You are on page 1of 18

ARTICLE IN PRESS

Journal of Environmental Management 81 (2006) 1–18


www.elsevier.com/locate/jenvman

Review

Modelling anaerobic biofilm reactors—A review


V. Saravanan, T.R. Sreekrishnan
Department of Biochemical Engineering and Biotechnology, Indian Institute of Technology Delhi, New Delhi-110 016, India
Received 7 October 2004; received in revised form 4 October 2005; accepted 5 October 2005
Available online 6 March 2006

Abstract

Anaerobic treatment has become a technically as well as economically feasible option for treatment of liquid effluents after the
development of reactors such as the upflow anaerobic sludge blanket (UASB) reactor, expanded granular sludge bed (EGSB) reactor,
anaerobic biofilter and anaerobic fluidized bed reactor (AFBR). Considerable effort has gone into developing mathematical models for
these reactors in order to optimize their design, design the process control systems used in their operation and enhance their operational
efficiency. This article presents a critical review of the different mathematical models available for these reactors. The unified anaerobic
digestion model (ADM1) and its application to anaerobic biofilm reactors are also outlined.
r 2006 Elsevier Ltd. All rights reserved.

Keywords: Mathematical model; UASB; AFBR; EGSB; Biofilter; Biofilm; Anaerobic

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Models for UASB reactors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1. Flow model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Reactor model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Proposed models for the structure of the biofilm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1. Multi-layer model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.2. Syntrophic microcolony model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3. Non-layered structure model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.4. Granule cluster structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4. Models for AFBR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.1. Bed fluidization model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.1.1. Terminal settling velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.1.2. Fluidization mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.1.3. Effect of gas production on hydrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.1.4. Bed stratification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.2. Kinetic and reactor sub-models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.2.1. Stratified biofilm models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
5. The EGSB reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
6. The anaerobic biofilter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
7. The unified model for anaerobic digestion (ADM1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
7.1. The extension of ADM1 to biofilm reactors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
8. Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

Corresponding author. Tel.: +91 11 26591014; fax: +91 11 26582282.


E-mail address: sreekrishnan_t_r@hotmail.com (T.R. Sreekrishnan).

0301-4797/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jenvman.2005.10.002
ARTICLE IN PRESS
2 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

1. Introduction Biogas

The term ‘‘biofilm reactor’’ refers to that class of Effluent


bioreactors where the biocatalyst exists in an anchored Weir
form, either on the surface of an inert ‘‘carrier’’ or attached Settler
to one another. The carrier could be the wall of the reactor, Baffles
baffles provided for this purpose or particles of some inert
material. Biocatalysts such as microorganisms could also Sludge Blanket
grow attached to one another, giving rise to a ‘‘biogra-
nule’’. The carrier or the biogranule could be stationary as
in a packed-bed or expanded bed system or mobile as in the Sludge granules
Sludge Bed
case of a fluidized bed system. Typically, in such reactors,
Influent
the rate of substrate conversion is limited by the rate of
transport of substrate into the biofilm. In an anaerobic
Fig. 1. Schematic diagram of UASB reactor.
biofilm reactor, the biocatalyst will include all the different
bacterial species responsible for the break down of complex
organic molecules to a final end product consisting of Biogas
methane and carbon dioxide. The anaerobic treatment, as
the main biological step in wastewater treatment systems,
Treated water
was scarce until the development of the upflow anaerobic
sludge blanket (UASB) reactor in the early 1970s (Lettinga
et al., 1980). UASB processes are based on the develop- Recycle line
ment of dense granules (1–4 mm) formed by the natural Biofilm
self-immobilization of the anaerobic bacteria. This kind of
Carrier
immobilization does not employ any support material such
as Raschig rings or clay in the reactor (Nicolella et al.,
2000). A schematic of a typical UASB reactor is shown in
Fig. 1. Wastewater enters the bottom of the reactor
through the inlet liquid distribution system and passes
upward through the dense anaerobic sludge bed. Because
of the high biomass concentration, it was demonstrated Wastewater feed
that volumetric organic loading rates as high as 50 kg
Fig. 2. Anaerobic fluidized bed reactor (AFBR).
chemical oxygen demand (COD) per m3 per day could be
employed (Hulshoff Pol, 1989). The liquid velocity inside
the reactor is usually in the range of 0.5–1.0 m/h. This are simple in concept and close to the physical situation
reactor consists of a sludge bed, a sludge blanket and a existing in these reactors. During the past few decades there
clarifier zone supplemented with a physical device called have been a number of studies on modelling of UASB and
the gas–solid separator. AFBR reactors. An effort is made in this article to present
In anaerobic fluidized bed reactor (AFBR), the liquid to the gist of different approaches used in developing
be treated is pumped through a bed of inert particles mathematical models for these reactors. This article also
(typically sand with a particle size range of 0.2–0.8 mm) at reviews different proposals available for modelling the
a velocity sufficient enough (10–20 m/h) to cause fluidiza- structure of the biological granules and biofilm-covered
tion (Nicolella et al., 2000). In the fluidized state, the media particles in these reactors.
provides a large surface for attached biological growth and
allows biomass concentrations to develop in the range of 2. Models for UASB reactors
10–40 kg/m3 (Cooper and Sutton, 1983). A typical flow
diagram of an AFBR is shown in Fig. 2. Compared to The formation of anaerobic granular sludge is consid-
other high rate anaerobic reactors such as UASB, the ered to be a necessary condition for the successful
fluidized bed system is claimed to have the following operation of a UASB reactor. The efficiency of the reactor
advantages: higher purification capacity, no clogging of depends mainly on active biomass concentration and the
the reactor (as in filters), no problem of sludge washout influent flow rate. As the reactor reaches steady state, a
(as in UASB systems if granular sludge is not obtained), dense bed composed of granulated sludge develops at the
and small volume and land area requirements (Heijnen bottom. Immediately above the sludge bed, a zone
et al., 1989). consisting of finely suspended particles called the sludge
To study the sensitivity of the process to various blanket forms. A clear zone over this sludge blanket
operating parameters and to optimize the design of these constitutes the settling zone. To keep the sludge granules in
reactors, it is necessary to have mathematical models which suspended condition inside the reactor, the inlet flow rate
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 3

should be below a threshold limit (normally below 1 m/h). By-pass flow


This low influent flow rate causes channelling problems in
the sludge bed zone, resulting in lowered efficiency of the
reactor. Hence to develop mathematical models for this CSTR Dispersed
reactor, it is important to analyse the flow pattern inside plug flow
the reactor and reaction kinetics within the biological
granules. In general, models for UASB reactors consist of Dead volume
two parts:
Fig. 3. Representation of hydraulic model (Wu and Hickey, 1997).

1. fluid flow model;


2. reactor model.
Effluent
2.1. Flow model
Settler
Fluid flow models for UASB reactors differ considerably
in their approach. The different zones of a UASB reactor
are modelled as continuos stirred tank reactors (CSTR) or
plug flow reactors (PFR) having dead volume with bypass
flow between the zones (Bolle et al., 1986; Van der Meer, Sludge blanket
1979; Wu and Hickey, 1997). The flow behaviour within a
zone mainly depends on the concentration and character-
istics of the biomass (Ojha and Singh, 2002). Short circuit
Wu and Hickey (1997) have developed a flow model to
describe the flow pattern in a UASB reactor. They modelled Dead space Sludge bed
the sludge bed and blanket as a non-ideal CSTR by using a
combination of an ideal CSTR along with a dead zone and
a bypass flow. This CSTR is in series with a dispersed plug
flow reactor (PFR) (non-ideal PFR) that represents the Influent
clarification zone above the sludge blanket (Fig. 3).
The equation of the flow model is given by Fig. 4. Block diagram of fluid flow pattern (Bolle et al., 1986).

dC
Vb ¼ V b EðtÞ  Qf CðtÞ, (2.1)
dt
Narnoli and Indu (1997) developed a model for the
EðtÞ ¼ VM f
b T in
if 0ptpT in ; sludge blanket of a UASB reactor. According to them, the
(2.2) maximum organic load which a reactor can assimilate
¼0 if T in p0;
depends on proportioning of the reactor height into the bed
where C(t) is the tracer concentration within the CSTR, Vb and the blanket sections. A condition of force equilibrium
the CSTR working volume, E(t) the input function for was considered between the rising gas bubbles from the
tracer impulse, Qf the low fraction that enters the working sludge bed and the adjacent water mass. The negative
volume, Mf the fraction of mass input that go through pressure behind the bubble attracts the water mass and
reactor’s working volume, and Tin the tracer injection time. leads to the formation of a wake. Using diffusion concepts,
Ojha and Singh (2002) analysed the flow distribution in solid particles moving along the wake due to the
different zones of the reactor in order to simulate the concentration gradient were equated to those settling down
UASB reactor performance. They used the flow resistance under the influence of gravity at steady state. The solid
approach to represent flow distribution. It was found that concentration along the blanket height was computed and
with an increase in flow resistance in the UASB reactor found to match the experimental observations. This model
system, the magnitude of short-circuiting flows at the could be used to optimize the reactor dimensions and the
reactor bed increased. The flow distribution at the blanket desludging schedule.
and clarifier was found to have an influence on the flow Singhal et al. (1998) reported that a simple two-zone
resistance. But no generalized model could be obtained. axial dispersion model adequately describes the fluid flow
Bolle et al. (1986) divided the reactor into three characteristics of UASB reactors. They found that the fluid
compartments (Fig. 4): sludge bed, sludge blanket and flow behaviour is very sensitive to the volume of zones and
settler. The liquid flow in the sludge bed and the sludge degree of bypassing between them but less sensitive to the
blanket were described by completely stirred tank reactor dead volume. The model assumed that some of the liquid
systems. Liquid flow in the internal settler was described by bypasses the first zone and enters directly into the second
a plug flow model. zone. The schematic representation is shown in Fig. 5.
ARTICLE IN PRESS
4 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

act as limiting substrates. Stewart (1956) and Agardy et al.


