Top Banner

of 18

zazcle.pdf

Jun 03, 2018

Download

Documents

raguerre
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • 8/12/2019 zazcle.pdf

    1/18

    Journal of Environmental Management 81 (2006) 118

    Review

    Modelling anaerobic biofilm reactorsA 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 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22.1. 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 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104.2.1. 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

    ARTICLE IN PRESS

    www.elsevier.com/locate/jenvman

    0301-4797/$ - see front matterr 2006 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.jenvman.2005.10.002

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

    E-mail address: [email protected] (T.R. Sreekrishnan).

    http://www.elsevier.com/locate/jenvmanhttp://www.elsevier.com/locate/jenvman
  • 8/12/2019 zazcle.pdf

    2/18

    1. Introduction

    The term biofilm reactor refers to that class of

    bioreactors where the biocatalyst exists in an anchored

    form, either on the surface of an inert carrier or attached

    to one another. The carrier could be the wall of the reactor,

    baffles provided for this purpose or particles of some inertmaterial. Biocatalysts such as microorganisms could also

    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

    case of a fluidized bed system. Typically, in such reactors,

    the rate of substrate conversion is limited by the rate of

    transport of substrate into the biofilm. In an anaerobic

    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

    methane and carbon dioxide. The anaerobic treatment, as

    the main biological step in wastewater treatment systems,

    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-

    ment of dense granules (14 mm) formed by the natural

    self-immobilization of the anaerobic bacteria. This kind of

    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

    that volumetric organic loading rates as high as 50 kgchemical 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 0.51.0 m/h. This

    reactor consists of a sludge bed, a sludge blanket and a

    clarifier zone supplemented with a physical device called

    the gassolid separator.

    In anaerobic fluidized bed reactor (AFBR), the liquid to

    be treated is pumped through a bed of inert particles

    (typically sand with a particle size range of 0.20.8 mm) at

    a velocity sufficient enough (1020 m/h) to cause fluidiza-

    tion (Nicolella et al., 2000). In the fluidized state, the media

    provides a large surface for attached biological growth and

    allows biomass concentrations to develop in the range of

    1040 kg/m3 (Cooper and Sutton, 1983). A typical flow

    diagram of an AFBR is shown in Fig. 2. Compared to

    other high rate anaerobic reactors such as UASB, the

    fluidized bed system is claimed to have the following

    advantages: higher purification capacity, no clogging of

    the reactor (as in filters), no problem of sludge washout

    (as in UASB systems if granular sludge is not obtained),

    and small volume and land area requirements (Heijnen

    et al., 1989).

    To study the sensitivity of the process to various

    operating parameters and to optimize the design of these

    reactors, it is necessary to have mathematical models which

    are simple in concept and close to the physical situation

    existing in these reactors. During the past few decades there

    have been a number of studies on modelling of UASB and

    AFBR reactors. An effort is made in this article to present

    the gist of different approaches used in developing

    mathematical models for these reactors. This article also

    reviews different proposals available for modelling the

    structure of the biological granules and biofilm-covered

    particles in these reactors.

    2. Models for UASB reactors

    The formation of anaerobic granular sludge is consid-

    ered to be a necessary condition for the successful

    operation of a UASB reactor. The efficiency of the reactor

    depends mainly on active biomass concentration and the

    influent flow rate. As the reactor reaches steady state, a

    dense bed composed of granulated sludge develops at the

    bottom. Immediately above the sludge bed, a zone

    consisting of finely suspended particles called the sludge

    blanket forms. A clear zone over this sludge blanket

    constitutes the settling zone. To keep the sludge granules in

    suspended condition inside the reactor, the inlet flow rate

    ARTICLE IN PRESS

    Biogas

    Effluent

    Influent

    Baffles

    Weir

    Settler

    Sludge Bed

    Sludge Blanket

    Sludge granules

    Fig. 1. Schematic diagram of UASB reactor.

    Wastewater feed

    Carrier

    Biofilm

    Recycle line

    Treated water

    Biogas

    Fig. 2. Anaerobic fluidized bed reactor (AFBR).

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1182

  • 8/12/2019 zazcle.pdf

    3/18

    should be below a threshold limit (normally below 1 m/h).

    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

    reactor, it is important to analyse the flow pattern inside

    the reactor and reaction kinetics within the biological

    granules. In general, models for UASB reactors consist oftwo parts:

    1. fluid flow model;

    2. reactor model.

    2.1. Flow model

    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,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).

    Wu and Hickey (1997) have developed a flow model to

    describe the flow pattern in a UASB reactor. They modelled

    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

    clarification zone above the sludge blanket (Fig. 3).

    The equation of the flow model is given by

    VbdCdt

    VbEt QfCt, (2.1)

    Et MfVbTin

    if 0ptpTin;

    0 ifTinp0; (2.2)

    whereC(t) is the tracer concentration within the CSTR, Vbthe CSTR working volume, E(t) the input function for

    tracer impulse, Qfthe low fraction that enters the working

    volume, Mf the fraction of mass input that go through

    reactors working volume, andTinthe tracer injection time.

    Ojha and Singh (2002) analysed the flow distribution in

    different zones of the reactor in order to simulate the

    UASB reactor performance. They used the flow resistance

    approach to represent flow distribution. It was found that

    with an increase in flow resistance in the UASB reactor

    system, the magnitude of short-circuiting flows at the

    reactor bed increased. The flow distribution at the blanket

    and clarifier was found to have an influence on the flow

    resistance. But no generalized model could be obtained.

    Bolle et al. (1986) divided the reactor into three

    compartments (Fig. 4): sludge bed, sludge blanket and

    settler. The liquid flow in the sludge bed and the sludge

    blanket were described by completely stirred tank reactor

    systems. Liquid flow in the internal settler was described by

    a plug flow model.

    Narnoli and Indu (1997) developed a model for the

    sludge blanket of a UASB reactor. According to them, the

    maximum organic load which a reactor can assimilate

    depends on proportioning of the reactor height into the bed

    and the blanket sections. A condition of force equilibrium

    was considered between the rising gas bubbles from the

    sludge bed and the adjacent water mass. The negative

    pressure behind the bubble attracts the water mass and

    leads to the formation of a wake. Using diffusion concepts,

    solid particles moving along the wake due to the

    concentration gradient were equated to those settling down

    under the influence of gravity at steady state. The solid

    concentration along the blanket height was computed and

    found to match the experimental observations. This model

    could be used to optimize the reactor dimensions and the

    desludging schedule.

    Singhal et al. (1998) reported that a simple two-zone

    axial dispersion model adequately describes the fluid flow

    characteristics of UASB reactors. They found that the fluid

    flow behaviour is very sensitive to the volume of zones and

    degree of bypassing between them but less sensitive to the

    dead volume. The model assumed that some of the liquid

    bypasses the first zone and enters directly into the second

    zone. The schematic representation is shown inFig. 5.

    ARTICLE IN PRESS

    Dead volume

    CSTR Dispersed

    plug flow

    By-pass flow

    Fig. 3. Representation of hydraulic model (Wu and Hickey, 1997).

    Effluent

    Settler

    Sludge blanket

    Short circuit

    Dead space Sludge bed

    Influent

    Fig. 4. Block diagram of fluid flow pattern (Bolle et al., 1986).

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 3

  • 8/12/2019 zazcle.pdf

    4/18

    They had verified the model efficacy through tracer

    studies. The step response of an axially dispersed tubular

    flow reactor for an inert tracer is described by the following

    partial differential equation:

    1

    Pe

    q2C

    qZ2

    qC

    qZ

    qC

    qy, (2.3)

    where C is the dimensionless concentration of the tracer

    (c/c0), Pe is the Peclet number (uL/D), Z is the dimension-

    less distance (z/L),y is the dimensionless time (t/t), u is the

    linear average velocity of the liquid in the reactor, L is the

    length of the reactor, and D is the axial dispersion

    coefficient. The simulation and experimental results gave

    a very good fit validating this flow model.

    All the above multi-compartment models were capable

    of fitting laboratory scale experimental data well. In the

    case of full-scale reactors, the feed distribution is very

    different from that in the laboratory scale reactor. Distinct

    zones of sludge bed, sludge blanket and clarifier may not be

    observed due to the large volume of the reactor and uneven

    flow distribution. This will make the task of quantifying theextent of non-ideality in the flow patterns difficult. It may

    not be correct to predict the fluid flow pattern based on lab

    scale studies alone. Hence, further investigations are

    required to test the models for full-scale reactors.

