Fundamental Principles of Ultrafiltration.pdf

Embed Size (px)

Citation preview

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    1/15

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    2/15

    68

    the membrane. For example, a membrane that is

    cleaned before use has quite a different performance

    than that of one that has not been cleaned.

    Real-life performance

    It may be argued that the real-life performance

    of a membrane can only be obtained by industrial,

    long-term, evaluations of the membrane. Theoretical

    models would thus be merely of scientific interest.

    We agree that experimental investigations are impor-

    tant, but we are also convinced that theoretical

    modelling of the ultrafiltration process is necessary if

    we are to understand, and predict, the performance

    of UF membranes.

    The goal is, of course, to be able to design and

    optimize systems where the ultrafiltration equipment

    is an essential, but unnoticed, part of the system, as

    for example in electrodeposition systems. In order to

    reach this goal we need theoretical models to explain

    the (too often unpredictable) behaviour of ultrafil-

    tration membranes.

    Influence of pressure

    Fluid flow through porous media is usually de-

    scribed by the well-known Darcy equation 91

    KAP

    J=-

    p AL

    (1)

    where J is the water flux, K is the specific permeabil-

    ity of the medium, AP is the pressure difference

    across the medium, p is the fluid viscosity and AL is

    the thickness of the medium.

    It is, however, difficult to measure the effective

    thickness of an asymmetric membrane. The thickness

    and the specific permeability of the membrane are

    therefore usually combined to form a medium-spe-

    cific parameter, the hydraulic resistance of the mem-

    brane, R, = AL /K.

    In ultrafiltration, solutes and/or particles are re-

    tained by the porous medium, the membrane. This

    makes the introduction of more involved relations

    than eqn. (1) necessary.

    Solute is transmitted by convection towards the

    membrane as soon as the transport of solvent

    through the membrane commences, and the concen-

    tration is increased on the feed side of the membrane

    as solute is retained. The osmotic pressure difference

    Ail across the membrane can then become substan-

    tial. The driving force of the fluid is in this case given

    by AP - c AD, according to the three-parameter

    model of Kedem and Katchalsky [ 10, 111. The reflec-

    tion coefficient 0 indicates the degree of permselectiv-

    ity of the membrane. When CJ = 1 the solute is totally

    retained and when 0 = 0 it is totally permeable.

    The resistance of the accumulated solute at the

    membrane surface is sometimes represented as a

    hydraulic resistance R,. If we introduce hydraulic

    resistances instead of permeability in Darcys equa-

    tion and take the osmotic pressure of the solute into

    Fig. 1. Alternative models of concentration polarization: a) the

    osmotic pressure model, b) the gel layer model and c) the

    resistance in series model. The concentrations c,, cb, cP, and ca

    denote the concentrations at the membrane wall, in the bulk

    solution, the permeate, and the gel, respectively. The hydraulic

    resistances are: the membrane resistance R,,,, the resistance of the

    gel layer

    R

    and the resistance of the solute

    R .

    consideration, the ultrafiltration flux may be de-

    scribed by the generalized equation

    AP--AII

    J= /4%+%)

    The theoretical models that will be treated in this

    review can all be related to eqn. (2). These models

    are the osmotic pressure model, the gel layer model

    and the resistance in series model. Figure 1 shows

    concentration polarization as represented by the

    different models.

    Theoretical models

    In the

    osmotic pressure model

    the solute hydraulic

    resistance R, is substituted by a continuous, steep,

    concentration gradient at the membrane, resulting in

    a substantial osmotic pressure.

    In the gel layer model it is assumed that the solute

    concentration at the membrane surface reaches a

    limiting value, the gel concentration cg.

    In the resistance in seri es model the resistance to

    flow is represented by a hydraulic resistance of the

    solute,

    R,.

    The osmotic pressure AII is neglected.

    The osmotic pr essure model

    At typical UF feed concentrations the osmotic

    pressure of macrosolutes is negligible. Osmotic

    effects are therefore frequently ignored. Very high

    concentrations at the membrane wall and subse-

    quently substantial osmotic pressures have, however,

    been demonstrated for dextran and whey protein

    solutions, for example [ 121.

    The concentration at the membrane surface de-

    pends on the ratio between the convective transport

    of material to the membrane and the diffusion of

    material back to the bulk stream. During normal

    ultrafiltration conditions the solute concentration at

    the membrane wall, c,,

    may rise rapidly from the

    bulk value, cb (see Fig. l(a)). The fluid flow through

    the boundary layer adjacent to the membrane is

    frequently described by the film model as

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    3/15

    69

    where c,, is the concentration in the permeate and k

    is the mass transfer coefficient.

    Taking the osmotic pressure at the membrane

    wall into account, Wijmans et

    al. [

    131 have derived a

    relation between pressure and flux. Their calculated

    results have been found to be in good accordance

    with experimental results [ 141.

    Wijmans et

    al. [

    131 used the following relationship

    between the osmotic pressure and the concentration

    at the membrane wall:

    I

    lT, =

    ac,

    (4)

    where a and II are solution-dependent constants.

    When the solute is totally retained (d = 1 and

    cp = 0), and the hydraulic resistance of the solute,

    R,,

    is

    neglected, combination of eqns. (2)-(4) gives the

    following expression relating flux and applied pres-

    sure [13]:

    J = AP -

    ac,, exp(nJ/k)

    11&

    (5)

    A limitation of eqn. (5) is the availability of osmotic

    pressure data for solutions containing high concen-

    trations of macromolecules.

    The pure water permeability (PWP) of ultrafiltra-

    tion membranes is given as the pure water flux

    (PWF) per unit pressure difference:

    PWP = (PWFIAP),, = = l/(pR,)

    (6)

    The flux as a function of operating pressure has been

    calculated for three membranes with varying pure

    water fluxes. The results are given in Fig. 2.

    Initially, the flux increases linearly with pressure,

    as can be seen in Fig. 2. As the flux increases further,

    more and more material is transported to the mem-

    brane wall and concentration polarization increases.

    The flux levels off as a consequence of the increasing

    concentration polarization (see point ( 1) under Gen-

    eral trends).

    250

    f 200

    g 150

    x 100

    50

    0

    0 0.2 0.4 0. 6 0. 8 1.0 1.2

    PRESSURE (MPa)

    Fig. 2. The effect of operating pressure on flux for three ultrafil-

    tration membranes with varying permeabilities, calculated using

    eqn. (5). The PWPs of the three membranes were 1000, 250 and

    1C01m-2h- MPa- at T=2OC (this corresponds to

    & = 3.6 X 102, 14.4 x 10 and 36.0 x 10tZm~, respectively).

    Lines without symbols represent the pure water fluxes of the three

    membranes and lines with symbols represent the flux of a

    1.0 wt.% solution (c, = O.Ol), which has an osmotic pressure of

    0.1 MPa at a concentration of 10 wt.% (a = 10 MPa and n = 2).

    This applies fairly well to solutions of dextran [ 141. The value of

    the mass transfer coefficient was 2 x 10m5 m SK.

    The inf luence of the pure water j?ux

    Compare, for example, the two membranes with

    PWP = 1000 and 100 1 m- h- MPa- in Fig. 2. As

    can be seen, the increase of the flux declines faster

    for the high permeability membrane than for the low

    permeability membrane. This explains why mem-

    branes with an initially high flux exhibit greater flux

    declines than low permeability membranes (see point

    (2) under General trends).

    The influence of concentration polarization is

    small for membranes with low pure water fluxes.

    They are, consequently, almost linearly dependent

    on pressure, as stated in Darcys equation.

    Membranes with high pure water fluxes (those

    that look so promising in membrane tables ) suffer

    from high osmotic pressures which are caused by

    high concentrations at the membrane wall. Owing to

    this high osmotic pressure, the flux of these mem-

    branes is much lower than the pure water flux.