(1963) have used COD as the limiting substrate in the
Sludge Clarifier development of their models for the anaerobic digestion
Blanket
process. Lawrence and McCarty (1967) used volatile acids
concentration as the limiting substrate, since it is generally
believed that the rate-limiting step in the anaerobic
Zone-1 Zone-2
decomposition of complex organics is the conversion of
Fig. 5. Schematic of the two-compartment axial dispersion model volatile acids to methane. Andrews (1969) has presented a
(Singhal et al., 1998). dynamic model for the anaerobic digestion process, which
considers only the methanogenesis step. It is commonly
observed in the field that a high volatile acids concentration
They had verified the model efficacy through tracer and low pH value leads to digester failure. To represent this
studies. The step response of an axially dispersed tubular phenomenon in a quantifiable way an inhibition function
flow reactor for an inert tracer is described by the following (Haldane type) was used in the model. The un-ionized form
partial differential equation: of volatile acids was considered as the rate limiting
substrate since it is a function of both pH and the total
1 q2 C qC qC
 ¼ , (2.3) volatile acids concentration.
Pe qZ2 qZ qy Rittman and co-workers have presented a biofilm model
where C is the dimensionless concentration of the tracer describing the biofilm growth and substrate flux under
(c/c0), Pe is the Peclet number (uL/D), Z is the dimension- steady-state conditions (Rittmann, 1982; Rittmann and
less distance (z/L), y is the dimensionless time (t/t), u is the McCarty, 1980a). UASB reactor models using methano-
linear average velocity of the liquid in the reactor, L is the genic biofilm models have focused on mass transfer
length of the reactor, and D is the axial dispersion limitations and single as well as multiple limiting substrates
coefficient. The simulation and experimental results gave (Alphenaar et al., 1993; Atkinson and Davies, 1974;
a very good fit validating this flow model. Atkinson and How, 1974; Buffiere and Steyer, 1995; de
All the above multi-compartment models were capable Beer et al., 1992; Lens et al., 1993; Lin, 1991).
of fitting laboratory scale experimental data well. In the Thick biofilms may give rise to mass-transfer limitations
case of full-scale reactors, the feed distribution is very for the substrate within the biofilm, resulting in an overall
different from that in the laboratory scale reactor. Distinct reduction in the substrate conversions achieved. Under this
zones of sludge bed, sludge blanket and clarifier may not be condition, the transport of substrate into the biofilm or
observed due to the large volume of the reactor and uneven transport of products out of the biofilm could become the
flow distribution. This will make the task of quantifying the rate-determining step. Contradictory results have been
extent of non-ideality in the flow patterns difficult. It may reported on the impact of internal (within the biofilm) and
not be correct to predict the fluid flow pattern based on lab external (in the bulk liquid phase) mass transport. While
scale studies alone. Hence, further investigations are some reports indicate that both internal and external
required to test the models for full-scale reactors. diffusion limitations influence the rate of substrate utiliza-
tion (Dolfing, 1985; Wu et al., 1995), others indicate that
2.2. Reactor model mass transport limitations are not observed in anaerobic
biofilms (de Beer et al., 1992; Schmidt and Ahring, 1991)
Substrate degradation in the bioreactor is dependant on even at film thicknesses as high as 2.6 mm (Droste and
the observed reaction rate. In a biofilm type of reactor set- Kennedy, 1986). In addition, it has been reported that pH
up, the observed reaction rate is normally less than the rate profiles inside anaerobic granules may affect the conversion
predicted by the reaction kinetics. The reason for this is rate more than mass transport limitations (de Beer et al.,
that the substrate concentration actually available to the 1992).
microorganisms is less than the substrate concentration Gonzalez-Gil et al. (2001a) studied the influence of
present in the bulk liquid. This is due to the control exerted external and internal mass transport in anaerobic granular
by the mechanism of molecular diffusion on the penetra- sludge. For this, kinetic properties of acetate degrading
tion of substrate into the biofilm. It is, therefore, important methanogenic sludge granules of different mean diameters
to analyse the rate-determining step while developing were assessed at different up-flow velocities. It was
models. Hence, the reactor model consists of 3 parts: experimentally found that external mass transport resis-
substrate utilization within the biofilm (granules), mass tance can be neglected and that biogas formation does not
transfer and transport in the granule bed and blanket, and influence the diffusion rates. The substrate uptake could be
mass transport within the clarifier. explained by biofilm (internal) diffusion.
Kinetic models are based on the relationship between The general approach to the modelling problem starts
limiting substrate concentration and growth rate as with development of the biofilm model. These are then
proposed by Monod (1950). James (1961) pointed out that used to find the substrate flux at the surface of each
in biological processes a wide variety of substances might granule. The biofilm model is then tailored with the flow
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 5

model to predict the overall performance of the reactor. In where Ar is the total granule surface area; Tin the substrate
general, the following assumptions were made in develop- injection time for the acetate impulse; kl Ar ðsb  sð0; tÞÞ
ing anaerobic biofilm models: describes the substrate transferring from the bulk solution
through the liquid boundary layer into the granules. Data
1. Granules are spherical. from an acetate impulse experiment was simulated using
2. Substrate utilization is described by the Monod model. these hydraulic- reaction–diffusion models with a good fit.
3. External mass transport is negligible. The results indicated that diffusion inside the granules,
4. Density of the biofilm is constant throughout the reaction kinetics and hydraulic behaviour play important
granule. roles in the performance of a UASB reactor.
5. The relative concentrations of the acidogenic and Bolle et al. (1986) developed an integrated dynamic
methanogenic bacteria are constant throughout an model for the UASB reactor. Their flow model has been
anaerobic granule. This remains true irrespective of described in Section 2.1. Substrate and biomass balance
the location in the reactor from where the granule was were done over the sludge bed and sludge blanket. Thus the
taken. model was developed by integrating the fluid flow pattern
in the reactor, the bacterial growth and substrate utiliza-
tion kinetics, and the mass transport between different
Many of the reported studies deal with steady-state
compartments and different phases. This model was able to
conditions. Unsteady state is the most critical situation for
predict the sludge bed height, the biomass concentration in
modelling, associated with real-time control strategies. Wu
the sludge blanket, the short-circuiting flows over the bed
and Hickey (1997) developed a dynamic model to describe
and the blanket, and the effluent COD concentrations as a
UASB reactors with methanogenic anaerobic granules.
function of the hydrodynamic load, COD load, pH and
They assumed that substrate diffuses into the granules’
settler efficiency.
outer layer upto a thickness of 100 mm. At the inner edge of
Skiadas and Ahring (2002) proposed a model for
this layer, the substrate gradient reaches zero. Growth of
UASB reactors based on the cellular automata (CA) concept.
acetate utilizing methanogens is neglected due to their low
A cellular automation is a simulation, which is discrete in
growth rate.
time, space and state. A CA model usually consists of an
By applying an acetate mass balance on anaerobic
array of compartments similar to the spaces in a game of
granules, the granular bed and the clarifier, respectively,
naughts and crosses. The CA theory has been applied to
the dynamic model equations are obtained.
predict the layer structure of the granules, which is high
For anaerobic granules:
acidogen and low methanogen concentrations at the outer
 2 
qs qs 2 @s km xm s granule layers and the reverse at the inner granule layers. It
¼D   . (2.4) has also predicted the granule diameter and granule
qt qx2 R  x @x ks þ s
microbial compositions as functions of the operational
Boundary conditions: parameters. The other models do not consider the effect of
operational parameters on granule size and composition.
1. D qs þ kl sb ¼ kl sb ; x ¼ 0, Details on automation models can be found in the excellent
qx review paper by Wimpenny and Colasanti (1997).
Even though different authors have presented good
2. qs ¼ 0; x ¼ d, experimental fit for their models, a unified approach is
qx lacking. The above kinetic models, in general, assume
external mass transfer to be negligible. Some of the
where R is the granule radius; km the specific substrate assumptions may lead to poor predictions. For example,
utilization rate; ks the half velocity constant; D the effective the assumption of Monod kinetics, spherical shape and
diffusion coefficient; kl the mass transfer coefficient; xm the uniform density of the granules may not be true in actual
active acetate utilizer biomass within the outer layer d; s the reactors. The lack of versatile models in the literature
acetate concentration within the biofilm. shows that a lot of improvements can be made in future
A mass balance on a granule bed involves three terms: models. Some of them could be
substrate entering the reactor, substrate leaving the reactor,
and substrate being taken up by the granules and
subsequently being utilized: 1. Incorporation of the variation of density within the
ds biofilm/granule.
Vb ¼ V b EðtÞ  Qf sðtÞ  kl Ar ðsb  sð0; tÞÞ. (2.5) 2. Kinetic expressions, which include inhibition terms.
dt
3. Kinetic expressions, which include the non-uniform pH
Initial condition: sb(0) ¼ sb0 profile inside the biofilm.
4. Diversity in the bacterial population distribution
EðtÞ ¼ VM f
b T in
if 0ptpT in for impulse;
(2.6) inside the granule in terms of the predominant species/
¼0 if T in p0; groups.
ARTICLE IN PRESS
6 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

