., 1994. Biofiltration of toluene vapor under steady-state and biofilters. Rev. Mex. Ing. Química 13, 749–755.
transient conditions: theory and experimental results. Chem. Eng. Sci. 49, Vergara-Fernández, A., Scott, F., Moreno-Casas, P., Díaz-Robles, L., Muñoz, R., Diaz-
4347–4360. http://dx.doi.org/10.1016/S0009-2509(05)80026-0. Robles, L., Muñoz, R., 2016. Elucidating the key role of the fungal mycelium on the
Smits, J.P., van Sonsbeek, H.M., Tramper, J., Knol, W., Geelhoed, W., Peeters, M., biodegradation of n-pentane as a model hydrophobic VOC. Chemosphere 157, 89–96.
Rinzema, A., 1999. Modelling fungal solid-state fermentation: the role of inactivation http://dx.doi.org/10.1016/j.chemosphere.2016.05.034.
kinetics. Bioprocess Eng. 20, 391. http://dx.doi.org/10.1007/s004490050607. Vergara-Fernández, A., Yánez, D., Morales, P., Scott, F., Aroca, G., Diaz-Robles, L.,
Spigno, G., De Marco Faveri, D., 2005. Modeling of a vapor-phase fungi bioreactor for the Moreno-Casas, P., 2018. Biofiltration of benzo[α]pyrene, toluene and formaldehyde
abatement of hexane: fluid dynamics and kinetic aspects. Biotechnol. Bioeng. 89, in air by a consortium of Rhodococcus erythropolis and Fusarium solani : effect of
319–328. http://dx.doi.org/10.1002/bit.20336. inlet loads, gas flow and temperature. Chem. Eng. J. 332, 702–710. http://dx.doi.
Spigno, G., Tronci, S., 2015. Development of hybrid models for a vapor-phase Fungi org/10.1016/j.cej.2017.09.095.
bioreactor. Math. Probl. Eng. 2015, 1–11. http://dx.doi.org/10.1155/2015/801213. Vigueras, G., Shirai, K., Martins, D., Franco, T.T., Fleuri, L.F., Revah, S., 2008. Toluene
Spigno, G., Pagella, C., Daria Fumi, M., Molteni, R., Marco De Faveri, D., 2003. VOCs gas phase biofiltration by Paecilomyces lilacinus and isolation and identification of a
removal from waste gases: gas-phase bioreactor for the abatement of hexane by hydrophobin protein produced thereof. Appl. Microbiol. Biotechnol. 80, 147–154.
Aspergillus niger. Chem. Eng. Sci. 58, 739–746. http://dx.doi.org/10.1016/S0009- http://dx.doi.org/10.1007/s00253-008-1490-6.
2509(02)00603-6. Vigueras, G., Arriaga, S., Shirai, K., Morales, M., Revah, S., 2009. Hydrophobic response
Sugai-Guérios, M.H., Balmant, W., Furigo, A., Krieger, N., Mitchell, D.A., 2015. Modeling of the fungus Rhinocladiella similis in the biofiltration with volatile organic com-
the Growth of Filamentous Fungi at the Particle Scale in Solid-State Fermentation pounds with different polarity. Biotechnol. Lett. 31, 1203–1209. http://dx.doi.org/
Systems. pp. 171–221. http://dx.doi.org/10.1007/10_2014_299. 10.1007/s10529-009-9987-3.
Swanson, W.J., Loehr, R.C., 1997. Biofiltration: fundamentals, design and operations Villeneuve, P.J., Jerrett, M., Su, J., Burnett, R.T., Chen, H., Brook, J., Wheeler, A.J.,
principles, and applications. J. Environ. Eng. 123, 538–546. http://dx.doi.org/10. Cakmak, S., Goldberg, M.S., 2013. A cohort study of intra-urban variations in volatile
1061/(ASCE)0733-9372(1997)123:6(538). organic compounds and mortality, Toronto, Canada. Environ. Pollut. 183, 30–39.
Szewczyk, K.W., Myszka, L., 1994. The effect of temperature on the growth of A. niger in http://dx.doi.org/10.1016/j.envpol.2012.12.022.
solid state fermentation. Bioprocess Eng. 10, 123–126. http://dx.doi.org/10.1007/ Wang, F., Li, C., Liu, W., Jin, Y., 2012. Effect of exposure to volatile organic compounds
BF00369467. (VOCs) on airway inflammatory response in mice. J. Toxicol. Sci. 37, 739–748.
Tang, M.J., Shiraiwa, M., Pöschl, U., Cox, R.A., Kalberer, M., 2015. Compilation and http://dx.doi.org/10.2131/jts.37.739.
evaluation of gas phase diffusion coefficients of reactive trace gases in the atmo- Wilke, C.R., Chang, P., 1955. Correlation of diffusion coefficients in dilute solutions.
sphere: volume 2. Diffusivities of organic compounds, pressure-normalised mean free AICHE J. 1, 264–270. http://dx.doi.org/10.1002/aic.690010222.
paths, and average Knudsen numbers for gas uptake calculations. Atmos. Chem. Phys. Win-Shwe, T.-T., Fujimaki, H., Arashidani, K., Kunugita, N., 2013. Indoor volatile organic
15, 5585–5598. http://dx.doi.org/10.5194/acp-15-5585-2015. compounds and chemical sensitivity reactions. Clin. Dev. Immunol. 2013, 1–8.
van Lith, C., Leson, G., Michelsen, R., vanLith, C., Leson, G., Michelsen, R., 1997. http://dx.doi.org/10.1155/2013/623812.