    A reasonable question is, of course, if we can ever

    reach the level of the pure water flux of the high

    permeability membranes. One way of accomplishing

    this is by increasing the mass transfer coefficient

    k.

    The influence of other parameters may be obtained

    from the derivative of eqn. (5). The derivative of this

    equation is

    [ 131

    dJ=L(l +A %)-

    d AP pR,

    The order of magnitude of the deviation from the

    pure water flux is given by the second term in eqn.

    (7), AfI

    n/pR,k.

    The product flux is affected by, for

    example, the permeability of the membrane (the

    effect of

    R,

    is demonstrated in Fig. 2), the tempera-

    ture of the solution (which affects p), the osmotic

    pressure (AD and n) and the cross-flow velocity

    (affects

    k).

    These parameters are discussed further

    later in this paper.

    The gel layer model

    The pressure independence of flux when the oper-

    ating pressure is increased was originally explained

    by Blatt

    et al. [

    151 as being due to the formation of

    a gel layer at the membrane surface. They stated that

    a limiting flux value is reached when the concentra-

    tion of solute in the boundary layer reaches the gel

    concentration cg. The flux-limiting value for a totally

    retained solute (cr = 0) at gel layer conditions is

    given by eqn. (3) as

    J=kln 2

    0

    (8)

    In the gel layer model the osmotic pressure is

    assumed to be zero. The fluid flow is then described

    by

    J=

    AP

    14Rm + Rs)

    (9)

    When the concentration at the wall has not

    yet reached the gel concentration, the polarized

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    4/15

    70

    boundary layer offers a resistance to the flux,H&,

    an d R, = R, ,

    .

    Under gel layer conditions the resis-

    tance of the gel layer represents the resistance of the

    solute, R, = R ,. When both concentration and gel

    polarization occur the resistance of the solute is

    R,= R,+Rb, [15].

    The gel layer model predicts the flux to be inde-

    pendent of operating pressure. An increased pressure

    merely results in a thicker gel layer (larger &),

    which retards the flux to its original value.

    The gel layer model has been frequently used to

    correlate experimental limiting fluxes [ 16- 181. The

    gel concentration may be obtained by extrapolation

    of a plot of J versus In cb. It has, however, been

    shown that the information obtained on gel concen-

    trations is not reliable. For identical solutions differ-

    ent authors have found widely varying values at cg

    [191.

    Experimental evidence [ 151 indicates that the vis-

    cous solution in the boundary layer approaches a

    close-packed configuration of low hydraulic perme-

    ability. It has been argued, however, that it is only

    materials such as agar, pectin, gelatin and some

    proteins that might denaturate at the membrane

    surface to give true gels [ 141. It has also been shown

    that feed solutions of various macrosolutes with

    concentration cb = cg did not give zero flux [20].

    Based on their analysis, Wijmans et a l . [ 131 con-

    cluded that osmotic pressure limitations would be

    expected in the ultrafiltration of macrosolutes with

    molecular weights of 10 000- 100 000, and gel layer

    limitations would be expected for larger molecules.

    Th e r es is tan ce n ser ies mod el

    The concentrated solution of macromolecules at

    the membrane wall provides a physical barrier to

    solvent and solute transport. The resistance of this

    boundary layer may be treated as one or several

    resistances in series with the membrane, as already

    discussed in the gel layer model.

    The resistance in series model has been shown to

    apply for dextran solutions [21]. The resistance of

    the solute in this investigation was estimated from

    independently measured sedimentation coefficients.

    The results were also in reasonable agreement with

    results obtained from the osmotic pressure model.

    The resistance in series model predicts a pressure-

    independent flux, as does the gel layer model. That is

    to say, an increase in the pressure results in a thicker

    gel layer and an increased hydraulic resistance.

    An advantage with the resistance in series model

    is that it makes it possible to distinguish between th0

    influence of different flux decline phenomena.

    The resistance to flow may be accounted for by a

    number of resistances: the resistance R, of the mem-

    brane, the boundary layer resistance R,,, the gel

    layer resistance R,,

    and the adsorbed layer resistance

    R,. Equation (2) may then be written as

    AP

    J =p(R,+R,+R,+Rh , )

    (10)

    01

    0 a

    b

    1000

    2000 3000

    4000

    ROTATION SPEED (rpm)

    Fig. 3. Determination of the influence of the membrane resis-

    tance, boundary layer resistance, gel layer resistance and adsorbed

    layer resistance for 1.0 wt.% BSA at 0.5 MPa. The numbers

    denote the order in which the measurements were performed. The

    rotary module is described in detail in ref. 22.

    The influence of different resistances on flux has

    been investigated for bovine serum albumin (BSA) in

    a rotary module [22]. Different flow resistance re-

    gions were distinguished by varying the rotation

    speed. Figure 3 shows the experimental results.

    First, the membrane resistance & was deter-

    mined by ultrafiltration of pure water at high rota-

    tion velocity. From the pure water flux the

    membrane resistance could be calculated as, in this

    case, R, = R , = R b, = 0.

    The pure water was then replaced by a 1.0% BSA

    solution. The flux was measured at decreasing

    rotation velocities. Concentration polarization and

    gel layer formation were assumed to be zero

    (R , = R,, = 0) as long as the flux remained constant.

    The flux began to decrease when the velocity was

    decreased to about 2400 rev min- as can be seen in

    Fig. 3. As long as it was possible to restore the initial

    flux value by increasing the rotation velocity to the

    original value, the influence of the gel layer was

    assumed to be zero, i.e.

    R, = 0.

    Down to about 1000 rev min- the original per-

    meate flux was almost completely restored when the

    velocity was increased. When it was no longer possi-

    ble to attain the original flux values when the veloc-

    ity was increased, it was deduced that an irreversible

    phenomenon had taken place, namely gel layer for-

    mation.

    Compac t i o n o f memb r ane an d gel l a yer

    The theoretical models described above all predict

    a limiting flux value at increasing pressure. They do

    not, however, explain why the flux may decrease

    when pressure is increased. This phenomenon may,

    however, be explained as being the result of com-

    paction of the membrane and/or the gel layer.

    The membrane pure water permeability K has

    been related to pressure for reverse osmosis cellulose

    acetate membranes by [23 ]

    K= exp( -uP)

    (11)

    where K,, is the specific membrane permeability at

    zero pressure and LX s a measure of the susceptibility

    of the membrane to compaction.

    It has been questioned whether the porosity of

    ultrafiltration membranes is influenced to any appre-

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    5/15

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    6/15

    72

    250

    200

    150

    100

    50

    0

    0 5 10

    15

    20

    CONCENTRATION IN BULK (w- )

    Fig. 6. The influence of bulk concentration on flux at two operat-

    ing pressures, 0.5 and 1.0 MPa. The values of the parameters in

    eon. (5) were: PWP-2501m-2h-MPa-1 at T=2OC

    (R, = 1.4 x lOI rn-), k = 2 x 10e5 m s-, (I = 10 MPa and

    n = 2.

    trends). The film model (eqn. (3)) predicts, for exam-

    ple, that the flux J varies proportionally with the

    logarithm of the bulk concentration ct,, and the gel

    layer model predicts that the intercept of a plot of J

    versus In ct, corresponds to the gel concentration cg.

    The variation of flux with concentration, as pre-

    dicted by the osmotic pressure model (eqn. (S)), is

    shown in Fig. 6.

    The flux is, as expected, higher at the high operat-

    ing pressure. But, although the pure water flux (at

    c,, = 0 in Fig. 6) is doubled when the pressure is

    doubled, this considerable difference between the two

    flux curves vanishes as the concentration increases.