3. Proposed models for the structure of the biofilm Acidogens

In an anaerobic reactor, under favourable conditions, Methanothrix


bacterial cells can attach to each other and grow as a
H2 producing Acetogens and
granule. This granule has a 3-D structure. When some inert
H2 consuming organisms
carrier particles or surfaces are introduced in the system,
cells grow on the carrier particles/surfaces forming a
biofilm. Depending upon the structure of the carrier Fig. 6. Three-layered structure of the bacterial aggregates (MacLeod
material the resulting biofilm will have a 2-D or 3-D et al., 1990).
structure. Microbial composition and their distribution
inside the granules play an important role in substrate
conversion. Conversion could be limited by the process of essential. The layered structure of UASB granules is
diffusion of the substrate within the biofilm as well as the supported by the works of Arching et al. (1993) and Lens
concentration of biomass bringing about the conversion. et al. (1995) with immunological and histological methods.
This information is very essential to develop the biofilm The dynamic model proposed by Arcand et al. (1994) also
model. Many authors have reported that the structure of supports this structure. Similar support for the layered
the granule/biofilm is mainly determined by the substrate structure is found in reports from Santegoeds et al. (1999)
degradation kinetics rather than the geometry with which it using microelectrodes. Sekiguchi et al. (1998, 1999) and
is formed (Batstone et al., 2004; Fang et al., 1995; Guiot et Tagawa et al. (2000) also confirmed this structure by
al., 1992). Batstone et al. (2004) have developed a biofilm fluorescence in situ hybridization using 16S rRNA targeted
structure model which predicts the structure of a 2-D oligonucleotides. A distinct layered structure was also
biofilm. This model has been formulated based on found in the methanogenic–sulfidogenic aggregates, with
substrate degradation and diffusion kinetics. They exam- sulfate-reducing bacteria in the outer 50–100 mm and
ined four different types of granules obtained from UASB methanogens in the inner part (Sekiguchi et al., 1998).
reactors treating wastewaters from a cannery, a slaughter- Unlike the initial multi-layer model proposed by MacLeod
house, and two breweries. The microbial structures of these et al. (1990) recent research showed that UASB granules
granules were assessed by fluorescence in situ hybridization have large, dark, non-staining centres, in which neither
probing with 16S rRNA-directed oligonucleotide probes, archaeal nor bacterial signals could be found (Rocheleau
scanning electron microscope and transmission electron et al., 1999). In fact, the non-staining centre in the UASB
microscope. The biofilm model could exactly predict the granules might be the result of the accumulation of
structures of different types of granules observed through metabolically inactive, decaying biomass and inorganic
the experiments when the model was simulated for the materials (Sekiguchi et al., 1998).
same operating conditions. This proves that the structure The importance of ECP in anaerobic granulation has
of the biofilm is influenced by the substrate kinetics and not been observed by Schmidt and Ahring (1994, 1996). They
by the geometry of the biofilm. Van Loosdrecht et al. conclude that ECP may play an important role in building
(2002) has shown that a 2-D model is sufficient to represent spatial structure and maintaining the stability of UASB
a 3-D structure. Picioreanu et al. (2001) also has supported granules, but are unsure of its contribution to the initiation
the same. Hence, the different models proposed for the of anaerobic granulation. A large quantity of ECP is
biofilm structure in this section are applicable to both unnecessary for making up active granules since it
granules and 2-D biofilms. decreases the porosity and thereby increases the diffusional
resistance for substrate to penetrate the granule. Too much
3.1. Multi-layer model ECP could even cause deterioration of floc formation
(Schmidt and Ahring, 1996).
MacLeod et al. (1990) and Guiot et al. (1992) proposed a
three-layered structure for the bacterial aggregates treating
carbohydrates in a UASB reactor. According to this 3.2. Syntrophic microcolony model
model, the microbiological composition of granules is
different in each layer. The inner layer mainly consists of According to the syntrophic microcolony model, a close
methanogens that may act as nucleation centres necessary synergistic relationship among different microbial groups is
for the initiation of granule development. H2-producing essential for efficiently breaking down the complex organic
and H2-utilizing bacteria are dominant species in the compounds. In fact, the syntrophic microcolonies provide
middle layer, and a mixed species including rods, cocci and the kinetic and thermodynamic requirements for inter-
filamentous bacteria takes the predominant position in the mediate transference and therefore efficient substrate
outermost layer (Fig. 6). conversion (Schink and Thauer, 1988). They state that
To convert a target organic compound to methane, the the synergistic requirements would drive bacteria to form
spatial organization of methanogens and other microbial granules, in which different species function in a synergistic
species in the granules (such as those in a UASB reactor) is way and can easily survive.
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 7

3.3. Non-layered structure model 3.4. Granule cluster structure

Contrary to the multi-layer model, anaerobic granules Recently the structure of anaerobic granules of an
with non-layered structure have also been reported (Fang et expanded granular sludge bed (EGSB) reactor was studied
al., 1995; Grotenhuis et al., 1991; Wu et al., 2001). by Gonzalez-Gil et al. (2001b). They found black spherical
There is evidence that a layered structure of the UASB granules having numerous whitish spots on their surfaces.
granules would be developed with carbohydrates and a non- Cross-sectioning these aggregates revealed that the whitish
layered one using substrates having a rate-limiting hydrolytic spots appeared to be white clusters embedded in a black
or fermentative step (e.g. proteins) (Fang, 2000; Fang et al., matrix (Fig. 7). High magnification electron microscopy
1995). This is probably due to different initial steps in the showed that the white clusters mainly consisted of acetate
degradation of carbohydrate as compared utilizing methanogens (Mathanosaeta spp.) and the black
to that of proteins. Degradation of carbohydrates to matrix consisted of syntrophic species and hydrogenotrophic
small molecules is faster than the subsequent degradation methanogens (Methanobacterium like and Methaospirillum
of the intermediates, whereas the initial step in protein like organisms). Fluorescent in situ hybridization using 16 s
degradation is a rate-limiting step. That is, the protein rRNA probes of crushed granules showed that 70% of the
degradation step is slower than the steps which follow it. cells belonged to the archaebacterial domain and 30% to
Results from fluorescence in situ hybridization com- eubacterial domain. The authors have provided a very logical
bined with confocal scanning laser microscopy clearly explanation as to why a cluster structure, as observed, is
showed that protein-fed granules possess non-layered advantageous over a layered structure. Acetate produced in
structure with a random distribution of Methanosaeta concilii the black zone is transported by random diffusion in all
(Rocheleau et al., 1999). However, granules, which differ in directions and thus penetrates the Methanosaeta clusters from
their composition in terms of the predominant microbial all sides. Hence, substrate depleted zones are circumvented,
species, can still be formed from the same substrate which allows the growth of more active biomass per unit area
(Daffonchio et al., 1995; Schmidt and Ahring, 1996). of aggregate.
Based on microscopic examination of the UASB Batstone et al. (2004) studied the influence of substrate
granules, recently Fang (2000) proposed that the microbial degradation kinetics on the microbial community structure
distribution of the UASB granules strongly depends on the in granular anaerobic biomass. The granules, which were
degradation thermodynamics and kinetics of individual grown in the effluent containing soluble as well as
substrate. Therefore, it appears that different dominating particulate protein, had homogeneous structure. The
catabolic pathways may give rise to granules, which are primary cause of this structure was assessed through
different in their structure. Spontaneous and sudden biofilm modelling. They postulate that the particulate
washout of the established granular sludge bed, as a result nature of the wastewater and the slow rate of particulate
of a change in wastewater composition, is a common hydrolysis, rather than the presence of proteins in the
problem encountered in the operation of UASB systems. wastewater, was responsible for the homogeneous structure
So far, none of the individual models for the granule of the granules. Because solids hydrolysis was rate limiting,
structure can explain this phenomenon. If a factor that is soluble substrate concentrations were very low (below
independent of the wastewater composition can initiate the Monod half-saturation concentration), which caused low
formation of UASB granules, a change in the wastewater growth rates.
composition should not lead to the washout of the entire From the above information it can be concluded that
granular sludge bed. Thus, it is a reasonable speculation there are two key factors which determine the structure of a
that there should be a substrate composition-associated granule (i.e. the organization of the microbial community
factor that highly contributes to the formation of UASB within a granule). They are
granules, but is not yet included in the present models of
granule structure (Liu et al., 2003). 1. The nature of organic compounds present in the
Liu et al. (2003) proposed a general model for anaerobic wastewater.
granulation in UASB reactors. This model consists of four 2. The kinetics of substrate degradation.
steps for granulation.
Methanosaeta clusters

1. Physical movement to initiate bacterium-to-bacterium Zone with syntrophic


contact or bacterial attachment onto nuclei. eubacteria and
hydrogenotrophic
2. Initial attractive forces to keep stable multi-cellular
methanogens
contacts, e.g. physical, chemical and biochemical forces.
3. Microbial forces to make cell aggregation mature, e.g. Seed aggregate
production of ECP.
4. Hydrodynamic shear force shaping 3-D structure of Fig. 7. Schematic representation of the architecture of anaerobic
microbial aggregates. aggregates (Gonzalez-Gil et al., 2001b).
ARTICLE IN PRESS
8 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

The relative concentrations of different microbial species Di Felice, 1995), have to be modified to take into account
within a granule is another important factor required in the effect of the biofilm layer on the rigid particles
developing a biofilm model. This seems to depend on the (Nicolella et al., 2000).
concentration of the substrate. Since there could be
concentration gradients with respect to substrate concen- 4.1.1. Terminal settling velocity
tration within the same reactor, the composition of the The terminal settling velocity of a single spherical
granule in terms of the microbial species may vary from particle in an infinite expanse of fluid is
granule to granule obtained from different parts of the  
same reactor. Hence a robust model, capable of incorpor- 4gðrs  rl Þ 0:5
ut ¼ . (4.1)
ating the structure as well as composition of the granules in 3C D rl
terms of the different microbial communities present, is Depending on the biofilm thickness and on the carrier
required. Such a model for the biofilm/granule can be used type, values for the equivalent density of biofilm particles
in developing reactor models which are more versatile as (rs ) range typically from 1100 to 1500 kg/m3. The drag
well as reproducible with a higher level of accuracy. coefficient CD is a function of the particle Reynolds
number defined as
4. Models for AFBR
rl d s ut
Ret ¼ . (4.2)
The AFBR uses inert carrier particles to provide ml
mechanical support for growth of the biofilm. These Biofilm particles are in the intermediate flow regime
biogranules are maintained in a fluidized state by using (1oReto100) for the vast majority of cases, obtained
the energy of the influent liquid. Such a fluidized bed when sand (0.5–1 mm) or carrier materials with densities in
system is free from the channelling problems encountered the same range are used as the inert support. In the
in other biofilm reactors. A serious operational problem intermediate flow regime, the drag coefficient for a smooth,
associated with the AFBR is that when the biofilm grows rigid sphere is (Perry and Green, 1997)
on the carrier surface, the composite density of the film-
covered particle decreases, ultimately resulting in the carry C D ¼ 18:5Ret0:6 . (4.3)
over of the film-covered particles out of the reactor. Full- This correlation cannot be used for carrier-supported
scale application of AFBRs is not common due to lack of granules, since they are neither smooth nor rigid. For this
sound design principles. Many attempts have been made to reason, empirical correlations have been suggested for the
study the process kinetics and the factors affecting process estimation of CD for biofilm particles:
performance.
In general, models for AFBRs include the following C D ¼ 17:1Ret0:47 ; 50oRet o100 (4.4)
elements: Hermanovicz and Ganczarczyk (1983),

1. A bed fluidization model which describes the size and C D ¼ 36:66Ret0:67 ; 40oRet o90 (4.5)
number of bioparticles per unit fluidized bed volume. Mulcahy and Shieh (1987),
2. A biofilm model which describes the rate of substrate
24
conversion per individual granule. CD ¼ þ 21:55Ret0:518 ; 15oRet o87 (4.6)
3. A reactor flow model, which links the biofilm and bed Ret
fluidization models to yield substrate concentration as a Ro and Neethling (1990),
function of axial position within the AFBR. 24
CD ¼ þ 14:55Ret0:48 ; 40oRet o90 (4.7)
Ret
4.1. Bed fluidization model Yu and Rittmann (1997),
C D ¼ 29:6Ret0:6 ; 7oRet o90 (4.8)
Hydrodynamic behaviour of carrier supported biogra-
nules plays an important role in designing AFBRs. When Nicolella et al. (1999).
the granules grow, their size, shape and composite density The drag coefficient of the biofilm covered carrier
change. This has an impact on the hydrodynamic particle is generally found to be more than that of smooth,
behaviour of the granules. Information on settling and rigid spheres. Surface roughness has generally been
fluidization characteristics, such as fluidized bed-height, as considered as the reason for the increase in the drag
a function of liquid velocity is required for the designing of coefficient of biofilm covered carrier particles (Hermano-
AFBRs. Information on fluidized-bed height is important vicz and Ganczarczyk, 1983; Mulcahy and Shieh, 1987).
because it establishes the solids residence time and the The above correlations are all empirical since they were
specific biofilm surface area in the biologically active zone. obtained by fitting experimental data generated by
The relationships valid for fluidization of rigid particles, different groups, who employed fluidized bed systems,
readily available in the chemical engineering literature (e.g. which differed from one another in one or more aspects.
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 9