    2.2. Reactor model

    Substrate degradation in the bioreactor is dependant on

    the observed reaction rate. In a biofilm type of reactor set-

    up, the observed reaction rate is normally less than the rate

    predicted by the reaction kinetics. The reason for this is

    that the substrate concentration actually available to the

    microorganisms is less than the substrate concentration

    present in the bulk liquid. This is due to the control exerted

    by the mechanism of molecular diffusion on the penetra-

    tion of substrate into the biofilm. It is, therefore, important

    to analyse the rate-determining step while developing

    models. Hence, the reactor model consists of 3 parts:

    substrate utilization within the biofilm (granules), mass

    transfer and transport in the granule bed and blanket, and

    mass transport within the clarifier.

    Kinetic models are based on the relationship between

    limiting substrate concentration and growth rate as

    proposed byMonod (1950).James (1961)pointed out that

    in biological processes a wide variety of substances might

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

    (1963) have used COD as the limiting substrate in the

    development of their models for the anaerobic digestion

    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

    decomposition of complex organics is the conversion ofvolatile acids to methane. Andrews (1969)has presented a

    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

    and low pH value leads to digester failure. To represent this

    phenomenon in a quantifiable way an inhibition function

    (Haldane type) was used in the model. The un-ionized form

    of volatile acids was considered as the rate limiting

    substrate since it is a function of both pH and the total

    volatile acids concentration.

    Rittman and co-workers have presented a biofilm model

    describing the biofilm growth and substrate flux under

    steady-state conditions (Rittmann, 1982; Rittmann and

    McCarty, 1980a). UASB reactor models using methano-

    genic biofilm models have focused on mass transfer

    limitations and single as well as multiple limiting substrates

    (Alphenaar et al., 1993; Atkinson and Davies, 1974;

    Atkinson and How, 1974; Buffiere and Steyer, 1995; de

    Beer et al., 1992; Lens et al., 1993;Lin, 1991).

    Thick biofilms may give rise to mass-transfer limitations

    for the substrate within the biofilm, resulting in an overall

    reduction in the substrate conversions achieved. Under this

    condition, the transport of substrate into the biofilm or

    transport of products out of the biofilm could become the

    rate-determining step. Contradictory results have beenreported on the impact of internal (within the biofilm) and

    external (in the bulk liquid phase) mass transport. While

    some reports indicate that both internal and external

    diffusion limitations influence the rate of substrate utiliza-

    tion (Dolfing, 1985; Wu et al., 1995), others indicate that

    mass transport limitations are not observed in anaerobic

    biofilms (de Beer et al., 1992; Schmidt and Ahring, 1991)

    even at film thicknesses as high as 2.6 mm (Droste and

    Kennedy, 1986). In addition, it has been reported that pH

    profiles inside anaerobic granules may affect the conversion

    rate more than mass transport limitations (de Beer et al.,

    1992).

    Gonzalez-Gil et al. (2001a) studied the influence of

    external and internal mass transport in anaerobic granular

    sludge. For this, kinetic properties of acetate degrading

    methanogenic sludge granules of different mean diameters

    were assessed at different up-flow velocities. It was

    experimentally found that external mass transport resis-

    tance can be neglected and that biogas formation does not

    influence the diffusion rates. The substrate uptake could be

    explained by biofilm (internal) diffusion.

    The general approach to the modelling problem starts

    with development of the biofilm model. These are then

    used to find the substrate flux at the surface of each

    granule. The biofilm model is then tailored with the flow

    ARTICLE IN PRESS

    Zone-1 Zone-2

    Sludge

    Blanket

    Clarifier

    Fig. 5. Schematic of the two-compartment axial dispersion model

    (Singhal et al., 1998).

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1184

  • 8/12/2019 zazcle.pdf

    5/18

    model to predict the overall performance of the reactor. In

    general, the following assumptions were made in develop-

    ing anaerobic biofilm models:

    1. Granules are spherical.

    2. Substrate utilization is described by the Monod model.

    3. External mass transport is negligible.4. Density of the biofilm is constant throughout the

    granule.

    5. The relative concentrations of the acidogenic and

    methanogenic bacteria are constant throughout an

    anaerobic granule. This remains true irrespective of

    the location in the reactor from where the granule was

    taken.

    Many of the reported studies deal with steady-state

    conditions. Unsteady state is the most critical situation for

    modelling, associated with real-time control strategies.Wu

    and Hickey (1997)developed a dynamic model to describeUASB reactors with methanogenic anaerobic granules.

    They assumed that substrate diffuses into the granules

    outer layer upto a thickness of 100 mm. At the inner edge of

    this layer, the substrate gradient reaches zero. Growth of

    acetate utilizing methanogens is neglected due to their low

    growth rate.

    By applying an acetate mass balance on anaerobic

    granules, the granular bed and the clarifier, respectively,

    the dynamic model equations are obtained.

    For anaerobic granules:

    qs

    qt D

    q2s

    qx2

    2

    Rx

    @s

    @x

    kmxms

    ks s . (2.4)

    Boundary conditions:

    1. Dqs

    qxklsb klsb; x 0,

    2. qs

    qx 0; x d,

    where R is the granule radius; km the specific substrate

    utilization rate;ksthe half velocity constant; D the effective

    diffusion coefficient;kl the mass transfer coefficient;xmthe

    active acetate utilizer biomass within the outer layer d;sthe

    acetate concentration within the biofilm.

    A mass balance on a granule bed involves three terms:

    substrate entering the reactor, substrate leaving the reactor,

    and substrate being taken up by the granules and

    subsequently being utilized:

    Vbds

    dt VbEt Qfst klArsb s0; t. (2.5)

    Initial condition: sb(0) sb0

    Et MfVbTin

    if 0ptpTinfor impulse;

    0 ifTinp0;

    (2.6)

    whereAris the total granule surface area; Tinthe substrate

    injection time for the acetate impulse; klArsb s0; tdescribes the substrate transferring from the bulk solution

    through the liquid boundary layer into the granules. Data

    from an acetate impulse experiment was simulated using

    these hydraulic- reactiondiffusion models with a good fit.

    The results indicated that diffusion inside the granules,reaction kinetics and hydraulic behaviour play important

    roles in the performance of a UASB reactor.

    Bolle et al. (1986) developed an integrated dynamic

    model for the UASB reactor. Their flow model has been

    described in Section 2.1. Substrate and biomass balance

    were done over the sludge bed and sludge blanket. Thus the

    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

    compartments and different phases. This model was able to

    predict the sludge bed height, the biomass concentration in

    the sludge blanket, the short-circuiting flows over the bed

    and the blanket, and the effluent COD concentrations as a

    function of the hydrodynamic load, COD load, pH and

    settler efficiency.

    Skiadas and Ahring (2002) proposed a model for

    UASB reactors based on the cellular automata (CA) concept.

    A cellular automation is a simulation, which is discrete in

    time, space and state. A CA model usually consists of an

    array of compartments similar to the spaces in a game of

    naughts and crosses. The CA theory has been applied to

    predict the layer structure of the granules, which is high

    acidogen and low methanogen concentrations at the outer

    granule layers and the reverse at the inner granule layers. It

    has also predicted the granule diameter and granulemicrobial compositions as functions of the operational

    parameters. The other models do not consider the effect of

    operational parameters on granule size and composition.

    Details on automation models can be found in the excellent

    review paper byWimpenny and Colasanti (1997).

    Even though different authors have presented good

    experimental fit for their models, a unified approach is

    lacking. The above kinetic models, in general, assume

    external mass transfer to be negligible. Some of the

    assumptions may lead to poor predictions. For example,

    the assumption of Monod kinetics, spherical shape and

    uniform density of the granules may not be true in actual

    reactors. The lack of versatile models in the literature

    shows that a lot of improvements can be made in future

    models. Some of them could be

    1. Incorporation of the variation of density within the

    biofilm/granule.

    2. Kinetic expressions, which include inhibition terms.

    3. Kinetic expressions, which include the non-uniform pH

    profile inside the biofilm.

    4. Diversity in the bacterial population distribution

    inside the granule in terms of the predominant species/

    groups.

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 5

  • 8/12/2019 zazcle.pdf

    6/18

    3. Proposed models for the structure of the biofilm

    In an anaerobic reactor, under favourable conditions,

    bacterial cells can attach to each other and grow as a

    granule. This granule has a 3-D structure. When some inert

    carrier particles or surfaces are introduced in the system,

    cells grow on the carrier particles/surfaces forming abiofilm. Depending upon the structure of the carrier

    material the resulting biofilm will have a 2-D or 3-D

    structure. Microbial composition and their distribution

    inside the granules play an important role in substrate

    conversion. Conversion could be limited by the process of

    diffusion of the substrate within the biofilm as well as the

    concentration of biomass bringing about the conversion.

