Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Properties of Filter Cake in Cake Filtration and Membrane Filtration†

Eiji Iritani
Department of Chemical Engineering,
Nagoya University*

Abstract

Filtration of liquid suspensions is a widely practiced process in many industries. In recent years,
membrane filtration of colloidal solutions has attracted a considerable amount of attention. The
appropriate control of both the filtration rate and the rejection (or the transmission) of the particles
and/or the macromolecules in the filtration process is of great interest in both industry and acade-
mia. Such filtration behaviors are strongly affected by the properties of the filter cake formed by the
accumulation of the particles and/or the macromolecules on the surface of the filter medium or the
membrane. This paper overviewed the author’s own contributions on the recent developments on the
behaviors of the filter cake in cake filtration and membrane filtration. The paper will mainly deal
with measurements and analysis of the internal structures of the filter cake, the role of the solution
environment in the properties of the filter cake, and filtration and fractionation mechanism of mix-
tures.

The appropriate control of both the filtration rate


1. Introduction
and the rejection (or the transmission) of the par-
Filtration is the operation of separating a dispersed ticles and/or the macromolecules in the filtration
phase of solid particles from a f luid by means of a fil- process is of great interest in both industry and acad-
ter medium which permits the passage of the f luid emia. While there exist many factors inf luencing the
but retains the particles. Filtration is probably one of filtration behaviors [1], the underlying phenomena
the oldest unit operation. The old forms of filtration are currently not well understood.
by straining through porous materials were described The filtration behaviors are strongly affected by the
by the earliest Chinese writers. A gravity filter used in properties of the filter cake formed by the accumula-
a chemical process industry was described in an tion of the particles and/or the macromolecules on
Egyptian papyrus which has its origin in about the the surface of the filter medium or the membrane [2].
third century A. D. Therefore, an understanding of the cake structure
In recent years, many developments have increased can serve as a basis for clarifying the real mechanism
the application of filtration. Filtration steps are re- of cake filtration and membrane filtration. In this arti-
quired in many important processes and in widely cle, role of filter cake in cake filtration and membrane
divergent industries. The importance of filtration filtration is overviewed.
techniques has been emphasized by the increased
need for protection of the environment and by the
2. Experimental Testing Procedures
increasingly critical need for larger supplies of en-
ergy. Recently, membrane filtration of colloids has It is extremely important from both theoretical and
become increasingly important in widely diversified industrial viewpoints to develop simple and precise fil-
fields. tration test methods. The most commonly used tests
are classified into three groups: vacuum filtration
tests, pressure filtration tests, and compression-per-
* Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
meability tests. Such overall filtration characteristics
Phone: +81-52-789-3374, Fax: +81-52-789-5300
E-mail: iritani@nuce.nagoya-u.ac.jp as the average specific filtration resistance α av and

Accepted: September 1, 2003 the average ratio m of wet to dry cake mass are mea-

KONA No.21 (2003) 19


sured by conducting vacuum or pressure filtration AIR PRESSURE INLET
experiments. In general, α av is calculated by using SLURRY INLET

both the constant pressure filtration coefficient Kv


determined from the slope of the Ruth plots [3] and
the m-value. The value of m has almost invariably SLURRY
been determined by weighing the filter cake before
CAKE SURFACE
and after the cake dries. However, visual determina- DISK WITH HOLE
tion of the end of filtration often leads to erroneous
values for m and also α av because of the indistinct FILTER CAKE
Dh
interface between the cake surface and the slurry.
The compression-permeability cell [4-6] is a device
in which a mechanical load is applied through a pis- D