The deformable nature and surface roughness of the A practical working expression was derived based on the
biofilms and the type of carrier particle used play a major above two expressions by Andrews and Tien (1979) as
role in determining the hydrodynamic nature of the biofilm follows:
covered particle. Hence, bed fluidization models should be H
able to incorporate those parameters which have an ¼ 1 þ ð1 þ BÞx̄, (4.18)
Hc
influence on the hydrodynamic behaviour of the biofilm
covered particles. It may not be feasible to have an where
analytical expression for this purpose, given the complexity     
ec D 1 þ ec 1  3A2 1  3A
and variability inherent to the system. While an empirical B¼ 3 þ x0 2 þ D  D¼ ,
3ð1  ec Þ 1  ec 1  3A 3n
relationship is acceptable, it should be able to cater to a
wider range of operational parameters normally encoun- where H is the bed height, e the bed porosity, x the film
tered rather than a narrow set of operating conditions. volume/clean particle volume, x̄ the mean value of
distribution function of bacterial film gðxÞ, A the buoyant
4.1.2. Fluidization mechanics density of bacterial film/buoyant density of clean particle, n
In a fluidized bed anaerobic reactor, the film-covered the exponent in the Richardson–Zaki correlation. The
particles are kept in a fluidized state by the incoming liquid. subscript ‘c’ denotes the bed of clean particles.
The bed porosity and biomass concentration in the bed are This expression was found to predict the experimental
determined by the mechanics of fluidization. Hence, a data well.
realistic mathematical expression for the bed fluidization is Mulcahy and La Motta (1978) have developed specific
necessary. correlations to determine ut and n for the case of
For a bed of uniform spherical particles, the following bioparticles in a fluidized bed as follows:
relation was proposed (Richardson and Zaki, 1954): " #0:75
ðrs  rl Þgd 1:67
us ¼ ui ðeÞn , (4.9) ut ¼
p
, (4.19)
27:5r0:33
l m0:67
where us is the superficial liquid velocity, e ¼ porosity and
ui ¼ ut 10d=D , (4.10) n ¼ 10:35 Re0:18
t ð40oRet o90Þ, (4.20)
where ut is the terminal settling velocity of particle, d the where rs is the density of bioparticles; rl the liquid density;
particle diameter and D the diameter of bed, n is a constant m the liquid viscosity.
given by Ngian and Martin (1980) found that the Richardson and
Zaki correlations gave a satisfactory estimate for ui for
n ¼ 4:65 þ 20d=DðRet o0:2Þ, (4.11)
small support particles whereas ui was 30–70% below the
n ¼ ð4:4 þ 18d=DÞðRet Þ0:03 ð0:2oRet o1Þ, (4.12) experimentally determined value for larger support parti-
cles. Nicolella et al. (1999) found ui to be only 80% of the
n ¼ ð4:4 þ 18d=DÞðRet Þ0:1 ð1oRet o200Þ, (4.13) unhindered settling velocity. They recommend that the
correlation should be used with caution while determining
n ¼ 4:4ðRet Þ0:1 ð200oRet o500Þ, (4.14) the constant ui.
In general, the Richardson and Zaki correlation is found
n ¼ 2:4ð500oRet Þ. (4.15) to describe the bed expansion characteristics of a fluidized
bed. But the values of the constants n and ui seem to be a
Whether this equation can be applied directly and in its
function of the property of the biofilm covered particles.
entirety to a biological fluidized bed is a controversial
The application of these correlations to a full-scale reactor
subject. Richardson and Zaki have derived the equation for
containing a wide size distribution of biofilm covered
a bed of uniform spherical particles, which are hard and
particles needs to be verified.
have a smooth surface. But the biological film covered
particles seldom have these characteristics.
Andrews and Tien (1979) related the bed height to the 4.1.3. Effect of gas production on hydrodynamics
amount of biomass in the bed. The fluidized bed tends to The effect of gas production on hydrodynamics of
stratify vertically based on the settling velocity of the fluidized beds is an important factor to be studied for
bioparticles. If no stratification is assumed, the bed height design and scale-up of AFBRs. Many investigations on the
is related to average biofilm thickness x̄ by flow pattern in an AFBR suggested that an axially
dispersed plug flow model can be used for the flow model
H ð1  ec Þð1 þ x̄Þ (Hirata et al., 2000; Seok and Komisar, 2003). In these
¼ . (4.16)
H c 1  ec ½ð1 þ x̄Þ1=3 =ð1 þ Ax̄Þ1=n studies the effect of gas production on the flow pattern was
In the case of complete stratification not considered. Diez-Blanco et al. (1995) have studied the
Z xmax effect of gas production on the hydrodynamic behaviour of
H ð1 þ xÞ an AFBR. In this study, the bed contraction due to biogas
¼ ð1  ec Þ gðxÞ dx. (4.17)
Hc 0 1e production in a fluidized bed of 6 m height was estimated to
ARTICLE IN PRESS
10 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

be less than 6%. Based on this, the investigators considered 4.2. Kinetic and reactor sub-models
the effect of biogas on the hydrodynamic behaviour to be
negligible. In contrast, Buffiere et al. (1998a) reported that Substrate removal within a biological film of uniform
the effect of biogas on bed height could be negligible. But thickness attached to a spherical particle is described by
the biogas effervescence affects the phase hold-ups (which  
D d 2 ds
affects the liquid solid contact time) and liquid mixing of r ¼ Rt , (4.21)
the reactor system. The experimental study of Buffiere et al. r2 dr dr
(1998b) compared the hydrodynamic behaviour of a gas where D is the effective diffusion coefficient of substrate
producing fluidized bed with a classical gas injected three within the biological film; r the radial coordinate measured
phase fluidized bed reactor. The overall gas hold up was from the centre of the support particle; s the substrate
found to be more in the case of the gas producing fluidized concentration within the biofilm and Rt the intrinsic rate of
bed than the gas injected one. They also observed axial substrate consumption per unit volume of biological film.
variation of phase hold-up. It is generally accepted by many authors that substrate
removal kinetics can be modelled using the microbial
growth model proposed by Monod. Many researchers
4.1.4. Bed stratification (Grady, 1982; Henze and Harremoes, 1983; Iwai and
Schreyer and Coughlin (1999) reported a stratification of Kitao, 1994) have suggested that depending on the values
biofilm coated sand particles in a fluidized bed reactor. of the model constants, reaction rate can be represented by
Stratification can be attributed to the influence of a biofilm first- or zero-order kinetics. For AFBRs, many authors
on a particle’s settling velocity. The presence of a biofilm demonstrated that zero-order kinetics provide an adequate
cover decreases a particle’s overall density, thereby description of substrate consumption (i.e. considering the
increasing its buoyancy. The biofilm also increases the acidogenesis and methanogenesis phases together) (La
particle’s size thereby increasing the drag force exerted on Motta and Patricio, 1996; Mulcahy and La Motta, 1978;
it by the liquid flowing upward. The particles in a fluidized Mulcahy et al., 1980; Shieh et al., 1982).
bed are expected to segregate according to size and mean For zero-order reaction, the reaction rate term is
density (Ro and Neethling, 1994; Safferman and Bishop,
1996; Trinet et al., 1991). Rt ¼ rk0 . (4.22)
Bed stratification has many negative effects on the The solution of Eq. (4.21) for complete substrate
performance of the reactor. Thicker biofilms pose diffusion penetration inside the biofilm and for the partial penetra-
limitation and wash out problems. Hence, to prevent tion has been presented by Mulcahy et al. (1980).
stratification and to maintain uniform particle size, efforts For fully penetrated biofilm:
have been made to remove excess biofilm from larger
particles (Ruggeri et al., 1994; Safferman and Bishop, 1996; Observed rate ¼ 43prk0 ðr3p  r3m Þ. (4.23)
Shieh et al., 1981; Trinet et al., 1991). Examples of such For partial substrate penetration:
efforts include external sand-biomass separators (screens),
operation of an impeller at the top of the bed and internal ðr3p  r3m Þ1:9
screen cleaning devices. Many researchers have examined Observed rate ¼ 1:76ðrk0 Þ1:45 ðr3p  r3m Þ1:9 s0:45
b 0:45
.
r1:8
p D
the effect of shear on biofilm and biofilm detachment rate,
mainly as a tool to control biofilm thickness (Chang and (4.24)
Rittmann, 1988; Chang et al., 1991; Gjaltema et al., 1997; rm is the radius of support particle; rp the radius of
Peyton and Characklis, 1992; Rittmann, 1982; Safferman biological particle; sb the concentration of substrate in the
and Bishop, 1996; Stewart, 1993; Trinet et al., 1991). bulk of the liquid within the fluidized bed.
Biofilm detachment rate appears to depend on turbulence A simple plug flow model was used to describe dissolved
and particle concentration; an increase in either increases substrate transport in the axial direction in a fluidized bed
detachment rate. Biofilm has been observed to be relatively reactor (Mulcahy and La Motta, 1978):
smoother and more homogeneous under conditions of high
dsb
shear (i.e., high liquid velocity) than under low shear U þ Rv ¼ 0. (4.25)
conditions (Lau, 1995; Zhang and Bishop, 1994). The study dz
by Schreyer and Coughlin (1999) showed that the With the boundary condition
introduction of a thinner to increase the shear resulted in
z ¼ 0; sb ¼ s0
a non-stratified bed.
Bed stratification could occur due to differences in the where U is the average liquid velocity in the longitudinal
biofilm thickness or differences in the carrier particle size or direction; sb the substrate concentration in the bulk of the
both. But bed stratification is most common when carrier liquid; s0 the influent substrate concentration; Rv the rate
particles are not of uniform size. A completely mixed bed of reaction.
(i.e. no stratification) was observed by Andrews and Tien For fully substrate penetrated biofilm, the concen-
(1979) when uniform carrier particles were used. tration profile along the reactor was given by Mulcahy
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 11