    The two flux curves in Fig. 6 approach zero at the

    bulk concentration at which the osmotic pressure

    equals the operating pressure. It is thus possible to

    increase the final concentration of the solution by

    increasing the operating pressure.

    The osmotic-pressure-concentration relationship

    of the solution not only determines the final retentate

    150

    125

    100

    E-

    2 75

    =

    s

    k

    50

    25

    CONCENTRATION IN BULK (W- )

    Fig. 7. The influence of bulk concentration on flux for two solutes

    with different osmotic pressure-concentration relationships. The

    high osmotic pressure solute (a = 23 and n = 2) corresponds to a

    solution of PEG. The low osmotic pressure solute (a = 10

    and n = 2) corresponds to a solution of dextran. The values of

    other parameters in eon. (5) were: PWP = 250 1 m mz h

    MPa

    at

    T = 20 C

    R, =

    1.4 x 10s mm), AP = 0.5 MPa and

    k =

    2 x 10VSm s-r.

    200

    150

    100

    50

    0

    CONCENTRATION IN BULK (W-/o)

    Fig. 8. The influence of bulk concentration on flux for two

    membranes with different pure water permeabihties, 1000 and

    250 1 rnpz hh MPa- at T=20C (&=3.6x IO* and

    1.4 x IOr m-, respectively). The values of the parameters in eqn.

    (5) were: AP = 0.5 MPa,

    k = x

    IO-srn s-r, a = 10 MPa and

    n = 2.

    concentration. It also determines the magnitude of

    the ultrafiltration flux. In Fig. 7 the flux of two

    solutions with different osmotic-pressure-concentra-

    tion relationships is shown. The high osmotic-pres-

    sure-concentration relationship applies fairly well to

    a solution of polyethylene glycol (PEG) and the low

    osmotic-pressure-concentration relationship applies

    to a solution of dextran [14]. As shown in the

    Figure, the flux of the solute with the higher osmotic

    pressure (PEG) was not only lower, but also de-

    creased faster.

    When the concentration in the feed increases, the

    flux becomes less and less affected by the PWP of the

    membrane, as can be seen in Fig. 8. Thus, the flux at

    the end of the concentration operation cannot be

    increased by the introduction of a membrane with a

    higher pure water flux.

    Wijmans et al. [ 131 have rearranged the osmotic

    pressure model relationship, eqn. (5), to give

    dJ

    k

    -= _

    d In ct, 1 + &, k/AII n

    (16)

    Equation ( 16) predicts a slope approaching -k for

    the J versus In c,, plot at high bulk concentrations

    (high AII values). This is also predicted by the gel

    layer model.

    The mass trader coefficient

    The flux is thus very sensitive to operational

    parameters that may affect the mass transfer co-

    efficient

    k. The

    mass transfer coefficient may be

    obtained from correlations of the form

    where

    d,,

    is the hydraulic diameter of the flow chan-

    nel,

    D

    the diffusivity, and Sh, Re and Sc are the

    Sherwood, Reynolds and Schmidt numbers. The

    constants

    A, a, b

    and c vary with flow conditions.

    The mass transfer coefficient has been shown to

    have a slight tendency to decrease with increasing

    pressure and bulk concentration [ 121. The decrease

    in flux with increasing pressure at elevated pressures

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    7/15

    73

    was suggested earlier in this paper to be the result of

    the compaction of the membrane and gel layer at

    high pressures. The decrease in the mass transfer

    coefficient with increasing pressure has been put

    forward as another explanation of this phenomenon

    [141.

    The diffusivity of macrosolutes is usually depen-

    dent on concentration. Owing to the concentration

    polarization, the concentration at the membrane sur-

    face may differ significantly from the concentration

    in the bulk, This may, of course, modify the effective

    mass transfer coefficient. The diffusivity is also in-

    creased by an increase in the temperature. However,

    the operational parameter that has the largest impact

    on the mass transfer coefficient is the cross-flow

    velocity.

    I nfl uence of the cross ow velocity

    In both the osmotic pressure and the gel layer

    models it is presumed that the mass transfer co-

    efficient is proportional to un, where u is the cross-

    flow velocity and a is the Reynolds number exponent

    in eqn. ( 17). The value of the exponent has been

    much debated. A thorough review of different mass

    transfer coefficients, together with their adaption to

    ultrafiltration, is given in ref. 25.

    In Figs. 9 and 10 a value of II = 0.75 according to

    the Chilton-Colburn model has been used. The in-

    fluence of the cross-flow velocity on the flux for two

    membranes of differing pure water flux (Fig. 9) and

    at varying bulk concentrations (Fig. 10) has been

    calculated from eqn. (5).

    The influence of the cross-flow velocity increases

    as the permeability of the membrane increases, as

    can be seen in Fig. 9. This is easily understood as we

    know that the concentration polarization is more

    pronounced for membranes with higher pure water

    fluxes.

    It is, perhaps, not that easy to anticipate how the

    cross-flow velocity influences the flux at various con-

    centrations. As a rule, we can say that flux becomes

    independent of cross-flow velocity at a much lower

    velocity for low concentration solutions. In Fig. 10

    it can be seen, for example, that the solution with

    2 5 0

    I

    CROSS FLOW VELOCITY (m/s)

    Fig. 9. The influence of the cross-flow velocity on flux for two

    membranes with different pure water permeabilities, 500 and

    250 1 m-* h- MPa-

    at T=20C (R,,,=7.2x IO* and

    1.4 x IO m-l). The values of the other parameters in eqn. (5)

    were: AP = 0.5 MPa,

    cb = 1.0 wt.%, k = 2 x 10-s m s-

    at

    U =4ms-1,

    cr = 10 MPa and n = 2.

    01

    8

    1

    0

    2

    4

    6

    8

    1

    CROSS FLOW VELOCITY (m/s)

    Fig. 10. The influence of the cross-flow velocity on flux at three

    different bulk concentrations, 0.1, 1.0 and 5.0 wt.%. The values

    of the parameters in eqn. (5) were: PWP = 500

    1mm*

    hh

    MPa-

    at T = 20 C (R, = 7.2 x 10 II-),

    AP = 0.5 MPa,

    k=2x10-5ms-atu=4ms-.a=10MPaandn=2.

    ci, = 0.1 wt.% has reached its maximum flux at a

    velocity of approximately 4 m s-l, while the flux of

    the 5.0 wt.%

    solution is still increasing at

    u =8ms-.

    If a gel layer is formed, the flux may become

    independent of cross-flow velocity at low velocities,

    even for high concentration solutions, because once

    a gel layer has been formed it is not possible to

    recover the flux by simply increasing the velocity

    (this was illustrated in Fig. 3). Thus, the velocity

    ought to be kept, if possible, at such a level that gel

    formation is avoided. If the cross-flow velocity is

    decreased temporarily during operation, the forma-

    tion of a gel layer may be avoided if the flux also is

    decreased, for example, by temporarily lowering the

    operating pressure.

    Rotating modules

    The ultrafiltration flux of a membrane is con-

    trolled by the rate at which the retained solutes can

    be transferred from the membrane wall back into the

    bulk fluid. Thus, operational variables that aid back-

    transport from the membrane will directly increase

    the flux.

    The shear rate at the membrane wall is, as has

    been clear from the above, the major depolarizing

    parameter. This fact initiated the development of

    rotating modules [26,27]. However, the commercial

    success of rotating modules has been limited (due to

    high investment costs).

    Positive results from tests with a modified plate

    and frame rotary module, the ABB CROT filter,

    have been reported recently. In pilot-plant equip-

    ment, bleaching effluent from a Swedish pulp mill

    has been treated successfully over a one-year period.

    A larger pilot plant (membrane area 200 m2) has

    recently been installed at the same mill.