    This information is very essential to develop the biofilm

    model. Many authors have reported that the structure of

    the granule/biofilm is mainly determined by the substrate

    degradation kinetics rather than the geometry with which it

    is formed (Batstone et al., 2004;Fang et al., 1995;Guiot et

    al., 1992). Batstone et al. (2004) have developed a biofilm

    structure model which predicts the structure of a 2-D

    biofilm. This model has been formulated based on

    substrate degradation and diffusion kinetics. They exam-

    ined four different types of granules obtained from UASB

    reactors treating wastewaters from a cannery, a slaughter-

    house, and two breweries. The microbial structures of these

    granules were assessed by fluorescence in situ hybridization

    probing with 16S rRNA-directed oligonucleotide probes,

    scanning electron microscope and transmission electron

    microscope. The biofilm model could exactly predict the

    structures of different types of granules observed through

    the experiments when the model was simulated for thesame operating conditions. This proves that the structure

    of the biofilm is influenced by the substrate kinetics and not

    by the geometry of the biofilm. Van Loosdrecht et al.

    (2002)has shown that a 2-D model is sufficient to represent

    a 3-D structure.Picioreanu et al. (2001)also has supported

    the same. Hence, the different models proposed for the

    biofilm structure in this section are applicable to both

    granules and 2-D biofilms.

    3.1. Multi-layer model

    MacLeod et al. (1990)andGuiot et al. (1992)proposed a

    three-layered structure for the bacterial aggregates treating

    carbohydrates in a UASB reactor. According to this

    model, the microbiological composition of granules is

    different in each layer. The inner layer mainly consists of

    methanogens that may act as nucleation centres necessary

    for the initiation of granule development. H2-producing

    and H2-utilizing bacteria are dominant species in the

    middle layer, and a mixed species including rods, cocci and

    filamentous bacteria takes the predominant position in the

    outermost layer (Fig. 6).

    To convert a target organic compound to methane, the

    spatial organization of methanogens and other microbial

    species in the granules (such as those in a UASB reactor) is

    essential. The layered structure of UASB granules is

    supported by the works ofArching et al. (1993) andLens

    et al. (1995) with immunological and histological methods.

    The dynamic model proposed by Arcand et al. (1994) also

    supports this structure. Similar support for the layered

    structure is found in reports fromSantegoeds et al. (1999)

    using microelectrodes. Sekiguchi et al. (1998, 1999) and

    Tagawa et al. (2000) also confirmed this structure by

    fluorescence in situ hybridization using 16S rRNA targeted

    oligonucleotides. A distinct layered structure was also

    found in the methanogenicsulfidogenic aggregates, with

    sulfate-reducing bacteria in the outer 50100mm and

    methanogens in the inner part (Sekiguchi et al., 1998).

    Unlike the initial multi-layer model proposed byMacLeod

    et al. (1990) recent research showed that UASB granules

    have large, dark, non-staining centres, in which neither

    archaeal nor bacterial signals could be found (Rocheleau

    et al., 1999). In fact, the non-staining centre in the UASB

    granules might be the result of the accumulation of

    metabolically inactive, decaying biomass and inorganic

    materials (Sekiguchi et al., 1998).The importance of ECP in anaerobic granulation has

    been observed by Schmidt and Ahring (1994, 1996). They

    conclude that ECP may play an important role in building

    spatial structure and maintaining the stability of UASB

    granules, but are unsure of its contribution to the initiation

    of anaerobic granulation. A large quantity of ECP is

    unnecessary for making up active granules since it

    decreases the porosity and thereby increases the diffusional

    resistance for substrate to penetrate the granule. Too much

    ECP could even cause deterioration of floc formation

    (Schmidt and Ahring, 1996).

    3.2. Syntrophic microcolony model

    According to the syntrophic microcolony model, a close

    synergistic relationship among different microbial groups is

    essential for efficiently breaking down the complex organic

    compounds. In fact, the syntrophic microcolonies provide

    the kinetic and thermodynamic requirements for inter-

    mediate transference and therefore efficient substrate

    conversion (Schink and Thauer, 1988). They state that

    the synergistic requirements would drive bacteria to form

    granules, in which different species function in a synergistic

    way and can easily survive.

    ARTICLE IN PRESS

    Acidogens

    Methanothrix

    H2producing Acetogens and

    H2consuming organisms

    Fig. 6. Three-layered structure of the bacterial aggregates (MacLeod

    et al., 1990).

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1186

  • 8/12/2019 zazcle.pdf

    7/18

    3.3. Non-layered structure model

    Contrary to the multi-layer model, anaerobic granules

    with non-layered structure have also been reported (Fang et

    al., 1995; Grotenhuis et al., 1991; Wu et al., 2001).

    There is evidence that a layered structure of the UASB

    granules would be developed with carbohydrates and a non-layered one using substrates having a rate-limiting hydrolytic

    or fermentative step (e.g. proteins) (Fang, 2000;Fang et al.,

    1995). This is probably due to different initial steps in the

    degradation of carbohydrate as compared

    to that of proteins. Degradation of carbohydrates to

    small molecules is faster than the subsequent degradation

    of the intermediates, whereas the initial step in protein

    degradation is a rate-limiting step. That is, the protein

    degradation step is slower than the steps which follow it.

    Results from fluorescence in situ hybridization com-

    bined with confocal scanning laser microscopy clearly

    showed that protein-fed granules possess non-layered

    structure with a random distribution ofMethanosaeta concilii

    (Rocheleau et al., 1999). However, granules, which differ in

    their composition in terms of the predominant microbial

    species, can still be formed from the same substrate

    (Daffonchio et al., 1995;Schmidt and Ahring, 1996).

    Based on microscopic examination of the UASB

    granules, recentlyFang (2000)proposed that the microbial

    distribution of the UASB granules strongly depends on the

    degradation thermodynamics and kinetics of individual

    substrate. Therefore, it appears that different dominating

    catabolic pathways may give rise to granules, which are

    different in their structure. Spontaneous and sudden

    washout of the established granular sludge bed, as a resultof a change in wastewater composition, is a common

    problem encountered in the operation of UASB systems.

    So far, none of the individual models for the granule

    structure can explain this phenomenon. If a factor that is

    independent of the wastewater composition can initiate the

    formation of UASB granules, a change in the wastewater

    composition should not lead to the washout of the entire

    granular sludge bed. Thus, it is a reasonable speculation

    that there should be a substrate composition-associated

    factor that highly contributes to the formation of UASB

    granules, but is not yet included in the present models of

    granule structure (Liu et al., 2003).

    Liu et al. (2003)proposed a general model for anaerobic

    granulation in UASB reactors. This model consists of four

    steps for granulation.

    1. Physical movement to initiate bacterium-to-bacterium

    contact or bacterial attachment onto nuclei.

    2. Initial attractive forces to keep stable multi-cellular

    contacts, e.g. physical, chemical and biochemical forces.

    3. Microbial forces to make cell aggregation mature, e.g.

    production of ECP.

    4. Hydrodynamic shear force shaping 3-D structure of

    microbial aggregates.

    3.4. Granule cluster structure

    Recently the structure of anaerobic granules of an

    expanded granular sludge bed (EGSB) reactor was studied

    byGonzalez-Gil et al. (2001b). They found black spherical

    granules having numerous whitish spots on their surfaces.

    Cross-sectioning these aggregates revealed that the whitishspots appeared to be white clusters embedded in a black

    matrix (Fig. 7). High magnification electron microscopy

    showed that the white clusters mainly consisted of acetate

    utilizing methanogens (Mathanosaeta spp.) and the black

    matrix consisted of syntrophic species and hydrogenotrophic

    methanogens (Methanobacterium like and Methaospirillum

    like organisms). Fluorescent in situ hybridization using 16 s

    rRNA probes of crushed granules showed that 70% of the

    cells belonged to the archaebacterial domain and 30% to

    eubacterial domain. The authors have provided a very logical

    explanation as to why a cluster structure, as observed, is

    advantageous over a layered structure. Acetate produced in

    the black zone is transported by random diffusion in all

    directions and thus penetrates the Methanosaeta clusters from

    all sides. Hence, substrate depleted zones are circumvented,

    which allows the growth of more active biomass per unit area

    of aggregate.

    Batstone et al. (2004) studied the influence of substrate

    degradation kinetics on the microbial community structure

    in granular anaerobic biomass. The granules, which were

    grown in the effluent containing soluble as well as

    particulate protein, had homogeneous structure. The

    primary cause of this structure was assessed through

    biofilm modelling. They postulate that the particulate

    nature of the wastewater and the slow rate of particulatehydrolysis, rather than the presence of proteins in the

    wastewater, was responsible for the homogeneous structure

    of the granules. Because solids hydrolysis was rate limiting,

    soluble substrate concentrations were very low (below

    Monod half-saturation concentration), which caused low

    growth rates.