et al. (1980) as Eq. (4.29) gave


" 3 #  
Vm rp sb sss
se ¼ s0  rk0 1 , (4.26) Rt ¼ K , (4.32)
Q rm km þ sss
where K ¼ ðrb dmmax =Y x=s Þ:
where se is the concentration of substrate in the effluent
Transforming the above equation using Lineweaver–
stream. La Motta and Patricio (1996) tested this model by
Burke linearization,
conducting experiments. This test showed a reasonable
agreement between the zero-order kinetic model with sb km 1 1
¼ þ . (4.33)
complete substrate penetration and the experimental data. Rt K sss K
Hirata et al. (2000) used a different approach to evaluate
Plotting sb/Rt versus 1/sss gave a straight line with slope
the kinetic parameters of the biochemical reactions taking
km/K and intercept 1/K. Using the above plot for the
place in a three phase fluidized bed reactor. Using the
known value of inlet substrate concentration, the steady-
substrate balance at steady state and assuming Monod
state substrate concentration could be found.
kinetics, an equation relating the substrate consumption
Buffiere et al. (1998c) studied the biofilm activity along
rate to substrate concentration (expressed as Biochemical
the reactor height. They developed a biofilm model
Oxygen Demand, BOD5) and total biofilm surface area was
assuming
established. The following assumptions were made in
formulating the model:
1. Homogeneous biofilm of uniform thickness.
2. Spherical support media of uniform size.
1. Reactor system is completely mixed.
3. Internal mass transfer described by Fick’s law.
2. Total organic carbon (TOC), which is expressed in terms
4. Liquid phase perfectly mixed with homogeneous con-
of BOD5, is the only rate-limiting substrate. Other
centration.
substrates are in excess.
5. No external mass transfer limitation.
3. The reaction follows Monod kinetics and substrate
inhibition is negligible.
4. Reaction occurs at constant volume. The mass balance for a substrate s in the biofilm is
expressed by
  X
Performing a substrate balance at steady state yielded D d 2 ds
r ¼ Vsi . (4.34)
1 r2 dr dr i
F ðsin  sss Þ ¼ rx V , (4.27)
Y x=s The boundary conditions are
where F is the inlet flow rate, V the reactor volume, rx the r ¼ rp ; s ¼ s0 ,
biofilm growth rate, Yx/s the yield coefficient ¼ mass of
biomass formed/mass of substrate consumed, Sss the ds
r ¼ rc ; ¼ 0,
steady-state value of the rate limiting substrate concentra- dr
tion inside the reactor, Sin the inlet substrate concentration.  
If the reaction followed Monod kinetics, then the proposed xs mmax s
Vs ¼ . (4.35)
rate equation was Y x=s ks þ s
 
mmax sss The right-hand side of (4.34) is the sum of all substrate
rx ¼ mx ¼ x, (4.28)
km þ sss uptake rates minus the sum of all substrate production
rates.
where m is the specific growth rate, mmax is the maximum
In the liquid phase, the mass balance for one substrate s
specific growth rate and km the Monod constant.
is given by
Substituting rx into the substrate balance equation
   !
1 m max sss ds ds
F ðsin  sss Þ ¼ V x ¼ Rt , (4.29) V L ¼ Qðsin  sÞ þ DAp   , (4.36)
Y x=s km þ sss dt dr r¼rp

where s is the concentration of component s; sin the inlet


V x ¼ rb dsb , (4.30)
concentration of s; Q the input flow rate; rp the bioparticle
where rb is the biomass dry density, d the effective biofilm radius; r the radial distance measured from bioparticle
thickness and sb the total biofilm surface area. The total centre; D the diffusivity of component s in the biofilm; mmax
surface area was computed as follows: the maximum specific growth rate for s-utilizing bacteria;
Vsi the reaction rate of s through reaction I; Ap the
sb ¼ pðDave Þ2 N, (4.31)
exchange area of the bioparticles; VL the liquid phase
where N is the total number of particles inside the reactor volume; xs the s-utilizing bacteria concentration; Y x=s the
and Dave is the average particle diameter. Modifying biomass yield factor for s-utilizing bacteria.
ARTICLE IN PRESS
12 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

Table 1 Bed contraction was found to reduce the liquid–solid


Biomass composition of an anaerobic granule contact by 10–25%:
Type of bacteria Substrates for the %Distribution in the es
¼ 1 þ 0:045U 0:4
g . (4.38)
bacteria granule es0
Acidogens Glucose 65 Experimental results of gas hold-up in the reactor gave
Acetogens Butyrate 2.5 the following correlation:
Propionate 2.0
Methanogens Acetate 7.0 eg ¼ ð13  1:2Þd 0:168
p U 0:7
g . (4.39)
H2 utilizing bacteria Hydrogen 23.5
An axially dispersed plug flow model was found to
describe the liquid mixing in the reactor. The mass balance
The substrate gradient at the surface of a bioparticle in of the reactor was given by the equation
Eq (4.36) is taken from the biofilm model. The biomass 1 d2 s ds s
composition used in this simulation was averaged from 2
¼ þ Da (4.40)
Pe dx dx lþs
several studies such as the modelling work of Mosey (1983)
with the following notations:
and the experimental investigation of Sanchez et al. (1994)
(Table 1). z s rmax Hel U lH
x¼ ; s ¼ ; Da ¼ ; and Pe ¼ .
A set of batch experiments was conducted using H ks ks U l el E zl
bioparticles taken from different heights. Glucose, acetate The axial dispersion co-efficient was found experimen-
and propionate were used as substrates. Substrate con- tally using the tracer injection method. The experimental
sumption with time was monitored. Simulation results were results were found to fit the correlation of Muroyama et al.
found to give reasonable fit for experimentally observed (from Fan, 1989) very well:
values. The biomass composition (i.e. the relative concen-
tration of acidogens and methanogens) within the biofilm Dc U l
¼ 1:01U 0:738
l U 0:167
g D0:583
c , (4.41)
was manipulated to fit the experimental results. This el z
indicated the crucial role of biomass composition in where dp is the particle diameter; Dc the column diameter;
substrate kinetics. The specific activity of the biofilm was Ezl the axial dispersion coefficient; g the gravitational
also measured. The following were the findings: acceleration; H the bed height; ks the half-saturation
concentration in Monod expression; rmax the maximal
1. Thicker film bioparticles were found on the top of the reaction rate in Monod expression; s the substrate
reactor and thinner in the bottom. concentration; Ug the gas superficial velocity; Ul the liquid
2. Glucotrophic activity decreased from bottom to top and superficial velocity; Ut the terminal velocity of particles; x
methanogenic activity increased from bottom to top. the reduced bed height; es the solid hold up; es0 the solid
3. This indicates the change in biomass composition with hold-up in the liquid–solid fluidized bed; eg the gas hold-up.
biofilm size. Thus by knowing hold-ups (from Eq. (4.38) and (4.39))
and Ez (from Eq. (4.41)), the performance Eq. (4.40) was
Buffiere et al. (1998a) developed a model for AFBRs solved. The model results were found to give a more
based on total carbon removal kinetics. They considered realistic picture.
the effects of gas production in their model which were of
two kinds: 4.2.1. Stratified biofilm models
A stratified biofilm model was presented by Canovas-
1. Gas production modified the degree of axial mixing, Diaz and Howell (1988). The anaerobic biofilm was
which is responsible for the establishment of a concen- modelled as two distinct layers with the inner layer
tration gradient in the reactor. consisting of methanogens and the outer layer consisting
2. Gas production is responsible for bed contraction, of acidogens. The substrate is converted to acids in the
which reduces the contact between the liquid and outer layers, and is subsequently converted to methane by
bioparticles. the bacteria in the inner layer.
The differential equations for substrate uptake in the two
The TOC removal kinetics were found to be in good layers are:
agreement with the Monod model. The TOC uptake rate
d2 G
can be expressed by D1 ¼ k1 x1 , (4.42)
dz2
S
rTOC ¼ rmax , (4.37) d2 F k2 x2
ks þ S D2 ¼ ak1 x1 þ , (4.43)
where S is the TOC concentration in the reactor (the dz2 1 þ F =ki
reactor is assumed to be perfectly mixed). Parameters rmax where D1, D2 is the diffusivities of substrate through the
and ks were found by plotting the inverse of rTOC and 1/S. acidogenic and methanogenic layer. G, F the concentration
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 13

of glucose and fatty acids, k1 the zero-order kinetic the granular sludge bed in an expanded condition
constant with respect to sugar uptake and k1 the zero- (Zoutberg and Frankin, 1996). Reports on models for
order rate constant k2 the kinetic constant for VFA uptake; EGSB reactors are very scarce. But based on the knowl-
ki the inhibition constant; z the distance through biofilm; edge of UASB reactors and AFBR models, modelling of
x1, x2 the mass volume densities of acidogens and EGSB reactor can be attempted.
methanogens.
The authors also explain why a stratified biofilm is 1. The biofilm model is similar to the UASB reactor
advantageous as compared to an un-stratified biofilm. biofilm model. The composition and structure of the
When the bulk VFA concentration is high, in the case of an biofilm is expected to be influenced by the nature as well
unstratified film, methanogens will face VFA inhibition. In as concentration of the substrate. However, there is no
the case of stratified biofilm, the inhibition level is reason to believe that this influence will be significantly
minimized by the presence of the outer layer. different from what has been observed for biogranules
Droste and Kennedy (1986) have given a model for in the UASB reactor (Section 3.4). The only significant
sequential substrate utilization in a biofilm. This model difference will be the reduced level of the substrate
assumes that no interactions occur between the two groups concentration gradient along the height of the reactor
of microorganisms that will cause kinetic or diffusion due to the effect of recirculation.
parameters to change from values associated with indivi- 2. The liquid flow pattern could be expected to be
dual cultures of each group. Unstratified biofilm and somewhere between completely mixed and dispersed
Monod-type kinetics were assumed. plug flow, the exact pattern depending on the recycle
The governing differential equations are: ratio employed. However, any flow model employed will
need validation through tracer studies or any other
d 2 s1 k1 x1 s1
D1 2
¼ , (4.44) suitable experimental studies.
dx K 1 þ s1
3. A fluidization model which can predict the variation of
d 2 s2 k 2 x 2 s2 Yk1 x1 s1 bed height with upflow liquid velocity has to be
D2 2
 ¼ , (4.45) developed based on experimentation.
dx K 2 þ s2 K 1 þ s1
where D is the diffusivity; k the maximum specific reaction Tailoring of the above three models could result in a
velocity; K the half-velocity constant; s the substrate complete model for an EGSB reactor.
concentration; x the distance; x the active microorganism
concentration; Y the acetate yield coefficient (g acetate/g
6. The anaerobic biofilter
primary substrate).
From numerical analysis of the governing equation, it
The anaerobic filter is an anaerobic packed-bed biofilm
was found that the production of intermediate substrate in
reactor. Though both upflow as well as down flow
the biofilm increased the conversion of primary substrate to
configurations are possible, the upflow mode of operation
ultimate product. The increase was not always significant.
is more common. The biomass forms a film on the surface of
To summarize, two ways of approaching the problem of
the packing media. The kinetic model of such a biofilm
modelling substrate uptake kinetics are explained in this
reactor has been studied quite extensively (Chang and
section. In the first one, TOC is considered as the rate-
Rittmann, 1987; Hamoda and Kennedy, 1987; Meunier
limiting substrate. In this case biomass composition,
and Williamson, 1981; Rittmann and McCarty, 1980a;
individual reaction steps and substrate diffusion limitations
Rittmann and McCarty, 1980b). Fig. 8 illustrates the ideal
are not considered. The parameters involved in the model
biofilm having uniform microbial density (Xf) and uniform
are evaluated by fitting the model to the experimental
thickness (Lf). When the substrate utilization follows the
results. Even though this approach seems to be simple, it
Monod relationship, the biofilm model incorporating the
does not have a sound theoretical explanation and remains
external mass transfer and internal simultaneous diffusion
empirical. In the second approach, the kinetic model is
and reaction can be derived as follows (Huang and Jih, 1997):
developed considering the individual substrate kinetics,
biofilm composition and diffusion limitations. This seems d2 S f kX f S f
¼ . (4.46)
to be a more realistic approach. Ultimately the validity of dZ 2 ½Df ðK s þ S f Þ
these models has to be verified for large-scale reactors.
Boundary conditions:
5. The EGSB reactor dS f Ds
Df ¼ ðS b  S s Þ ¼ J at Z ¼ 0,
dZ L
The EGSB reactor comes under the family of UASB
reactors. The use of effluent recirculation in a UASB (or a dS f
¼ 0 at Z ¼ Lf ,
high height/diameter ratio), resulted in the EGSB reactor dZ
(Seghezzo et al., 1998). Here also, the biomass is present in where S is the biofilm substrate concentration, Z is the
a granular form. The higher upflow liquid velocity keeps distance normal to the biofilm surface, D is the diffusivity, Ds
ARTICLE IN PRESS
14 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