    During ultrafiltration of oil emulsions in a rotary

    module it was found that the flux could be increased

    further if the roughness of the module wall was

    increased [28]. The increase in the wall roughness

    was accomplished by the introduction of a net. The

    increase in the flux was ascribed to an increase in the

    mass transfer coefficient due to the presence of the

    net, acting as a static convection promoter.

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    8/15

    74

    It has also been shown that the mass transfer

    coefficient can be increased, not only by increasing

    the wall roughness, but also by using corrugated

    membranes [29]. It was found that corrugations in-

    creased the mass transfer more effectively than an

    increase in the cross-flow velocity. These tests were,

    however, performed with reverse osmosis mem-

    branes, but a similar type of effect of membrane

    surface roughness has been demonstrated for ultrafil-

    tration membranes [301.

    I nfl uence of suspended soli ds

    The presence of suspended solids may enhance the

    flux. There are several examples of this phenomenon

    [31]. In such cases of flux enhancement it is thought

    that the solids intermittently disturb the polarized

    layer, sweeping away deposited solutes [32, 331.

    Influence of time

    Filtration theory predicts that flux decreases as

    t I* under unstirred conditions [34]. This is also

    predicted by the gel layer [35] and osmotic pressure

    models [361.

    The influence of time under cross-flow conditions

    is more complicated. The film mode1 predicts that a

    rapid drop in flux occurs as the boundary layer is

    built up. Several workers have studied the initial

    polarization period experimentally. Polarization time

    varied from about 5 to 50 s [ 181. The flux at the end

    of the initial polarization period should be the limit-

    ing flux predicted by eqn. (8).

    The gel layer model, as well as the osmotic pres-

    sure model, predicts that, after the initial polariza-

    tion period, flux remains constant with time.

    However, a gradual, long-term decay in flux is noted

    in many applications. This flux decline is due to

    fouling. The causes of fouling are very disparate. It is

    thus difficult to derive a general theoretical expres-

    sion for the flux-time relationship.

    A simple relation between flux and time is the

    empirical relationship

    J = Jot

    (18)

    where JO is the initial flux and n is an exponent less

    than zero. Equation (18) has been much used to

    predict flux decline in reverse osmosis systems [37].

    Equation ( 18) gives, however, no information about

    the influence of the operational parameters. Nor is

    the influence of the properties of the solution taken

    into account.

    In the remaining part of this paper we will discuss

    the influence of the nature of the solutes, the mem-

    brane material, pH and ionic strength on fouling.

    We will also discuss how fouling can be controlled.

    Fouling

    It is very risky to make general statements about

    the influence of different parameters on fouling. The

    (a)

    :,:,:>

    D

    \I

    \\\

    ,.;

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    9/15

    Characteristic pore sizes were determined for

    clean and fouled membranes by Hanemaaijer et al.

    [45]. The pore size changes caused by solute adsorp-

    tion were derived from saccharide retention data. It

    was shown that for a hydrophilic low protein ad-

    sorbing membrane, the calculated pore size was not

    affected by protein adsorption, whereas the pore size

    of hydrophobic membranes was reduced.

    Adsorption of one of the solutes in a mixture may

    thus also affect the retention of other compounds.

    Solutes that are not retained by the membrane when

    tested alone may very well be retained due to solute-

    membrane and solute-solute interactions of other

    compounds in the solution. It is thus advisable to

    confirm the expected retention characteristics of a

    membrane by tests with real solutions and not base

    predictions of membrane performance on tests with

    idealized solutions. But even predictions based on

    tests with real solutions may be unreliable as the

    composition of many process streams may occasion-

    ally differ.

    I nfl uence of membrane materi al

    It is commonly recognized that hydrophobic

    membranes have a larger fouling tendency than hy-

    drophilic membranes. Even the pure water flux may

    be affected by the nature of the membrane material

    as absolutely pure water is very hard to find (a

    well-known problem for anyone who works with

    membranes).

    Investigations

    [

    3 1,46,47] have shown that the

    pure water flux of hydrophobic membranes decreases

    with time, while the pure water flux of hydrophilic

    membranes is less time dependent. This phenomenon

    has been attributed to contamination of the pure

    water by bacteria and trace colloids [31] and to trace

    amounts of waste products of micro-organisms [46]

    present in the pure water.

    I nf ruence of membrane pore size

    Membrane performance cannot, however, be re-

    lated simply to the membrane material. For example,

    the flux decline of three membranes, all made of

    polysulphone, was noticeably different when the pH

    of a BSA solution was altered [48].

    The same disparate behaviour was found for three

    membranes with varying pore sizes used for the

    treatment of caustic bleach plant effluent [49]. The

    effect on the flux of the three membranes when the

    pH was lowered differed considerably, as can be seen

    in Fig. 12.

    At the higher pH the flux of all three membranes

    was almost identical, while when the pH was lowered

    to below the isoelectric pH of the solution, the

    magnitude of the flux followed the cut-off of the

    membranes, that is, the densest membrane experi-

    enced the largest flux decline, while the flux of the

    membrane with the largest pores was unchanged.

    It was thus the membrane with the smallest pores

    that exhibited the greatest flux decline in this case.

    -0 20 40 80 100

    a)

    TIME

    (h)

    E 200

    2 150

    zj 100

    ii 50

    0

    0

    20 40

    TIME (:;

    80 100

    b)

    Fig. 12. Flux of three membranes with varying cut-off at different

    values of pH: a) pH 11 and b) pH 4.6. Two membranes were

    made of polysulphone, PU120 with a cut-off of 20 000 and PU608

    with a cut-off of 8000, and one was made of polyethersulphone,

    ES404 with a cut-off of 4000. All three membranes were manufac-

    tured by PC1 Membrane Systems. The operating pressure was

    0.8 MPa and the temperature 55 C. Data from ref. 49.)

    Usually, however, the relative flux decline is greatest

    for the membrane with the highest pure water flux

    (which is often the membrane with the largest pores).

    I nfl uence of surfactants

    Surface chemical phenomena play an important

    role in the fouling of membranes. It is well known,

    for example, that hydrophobic solutes (e.g. BSA) are

    more readily adsorbed onto the membrane surface

    than hydrophilic solutes (e.g. dextran).

    Antifoams used in fermentation can cause consid-

    erable fouling of membranes during downstream

    processing [50-541. The membranes in some com-

    mercial membrane plants have even been destroyed

    when an antifoam agent has been exchanged. In such

    cases irreversible adsorption of the new antifoam

    agent has caused a nearly zero flux and made it

    necessary to replace all the membranes.

    Despite some frightening examples of the influ-

    ence of surfactants, as a class of materials they do

    not lead to fouling problems. It has been shown, for

    example, that the flux of ultrafiltration membranes

    can be enhanced if the membranes are pretreated

    with non-ionic surfactants before use [55]. There is

    also evidence that cleaning, due to the surfactants in

    the cleaning solution, may not be considered merely

    as a cleaning operation, but also as an in situ surface

    modification [49].

    Some surfactants, however, may interact with the

    membrane and cause an irreversible flux decline.

    Many parameters may affect the flux decline of

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    10/15

    76

    surfactant solutions, for example, the membrane ma-

    terial, the pore size of the membrane and, of course,

    the nature of the surfactant.

    Surf actants causing fou li ng problems

    The flux of antifoams with cloud points has been

    shown to be virtually zero for polysulphone mem-

    branes at, and above, the cloud-point temperature

    [54]. When the temperature was lowered the flux

    returned approximately to the original level. No flux

    decrease was observed for a hydrophilic membrane.

    This type of flux decline was not observed for non-

    cloud-point antifoams.