    From the above information it can be concluded that

    there are two key factors which determine the structure of a

    granule (i.e. the organization of the microbial community

    within a granule). They are

    1. The nature of organic compounds present in the

    wastewater.

    2. The kinetics of substrate degradation.

    ARTICLE IN PRESS

    Methanosaetaclusters

    Zone with syntrophic

    eubacteria and

    hydrogenotrophic

    methanogens

    Seed aggregate

    Fig. 7. Schematic representation of the architecture of anaerobic

    aggregates (Gonzalez-Gil et al., 2001b).

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 7

  • 8/12/2019 zazcle.pdf

    8/18

    The relative concentrations of different microbial species

    within a granule is another important factor required in

    developing a biofilm model. This seems to depend on the

    concentration of the substrate. Since there could be

    concentration gradients with respect to substrate concen-

    tration within the same reactor, the composition of the

    granule in terms of the microbial species may vary fromgranule to granule obtained from different parts of the

    same reactor. Hence a robust model, capable of incorpor-

    ating the structure as well as composition of the granules in

    terms of the different microbial communities present, is

    required. Such a model for the biofilm/granule can be used

    in developing reactor models which are more versatile as

    well as reproducible with a higher level of accuracy.

    4. Models for AFBR

    The AFBR uses inert carrier particles to provide

    mechanical support for growth of the biofilm. These

    biogranules are maintained in a fluidized state by using

    the energy of the influent liquid. Such a fluidized bed

    system is free from the channelling problems encountered

    in other biofilm reactors. A serious operational problem

    associated with the AFBR is that when the biofilm grows

    on the carrier surface, the composite density of the film-

    covered particle decreases, ultimately resulting in the carry

    over of the film-covered particles out of the reactor. Full-

    scale application of AFBRs is not common due to lack of

    sound design principles. Many attempts have been made to

    study the process kinetics and the factors affecting process

    performance.

    In general, models for AFBRs include the followingelements:

    1. A bed fluidization model which describes the size and

    number of bioparticles per unit fluidized bed volume.

    2. A biofilm model which describes the rate of substrate

    conversion per individual granule.

    3. A reactor flow model, which links the biofilm and bed

    fluidization models to yield substrate concentration as a

    function of axial position within the AFBR.

    4.1. Bed fluidization model

    Hydrodynamic behaviour of carrier supported biogra-

    nules plays an important role in designing AFBRs. When

    the granules grow, their size, shape and composite density

    change. This has an impact on the hydrodynamic

    behaviour of the granules. Information on settling and

    fluidization characteristics, such as fluidized bed-height, as

    a function of liquid velocity is required for the designing of

    AFBRs. Information on fluidized-bed height is important

    because it establishes the solids residence time and the

    specific biofilm surface area in the biologically active zone.

    The relationships valid for fluidization of rigid particles,

    readily available in the chemical engineering literature (e.g.

    Di Felice, 1995), have to be modified to take into account

    the effect of the biofilm layer on the rigid particles

    (Nicolella et al., 2000).

    4.1.1. Terminal settling velocity

    The terminal settling velocity of a single spherical

    particle in an infinite expanse of fluid is

    ut 4grs rl

    3CDrl

    0:5. (4.1)

    Depending on the biofilm thickness and on the carrier

    type, values for the equivalent density of biofilm particles

    (rs) range typically from 1100 to 1500 kg/m3. The drag

    coefficient CD is a function of the particle Reynolds

    number defined as

    Ret rldsut

    ml. (4.2)

    Biofilm particles are in the intermediate flow regime(1oReto100) for the vast majority of cases, obtained

    when sand (0.51 mm) or carrier materials with densities in

    the same range are used as the inert support. In the

    intermediate flow regime, the drag coefficient for a smooth,

    rigid sphere is (Perry and Green, 1997)

    CD 18:5Re0:6t . (4.3)

    This correlation cannot be used for carrier-supported

    granules, since they are neither smooth nor rigid. For this

    reason, empirical correlations have been suggested for the

    estimation ofCD for biofilm particles:

    CD 17:1Re0:47t ; 50oReto100 (4.4)

    Hermanovicz and Ganczarczyk (1983),

    CD 36:66Re0:67t ; 40oReto90 (4.5)

    Mulcahy and Shieh (1987),

    CD 24

    Ret21:55Re0:518t ; 15oReto87 (4.6)

    Ro and Neethling (1990),

    CD 24

    Ret14:55Re0:48t ; 40oReto90 (4.7)

    Yu and Rittmann (1997),

    CD 29:6Re0:6t ; 7oReto90 (4.8)

    Nicolella et al. (1999).

    The drag coefficient of the biofilm covered carrier

    particle is generally found to be more than that of smooth,

    rigid spheres. Surface roughness has generally been

    considered as the reason for the increase in the drag

    coefficient of biofilm covered carrier particles (Hermano-

    vicz and Ganczarczyk, 1983;Mulcahy and Shieh, 1987).

    The above correlations are all empirical since they were

    obtained by fitting experimental data generated by

    different groups, who employed fluidized bed systems,

    which differed from one another in one or more aspects.

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 1188

  • 8/12/2019 zazcle.pdf

    9/18

    The deformable nature and surface roughness of the

    biofilms and the type of carrier particle used play a major

    role in determining the hydrodynamic nature of the biofilm

    covered particle. Hence, bed fluidization models should be

    able to incorporate those parameters which have an

    influence on the hydrodynamic behaviour of the biofilm

    covered particles. It may not be feasible to have ananalytical expression for this purpose, given the complexity

    and variability inherent to the system. While an empirical

    relationship is acceptable, it should be able to cater to a

    wider range of operational parameters normally encoun-

    tered rather than a narrow set of operating conditions.

    4.1.2. Fluidization mechanics

    In a fluidized bed anaerobic reactor, the film-covered

    particles are kept in a fluidized state by the incoming liquid.

    The bed porosity and biomass concentration in the bed are

    determined by the mechanics of fluidization. Hence, a

    realistic mathematical expression for the bed fluidization is

    necessary.

    For a bed of uniform spherical particles, the following

    relation was proposed (Richardson and Zaki, 1954):

    us uien, (4.9)

    where us is the superficial liquid velocity, e porosity and

    ui ut10d=D, (4.10)

    where ut is the terminal settling velocity of particle, d the

    particle diameter and D the diameter of bed, n is a constant

    given by

    n 4:6520d=DReto0:2, (4.11)

    n 4:418d=DRet0:03 0:2oReto1, (4.12)

    n 4:418d=DRet0:1 1oReto200, (4.13)

    n 4:4Ret0:1 200oReto500, (4.14)

    n 2:4500oRet. (4.15)

    Whether this equation can be applied directly and in its

    entirety to a biological fluidized bed is a controversial

    subject. Richardson and Zaki have derived the equation for

    a bed of uniform spherical particles, which are hard and

    have a smooth surface. But the biological film coveredparticles seldom have these characteristics.

    Andrews and Tien (1979) related the bed height to the

    amount of biomass in the bed. The fluidized bed tends to

    stratify vertically based on the settling velocity of the

    bioparticles. If no stratification is assumed, the bed height

    is related to average biofilm thickness x by

    H

    Hc

    1ec1 x

    1ec1 x1=3=1A x1=n

    . (4.16)

    In the case of complete stratification

    H

    Hc 1ec Z

    xmax

    0

    1x

    1e

    gx dx. (4.17)

    A practical working expression was derived based on the

    above two expressions by Andrews and Tien (1979) as

    follows:

    H

    Hc1 1B x, (4.18)

    where

    B ecD

    31ec 3x0 2D

    1ec

    1ec

    13A2

    13A

    D

    1 3A

    3n ,

    where H is the bed height, e the bed porosity, x the film

    volume/clean particle volume, x the mean value of

    distribution function of bacterial film gx, A the buoyant

    density of bacterial film/buoyant density of clean particle, n

    the exponent in the RichardsonZaki correlation. The

    subscript c denotes the bed of clean particles.

    This expression was found to predict the experimental

    data well.

    Mulcahy and La Motta (1978) have developed specific

    correlations to determine ut and n for the case of

    bioparticles in a fluidized bed as follows:

    ut rsrlgd

    1:67p

    27:5r0:33l m0:67

    " #0:75, (4.19)

    n 10:35Re0:18t 40oReto90, (4.20)

    wherersis the density of bioparticles; rl the liquid density;

    m the liquid viscosity.

    Ngian and Martin (1980)found that the Richardson and

    Zaki correlations gave a satisfactory estimate for ui for

    small support particles whereas ui was 3070% below theexperimentally determined value for larger support parti-

    cles.Nicolella et al. (1999) found u i to be only 80% of the

    unhindered settling velocity. They recommend that the

    correlation should be used with caution while determining

    the constant ui.