Outlet

Ss Biofilm

Sampling port Liquid


Gas phase Phase Sb Carrier
material

Lf

Inlet

Fig. 8. Schematic diagram of an anaerobic filter showing the differential section.

is the axial dispersion coefficient, Xf is the microbial density of Huang and Jih (1997) confirmed that the flow regime is
in the biofilm, Ks is the Monod constant, k is the maximum close to completely mixed.
specific substrate utilization rate, J is the substrate flux.
Here, the subscripts ‘f’ and ‘b’ denote biofilm and bulk 7. The unified model for anaerobic digestion (ADM1)
liquid, respectively.
The greatest difficulty in applying the model is to The IWA Anaerobic Digestion Modeling Task Group
measure the biofilm thickness (Lf). Rittmann and McCarty developed a generalized anaerobic digestion model (Bat-
(1978) and Suidan (1986) imposed another boundary stone et al., 2002). The biochemical as well as physico-
condition to the model, i.e.S f ¼ 0 at Z ¼ Lf chemical processes were included in the model. The
This model is called the deep-biofilm model. Another biochemical steps include (i) disintegration from homo-
reported approach for indirectly estimating the steady-state geneous particulates to carbohydrates, proteins and lipids,
biofilm thickness was by assuming that the biomass growth (ii) the extracellular hydrolysis of these particulate sub-
equals the biomass decay and/or shear losses in biofilm strates to sugars, amino acids, and long chain fatty acids
reactors. For instance, Rittmann and McCarty (1980a) (LCFA), respectively, (iii) acidogenesis from sugars and
computed the steady-state biofilm thickness using para- amino acids to volatile fatty acids (VFAs) and hydrogen
meters of the substrate flux, microbial growth, biomass (iv) acetogenesis of LCFA and VFAs to acetate and (v)
decay and sloughing rate (Eq. (4.47)). The model obtained, separate methanogenesis steps from acetate and hydrogen/
called the steady-state biofilm model, is given by CO2. The physico-chemical equations describe ion associa-
tion and dissociation, and gas–liquid transfer. The inhibi-
JY tion kinetics have been incorporated in the biochemical
Lf ¼ , (4.47) process.
bX f

where Y is the biomass yield coefficient and b is the 7.1. The extension of ADM1 to biofilm reactors
sloughing rate.
The models mentioned above have been experimentally The ADM1 model gives a unified approach to anaerobic
verified or simulated using the results reported in the digestion. This model was successfully implemented for a
literature (Chang and Rittmann, 1987; Liu et al., 1991; CSTR (Batstone et al., 2002). ADM1 was developed for a
Meunier and Williamson, 1981; Rittmann and McCarty, suspended cell system, where there is no mass transfer
1980b, Wang et al., 1987). The kinetic parameters included limitation for movement of the substrate from the bulk
in the models were either determined by independent liquid phase to the cells. On the other hand, in biofilm
experiments or obtained from published papers. The liquid reactors, the rate of transport of substrate from the bulk
flow regime was assumed to be completely mixed or plug- liquid to the microbial population is controlled by diffusion
flow with axial dispersion (dispersion coefficients were of substrate within the biofilm. Hence to extend the ADM1
calculated using empirical equations). An axial dispersion for biofilm systems, the substrate utilization kinetics of the
model coupled with deep-biofilm kinetics was used by single cell system must be replaced with a biofilm model.
Huang and Jih (1997). The experimental results and the The biofilm models for different biofilm reactor systems
calculated values showed good agreement. The tracer study have been extensively reported in the previous sections. In
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 15

a similar way the reactor flow system of ADM1 (i.e. CSTR granule is more than that of a rigid particle of the same
system) also has to be modified. In the case of a UASB size. The experimental evidence published so far (Andrews
reactor, the reactor flow system will have a combination of and Tien, 1979; Hermanovicz and Cheng, 1990; Mulcahy
CSTRs and a plug flow system as explained in Section 2.1. and Shieh, 1987; Ngian and Martin, 1980; Nicolella et al.,
AFBR and EGSB reactors will have axially dispersed plug 1999) suggests that the fluidized bed expansion pattern is
flow systems as shown in Sections 4.2 and 5. A biofilter observed to follow the Richardson and Zaki correlation.
flow system is similar to that of ADM1. The physico- The studies show that the expansion index can be
chemical process system of ADM1 can be directly extended calculated through the Richardson and Zaki correlation
to the biofilm reactor systems. Thus, the modification of (Nicolella et al., 1999). But the parameter ui was found to
the biochemical processes and the reactor flow system of be 30–80% of the experimentally determined ut value.
ADM1 with that of the biofilm system could result in a All these studies were conducted with uniform biogranule
unified model for biofilm reactors. sizes.
Many authors (Ro and Neethling, 1994; Safferman and
8. Summary and conclusions Bishop, 1996; Schreyer and Coughlin, 1999; Trinet et al.,
1991) observed that the bed becomes stratified due to the
Development of biofilm reactors has made anaerobic presence of granules with different biofilm thicknesses. In
treatment an attractive option to treat wastewaters. For such a context, the application of the Richardson and Zaki
optimum design and scale-up of these reactors, mathema- correlation remains dubious. Hence models considering
tical models are required. In this paper, various parameters expansion characteristics of stratified beds are necessary
affecting the performance of anaerobic biofilm reactors for proper process design.
were reviewed. The important parameters are: Ideal plug flow and CSTR models were used to describe
the flow pattern inside a laboratory scale AFBR. The use
1. the effect of hydrodynamics/flow pattern on reactor of this assumption in large scale reactors has to be
performance; investigated. Similar to UASB reactors, while developing
2. the mass transfer within granules/biofilms; kinetic models for AFBRs, the influence of pH, substrate
3. the kinetic effects; and product inhibition, validity of Monod kinetics,
4. the structure and composition of biogranules. microbial composition and location and sphericity of the
granules have to be studied. In the case of anaerobic filters,
In UASB reactors the different zones are idealized with the deep-biofilm model seems to represent the laboratory
different flow patterns. Sludge blankets and sludge beds are scale reactors. But, the assumption of the substrate
mostly described by a CSTR flow pattern with bypass. The concentration reaching zero at the filter media has to be
clarifier zone is described by the axially dispersed plug flow verified in the large-scale reactor. A realistic model
model. The application of these models for actual full-scale considering all the above-mentioned short comings is
reactors is yet to be studied because these models do not necessary for proper design and scale up of such reactors.
consider the existence of non-ideal conditions in full-scale The integration of the flow model and biofilm model for
reactors. These non-ideal conditions may include non- these biofilm reactors with ADM1 can result in a robust
existence of different zones, improper flow distribution and model, which can be a tool for design purposes.
dead zones.
Reactor models for UASB reactors consider many References
questionable assumptions such as spherical granules,
steady-state operation, description of substrate degrada- Agardy, F.J., Cole, R.D., Pearson, E.A., 1963. Kinetic and activity
tion by simple Monod kinetics, no substrate/product parameters of anaerobic fermentation systems. Sanitary Engineering
inhibition and the effect of pH. Further studies are Research Laboratory Report, University of California, Berkeley.
Alphenaar, P.A., Perez, M.C., Lettinga, G., 1993. The influence of
required to develop models which do not depend too
substrate transport limitation on porosity and methanogenic activity
much on these assumptions for their development. of anaerobic sludge granules. Applied Microbiology and Biotechnol-
In developing kinetic models for both UASBs and ogy 39, 276–280.
AFBRs, granule structure plays an important role. Studies Andrews, J.F., 1969. A mathematical model for the continuous culture of
show that the structure of the granules and bacterial microorganisms utilizing inhibitory substrate. Biotechnology and
composition depends on the type of effluent being treated. Bioengineering 10, 707–723.
Andrews, G.F., Tien, C., 1979. The expansion of a fluidized bed
Various theories are provided to support the layered and containing biomass. American Institute of Chemical Engineers Journal
un-layered structures of the granules. The variation of 25, 720–723.
granule structure within the same reactor remains un- Arcand, Y., Guitot, S.R., Desrochers, M., Chavarie, C., 1994. Impact of
explained as of today. Incorporation of this variation in the the reactor hydrodynamics and organic loading on the size and activity
models poses a challenge for the modellers. of anaerobic granules. The Chemical Engineering Journal 56, 23–35.
Arching, B.K., Schmidt, J.E., Winther-Nielsen, M., Macario, A.J.L., de
The hydrodynamics and bed expansion characteristics of Macario, E.C., 1993. Effect of medium composition and sludge
AFBRs have been reported in the literature. It is generally removal on the production, composition and architecture of thermo-
found that the drag co-efficient of a biofilm covered philic (55 1C acetate-utilizing granules from an upflow anaerobic
ARTICLE IN PRESS
16 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

