You are on page 1of 4

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

You might also like