    In general, the Richardson and Zaki correlation is found

    to describe the bed expansion characteristics of a fluidized

    bed. But the values of the constants n and ui seem to be a

    function of the property of the biofilm covered particles.

    The application of these correlations to a full-scale reactor

    containing a wide size distribution of biofilm covered

    particles needs to be verified.

    4.1.3. Effect of gas production on hydrodynamics

    The effect of gas production on hydrodynamics of

    fluidized beds is an important factor to be studied for

    design and scale-up of AFBRs. Many investigations on the

    flow pattern in an AFBR suggested that an axially

    dispersed plug flow model can be used for the flow model

    (Hirata et al., 2000; Seok and Komisar, 2003). In these

    studies the effect of gas production on the flow pattern was

    not considered.Diez-Blanco et al. (1995) have studied the

    effect of gas production on the hydrodynamic behaviour of

    an AFBR. In this study, the bed contraction due to biogas

    production in a fluidized bed of 6 m height was estimated to

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 9

  • 8/12/2019 zazcle.pdf

    10/18

    be less than 6%. Based on this, the investigators considered

    the effect of biogas on the hydrodynamic behaviour to be

    negligible. In contrast,Buffiere et al. (1998a) reported that

    the effect of biogas on bed height could be negligible. But

    the biogas effervescence affects the phase hold-ups (which

    affects the liquid solid contact time) and liquid mixing of

    the reactor system. The experimental study ofBuffiere et al.(1998b) compared the hydrodynamic behaviour of a gas

    producing fluidized bed with a classical gas injected three

    phase fluidized bed reactor. The overall gas hold up was

    found to be more in the case of the gas producing fluidized

    bed than the gas injected one. They also observed axial

    variation of phase hold-up.

    4.1.4. Bed stratification

    Schreyer and Coughlin (1999)reported a stratification of

    biofilm coated sand particles in a fluidized bed reactor.

    Stratification can be attributed to the influence of a biofilm

    on a particles settling velocity. The presence of a biofilm

    cover decreases a particles overall density, thereby

    increasing its buoyancy. The biofilm also increases the

    particles size thereby increasing the drag force exerted on

    it by the liquid flowing upward. The particles in a fluidized

    bed are expected to segregate according to size and mean

    density (Ro and Neethling, 1994; Safferman and Bishop,

    1996;Trinet et al., 1991).

    Bed stratification has many negative effects on the

    performance of the reactor. Thicker biofilms pose diffusion

    limitation and wash out problems. Hence, to prevent

    stratification and to maintain uniform particle size, efforts

    have been made to remove excess biofilm from largerparticles (Ruggeri et al., 1994;Safferman and Bishop, 1996;

    Shieh et al., 1981; Trinet et al., 1991). Examples of such

    efforts include external sand-biomass separators (screens),

    operation of an impeller at the top of the bed and internal

    screen cleaning devices. Many researchers have examined

    the effect of shear on biofilm and biofilm detachment rate,

    mainly as a tool to control biofilm thickness (Chang and

    Rittmann, 1988;Chang et al., 1991;Gjaltema et al., 1997;

    Peyton and Characklis, 1992; Rittmann, 1982; Safferman

    and Bishop, 1996; Stewart, 1993; Trinet et al., 1991).

    Biofilm detachment rate appears to depend on turbulence

    and particle concentration; an increase in either increases

    detachment rate. Biofilm has been observed to be relatively

    smoother and more homogeneous under conditions of high

    shear (i.e., high liquid velocity) than under low shear

    conditions (Lau, 1995;Zhang and Bishop, 1994). The study

    by Schreyer and Coughlin (1999) showed that the

    introduction of a thinner to increase the shear resulted in

    a non-stratified bed.

    Bed stratification could occur due to differences in the

    biofilm thickness or differences in the carrier particle size or

    both. But bed stratification is most common when carrier

    particles are not of uniform size. A completely mixed bed

    (i.e. no stratification) was observed by Andrews and Tien

    (1979)when uniform carrier particles were used.

    4.2. Kinetic and reactor sub-models

    Substrate removal within a biological film of uniform

    thickness attached to a spherical particle is described by

    D

    r2d

    dr r2

    ds

    dr Rt, (4.21)where D is the effective diffusion coefficient of substrate

    within the biological film; r the radial coordinate measured

    from the centre of the support particle; s the substrate

    concentration within the biofilm and Rtthe intrinsic rate of

    substrate consumption per unit volume of biological film.

    It is generally accepted by many authors that substrate

    removal kinetics can be modelled using the microbial

    growth model proposed by Monod. Many researchers

    (Grady, 1982; Henze and Harremoes, 1983; Iwai and

    Kitao, 1994) have suggested that depending on the values

    of the model constants, reaction rate can be represented by

    first- or zero-order kinetics. For AFBRs, many authors

    demonstrated that zero-order kinetics provide an adequate

    description of substrate consumption (i.e. considering the

    acidogenesis and methanogenesis phases together) (La

    Motta and Patricio, 1996; Mulcahy and La Motta, 1978;

    Mulcahy et al., 1980;Shieh et al., 1982).

    For zero-order reaction, the reaction rate term is

    Rt rk0. (4.22)

    The solution of Eq. (4.21) for complete substrate

    penetration inside the biofilm and for the partial penetra-

    tion has been presented by Mulcahy et al. (1980).

    For fully penetrated biofilm:

    Observed rate 43prk0r3p r

    3m. (4.23)

    For partial substrate penetration:

    Observed rate 1:76rk01:45r3p r

    3m

    1:9s0:45br3pr

    3m

    1:9

    r1:8p D0:45

    .

    (4.24)

    rm is the radius of support particle; rp the radius of

    biological particle; sb the concentration of substrate in the

    bulk of the liquid within the fluidized bed.

    A simple plug flow model was used to describe dissolved

    substrate transport in the axial direction in a fluidized bedreactor (Mulcahy and La Motta, 1978):

    Udsb

    dz Rv 0. (4.25)

    With the boundary condition

    z 0; sb s0

    where U is the average liquid velocity in the longitudinal

    direction; sb the substrate concentration in the bulk of the

    liquid; s0 the influent substrate concentration; Rv the rate

    of reaction.

    For fully substrate penetrated biofilm, the concen-

    tration profile along the reactor was given by Mulcahy

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 11810

  • 8/12/2019 zazcle.pdf

    11/18

    et al. (1980)as

    se s0rk0Vm

    Q

    rp

    rm

    3 1

    " #, (4.26)

    where se is the concentration of substrate in the effluent

    stream.La Motta and Patricio (1996)tested this model by

    conducting experiments. This test showed a reasonable

    agreement between the zero-order kinetic model with

    complete substrate penetration and the experimental data.

    Hirata et al. (2000)used a different approach to evaluate

    the kinetic parameters of the biochemical reactions taking

    place in a three phase fluidized bed reactor. Using the

    substrate balance at steady state and assuming Monod

    kinetics, an equation relating the substrate consumption

    rate to substrate concentration (expressed as Biochemical

    Oxygen Demand, BOD5) and total biofilm surface area was

    established. The following assumptions were made in

    formulating the model:

    1. Reactor system is completely mixed.

    2. Total organic carbon (TOC), which is expressed in terms

    of BOD5, is the only rate-limiting substrate. Other

    substrates are in excess.

    3. The reaction follows Monod kinetics and substrate

    inhibition is negligible.

    4. Reaction occurs at constant volume.

    Performing a substrate balance at steady state yielded

    Fsin sss 1

    Yx=srxV, (4.27)

    where F is the inlet flow rate, Vthe reactor volume, rxthe

    biofilm growth rate, Yx/s the yield coefficient mass of

    biomass formed/mass of substrate consumed, Sss the

    steady-state value of the rate limiting substrate concentra-

    tion inside the reactor, Sinthe inlet substrate concentration.

    If the reaction followed Monod kinetics, then the proposed

    rate equation was

    rx mx mmaxsss

    km sss

    x, (4.28)

    where m is the specific growth rate, mmax is the maximum

    specific growth rate and km the Monod constant.

    Substituting rx into the substrate balance equation

    Fsin sss 1

    Yx=s

    mmax sss

    km sss

    Vx Rt, (4.29)

    Vx rbdsb, (4.30)

    where rb is the biomass dry density, d the effective biofilm

    thickness and sb the total biofilm surface area. The total

    surface area was computed as follows:

    sb pDave2N, (4.31)

    where N is the total number of particles inside the reactor

    and Dave is the average particle diameter. Modifying

    Eq. (4.29) gave

    Rt K sbsss

    kmsss

    , (4.32)

    where K rbdmmax=Yx=s:Transforming the above equation using Lineweaver

    Burke linearization,

    sb

    Rt

    km

    K

    1

    sss

    1

    K. (4.33)

    Plotting sb/Rt versus 1/sss gave a straight line with slope

    km/K and intercept 1/K. Using the above plot for the

    known value of inlet substrate concentration, the steady-

    state substrate concentration could be found.