    Irreversible fouling is often observed when

    cationic polymers and surfactants are in contact with

    negatively charged membranes. Cationic surfactants

    do not always

    cause irreversible flux declines, how-

    ever. The initial flux of cellulose acetate membranes

    was restored when the membranes were flushed with

    pure water after ultrafiltration of the cationic surfac-

    tant hexadecyl trimethylammonium bromide [47]. A

    slight, irreversible flux decline was observed for

    membranes of polysulphone and polyvinyldifluoride

    when exposed to the same surfactant.

    Critical

    micelle concentration

    The flux decline on the addition of a surfactant is

    usually very fast, within minutes, and even small

    amounts of surfactant are enough to cause a consid-

    erable drop in flux. At the critical micelle concentra-

    tion (CMC) micelles are formed and the retention

    then often increases as the micelles are too large to

    pass freely through the membrane pores.

    At CMC the flux often stabilizes at a constant

    level, as can be seen in Fig. 13. The Figure shows the

    influence of a non-ionic surfactant on the flux of a

    600

    506

    406

    300

    266

    160

    0

    100

    200 300

    400 500

    TIME

    (h)

    Fig. 13. Variation of flux during addition of the non-ionic surfac-

    tant Triton X-100. The polysulphone membrane, DDS GRIO, has

    a nominal cut-off of 500 000. The line with the open circles shows

    the Aux decline of a membrane that has only been conditioned

    with pure water. The line with the solid dots shows the flux decline

    of the same membrane, but after cleaning of the membrane with

    a commercial alkaline cleaning agent, Ultrasil 10 from Henkel.

    Surfactant was added stepwise, the concentration 0.1 CMC, for

    example, indicates that the concentration was 10 of the critical

    micelle concentration. The operating pressure was 0.1 MPa and

    the temperature 30 C.

    polysulphone membrane, before and after cleaning

    of the membrane. After cleaning, the pure water

    flux of the membrane increased from 140 to

    540 1m- h-.

    The tlux increase after cleaning, shown in Fig. 13,

    makes it easy to understand why cleaning is such a

    vital part of membrane operations. This will, how-

    ever, be treated in more detail in a following section.

    We will now take a look at two other properties that

    may influence the fouling tendency of solutes,

    namely the solution pH and the ionic strength.

    Influence

    of

    pH and ionic strength

    The solution pH and ionic strength affect the

    charge, stability and tendency to aggregate of the

    solute molecules and this influences, of course, the

    membrane performance.

    A reduction in the flux at the isoelectric pH has

    been found for many solutes (BSA [20,31,56] and

    whey [57,58], for example). At the isoelectric pH the

    net charge is zero. Thus, a solute which under nor-

    mal conditions is not adsorbed onto the membrane

    due to electrostatic repulsion (Fig. 14(a)) is no

    longer repelled at the isoelectric pH (Fig. 14(b)).

    Alteration of pH does not only influence the

    solute-membrane interaction, but also the solute-

    solute interaction. If the electrostatic repulsion be-

    tween solute molecules ceases this may result in a

    more closely packed and denser gel layer. The same

    result may be obtained if salt is added to the solu-

    tion.

    It has been shown that the addition of salt to a

    starch solution results in a drastic decrease in flux

    [59]. It was suggested that the cations shielded the

    charges of ionized carboxyl groups on the oxidized

    starch chain. This shielding caused contraction of the

    starch molecules which resulted in lower viscosity

    and stability of starch in the solution. Densification

    of the starch gel increased the hydraulic resistance of

    the gel layer and the flux was thus decreased. No

    interactions were observed for completely un-ionized

    pearl starch.

    t hydrophob ic t ll

    (b)

    Fig. 14. (a) A charged solute is repelled by a membrane of the

    same charge. (b) At the isoelectric pH the solute is uncharged and

    may be adsorbed onto the membrane.

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    11/15

    77

    During ultrafiltration of kraft black liquor it was

    found that large molecules, that were able to perme-

    ate the membrane at a high ionic strength, were

    almost completely retained at low ionic strength [60].

    The influence of pH may be altered if salt is

    present in the solution. This has been demonstrated

    for BSA [LX]. In the absence of salt the flux was at its

    minimum at the isoelectric pH. In the presence of

    0.2 M NaCl the flux was lowest at pH 2 and in-

    creased monotonically with increasing pH.

    The zeta potenti al of the membrane

    The pH may, however, affect not only the charge

    of the solute, but also the net charge of the mem-

    brane, the zeta potential. It has been shown [61] that

    a polycarbonate membrane has an isoelectric point

    at pH 4. Above this pH the membrane had a slight

    negative surface charge, which increased in magni-

    tude with increasing pH. The surface was found to

    be fully charged above a pH of about 6. The surface

    was neutral below a pH of 3. The same dependence

    on pH has been shown for membranes of polysul-

    phone and polyamide [62].

    Most natural substances are negatively charged.

    The maximum flux of a negatively charged mem-

    brane ought then to be found at a high pH, where

    the electrostatic repulsion is at its maximum. This

    supposition has been confirmed in many investiga-

    tions. For example, for bleach plant effluent [49] and

    for BSA solutions with salt present [56].

    The flux may, however, also be higher at a low

    pH. This has been found, for example, during ultra-

    filtration of soybean extract [63] and for cheddar

    cheese whey [64]. Of course, not only the pH may

    change the charge of the membrane. If solutes are

    adsorbed on the membrane this may change not only

    the charge, but also the hydrophobic/hydrophilic

    properties of the membrane.

    Pretreatment

    The fouling ability of a solution may be dimin-

    ished by suitable pretreatment. The pretreatment can

    be mechanical, thermal or chemical. Modifications to

    the feed solution include adjustment of pH, removal

    of fibres, fines, etc. and heat treatment of the feed

    solution.

    Mechanical pretreatment is especially important

    when using thin-channel membrane modules. The

    danger of plugging the flow channels is greatest for

    spiral, hollow-fibre and plate and frame modules.

    Tubular modules also require some solution pre-

    treatment. When treating effluents in the pulp and

    paper industry a continuous sand filter is often used

    as a prefilter for plate and frame modules [65], while

    for tubular membranes a 1 mm screen is sufficient

    t 351.

    Centrifugation is often used, for instance to re-

    move fines from whey and fibres from starch pro-

    cessing waters. Another example is centrifugation of

    blood for the fractionation of red blood cells from

    blood serum proteins.

    Thermal pretreatment is often necessary in order

    to avoid undesirable changes, especially when pro-

    cessing biological materials which are sensitive to

    biodegradation, such as liquid foods, food waste

    waters and fermentation broths. The pH of whey, for

    instance, decreases rapidly without heat pretreat-

    ment.

    Calcium phosphate is known to cause a flux de-

    cline during the ultrafiltration of milk and whey

    when the temperature and pH are not carefully

    controlled [39]. Milk and whey are therefore nor-

    mally heated to 55-60 C and held at this tempera-

    ture for up to half an hour before ultrafiltration at

    50-55 C. As the solubility of calcium phosphate

    decreases with increasing temperature the salt precip-

    itates during the preheating operation. If the milk, or

    whey, is not preheated, calcium phosphate may pre-

    cipitate at the membrane surface and in the pores.

    This phenomenon, known as scaling, can also occur

    in the porous support. The deposited salt is very

    difficult to remove. However, no significant flux de-

    cline is observed when the salt is precipitated before

    the ultrafiltration.

    Chemical pretreatment,

    can, in many cases, im-

    prove flux. The reduction of flux of protein solutions

    at the isoelectric pH has already been discussed in

    this paper. This reduction is avoided if the pH is

    adjusted before the protein solution is concentrated.