sludge blanket reactor). Applied and Environmental Microbiology 59, Fang, H.H.P., Chui, H.K., Li, Y.Y., 1995. Effect of degradation kinetics
2538–2545. on the microstructure of anaerobic biogranules. Water Science and
Atkinson, B., Davies, I.J., 1974. The overall rate of substrate uptake Technology 32 (8), 165–172.
(reaction by microbial films). Part I. A biological rate equation. Gjaltema, A., Vinke, J.L., van Loosdrecht, M.C.M., Heijnen, J.J., 1997.
Transactions of the Institutions of Chemical Engineers 52, 248–259. Abrasion of suspended biofilm pellets in airlift reactors: importance of
Atkinson, B., How, S.Y., 1974. The overall rate of substrate uptake shape, structure and particle concentrations. Biotechnology and
(reaction by microbial films). Transactions of Institute of Chemical Bioengineering 53, 88–99.
Engineers 52, 260–268. Gonzalez-Gil, G., Seghezzo, L., Lettinga, G., Kleerebezem, R., 2001a.
Batstone, D.J., Keller, J., Angelidaki, I., Kalyuzhnyi, S., Pavlostathis, Kinetics and mass-transfer phenomena in anaerobic granular sludge.
S.G., Rozzi, A., Sanders, W., Siegrist, H., Vavilin, V., 2002. (IWA Biotechnology and Bioengineering 73, 125–134.
Task Group on Modeling of Anaerobic Digestion Processes). IWA Gonzalez-Gil, G., Lens, P.N.L., Van Aelst, A., Van As, H., Versprille,
Publishing, London. A.I., Lettinga, G., 2001b. Cluster structure of anaerobic aggregates of
Batstone, D.J., Keller, J., Blackall, L.L., 2004. The influence of substrate an expanded granular sludge bed reactor. Applied and Environmental
kinetics on the microbial community structure in granular anaerobic Microbiology 67, 3683–3692.
biomass. Water Research 38, 1390–1404. Grady, C.P.L., 1982. Modeling of biological fixed films—a stat of art
Bolle, W.L., van Breugel, J., van Eybergen, G.C., Kossen, N.W.F., van review. In: Proceedings of the First International Conference on Fixed-
Gils, W., 1986. An integrated dynamic model for the UASB reactor. film Biological Processes, Kingsland, OH.
Biotechnology and Bioengineering 28, 1621–1636. Grotenhuis, J.T.C., Smit, M., Plugge, C.M., Yuansheng, X., van
Buffiere, P., Steyer, J.P., 1995. Comprehensive modeling of methanogenic Lammeren, A.A.M., Stams, A.J.M., Zehnder, A.J.B., 1991. Bacterial
biofilms in fluidized bed systems: mass transfer limitations and composition and structure of granular sludge adapted to different
mutisubstrate aspects. Biotechnology and Bioengineering 48, 725–736. substrates. Applied and Environmental Microbiology 57, 1942–1949.
Buffiere, P., Fonade, C., Moletta, R., 1998a. Mixing and phase hold-ups Guiot, S.R., Pauss, A., Costerton, J.W., 1992. A structured model of the
variations due to gas production in anaerobic fluidized-bed digesters: anaerobic granule consortium. Water Science and Technology 25 (7),
influence on reactor performance. Biotechnology and Bioengineering 1–10.
60, 36–43. Hamoda, M.F., Kennedy, K.J., 1987. Biomass retention and performance
Buffiere, P., Fonade, C., Moletta, R., 1998b. Liquid mixing and phase of anaerobic fixed-film reactors treating acetic acid wastewater.
hold-ups variations in gas producing fluidized bed bioreactors. Biotechnology and Bioengineering 30, 272–281.
Chemical Engineering Science 53, 617–627. Heijnen, J.J., Mulder, A., Enger, W., Hoeks, F., 1989. Review on the
Buffiere, P., Fonade, C., Moletta, R., 1998c. Modeling and experiments on application of anaerobic fluidized bed reactors in wastewater treat-
the influence of biofilm size and mass transfer in fluidized bed reactor ment. Chemical Engineering Journal 4, B37–B50.
for anaerobic digestion. Water Research 32, 657–668. Henze, M., Harremoes, P., 1983. Anaerobic treatment of wastewater in
Canovas-Diaz, M., Howell, J.A., 1988. Stratified mixed-culture biofilm fixed film reactors—a literature review. Water Science and Technology
model for anaerobic digestion. Biotechnology and Bioengineering 32, 15 (8/9), 1–101.
348–355. Hermanovicz, S.W., Cheng, Y.W., 1990. Biological fluidized bed reactor:
Chang, H.T., Rittmann, B.E., 1987. Verification of the model of biofilm hydrodynamics, biomass distribution and performance. Water Science
on activated carbon. Environmental Science and Technology 21, and Technology 22 (1-2), 193–202.
280–288. Hermanovicz, S.W., Ganczarczyk, J.J., 1983. Some fluidization character-
Chang, H.T., Rittmann, B.E., 1988. Comparative study of biofilm shear istics of biological bed. Biotechnology and Bioengineering 25,
loss on different adsorptive media. Journal of Water Pollution and 1321–1330.
Control Federation 60, 361–368. Hirata, A., Takemoto, T., Ogawa, K., Auresenia, J., Tsuneda, S., 2000.
Chang, H.T., Rittmann, B.E., Amar, D., Heim, R., Ehlinger, O., Lesty, Evaluation of kinetic parameters of biochemical reaction in three-
Y., 1991. Biofilm detachment mechanisms in a liquid fluidized bed. phase fluidized bed biofilm reactor for wastewater treatment.
Biotechnology and Bioengineering 38, 499–506. Biochemical Engineering Journal 5, 165–171.
Cooper, P.F., Sutton, P.M., 1983. Treatment of wastewaters using Huang, J.S., Jih, C.G., 1997. Deep-biofilm kinetics of substrate utilization
biological fluidized beds. Chemical Engineering 393, 392–405. in anaerobic filters. Water Research 39, 2309–2317.
Daffonchio, D., Thavessri, J., Verstraete, W., 1995. Contact angle Hulshoff Pol, L.W., 1989. The phenomenon of granulation of anaerobic
measurement and cell hydrophobicity of granular sludge upflow sludge. Ph.D. Thesis, Agricultural University Wageningen, The
anaerobic sludge bed reactors. Applied and Environmental Micro- Netherlands.
biology 61, 3676–3680. Iwai, S., Kitao, T., 1994. Wastewater Treatment with Microbial Films.
de Beer, D., Huisman, J.W., van den Heuvel, J.C., Ottengraf, S.P.P., 1992. Technomic Publishing Co., Lancaster, PA.
The effect of pH profiles in methanogenic aggregate on the kinetics of James, T.W., 1961. Continuous culture of microorganisms. Annual
acetate conversion. Water Research 26, 1329–1336. Review of Microbiology 15, 27–46.
Di Felice, R., 1995. Hydrodynamics of liquid fluidization. Chemical La Motta, E.J., Patricio, C., 1996. Substrate consumption kinetics in
Engineering Science 50, 1213–1245. anaerobic biofilm fluidized bed reactor. Journal of Environmental
Diez-Blanco, V., Garcia-Encina, P.A., Fernandez-Polnco, F., 1995. Effect Engineering 122, 198–204.
of biofilm growth, gas and liquid upflow velocities on the expansion of Lau, Y.L., 1995. Relative importance of mean velocity and bed shear on
an anaerobic fluidized bed reactor (AFBR). Water Research 29, biofilm accumulation in open-channel flows. Water Science and
1649–1654. Technology 32 (8), 193–198.
Dolfing, J., 1985. Kinetics of methane formation by granular sludge at low Lawrence, A.W., McCarty, P.L., 1967. Kinetics of methane fermentation
substrate concentrations. Applied Microbiology and Biotechnology in anaerobic waste treatment. Technical Report No. 75, Civil
22, 77–81. Engineering Department, Stanford University, Palo Alto, CA.
Droste, R.L., Kennedy, K.L., 1986. Sequential substrate utilization and Lens, P.N.L., de Beer, D., Cronenberg, C.C.H., Houwen, F.P., Ottengraf,
effectiveness factor in fixed biofilms. Biotechnology and Bioengineer- S.P.P., Verstraete, W.H., 1993. Heterogeneous distribution of micro-
ing 28, 1713–1720. bial activity in methanogenic aggregates: pH and glucose micro
Fan, L.-S., 1989. Gas–Liquid–Solid Fluidization Engineering. Butter- profiles. Applied and Environmental Microbiology 59, 3803–3815.
worth, London. Lens, P.N.L., de Beer, D., Cronenberg, C., Ottengraf, S., Verstraete, W.,
Fang, H.H.P., 2000. Microbial distribution in UASB granules and its 1995. The use microsensors to determine distributions in UASB
resulting effects. Water Science and Technology 42, 201–208. aggregates. Water Science and Technology 31 (1), 273–280.
ARTICLE IN PRESS
V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18 17