    Buffiere et al. (1998c) studied the biofilm activity along

    the reactor height. They developed a biofilm model

    assuming

    1. Homogeneous biofilm of uniform thickness.

    2. Spherical support media of uniform size.

    3. Internal mass transfer described by Ficks law.

    4. Liquid phase perfectly mixed with homogeneous con-

    centration.

    5. No external mass transfer limitation.

    The mass balance for a substrate s in the biofilm is

    expressed by

    D

    r2d

    dr r2

    ds

    dr

    Xi

    Vsi. (4.34)

    The boundary conditions are

    r rp; s s0,

    r rc; ds

    dr 0,

    Vs xs

    Yx=s

    mmaxs

    kss

    . (4.35)

    The right-hand side of (4.34) is the sum of all substrate

    uptake rates minus the sum of all substrate production

    rates.

    In the liquid phase, the mass balance for one substrate s

    is given by

    VLds

    dt Qsins DAp

    ds

    dr

    rrp

    !, (4.36)

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

    concentration ofs;Q the input flow rate; rp the bioparticle

    radius; r the radial distance measured from bioparticle

    centre;D the diffusivity of component s in the biofilm;mmaxthe maximum specific growth rate for s-utilizing bacteria;

    Vsi the reaction rate of s through reaction I; Ap the

    exchange area of the bioparticles; VL the liquid phase

    volume; xs the s-utilizing bacteria concentration; Yx=s the

    biomass yield factor for s-utilizing bacteria.

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 11

  • 8/12/2019 zazcle.pdf

    12/18

    The substrate gradient at the surface of a bioparticle in

    Eq (4.36) is taken from the biofilm model. The biomass

    composition used in this simulation was averaged from

    several studies such as the modelling work ofMosey (1983)

    and the experimental investigation ofSanchez et al. (1994)

    (Table 1).

    A set of batch experiments was conducted using

    bioparticles taken from different heights. Glucose, acetate

    and propionate were used as substrates. Substrate con-

    sumption with time was monitored. Simulation results were

    found to give reasonable fit for experimentally observed

    values. The biomass composition (i.e. the relative concen-

    tration of acidogens and methanogens) within the biofilm

    was manipulated to fit the experimental results. This

    indicated the crucial role of biomass composition in

    substrate kinetics. The specific activity of the biofilm was

    also measured. The following were the findings:

    1. Thicker film bioparticles were found on the top of the

    reactor and thinner in the bottom.

    2. Glucotrophic activity decreased from bottom to top andmethanogenic activity increased from bottom to top.

    3. This indicates the change in biomass composition with

    biofilm size.

    Buffiere et al. (1998a) developed a model for AFBRs

    based on total carbon removal kinetics. They considered

    the effects of gas production in their model which were of

    two kinds:

    1. Gas production modified the degree of axial mixing,

    which is responsible for the establishment of a concen-

    tration gradient in the reactor.2. Gas production is responsible for bed contraction,

    which reduces the contact between the liquid and

    bioparticles.

    The TOC removal kinetics were found to be in good

    agreement with the Monod model. The TOC uptake rate

    can be expressed by

    rTOC rmaxS

    ksS, (4.37)

    where S is the TOC concentration in the reactor (the

    reactor is assumed to be perfectly mixed). Parameters rmax

    and ks were found by plotting the inverse ofrTOC and 1/S.

    Bed contraction was found to reduce the liquidsolid

    contact by 1025%:

    es

    es01 0:045U0:4g . (4.38)

    Experimental results of gas hold-up in the reactor gave

    the following correlation:

    eg 13 1:2d0:168p U

    0:7g . (4.39)

    An axially dispersed plug flow model was found to

    describe the liquid mixing in the reactor. The mass balance

    of the reactor was given by the equation

    1

    Pe

    d2s

    dx2

    ds

    dxDa

    s

    ls (4.40)

    with the following notations:

    x z

    H; s

    s

    ks; Da

    rmax

    ks

    Hel

    Ul; andPe

    UlH

    elEzl.

    The axial dispersion co-efficient was found experimen-tally using the tracer injection method. The experimental

    results were found to fit the correlation of Muroyama et al.

    (fromFan, 1989) very well:

    DcUl

    elz 1:01U0:738l U

    0:167g D

    0:583c , (4.41)

    where dp is the particle diameter; Dc the column diameter;

    Ezl the axial dispersion coefficient; g the gravitational

    acceleration; H the bed height; ks the half-saturation

    concentration in Monod expression; rmax the maximal

    reaction rate in Monod expression; s the substrate

    concentration;Ug the gas superficial velocity; Ul the liquid

    superficial velocity; Ut the terminal velocity of particles; xthe reduced bed height; es the solid hold up; es0 the solid

    hold-up in the liquidsolid fluidized bed;eg the gas hold-up.

    Thus by knowing hold-ups (from Eq. (4.38) and (4.39))

    and Ez (from Eq. (4.41)), the performance Eq. (4.40) was

    solved. The model results were found to give a more

    realistic picture.

    4.2.1. Stratified biofilm models

    A stratified biofilm model was presented by Canovas-

    Diaz and Howell (1988). The anaerobic biofilm was

    modelled as two distinct layers with the inner layer

    consisting of methanogens and the outer layer consistingof acidogens. The substrate is converted to acids in the

    outer layers, and is subsequently converted to methane by

    the bacteria in the inner layer.

    The differential equations for substrate uptake in the two

    layers are:

    D1d2G

    dz2 k1x1, (4.42)

    D2d2F

    dz2 ak1x1

    k2x2

    1F=ki, (4.43)

    where D1, D2 is the diffusivities of substrate through the

    acidogenic and methanogenic layer. G,Fthe concentration

    ARTICLE IN PRESS

    Table 1

    Biomass composition of an anaerobic granule

    Type of bacteria Substrates for the

    bacteria

    %Distribution in the

    granule

    Acidogens Glucose 65

    Acetogens Butyrate 2.5

    Propionate 2.0

    Methanogens Acetate 7.0

    H2 utilizing bacteria Hydrogen 23.5

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 11812

  • 8/12/2019 zazcle.pdf

    13/18

    of glucose and fatty acids, k1 the zero-order kinetic

    constant with respect to sugar uptake and k1 the zero-

    order rate constantk2the kinetic constant for VFA uptake;

    ki the inhibition constant; z the distance through biofilm;

    x1, x2 the mass volume densities of acidogens and

    methanogens.

    The authors also explain why a stratified biofilm isadvantageous as compared to an un-stratified biofilm.

    When the bulk VFA concentration is high, in the case of an

    unstratified film, methanogens will face VFA inhibition. In

    the case of stratified biofilm, the inhibition level is

    minimized by the presence of the outer layer.

    Droste and Kennedy (1986) have given a model for

    sequential substrate utilization in a biofilm. This model

    assumes that no interactions occur between the two groups

    of microorganisms that will cause kinetic or diffusion

    parameters to change from values associated with indivi-

    dual cultures of each group. Unstratified biofilm and

    Monod-type kinetics were assumed.

    The governing differential equations are:

    D1d2s1

    dx2

    k1x1s1

    K1 s1, (4.44)

    D2d2s2

    dx2

    k2x2s2

    K2 s2

    Yk1x1s1

    K1 s1, (4.45)

    whereD is the diffusivity;k the maximum specific reaction

    velocity; K the half-velocity constant; s the substrate

    concentration; x the distance; x the active microorganism

    concentration; Y the acetate yield coefficient (g acetate/g

    primary substrate).

    From numerical analysis of the governing equation, itwas found that the production of intermediate substrate in

    the biofilm increased the conversion of primary substrate to

    ultimate product. The increase was not always significant.

    To summarize, two ways of approaching the problem of

    modelling substrate uptake kinetics are explained in this

    section. In the first one, TOC is considered as the rate-

    limiting substrate. In this case biomass composition,

    individual reaction steps and substrate diffusion limitations

    are not considered. The parameters involved in the model

    are evaluated by fitting the model to the experimental

    results. Even though this approach seems to be simple, it

    does not have a sound theoretical explanation and remains

    empirical. In the second approach, the kinetic model is

    developed considering the individual substrate kinetics,

    biofilm composition and diffusion limitations. This seems

    to be a more realistic approach. Ultimately the validity of

    these models has to be verified for large-scale reactors.