    Cleaning

    Cleaning and disinfection of the membrane plant

    are very important operations, especially in food

    processing. Fouling of the equipment, as well as the

    presence and growth of micro-organisms, necessi-

    tates regular cleaning and disinfection cycles. Mem-

    branes used in food plants are generally cleaned at

    least once a day, while those used for the treatment

    of electrodeposition paints, or for the production of

    desalinated water, need to be cleaned less frequently,

    usually no more than twice a year.

    A cleaning cycle generally includes the following

    stages: removal of product from the system, followed

    by rinsing the system with water; cleaning in one or

    several steps, followed by rinsing the system with

    water; disinfection of the system.

    The product should be removed at the same tem-

    perature as that used in the process. This is impor-

    tant, for instance, when dealing with products which

    tend to form gels at low temperatures.

    Both retentate and permeate should be discharged

    when the system is

    rinsed.

    Rinsing should continue

    until both the retentate and permeate streams are

    totally clear and neutral.

    Cleaning solutions

    A large number of

    cleaning

    recommendations

    are reported in the literature. Cleaning studies are

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    12/15

    usually trial and error investigations, whereas sys-

    tematic studies are very rare. Temperature, time,

    concentration and type of surface active agent are

    important cleaning parameters that vary with fou-

    lants and membrane material.

    As a rule, mineral deposits are removed by acids

    and proteins by alkaline solutions. For some prod-

    ucts, it is necessary to use different chemicals in

    succession to obtain a satisfactory cleaning result.

    For milk, where protein deposits dominate, an alka-

    line formulation is used first, sometimes followed by

    acid treatment to remove mineral deposits. If poly-

    sulphone membranes are used, a final alkaline clean-

    ing is often carried out in order to improve the flux.

    For whey, where mineral deposits dominate, acid

    cleaning is often performed first, and is then fol-

    lowed by alkaline cleaning.

    Some care is recommended when selecting a

    cleaning procedure since an injudicious choice could

    be catastrophic. As an example, in ref. 67 it is shown

    that pectine, present in fruit juices, presents difficult

    fouling problems. Especially alkaline detergents, and

    also unsuitable acid products, may lead to complete

    clogging of the membranes.

    In order to obtain a good mechanical cleaning

    effect, the cross-flow rate is often higher and the

    pressure lower during cleaning than during normal

    operating conditions.

    Cleaning solutions are often a mixture of chemi-

    cals. Alkaline cleaning solutions usually contain

    sodium hydroxide, phosphate, sequestering agents

    and surface active agents. Examples of some cleaning

    chemicals and the concentrations used are given in

    Table 1.

    The choice of surface active agent is very impor-

    tant, since some types may be adsorbed onto the

    membrane surface, resulting in a flux reduction. On

    the other hand, with a suitable choice of surfactant

    in the cleaning solution considerably enhanced fluxes

    may be obtained. Thus, as stated previously, it has

    been suggested [49] that cleaning may be considered

    not merely as a cleaning operation, but also as an in

    situ surface modification.

    The influence of the surfactant on the flux was

    demonstrated in an investigation where membranes

    fouled during ultrafiltration of whey were cleaned

    with different types of surfactants [68]. Considerably

    improved whey fluxes were experienced after clean-

    ing with a solution containing an anionic surfactant,

    TABLE I. Examples of concentrations of chemicals used in

    cleaning and disinfection solutions

    Cleaning/disinfecting agent

    Concentration (%)

    Caustic soda 0.5-1.0

    EDTA-Na, 0.5-1.0

    Nitric acid 0.3-0.5

    Chlorine 0.002-0.02

    Hydrogen peroxide 0.1

    Sodium bisulphite 0.25

    for example. Another example of the positive effect

    of cleaning is demonstrated in Fig. 13.

    Disinfection

    A diluted solution of hypochlorite, hydrogen per-

    oxide or sodium bisulphite is often used for

    disinfec-

    tion.

    It is important to clean and disinfect not only

    the active side of the membrane, but also the perme-

    ate side. This fact has to be considered especially for

    reverse osmosis membranes, since these dense mem-

    branes retain the cleaning and disinfecting agents.

    Membrane plants are often run discontinuously.

    When idle, the modules should be preserved in a

    diluted disinfection solution, for example sodium

    bisulphite, in order to prevent microbial growth in

    the membrane and/or membrane support.

    Means of facil itating cleaning

    Generally, the effect of the cleaning operation is

    checked by measuring the water flux after cleaning at

    a defined pressure, temperature and cross-flow veloc-

    ity. This is not a reliable measure, however. A high

    water flux does not guarantee a good product flux,

    whereas a low water flux indicates that the cleaning

    is not sufficient. The product flux in the following

    run is a better indication of whether the membranes

    have been cleaned satisfactorily [69].

    The cleaning operation is facilitated if the mem-

    brane fouling is reduced as much as possible. Factors

    which must then be considered are:

    _

    membrane properties (choose the most suitable

    type of membrane and module for each application);

    _

    solution pretreatment (before ultrafiltration, re-

    move salts that can cause severe scaling, for exam-

    ple) ;

    - flow velocity (try to keep the cross-flow velocity at

    such a level that gel formation is avoided);

    _

    rinsing water quality (be aware of the quality of

    the rinsing water).

    The presence of iron, silica, calcium and other

    inorganic ions in the rinsing water can cause precipi-

    tation of salts which are difficult, or impossible, to

    remove. The rinsing water quality required is as

    follows [701:

    Iron < 0.05 ppm

    Manganese < 0.02 ppm

    Silicate ( SiOz)

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    13/15

    understood. A vast number of empirical investiga-

    tions show that it is very hazardous to make general

    statements, especially about the influence of different

    parameters on fouling.

    Experience indicates that if we are to be able

    to describe the dynamics of ultrafiltration it will

    be necessary to combine knowledge from different

    areas. The theoretical models must combine knowl-

    edge and experience of fluid mechanics, mass trans-

    fer and surface chemistry phenomena.

    Nomenclature

    A*

    A

    a

    a

    c

    cb

    57

    CP

    :

    4

    4

    J

    JO

    K

    Kg

    KO

    k

    L

    AL

    AL,

    AP

    PWF

    PWP

    n

    n

    R

    I

    4

    RS

    Re

    rP

    SP

    S

    Sh

    t

    u

    0:

    &

    L-I

    mv

    d

    proportionality constant in eqn. (13)

    constant in eqn. ( 17)

    solution-dependent constant in eqn. (4), Pa

    constant in eqn. (17)

    constant in eqn. (17)

    constant in eqn. (17)

    bulk concentration, wt.

    gel concentration, wt.

    permeate concentration, wt.

    concentration at membrane wall, wt.

    diffusivity, m2 s-

    hydraulic diameter of flow channel, m

    diameter of solute molecules, m

    permeate flux, 1 m-* h-

    initial flux, 1 m-* h-

    specific permeability, mz

    permeability of a gel layer, m*

    specific permeability at zero pressure, m*

    mass transfer coefficient, m s-l

    length of flow channel, m

    thickness, m

    thickness of gel layer, m

    pressure difference across membrane, Pa

    pure water flux, 1 mm* h

    pure water permeability, 1 m-*

    hh

    Pa-

    solution-dependent constant in eqn. (4)

    exponent in eqn. ( 18)

    adsorbed layer hydraulic resistance, m ~

    boundary layer hydraulic resistance, m-

    gel layer hydraulic resistance, m -

    membrane hydraulic resistance, m ~

    solute hydraulic resistance, m-

    Reynolds number

    pore radius, m

    membrane surface porosity

    Schmidt number

    Sherwood number

    time, s

    cross-flow velocity, m s-

    compaction constant in eqn. (1 l), Paa

    porosity of gel

    viscosity, Pa s

    osmotic pressure difference across mem-

    brane, Pa

    osmotic pressure at membrane wall, Pa

    reflection coefficient

    79

    Refereoces

    1 I. Zahka and L. Mir, Ultratiltration of cathodic electrodepo-

    sition paints, Pluring Surj: Fiairhing, (Nov.) (1979) 34-39.