h
ton to a cake resting on a filter medium. Both overall
filtration characteristics and internal structures of the
INSERTED
filter cake are analyzed on the basis of experimental CYLINDER
data of the equilibrium porosity and the specific f low FILTER PAPER &
resistance of the compressed cake in the compres- FILTRATE PERFORATED PLATE
sion-permeability cell. However, this technique may
be rather tedious and time-consuming for industrial Fig. 1 Schematic diagram of filtration apparatus having sudden
practice. reduction in its filtration area.
A simple alternative procedure recently developed
depends on measurement of the capillary suction
time (CST), utilizing the very small suction pressure
applied to the slurry by the capillary action of an
absorbent filter paper [7]. However, little attempt has trate volume collected per unit effective medium area,
been made to determine filtration characteristics pre- and Kv is the Ruth coefficient of constant pressure fil-
cisely and easily under relatively high filtration pres- tration defined by
sure conditions.
2p(1ms)
A method has been developed for evaluating rigor- Kv (2)
mrsaav
ously the properties of the filter cake, such as the
average porosity εav and the average specific filtration where p is the applied filtration pressure, s is the
resistance α av [8]. It utilizes the sudden reduction in mass fraction of solids in the slurry, µ is the viscosity
filtration area of the cake surface. The specially de- of the filtrate, and ρ is the density of the filtrate.
signed apparatus is schematically shown in Fig. 1. Once the filter cake builds up to the underside of
A close-fitting cylinder with an inner diameter D of 4 the disk, the subsequent filter cake can form only
cm is inserted in the cylindrical brass filter. Inserted inside the hole in the disk. Consequently, the filtra-
cylinders having heights h of 5, 10 and 15 mm are tion area of the cake surface is reduced suddenly, and
used. A disk with a hole having a diameter Dh of 6 the filtration rate decreases markedly in accord with
mm is placed on top of the inserted cylinder, and the the decrease in formation rate of the filter cake. After
part below this constitutes the filter chamber. the filtrate volume v is beyond the critical volume vt at
Filter cake steadily builds up on the filter medium the transition point, the reciprocal filtration rate
as soon as the filtration process starts. The surface (dθ/dv) vs. v deviates remarkably from the relation
area of the growing filter cake equals exactly the area represented by Eq. (1).
of the filter medium. At this first stage, on the as- From the value of vt, the average porosity εav of the
sumption of negligible medium resistance, the recip- filter cake can be calculated using an overall mass bal-
rocal filtration rate (dθ/dv) is represented by the ance of dead-end filtration, to give
well-known Ruth equation for constant pressure filtra-
rs h(1s)rsvt
tion in the form [3] εav (3)
rs h(1s)rsh
dq 2
 v (1) where ρs is the true density of solids. The ratio m of
dv Kv
wet to dry cake mass in Eq. (2) is related to the aver-
where θ is the filtration time, v is the cumulative fil- age porosity εav of the filter cake by

20 KONA No.21 (2003)


103
reav 6
m1 (4)
rs (1eav) KOREAN KAOLIN
s0.393
Thus, the average specific filtration resistance α av can 5
h1 cm
p98 kPa
be evaluated accurately from Eqs. (1), (2), and (4) by p196
using the slope of the plot of dθ/dv against v and the p294
p392
value of m. 4
The values of m and α av obtained by this method

dθ/dv [s/cm]
can be predicted by using the compression-perme- TRANSITION POINT

ability cell data. On the basis of the compressible 3


cake filtration model [9, 10], the distributions of the
local porosity ε and the local specific filtration resis-
2
tance a and the apparent liquid velocity u relative to
solids in the filter cake can be estimated by using the
compression-permeability cell data in the form of e
1
and a as functions of the local solid compressive pres-
sure ps. Consequently, the values of m and a av can be
calculated from Eqs. (5) and (6), respectively. 0
0 0.5 1.0 1.5

  
w 1 v [cm]
r ed
0 N0
w
m1 (5)
 (1e)d wN 
1 w
rs Fig. 2 Relation between reciprocal filtration rate and filtrate vol-
0 0 ume per unit medium area for various applied filtration
pressures.

 u  d  w ·  pp1 d p
1 u w
a av p pm
m
(6)
0 1 0
s
0 N
a
1.7
where w is the net solid volume per unit medium area
lying from the medium up to an arbitrary position in KOREAN KAOLIN
s0.393
the cake, w 0 is the net solid volume of the entire cake
per unit medium area, and u1 is the filtration velocity. 1.6
m []