    5. The EGSB reactor

    The EGSB reactor comes under the family of UASB

    reactors. The use of effluent recirculation in a UASB (or a

    high height/diameter ratio), resulted in the EGSB reactor

    (Seghezzo et al., 1998). Here also, the biomass is present in

    a granular form. The higher upflow liquid velocity keeps

    the granular sludge bed in an expanded condition

    (Zoutberg and Frankin, 1996). Reports on models for

    EGSB reactors are very scarce. But based on the knowl-

    edge of UASB reactors and AFBR models, modelling of

    EGSB reactor can be attempted.

    1. The biofilm model is similar to the UASB reactorbiofilm model. The composition and structure of the

    biofilm is expected to be influenced by the nature as well

    as concentration of the substrate. However, there is no

    reason to believe that this influence will be significantly

    different from what has been observed for biogranules

    in the UASB reactor (Section 3.4). The only significant

    difference will be the reduced level of the substrate

    concentration gradient along the height of the reactor

    due to the effect of recirculation.

    2. The liquid flow pattern could be expected to be

    somewhere between completely mixed and dispersed

    plug flow, the exact pattern depending on the recycle

    ratio employed. However, any flow model employed will

    need validation through tracer studies or any other

    suitable experimental studies.

    3. A fluidization model which can predict the variation of

    bed height with upflow liquid velocity has to be

    developed based on experimentation.

    Tailoring of the above three models could result in a

    complete model for an EGSB reactor.

    6. The anaerobic biofilter

    The anaerobic filter is an anaerobic packed-bed biofilm

    reactor. Though both upflow as well as down flow

    configurations are possible, the upflow mode of operation

    is more common. The biomass forms a film on the surface of

    the packing media. The kinetic model of such a biofilm

    reactor has been studied quite extensively (Chang and

    Rittmann, 1987; Hamoda and Kennedy, 1987; Meunier

    and Williamson, 1981; Rittmann and McCarty, 1980a;

    Rittmann and McCarty, 1980b). Fig. 8 illustrates the ideal

    biofilm having uniform microbial density (Xf) and uniform

    thickness (Lf). When the substrate utilization follows the

    Monod relationship, the biofilm model incorporating the

    external mass transfer and internal simultaneous diffusionand reaction can be derived as follows (Huang and Jih, 1997):

    d2Sf

    dZ2

    kXfSf

    DfKs Sf. (4.46)

    Boundary conditions:

    DfdSf

    dZ

    Ds

    L SbSs Jat Z 0,

    dSf

    dZ 0 atZ Lf,

    where S is the biofilm substrate concentration, Z is the

    distance normal to the biofilm surface, D is the diffusivity,Ds

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 13

  • 8/12/2019 zazcle.pdf

    14/18

    is the axial dispersion coefficient, Xfis the microbial density

    in the biofilm, Ks is the Monod constant, kis the maximum

    specific substrate utilization rate, Jis the substrate flux.

    Here, the subscripts f and b denote biofilm and bulk

    liquid, respectively.

    The greatest difficulty in applying the model is to

    measure the biofilm thickness (Lf).Rittmann and McCarty

    (1978) and Suidan (1986) imposed another boundary

    condition to the model, i.e.Sf 0 atZ LfThis model is called the deep-biofilm model. Another

    reported approach for indirectly estimating the steady-state

    biofilm thickness was by assuming that the biomass growthequals the biomass decay and/or shear losses in biofilm

    reactors. For instance, Rittmann and McCarty (1980a)

    computed the steady-state biofilm thickness using para-

    meters of the substrate flux, microbial growth, biomass

    decay and sloughing rate (Eq. (4.47)). The model obtained,

    called the steady-state biofilm model, is given by

    Lf JY

    bXf, (4.47)

    where Y is the biomass yield coefficient and b is the

    sloughing rate.

    The models mentioned above have been experimentally

    verified or simulated using the results reported in the

    literature (Chang and Rittmann, 1987; Liu et al., 1991;

    Meunier and Williamson, 1981; Rittmann and McCarty,

    1980b,Wang et al., 1987). The kinetic parameters included

    in the models were either determined by independent

    experiments or obtained from published papers. The liquid

    flow regime was assumed to be completely mixed or plug-

    flow with axial dispersion (dispersion coefficients were

    calculated using empirical equations). An axial dispersion

    model coupled with deep-biofilm kinetics was used by

    Huang and Jih (1997). The experimental results and the

    calculated values showed good agreement. The tracer study

    ofHuang and Jih (1997)confirmed that the flow regime is

    close to completely mixed.

    7. The unified model for anaerobic digestion (ADM1)

    The IWA Anaerobic Digestion Modeling Task Group

    developed a generalized anaerobic digestion model (Bat-

    stone et al., 2002). The biochemical as well as physico-

    chemical processes were included in the model. The

    biochemical steps include (i) disintegration from homo-

    geneous particulates to carbohydrates, proteins and lipids,

    (ii) the extracellular hydrolysis of these particulate sub-strates to sugars, amino acids, and long chain fatty acids

    (LCFA), respectively, (iii) acidogenesis from sugars and

    amino acids to volatile fatty acids (VFAs) and hydrogen

    (iv) acetogenesis of LCFA and VFAs to acetate and (v)

    separate methanogenesis steps from acetate and hydrogen/

    CO2. The physico-chemical equations describe ion associa-

    tion and dissociation, and gasliquid transfer. The inhibi-

    tion kinetics have been incorporated in the biochemical

    process.

    7.1. The extension of ADM1 to biofilm reactors

    The ADM1 model gives a unified approach to anaerobic

    digestion. This model was successfully implemented for a

    CSTR (Batstone et al., 2002). ADM1 was developed for a

    suspended cell system, where there is no mass transfer

    limitation for movement of the substrate from the bulk

    liquid phase to the cells. On the other hand, in biofilm

    reactors, the rate of transport of substrate from the bulk

    liquid to the microbial population is controlled by diffusion

    of substrate within the biofilm. Hence to extend the ADM1

    for biofilm systems, the substrate utilization kinetics of the

    single cell system must be replaced with a biofilm model.

    The biofilm models for different biofilm reactor systems

    have been extensively reported in the previous sections. In

    ARTICLE IN PRESS

    Sampling port

    Inlet

    Outlet

    Liquid

    Phase Sb

    Biofilm

    Carrier

    materialGas phase

    Ss

    Lf

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

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 11814

  • 8/12/2019 zazcle.pdf

    15/18

    a similar way the reactor flow system of ADM1 (i.e. CSTR

    system) also has to be modified. In the case of a UASB

    reactor, the reactor flow system will have a combination of

    CSTRs and a plug flow system as explained in Section 2.1.

    AFBR and EGSB reactors will have axially dispersed plug

    flow systems as shown in Sections 4.2 and 5. A biofilter

    flow system is similar to that of ADM1. The physico-chemical process system of ADM1 can be directly extended

    to the biofilm reactor systems. Thus, the modification of

    the biochemical processes and the reactor flow system of

    ADM1 with that of the biofilm system could result in a

    unified model for biofilm reactors.

    8. Summary and conclusions

    Development of biofilm reactors has made anaerobic

    treatment an attractive option to treat wastewaters. For

    optimum design and scale-up of these reactors, mathema-

    tical models are required. In this paper, various parameters

    affecting the performance of anaerobic biofilm reactors

    were reviewed. The important parameters are:

    1. the effect of hydrodynamics/flow pattern on reactor

    performance;

    2. the mass transfer within granules/biofilms;

    3. the kinetic effects;

    4. the structure and composition of biogranules.

    In UASB reactors the different zones are idealized with

    different flow patterns. Sludge blankets and sludge beds are

    mostly described by a CSTR flow pattern with bypass. The

    clarifier zone is described by the axially dispersed plug flowmodel. The application of these models for actual full-scale

    reactors is yet to be studied because these models do not

    consider the existence of non-ideal conditions in full-scale

    reactors. These non-ideal conditions may include non-

    existence of different zones, improper flow distribution and

    dead zones.

    Reactor models for UASB reactors consider many

    questionable assumptions such as spherical granules,

    steady-state operation, description of substrate degrada-

    tion by simple Monod kinetics, no substrate/product

    inhibition and the effect of pH. Further studies are

    required to develop models which do not depend too

    much on these assumptions for their development.

    In developing kinetic models for both UASBs and

    AFBRs, granule structure plays an important role. Studies

    show that the structure of the granules and bacterial

    composition depends on the type of effluent being treated.

    Various theories are provided to support the layered and

    un-layered structures of the granules. The variation of

    granule structure within the same reactor remains un-

    explained as of today. Incorporation of this variation in the

    models poses a challenge for the modellers.