    2 B. R. Breslau, A. J. Testa, B. A. Milnes and G. Medjanis,

    Hollow fiber ultrafihration technology, ti. WurerEng., 17

    (1980) 20-26.

    3 I. H. Hanemaaijer, Toepassing van membranfdtratie in de

    zuivelindustrie,

    Voedingsmiddelen Technol., I 8 (4) (1985) 26-

    29.

    4

    V. Gekas, B. Hallstrom and G. Tr& rdh, Food and dairy

    applications: the state of the art, Desalinufion, 53 (1985)

    95-127.

    5 J.-L. Maubois, Recent developments of ultra8ltration in dairy

    industries, in E. Drioli and M. Nakagaki (eds.), Membranes

    and Membrane Processes,

    Plenum Press, New York, 1986, pp.

    255-262.

    6 R. M. Hedges and P. Pepper, Reverse osmosis and ultraliltra-

    tion-advances in the fruit and vegetable juice industries,

    Proc. Symp. on Fr uit jui ces for Eur ope, Den Haag, The

    Netherhzndv,

    Int. Fed. Fruit Juice Producers, 1986, pp. 253-

    260.

    7 I. R. Wahl, R. C. Hayes, M. H. Kleper and S. D. Pinto,

    Ultrafiltration for todays oily wastewaters: a survey of cur-

    rent ultrafiltration systems,

    Proc. Conf: on I ndustrial Wusfe,

    1980, pp. 719-729.

    8 A.-S. Jdnsson and R. Wimmerstedt, The application of mem-

    brane technology in the pulp and paper industry, Desalina-

    rion,

    53 (1985) 181-196.

    9 H. P. G. Darcy, Les

    Fontai nes Publi ques de la Vi lle de Duon,

    Victor Dalmont, Paris, 1856.

    10 0. Kedem and A. Katchalsky, Thermodynamic analysis of

    the permeability of biological membranes to non-electrolytes,

    J. Chromatogr., 27 (1958) 229-246.

    11 K. S. Spiegler and 0. Kedem, Thermodynamics of hyperf%l-

    tration (reverse osmosis): criteria for efficient membranes,

    Desalination, I (1966) 31 l-326.

    12 G. Jonsson, Boundary layer phenomena during ultrafiltration

    of dextran and whey protein solutions,

    Desalination, 51

    (1984) 61-77.

    13 J. G. Wijmans, S. Nakao and C. A. Smolders, Flux limitation

    in ultrafiltration: osmotic pressure model and gel layer model,

    J. Membr.

    Sci., 20

    (1984) 115-124.

    14. G. Jonsson, Characterization of uhrafiltration membranes:

    15

    16

    17

    18

    19

    20

    the osmotic pressure model, in G. Trlg&rdh (ed.), Proc. Conf.

    on Characterization of Ultrafi ltration Membranes, & en&

    Sforr, Sweden, Studentlitteratur, Lund, 1988, pp. 125- 140.

    W. F. Blatt, A. Dravid, A. S. Michaels and L. Nelsen, Solute

    polarization and cake formation in membrane ultrafiltration:

    causes, consequences, and control techniques, in J. E. Flinn

    (ed.), Membrane Science and Technol ogy, Plenum Press, New

    York, 1970, pp. 47-97.

    M. C. Porter, Concentration polarization with membrane ul-

    trafiltration,

    I nd. Eng. Chem., Pr od. Res. Dev., II (1972)

    234-248.

    A. G. Fane, C. J. D. Fell and A. G. Waters, The relation-

    ship between membrane surface pore characteristics and flux

    for ultrafiltration membranes, J. Membr. Sci., 9 (1981) 245-

    262.

    M. W. Chudacek and A. G. Fane, The dynamics of polarisa-

    tion in unstirred and stirred ultrafiltration, J. Membr. Sci., 22

    (1984) 145-160.

    P. De.jmek, Concentration polarization in ultrafiltration of

    macromolecules, Ph.D. Thesis,

    Dept.

    Food Engineering, Lund

    Univ., 1975.

    S. I. Nakao, T. Nomura and S. Kimura, Characteristics of

    macromolecular gel layer formed on ultrafiltration tubular

    membrane,

    AIChE J., 25 (1979) 615-622.

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    14/15

    80

    21 J. Cl. Wijmans, S. Nakao, J. W. A. van den Berg, F. R.

    Troelstra and C. A. Smolders, Hydrodynamic resistance of

    concentration polarization boundary layers in ultrafiltration, J.

    Me&r. Sci., 22 (1985) 117-135.

    22 E. Matthiasson, Macromolecular adsorption and fouling in

    ultrafiltration and their relationships to concentration polar-

    ization,

    Ph.D. Thesis,

    Dept. Food Engineering, Lund Univ.,

    1984.

    23 S. Sourirajan and T. Matsuura, Reverse osmosis/ultrafiltration

    process principles,

    NRCC Rep. No. 24188,

    Nat. Res. Council of

    Canada, 1985.

    24 R. B. Bird, W. E. Steward and E. N. Lightfoot,

    Transport

    Phenomena,

    Wiley, New York, 1960, Chap. 3.

    25 V. Gekas and B. Hallstrom, Mass transfer in the membrane

    concentration polarization layer under turbulent cross flow. I.

    Critical literature review and adaption of existing Sherwood

    correlations to membrane operations, J. Me&r. Sci., 3O( 1987)

    153-170.

    26 M. Lopez-Leiva, Ultrafiltration in rotary annular flow,

    Ph.D.

    Thesis,

    Dept. Food Engineering, Lund Univ., 1979.

    27 M. Lopez-Leiva, Ultrafiltration at low degrees of concentration

    polarization: technical possibilities,

    Desalin ation, 35 (1980)

    115-128.

    28

    F. Vigo, C. Uliana and P. Lupino, The performance of a

    rotating module in oily emulsions ultrafiltration,

    Sep. Sci.

    Tech&., 20 (1985) 213-230.

    29 I. G. R&z, J. G. Wassink and R. Klaassen, Mass transfer, fluid

    flow and membrane properties in flat and corrugated plate

    hyperfiltration modules,

    Desalin ation, 60 (1986) 213-222.

    30

    V. Gekas and K. Olund, Mass transfer in the membrane

    concentration polarization layer under turbulent cross flow. II.

    Application to the characterization of ultrafiltration mem-

    branes,

    J. Membr. Sci., 37( 1988) 145-163.

    31

    A. G. Fane, Ultrafiltration: factors influencing flux and rejec-

    tion, in R. J. Wakeman (ed.),

    Progress in Fi ltr ation and

    Separation,

    Elsevier, Amsterdam, 1986, pp. 101-179.

    32 A. G. Fane, C. J. D. Fell and M. T. Nor, Ultrafiltration in the

    presence of suspended matter,

    Inst.Chem. Eng. Symp. Ser. No.

    73, (1982) Cl -ClZ.

    33 R. L. Goldsmith, R. P. de Filippi and S. Hossain, New

    membrane process applications,

    AI ChE Symp. Ser. No. 120, 68

    (1972) 7-14.

    34

    H. Reihanian, C. R. Robertson and A. S. Michaels, Mecha-

    nisms of polarization and fouling of ultrafiltration membranes

    by proteins,

    J. Membr. Sci., 16 (1983) 237-258.

    35

    P. DeJardin, C. Toledo, E. Pefferkorn and R. Varoqui, Flow

    rates of solutions through ultrafiltration membranes monitored

    by the structure of adsorbed flexible polymers, in A. R. Cooper

    (ed.),

    Ultr ajil trotion Membranes and Appli cations,

    Plenum

    Press, New York, 1980, pp. 203-247.