Evaluating design options for biofilters. J. Air Waste Manage. Assoc. 47, 37–48. Woertz, J.R., Kinney, K.A., Szaniszlo, P.J., 2001. A fungal vapor-phase bioreactor for the
http://dx.doi.org/10.1080/10473289.1997.10464410. removal of nitric oxide from waste gas streams. J. Air Waste Manage. Assoc. 51,
Vergara-Fernández, A., Van Haaren, B., Revah, S., 2006. Phase partition of gaseous 895–902. http://dx.doi.org/10.1080/10473289.2001.10464321.
hexane and surface hydrophobicity of Fusarium solani when grown in liquid and Wösten, H.A., Schuren, F.H., Wessels, J.G., 1994. Interfacial self-assembly of a hydro-
solid media with hexanol and hexane. Biotechnol. Lett. 28, 2011–2017. http://dx. phobin into an amphipathic protein membrane mediates fungal attachment to hy-
doi.org/10.1007/s10529-006-9186-4. drophobic surfaces. EMBO J. 13, 5848–5854.
Vergara-Fernández, A., Hernández, S., Revah, S., 2008. Phenomenological model of Wösten, H.A.B., Van Wetter, M.A., Lugones, L.G., Van der Mei, H.C., Busscher, H.J.,
fungal biofilters for the abatement of hydrophobic VOCs. Biotechnol. Bioeng. 101, Wessels, J.G.H., 1999. How a fungus escapes the water to grow into the air. Curr.
1182–1192. http://dx.doi.org/10.1002/bit.21989. Biol. 9, 85–88. http://dx.doi.org/10.1016/S0960-9822(99)80019-0.
Vergara-Fernández, A., Hernández, S., Revah, S., 2011. Elimination of hydrophobic vo- Xi, J., Kang, I., Hu, H., Zhang, X., 2015. A biofilter model for simultaneous simulation of
latile organic compounds in fungal biofilters: reducing start-up time using different toluene removal and bed pressure drop under varied inlet loadings. Front. Environ.
carbon sources. Biotechnol. Bioeng. 108, 758–765. http://dx.doi.org/10.1002/bit. Sci. Eng. 9, 554–562. http://dx.doi.org/10.1007/s11783-014-0671-z.
23003. Yang, C., Yu, G., Zeng, G., Yang, H., Chen, F., Jin, C., 2011. Performance of biotrickling
Vergara-Fernandez, A., Hernández, S., San Martín-Davison, J., Revah, S., Vergara- filters packed with structured or cubic polyurethane sponges for VOC removal. J.
Fernández, A., Hernández, S., San Martín-Davison, J., Revah, S., Vergara-Fernandez, Environ. Sci. 23, 1325–1333. http://dx.doi.org/10.1016/S1001-0742(10)60565-7.
A., Hernández, S., San Martín-Davison, J., Revah, S., 2011. Morphological char- Zamir, S.M., Halladj, R., Nasernejad, B., 2011a. Removal of toluene vapors using a fungal
acterization of aerial hyphae and simulation growth of Fusarium solani under dif- biofilter under intermittent loading. Process. Saf. Environ. Prot. 89, 8–14. http://dx.
ferent carbon source for application in the hydrophobic VOCs biofiltration. Rev. Mex. doi.org/10.1016/j.psep.2010.10.001.
Ing. Química 10, 225–233. Zamir, M., Halladj, R., Saber, M., Ferdowsi, M., Nasernejad, B., 2011b. Biofiltration of
Vergara-Fernandez, A., Hernandez, S., Munoz, R., Revah, S., Vergara-Fernández, A., hexane vapor: experimental and neural model analysis. CLEAN - Soil, Air, Water 39,
Hernández, S., Muñoz, R., Revah, S., Vergara-Fernandez, A., Hernandez, S., Munoz, 813–819. http://dx.doi.org/10.1002/clen.201000525.
R., Revah, S., 2012. Influence of the inlet load, EBRT and mineral medium addition Zhu, X., Suidan, M.T., Pruden, A., Yang, C., Alonso, C., Kim, B.R.J., Kim, B.R.J., 2004.
on spore emission by Fusarium solani in the fungal biofiltration of hydrophobic Effect of substrate Henry's constant on Biofilter performance. J. Air Waste Manage.
VOCs. J. Chem. Technol. Biotechnol. 87, 778–784. http://dx.doi.org/10.1002/jctb. Assoc. 54, 409–418. http://dx.doi.org/10.1080/10473289.2004.10470918.
3762. Znad, H., Katoh, K., Kawase, Y., 2007. High loading toluene treatment in a compost based
Vergara-Fernández, A., Sóto-Sanchez, O., Vásquez-Bestagno, J., Vergara-Fernandez, A., biofilter using up-flow and down-flow swing operation. J. Hazard. Mater. 141,
Soto-Sanchez, O., Vasquez, J., 2012b. Effects of packing material type on n-pentane/ 745–752. http://dx.doi.org/10.1016/j.jhazmat.2006.07.039.
biomass partition coefficient for use in fungal Biofilters. Chem. Biochem. Eng. Q. 25,

1093

You might also like