The quantity pm is the pressure loss through the filter


medium, and can be neglected in this article.
In Fig. 2, the results obtained with the setup shown 1.5
in Fig. 1 are plotted in the form of the reciprocal flow THIS METHOD
CONVENTIONAL METHOD
rate (dq/dv) vs. the filtrate volume v per unit medium C-P CELL DATA
area. In the first stage of the operation, each curve 1.4
yields a straight line in accordance with Eq. (1). As 0 100 200 300 400
p [kPa]
soon as the cake builds up to the underside of the
disk with the hole, the value of dq/dv increases re-
markably. Fig. 3 Relation between ratio of wet to dry cake mass and filtra-
tion pressure.
The value of m can be calculated from Eqs. (3) and
(4) by using the values of the thickness h of the filter
chamber and the filtrate volume vt at the transition
point determined from Fig. 2. In Fig. 3, m is plotted
with respect to the filtration pressure p. The value of parison, the values of m determined by the conven-
m decreases with pressure. The experimental results tional method of weighing the filter cake before and
are fairly consistent with the calculations based on the after being dried are also included in the same figure.
compression-permeability cell data. The discrepancy The values thus obtained are rather large compared
in the low-pressure region may be due to the inf lu- with the calculations because of the indistinct inter-
ence of sedimentation, which becomes important in face between the cake surface and the slurry. In
the case of small formation rate of the cake. For com- Fig. 4, a av is plotted against pressure p. The resis-

KONA No.21 (2003) 21


1011 described by [11]
10
1 ∂ps
KOREAN KAOLIN Jw · (8)
8 s0.393 mars ∂w

On the assumption that the solvent velocity Jω is con-


αav [cm/g]

6
stant throughout the entire cake, on the basis of Eqs.
(7) and (8), one obtains [11]
4

N
dp
ps
s
THIS METHOD w 0a
2 CONVENTIONAL METHOD 1 (9)
N
dp
w0 p
C-P CELL DATA s
a
0
0
0 100 200 300 400
p [kPa] a av
a (10)
d( ln α av)
1
Nd(lnNp)
Fig. 4 Relation between average specific filtration resistance and
applied filtration pressure.

 
d( ln eav)
Nd(lnN p)
eeav 1 (11)
d( ln a av)
1
tance a av increases with pressure. The experimental Nd(lnN p)
results obtained by this method are fairly consistent
with the calculations, whereas the results obtained where w 0 (w0/rs) is the net solid volume of the
by the conventional method show poor agreement entire filter cake per unit effective medium area, and
because of incorrect values of m. It is demonstrated e (e/(1e)) is the local void ratio. Therefore, if the
from the results of Figs. 3 and 4 that this method is average specific filtration resistance a av and the aver-
more accurate compared with the conventional one. age void ratio eav (eav/(1eav)) of the filter cake are
represented as functions of the applied filtration pres-
sure p, then the variations of the local solid compres-
3. Compressible Cake Filtration Model
sive pressure ps, the local specific filtration resistance
The compressible cake filtration model [9, 10] is a and the local void ratio e with w can be evaluated
used to evaluate such internal quantities as the solid from Eqs. (9)(11). Thus, the variation of the mass
concentration, the compressive pressure acting on fraction c of the solid in the filter cake with the dis-
the solids and the local specific filtration resistance tance x from the medium surface can be also calcu-
within the filter cake exhibiting compressible behav- lated from the result of e versus w.
ior. The build-up of the filter cake increases the For a practical standpoint, the following empirical
hydraulic resistance to f low, thereby reducing the fil- functions of ps may be convenient for simplified evalu-
tration rate. Basically, the compressible cake filtration ations of filtration characteristic values [12, 13].
model can evaluate the reduction of the filtration rate
aa1 psn (12)
due to the increase in the hydraulic resistance caused
by the filter cake.
eE0Cc ln ps (13)
According to the compressible cake filtration mod-
el, the filtration rate J (dv/dq) in dead-end filtration where a1, n, E0, and Cc are the empirical constants.
with negligible medium resistance compared with the With the aid of Eqs. (12) and (13), Eqs. (9)(11) can
resistance of the filter cake is represented by [3] be written as [11]
1