Lettinga, G., van Velsen, A.F.M., Homba, S.W., de Zeeuw, W., Klapwijk, Ro, K.S., Neethling, J.B., 1994. Biological fluidized beds containing
A., 1980. Use of the upflow sludge blanket reactor concept for widely different bioparticles. Journal of Environmental Engineering
biological wastewater treatment especially for anaerobic treatment. 120, 1416–1426.
Biotechnology and Bioengineering 22, 699–734. Rocheleau, S., Greer, C.W., Lawrence, J.R., Cantin, C., Laramee, L.,
Lin, S.H., 1991. A mathematical model for a biological fluidized bed Guiot, S.R., 1999. Differentiation of Methanosaeta concilli and
reactor. Journal of Chemical Technology and Biotechnology 51, Methanosarcina barkeri in anaerobic mesophilic granular sludge by
473–482. in situ hybridization and confocal scanning laser microscopy. Applied
Liu, B.Y.M., Pfeffer, J.T., Suidan, M.T., 1991. Loading capacity of a and Environmental Microbiology 65, 2222–2229.
packed bed anaerobic reactor. Research Journal of Water Pollution Ruggeri, B., Caire, G., Specchia, V., Sassi, G., Bosco, F., Gianetto, A.,
Control Federation 63, 145–152. 1994. Determination of optimal biofilm activity in a biological
Liu, Y., Xu, H.L., Yang, S.F., Tay, J.H., 2003. Mechanisms and models fluidized reactor. Water Science and Technology 29 (10-11),
for anaerobic granulation in upflow anaerobic sludge blanket reactor. 347–351.
Water Research 37, 661–673. Safferman, S.I., Bishop, P.L., 1996. Aerobic fluidized bed reactor with
MaCLeod, F.A., Guiot, S.R., Costerton, J.W., 1990. Layered structure of internal media cleaning. Journal of Environmental Engineering 122,
bacterial aggregates produced in an upflow anaerobic sludge bed and 284–291.
filter reactor. Applied and Environmental Microbiology 56,
Sanchez, J.M., Arijo, S., Munoz, M.A., Morinigo, M.A., Borrego, J.J.,
1598–1607.
1994. Microbial colonization of different support materials used to
Meunier, A.D., Williamson, K.J., 1981. Packed bed biofilm reactors:
enhance the methanogenic process. Applied Microbiology and
design. Journal of Environmental Engineering 107, 319–337.
Biotechnology 41, 480–486.
Monod, J., 1950. La technique de culture continue, theorie et applications.
Santegoeds, C.M., Damgaard, L.R., Hesselink, G., Zopfi, J., Lens, P.,
Annals de L’Institute Pasteur 79, 390–410.
Muyzer, G., De Beer, D., 1999. Distribution of sulfate-reducing and
Mosey, F., 1983. Mathematical modeling of the anaerobic digestion
methanogenic bacteria in anaerobic aggregates determined by micro-
process: regulatory mechanisms for the formation of short-chain
sensor and molecular analysis. Applied and Environmental Micro-
volatile acids from glucose. Water Science and Technology 15 (8/9),
biology 65, 4618–4629.
209–232.
Mulcahy, L.T., La Motta, E.J., 1978. Mathematical model of the fluidized Schink, B., Thauer, R., 1988. Energetic of syntrophic methane formation
bed biofilm reactor. In: Proceedings of 51st Annual Conference, Water and the influence of aggregation In: Proceedings of the Granular
Pollution Control Federation, Anaheim, CA. Anaerobic Sludge-Microbiology and Technology Workshop, Pudoc,
Mulcahy, L.T., Shieh, W.K., 1987. Fluidization and reactor biomass Wageningen, The Netherlands, pp. 5–17.
characteristics of the denitrification fluidized bed biofilm reactor. Schmidt, J.E., Ahring, B.K., 1991. Acetate and hydrogen metabolism
Water Research 21, 451–458. in intact and disintegrated granules from an acetate-fed, 55 1C,
Mulcahy, L.T., Shieh, W.K., La Motta, E.J., 1980. Kinetics model of UASB reactor. Applied Microbiology and Biotechnology 28,
biological denitrification in a fluidized biofilm reactor FBBR. Progress 1753–1760.
in Water Technology 12, 143–157. Schmidt, J.E., Ahring, B.K., 1994. Extracellular polymers in granular
Narnoli, S.K., Indu, M., 1997. Sludge blanket of UASB reactor: sludge from different upflow anaerobic sludge blanket UASB reactors.
mathematical simulation. Water Research 31, 715–726. Applied Microbiology and Biotechnology 42, 457–462.
Ngian, K.F., Martin, W.R.B., 1980. Bed expansion characteristics of Schmidt, J.E., Ahring, B.K., 1996. Granular sludge formation in upflow
liquid fluidized particles with attached microbial growth. Biotechnol- anaerobic sludge blanket UASB reactors. Biotechnology and Bioengi-
ogy and Bioengineering 22, 1843–1856. neering 49, 229–246.
Nicolella, C., van Loosdrecht, M.C.M., Di Felice, R., Rovatti, M., 1999. Schreyer, H.B., Coughlin, R.W., 1999. Effects of stratification in a
Terminal settling velocity and bed expansion characteristics of biofilm- fluidized bed bioreactor during treatment of metalworking wastewater.
coated particles. Biotechnology and Bioengineering 62, 63–70. Biotechnology and Bioengineering 63, 129–140.
Nicolella, C., van Loosdrecht, M.C.M., Heijnen, J.J., 2000. Wastewater Seghezzo, L., Zeeman, G., van Lier, J.B., Hamelers, H.V.M., Lettinga, G.,
treatment with particulate biofilm reactors. Journal of Biotechnology 1998. A review: the anaerobic treatment of sewage in UASB and
80, 1–33. EGSB reactors. Bioresource Technology 65, 175–190.
Ojha, C.S.P., Singh, R.P., 2002. Flow distribution parameters in relation
Sekiguchi, Y., Kamagata, Y., Syutsubo, K., Ohashi, A., Harada, H.,
to flow resistance in an upflow anaerobic sludge blanket reactor
Nakmura, K., 1998. Diversity of mesophilic and thermophilic granular
system. Journal of Environmental Engineering 128, 196–200.
sludge determined by 16s rRNA gene analysis. Microbiology 65,
Perry, R.H., Green, D.W., 1997. Chemical Engineers’ Hand-book.
4618–4629.
McGraw-Hill, New York.
Sekiguchi, Y., Kamagata, Y., Nakmura, K., Ohashi, A., Harada, H.,
Peyton, B.M., Characklis, W.G., 1992. Kinetics of biofilm detachment.
1999. Fluorescence in situ hybridization using 16S rRNA-
Water Science and Technology 26 (9/11), 1995–1998.
targeted oligonucleaotides reveals localization of methanogenes and
Picioreanu, C., Van Loosdrecht, M.C.M., Heijnen, J.J., 2001. Two-
selected uncultured bacteria in mesophilic and thermophilic
dimensional model of biofilm detachment caused by internal stress
sludge granules. Applied and Environmental Microbiology 65,
from liquid flow. Biotechnology and Bioengineering 72, 205–218.
1280–1288.
Richardson, J.F., Zaki, W.N., 1954. Sedimentation and fluidization: Part
I. Transactions of Institute of Chemical Engineers 32, 35–53. Seok, J., Komisar, S.M., 2003. Integrated modeling of anaerobic fluidized
Rittmann, B.E., 1982. The effect of shear stress on biofilm loss rate. bed bioreactor for deicing waste treatment I: Model derivation.
Biotechnology and Bioengineering 24, 501–506. Journal of Environmental Engineering 129, 100–109.
Rittmann, B.E., McCarty, P.L., 1978. Variable-order model of bacterial- Shieh, W.K., Sutton, P.M., Kos, P., 1981. Prediction reactor biomass
film kinetics. Journal of Environmental Engineering 104, 889–900. concentration in a fluidized bed system. Journal of Water Pollution
Rittmann, B.E., McCarty, P.L., 1980a. Model of steady state biofilm and Control Federation 53, 1574–1584.
kinetics. Biotechnology and Bioengineering 22, 2343–2357. Shieh, W.K., Mulcahy, L.T., LaMotta, E.J., 1982. Mathematical model
Rittmann, B.E., McCarty, P.L., 1980b. Evaluation of steady state biofilm for the fluidized bed biofilm reactor. Enzyme and Microbiol
kinetics 2. Biotechnology and Bioengineering 22, 2359–2373. Technology 4, 269–275.
Ro, K.S., Neethling, J.B., 1990. Terminal settling characteristics of Singhal, A., James, G., Praveen, V.V., Ramachandran, K.B., 1998. Axial
bioparticles. Research Journal of the Water Pollution Control dispersion model for upflow anaerobic sludge blanket reactors.
Federation 62, 901–906. Biotechnology Progress 14, 645–648.
ARTICLE IN PRESS
18 V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1–18

Skiadas, I.V., Ahring, B.K., 2002. A new model for anaerobic processes of Wang, Y.T., Suidan, M.T., Rittmann, B.E., 1987. Model biofilm kinetics
up-flow anaerobic sludge blanket reactors based on cellular automata. for a low-loaded expanded bed anaerobic reactor. Biotechnology and
Water Science and Technology 45 (10), 87–92. Bioengineering 30, 15–21.
Stewart, M.J., 1956. Reaction kinetics and operational parameters of Wimpenny, J.W.T., Colasanti, R., 1997. A unifying hypothesis for the
continuous flow anaerobic fermentation processes. Sanitary Engineer- structure of microbial biofilms based on cellular automation model.
ing Research Laboratory Report, University of California, Berkeley. FEMS Microbiology and Ecology 22, 1–16.
Stewart, P.S., 1993. A model of biofilm detachment. Biotechnology and Wu, M.M., Hickey, R.F., 1997. Dynamic model for UASB reactor
Bioengineering 41, 111–117. including reactor hydraulics, reaction and diffusion. Journal of
Suidan, M.T., 1986. Performance of deep biofilm reactors. Journal of Environmental Engineering 123, 244–252.
Environmental Engineering 112, 78–93. Wu, M.M., Criddle, C., Hickey, R.F., 1995. Mass transfer and
Tagawa, T., Syutsubo, K., Ohashi, A., Harada, H., 2000. Quantification temperature effects on substrate utilization by brewery granules.
of methanogen cell density in anaerobic granular sludge consortia by Biotechnology and Bioengineering 46, 465–475.
fluorescence in-situ hybridization. Water Science and Technology 42 Wu, J.H., Lui, W.T., Tseng, I.C., Cheng, S.S., 2001. Characterization of
(3/4), 77–82. microbial consortia in a terphthalate-degrading anaerobic granular
Trinet, F., Heim, R., Amar, D., Chang, H.T., Rittmann, B.E., 1991. Study sludge system. Microbiology 147, 373–382.
of biofilm and fluidisation of bioparticles in a three-phase liquid fluidized Yu, H., Rittmann, B.E., 1997. Predicting bed expansion and phase
bed reactor. Water Science and Technology 23 (7/9), 1347–1354. holdups for three-phase fluidized bed reactors with and without
Van der Meer, R.R., 1979. Anaerobic treatment of wastewater containing biofilm. Water Research 31, 2604–2616.
fatty acids in upflow reactors. Ph.D. Thesis, Delft University, Delft, Zhang, T.C., Bishop, P.L., 1994. Structure, activity and composition of
The Netherlands. biofilms. Water Science and Technology 29 (7), 35–344.
Van Loosdrecht, M.C.M., Heijnen, J.J., Eberl, H., Kreft, J., Picioreanu, Zoutberg, G.R., Frankin, R., 1996. Anaerobic treatment of chemical and
C., 2002. Mathematical modelling of biofilm structures. Antonie van brewer waste water with a new type of anaerobic reactor: the Biobed
Leeuwenhoek 81, 245–256. EGSB reactor. Water Science and Technology 34 (5-6), 375–381.

You might also like