    36 V. L. Vilker, C. K. Colton and K. A. Smith, Concentration

    polarization in protein ultrafiltration. Part 1: An optical shad-

    owgraph technique for measuring concentration profiles

    near a solution-membrane interface,

    AI ChE J., 27 (1981)

    632-645.

    37

    D. G. Thomas and W. R. Mixon, Effect of axial velocity and

    initial flux on flux decline of cellulose acetate membranes in

    hyperfiltration of primary sewage eflhients,

    Ind. Eng. Chem.,

    Process Des. Dev., I I (1972) 339-343.

    38

    A. S. Michaels, Ultrafiltration: an adolescent technology,

    CHEMTECH,

    (Jan.) (1981) 36-43.

    39 G. B. van den Berg and C. A. Smolders, Flux decline in

    membrane processes,

    Filtr. Sep., 25

    (1988) 115-121.

    40 E. Matthiasson and B. Sivik, Concentration polarization and

    fouling,

    Desalin ation, 35 (1980) 59- 103.

    41 G.

    Trlgardh (ed.),

    Proc. Corzf. on Characteri zation of Ultr o-

    fi ltr ation Membranes, & etis Slott, Sweden,

    Studentlitteratur,

    Lund, 1988.

    42 A. G. Fane, C. J. D. Fell and A. G. Waters, Ultrafiltration of

    protein solutions through partially permeable membranes-the

    effect of adsorption and solution environment,

    J.

    Membr.

    Sci.,

    16 (1983) 21 l-224.

    43 M. S. Lee, Membrane Ultrafiltration fouling and treatment,

    Ph.D. Thesis,

    Univ. College of Swansea, Wales, 1982.

    44 S. Munari, Characterization of ultrafiltration polymeric mem-

    branes, in G. Tragtirdh (ed.),

    Proc. Conf. on Characteri zation

    of Ultrafi ltration Membranes, & en& Slott, & e&n,

    Studentlit-

    teratur, Lund, 1988, pp. 81- 114.

    45 J. H. Hanemaaijer, T. Robbertsen, Th. van den Boomgaard, C.

    Olieman, P. Both and D. G. Schmidt, Characterization of clean

    and fouled ultrafiltration membranes,

    Desalin ation, 68 (1988)

    93-108.

    46

    L. A. Errede and P. D. Martinucci, Flow rate of water through

    porous membranes as affected by surface modification on the

    lower-pressure side of the membrane,

    Ind. Eng.

    Chem.,

    Prod.

    Res. Dew., 19 (1980) 573-580.

    47

    A.-S. Jonsson, The influence of surfactants on ultrafiltration

    membranes, in preparation.

    48 A. Suki, A. G. Fane and C. J. D. Fell, Flux decline in protein

    ultrafiltration,

    J. Membr. Sci., 21 (1984) 269-283.

    49

    A.-S. Jiinsson, Y. Blomgren and E. Petersson, Influence of pH

    and surfactants on ultrafiltration membranes during treatment

    of bleach plant effluent,

    Nor dic Pul p Paper Res. J., 3 (1988)

    159-165.

    50 N. M.

    Fish and

    M.

    D. Lilly, The interactions between fermen-

    tation and protein recovery,

    Biotechnology, 2 (1984) 623-627.

    51

    J. M. S. Cabral, B. Casale and C. L. Cooney, Effect of

    antifoam agents and efficiency of cleaning procedures on the

    cross-flow filtration of microbial suspensions,

    Biotechnol.

    L& t., 7 (1985) 749-752.

    52 K.

    H. Kroner, W. Hummel, J. Viilkel and M.-R. Kula, Effects

    of antifoams on cross-flow filtration of microbial suspensions,

    in E. Drioli and M. Nakagaki (eds.),

    Membranes and Mem-

    brane Processes,

    Plenum Press, New York, 1986, pp. 223-232.

    53 W. Han&h, Cell harvesting with membranes, in W. C.

    McGregor (ed.),

    Membrane Separations in Biotechnology,

    Marcel Dekker, New York, 1986, pp. 61-88.

    54 W. C. McGregor, J. F. Weaver and S. P. Tansey, Antifoam

    effects

    on ultrafiltration,

    Biotechnol. Bioeng., 31 (1988) 385-

    389.

    55

    A. G. Fane, C. J. D. Fell and K. J. Kim, The effect

    of surfactant pretreatment on the ultrafiltration of proteins,

    Desalin ation, 53 (1985) 37-55.

    56

    A. G. Fane, C. J. D. Fell and A. Suki, The effect of pH and

    ionic environment on the ultrafiltration of protein solutions

    with retentive membranes,

    J.

    Membr.

    Sci.,

    16( 1983) 195-210.

    57 J. F. Hayes, J. A. Dunkerley, L. L. Muller and A. T. Griffin,

    Studies on whey processing by ultrafiltration. II. Improving

    permeation rates by preventing fouling,

    Aust. J. Dair y Tech-

    nol.,

    (Sept.) (1974) 132-140.

    58 F. Forbes, Considerations in the optimisation of ultrafiltra-

    tion,

    Chem. Eng. (Lundon),

    (Jan.) (1972) 29-34.

    59 H. B. Hopfenberg, V. T. Stannet and M. W. Bailey, Solute-

    solute interactions in ultrafiltration treatment of paper mill

    wastes,

    AI ChE Symp. Ser. No. 139, 70

    (1974) 1- 10.

    60 D. Woemer and J. L. McCarthy, Ultrafiltration of pulp mill

    liquors,

    Tappi, 70 (1987) 126-129.

    61

    P. D. Bisio, J. G. Cartledge, W. H. Keesom and C. J. Radke,

    Molecular orientation of aqueous surfactants on a hydropho-

    bic solid,

    J. Colloi d I nterface Sci., 78 (1980) 225-234.

    62

    J. Nassauer, Adsorption und haftung an oberIHchen und

    membranen,

    Ph.D. Thesis,

    Inst. fiir Milchwissenschaft und

    Lebensmittelverfahrenstechnik, Tech. Univ. Munich, 1985.

    63 0. Omosaiye, M. Cheryan and M. E. Matthews, Removal of

    oligosaccharides from soybean water extracts by ultrafiltra-

    tion,

    J. Food Sci., 43 (1978) 354-360.

  • 7/21/2019 Fundamental Principles of Ultrafiltration.pdf

    15/15

    81

    64 L. L. Muller, J. F. Hayes and A. T. Griffin, Studies on whey

    processing by ultrafiltration. 1. Comparative performance of

    various ultrafiltration modules on whey from hydrochloric acid

    casein and cheddar cheese,

    Aust. .l. Dair y Technol.,

    (June)

    (1973) 70-77.

    65 U. H. Haagensen, Case Sanyo Pulp, Iwakuni, Japan, 1924GE-

    0482-50, De Danske Sukkerfabrikker, 1982.

    66 Pulp deresination by ultrafiltration at MoDoCell,

    Leaflet,

    F CI

    Membrane Systems, Laverstoke Mill, Whitchurch,

    U.K., 1987.

    67 S. Bragula and K. Litner, Cleaning and disinfection of mem-

    brane plants: theory and practice,

    Abstr. I nt. Congr. und Messe

    Nahr ungsmitteltechnik, K& z, 1986.

    68 G.

    Tragigirdh, in preparation.

    69 G. Tr& rdh, Membrane cleaning,

    Desalin ation, 71 (1989)

    325-335.

    70

    Manual cleaning and disinfection, 2181-CB-088523, De

    Danske Sukkerfabrikker, 1985.