J
p
ma avw
(7)
ps
p 
 1
w
w0  1n
(14)

where w0 is the net mass of deposited solids per unit a ava1(1n)pn (15)
effective medium area. By accounting for the effects
Cc
of non-homogeneity and compressibility of the filter eavE0 Cc ln p (16)
1n
cake, the apparent solvent velocity Jω relative to solids
at an arbitrary position w in the filter cake can be On the basis of the model of solutions in which the

22 KONA No.21 (2003)


macromolecules appear as solid particles [14], it is volume v calculated from Eqs. (14) and (12) are illus-
assumed that the model which has been employed in trated in Figs. 6 and 7, respectively [16]. The dis-
filtration of particulate suspensions can be applied to tance x from the membrane surface in the figures
the f low of solvent through the filter cake in ultra- may be calculated from w by
filtration of the solution. In Fig. 5, the average spe-
x 1eav,w w
cific filtration resistance a av and the average void ratio  · (17)
L 1eav w 0
eav of the filter cake in protein ultrafiltration measured
with a filter which has a sudden reduction in its filtra-
tion area are plotted as functions of the filtration pres-
sure p [15]. The macrosolute used in the experiments
100
was bovine serum albumin (BSA) with a molecular
weight of 67,000 Da and with an isoelectric point of BSA
s5103
4.9. Increasing the filtration pressure causes a reduc- 80 pH 7.0
tion in the void ratio of the filter cake, leading to an p98 kPa
increase of the specific filtration resistance. It is
apparent that eav around the isoelectric point (pH 4.9) 60

ps [kPa]
is much smaller than that at pH 7.0. The two types of
plots show a linear relationship over the entire range
v5 cm
of data in accordance with Eqs. (15) and (16), respec- 40
v10 cm
tively. The parameters a1 and n in Eq. (12) can be cal-
v20 cm
culated by fitting Eq. (15) to the logarithmic plot of
20
a av vs. p. The value of E0 and Cc in Eq. (13) can be
obtained from a plot of eav vs. ln p in accordance with
Eq. (16), using the predetermined n value. The distri- 0
butions of ps within the filter cake can be obtained 0 0.5 1.0 1.5
from Eq. (14). With the aid of Eq. (14), Eqs. (12) and x [cm]

(13) provide the distributions of a and e within the fil-


ter cake, respectively. Fig. 6 Distributions of local solute compressive pressure in filter
cake for different filtrate volumes.
The variations of the local solute compressive pres-
sure ps and the local specific filtration resistance a
across the filter cake for different values of the filtrate
1015
10

BSA
4 20
s5103
8 pH 7.0
p98 kPa
2
BSA
15
s5103 6
, pH 4.9
α [m/kg]

1016 , pH 7.0
αav [m/kg]

8
eav []

6 10 v5 cm
4
4 v10 cm
v20 cm
5 2
2

1015 0 0
101 2 4 6 8 102 2 4 6 8 103 0 0.5 1.0 1.5
p [kPa] x [cm]

Fig. 5 Effect of applied filtration pressure on average specific fil- Fig. 7 Distributions of local specific filtration resistance in filter
tration resistance and average void ratio of filter cake for cake for different filtrate volumes.
different values of pH.