    The hydrodynamics and bed expansion characteristics of

    AFBRs have been reported in the literature. It is generally

    found that the drag co-efficient of a biofilm covered

    granule is more than that of a rigid particle of the same

    size. The experimental evidence published so far (Andrews

    and Tien, 1979; Hermanovicz and Cheng, 1990; Mulcahy

    and Shieh, 1987;Ngian and Martin, 1980;Nicolella et al.,

    1999) suggests that the fluidized bed expansion pattern is

    observed to follow the Richardson and Zaki correlation.

    The studies show that the expansion index can becalculated through the Richardson and Zaki correlation

    (Nicolella et al., 1999). But the parameter u i was found to

    be 3080% of the experimentally determined ut value.

    All these studies were conducted with uniform biogranule

    sizes.

    Many authors (Ro and Neethling, 1994; Safferman and

    Bishop, 1996; Schreyer and Coughlin, 1999; Trinet et al.,

    1991) observed that the bed becomes stratified due to the

    presence of granules with different biofilm thicknesses. In

    such a context, the application of the Richardson and Zaki

    correlation remains dubious. Hence models considering

    expansion characteristics of stratified beds are necessary

    for proper process design.

    Ideal plug flow and CSTR models were used to describe

    the flow pattern inside a laboratory scale AFBR. The use

    of this assumption in large scale reactors has to be

    investigated. Similar to UASB reactors, while developing

    kinetic models for AFBRs, the influence of pH, substrate

    and product inhibition, validity of Monod kinetics,

    microbial composition and location and sphericity of the

    granules have to be studied. In the case of anaerobic filters,

    the deep-biofilm model seems to represent the laboratory

    scale reactors. But, the assumption of the substrate

    concentration reaching zero at the filter media has to be

    verified in the large-scale reactor. A realistic modelconsidering all the above-mentioned short comings is

    necessary for proper design and scale up of such reactors.

    The integration of the flow model and biofilm model for

    these biofilm reactors with ADM1 can result in a robust

    model, which can be a tool for design purposes.

    References

    Agardy, F.J., Cole, R.D., Pearson, E.A., 1963. Kinetic and activity

    parameters of anaerobic fermentation systems. Sanitary Engineering

    Research Laboratory Report, University of California, Berkeley.

    Alphenaar, P.A., Perez, M.C., Lettinga, G., 1993. The influence of

    substrate transport limitation on porosity and methanogenic activityof anaerobic sludge granules. Applied Microbiology and Biotechnol-

    ogy 39, 276280.

    Andrews, J.F., 1969. A mathematical model for the continuous culture of

    microorganisms utilizing inhibitory substrate. Biotechnology and

    Bioengineering 10, 707723.

    Andrews, G.F., Tien, C., 1979. The expansion of a fluidized bed

    containing biomass. American Institute of Chemical Engineers Journal

    25, 720723.

    Arcand, Y., Guitot, S.R., Desrochers, M., Chavarie, C., 1994. Impact of

    the reactor hydrodynamics and organic loading on the size and activity

    of anaerobic granules. The Chemical Engineering Journal 56, 2335.

    Arching, B.K., Schmidt, J.E., Winther-Nielsen, M., Macario, A.J.L., de

    Macario, E.C., 1993. Effect of medium composition and sludge

    removal on the production, composition and architecture of thermo-

    philic (551C acetate-utilizing granules from an upflow anaerobic

    ARTICLE IN PRESS

    V. Saravanan, T.R. Sreekrishnan / Journal of Environmental Management 81 (2006) 118 15

  • 8/12/2019 zazcle.pdf

    16/18

    sludge blanket reactor). Applied and Environmental Microbiology 59,

    25382545.

    Atkinson, B., Davies, I.J., 1974. The overall rate of substrate uptake

    (reaction by microbial films). Part I. A biological rate equation.

    Transactions of the Institutions of Chemical Engineers 52, 248259.

    Atkinson, B., How, S.Y., 1974. The overall rate of substrate uptake

    (reaction by microbial films). Transactions of Institute of Chemical

    Engineers 52, 260268.Batstone, D.J., Keller, J., Angelidaki, I., Kalyuzhnyi, S., Pavlostathis,

    S.G., Rozzi, A., Sanders, W., Siegrist, H., Vavilin, V., 2002. (IWA

    Task Group on Modeling of Anaerobic Digestion Processes). IWA

    Publishing, London.

    Batstone, D.J., Keller, J., Blackall, L.L., 2004. The influence of substrate

    kinetics on the microbial community structure in granular anaerobic

    biomass. Water Research 38, 13901404.

    Bolle, W.L., van Breugel, J., van Eybergen, G.C., Kossen, N.W.F., van

    Gils, W., 1986. An integrated dynamic model for the UASB reactor.

    Biotechnology and Bioengineering 28, 16211636.

    Buffiere, P., Steyer, J.P., 1995. Comprehensive modeling of methanogenic

    biofilms in fluidized bed systems: mass transfer limitations and

    mutisubstrate aspects. Biotechnology and Bioengineering 48, 725736.

    Buffiere, P., Fonade, C., Moletta, R., 1998a. Mixing and phase hold-ups

    variations due to gas production in anaerobic fluidized-bed digesters:influence on reactor performance. Biotechnology and Bioengineering

    60, 3643.

    Buffiere, P., Fonade, C., Moletta, R., 1998b. Liquid mixing and phase

    hold-ups variations in gas producing fluidized bed bioreactors.

    Chemical Engineering Science 53, 617627.

    Buffiere, P., Fonade, C., Moletta, R., 1998c. Modeling and experiments on

    the influence of biofilm size and mass transfer in fluidized bed reactor

    for anaerobic digestion. Water Research 32, 657668.

    Canovas-Diaz, M., Howell, J.A., 1988. Stratified mixed-culture biofilm

    model for anaerobic digestion. Biotechnology and Bioengineering 32,

    348355.

    Chang, H.T., Rittmann, B.E., 1987. Verification of the model of biofilm

    on activated carbon. Environmental Science and Technology 21,

    280288.

    Chang, H.T., Rittmann, B.E., 1988. Comparative study of biofilm shearloss on different adsorptive media. Journal of Water Pollution and

    Control Federation 60, 361368.

    Chang, H.T., Rittmann, B.E., Amar, D., Heim, R., Ehlinger, O., Lesty,

    Y., 1991. Biofilm detachment mechanisms in a liquid fluidized bed.

    Biotechnology and Bioengineering 38, 499506.

    Cooper, P.F., Sutton, P.M., 1983. Treatment of wastewaters using

    biological fluidized beds. Chemical Engineering 393, 392405.

    Daffonchio, D., Thavessri, J., Verstraete, W., 1995. Contact angle

    measurement and cell hydrophobicity of granular sludge upflow

    anaerobic sludge bed reactors. Applied and Environmental Micro-

    biology 61, 36763680.

    de Beer, D., Huisman, J.W., van den Heuvel, J.C., Ottengraf, S.P.P., 1992.

    The effect of pH profiles in methanogenic aggregate on the kinetics of

    acetate conversion. Water Research 26, 13291336.

    Di Felice, R., 1995. Hydrodynamics of liquid fluidization. ChemicalEngineering Science 50, 12131245.

    Diez-Blanco, V., Garcia-Encina, P.A., Fernandez-Polnco, F., 1995. Effect

    of biofilm growth, gas and liquid upflow velocities on the expansion of

    an anaerobic fluidized bed reactor (AFBR). Water Research 29,

    16491654.

    Dolfing, J., 1985. Kinetics of methane formation by granular sludge at low

    substrate concentrations. Applied Microbiology and Biotechnology

    22, 7781.

    Droste, R.L., Kennedy, K.L., 1986. Sequential substrate utilization and

    effectiveness factor in fixed biofilms. Biotechnology and Bioengineer-

    ing 28, 17131720.

    Fan, L.-S., 1989. GasLiquidSolid Fluidization Engineering. Butter-

    worth, London.

    Fang, H.H.P., 2000. Microbial distribution in UASB granules and its

    resulting effects. Water Science and Technology 42, 201208.

    Fang, H.H.P., Chui, H.K., Li, Y.Y., 1995. Effect of degradation kinetics

    on the microstructure of anaerobic biogranules. Water Science and

    Technology 32 (8), 165172.

    Gjaltema, A., Vinke, J.L., van Loosdrecht, M.C.M., Heijnen, J.J., 1997.

    Abrasion of suspended biofilm pellets in airlift reactors: importance of

    shape, structure and particle concentrations. Biotechnology and

    Bioengineering 53, 8899.

    Gonzalez-Gil, G., Seghezzo, L., Lettinga, G., Kleerebezem, R., 2001a.Kinetics and mass-transfer phenomena in anaerobic granular sludge.

    Biotechnology and Bioengineering 73, 125134.

    Gonzalez-Gil, G.,