KONA No.21 (2003) 23


where eav,ω is the average void ratio from the mem- The rotor speed ranged from 45,000 to 60,000 min1.
brane surface to the distance w from the membrane. The distributions of the concentration, the concentra-
The thickness L of the protein filter cake is expressed tion gradient, and the refractive index gradient of the
based on the material balance as solutions in a 1.5 mm double-sector centerpiece were
measured over time using both the ultraviolet scan-
rs(1eav)
L v (18) ner absorption system and Schlieren optics. In the
rs(1s)rseav
sedimentation velocity experiment, the sedimentation
As shown in the figure, high specific filtration resis- velocity was determined from the displacement of the
tance can be obtained although the porosity in the sedimentation boundary. In the high-speed sedimen-
cake is not so small [2, 17]. The figure obviously tation equilibrium experiment, the equilibrium thick-
demonstrates that both ps and a decrease with the ness of the sediment of macromolecular solutes was
distance x. As the filtrate volume v increases, ps and a measured after equilibrium was reached at a constant
increase. The solute compressive pressure at the sur- rotor speed.
face of the protein filter cake is essentially zero as no At relatively high solution concentrations, there is
drag on the solutes has developed. The frictional drag an analogy between the sedimentation of a macromol-
on each solute adds to the drag on the previous ecule in a solvent and the permeation of a solvent
solutes as the solvent passes frictionally along the through the filter cake of macromolecules. The local
solutes in the compressible filter cake. Consequently, specific flow resistance a can be calculated as [15, 19]
the net compressive pressure increases as the mem-
(r sr)r i Ω2 r sr
brane surface is approached, resulting in the increas- a  (20)
mrsv0 mrs S
ing resistance to f low. At the interface of the filter
cake and the membrane, ps has risen to its maximum where ri is the radial distance of the sedimentation
value and is equal to the applied filtration pressure p boundary from the center of rotation, W is the angu-
for the negligible membrane resistance. Therefore, lar velocity of the rotor, v0 is the sedimentation veloc-
the protein filter cake tends to have a much more ity, and S is the sedimentation coefficient.
compact structure at the membrane exhibiting a large Thus, the relation between a and the volume frac-
resistance to f low in comparison to a relatively loose tion (1e) of the solute may be written as [15, 19, 21]
condition at the surface because the cake is com-
 
1/4.65
rs df 2
pressible. In the compressible filter cake formed by (1/a)1/4.65 {1C h(1e)} (21)
18C h
ultrafiltration of BSA solutions, it is demonstrated that
a pronounced variation of characteristic values of the where df is the average equivalent spherical diameter
filter cake can be seen. This is in agreement with the of solutes, Ch is the ratio of the volume of the hydrous
results obtained for filtration of particulate suspen- protein molecule to volume of anhydrous protein mol-
sions [9, 18]. ecule, and e is the porosity of the solution.
The local mass fraction c of the solute in the filter Figure 8 shows the permeability data obtained by
cake may be calculated from the local void ratio e in measuring the sedimentation velocities for a number
the form of different solution concentrations. The plots are vir-
tually linear as would be expected from Eq. (21). It is
rs
c (19) apparent that a around the isoelectric point is much
rsre
smaller than that at pH 7 with the same solution con-
Thus, the concentration distributions in the protein centration. On the other hand, the average specific fil-
filter cake may be predicted from Eqs. (13), (14), tration resistance a av of the cake in the ultrafiltration
(17)(19). of BSA solutions reaches a definite maximum around
the isoelectric point [2]. In order to account for this
discrepancy, a high-speed sedimentation equilibrium
4. Analysis of Ultrafiltration Behaviors Based
experiment was conducted.
on Ultracentrifugation Experiment
If the local solute concentration of the cake formed
The f lux decline and the cake structure in ultra- in ultrafiltration is known, then the specific f low resis-
filtration of protein solutions can be evaluated from tance a of the cake can be evaluated by using the per-
analytical ultracentrifugation experiments [15, 19, 20]. meability data. Therefore, it is necessary to evaluate
The experiments were performed in a Hitachi Model the local solute concentration of the cake. We thus
282 ultracentrifuge equipped with an optical system. determined the compression data representing the

24 KONA No.21 (2003)


104 103
12 4.0

11 3.8
BSA
pH 5.1
Ω6283 rad/s 3.6
10
(1/α)1/4.65 [(g/cm)1/4.65]

pH 7.0
Ω5760 rad/s
3.4
9

3.2

Rrs [m]
8

3.0
7

6 2.8
BSA
pH 5.1, s0.1
5
0 0.1 0.2 0.3 2.6 ω00.859 mm
pH 7.0, s0.097
1ε []
ω00.829 mm

2.4
Fig. 8 Relation between local specific f low resistance and poros- 106 2 3 4
ity of solution. RΩ2 [m·rad2/s2]

Fig. 9 Relation between equilibrium thickness of sediment and


centrifugal acceleration.

relation between the local porosity e and the local


solute compressive pressure ps from high-speed sedi-
mentation equilibrium experiments. If e is related to
ps by Eq. (22), the equilibrium thickness (Rrs) of the figure using Eqs. (22) and (23). On the basis of these
sediment is related to the centrifugal acceleration relations, the variations of the filtration rate in ultra-
RW 2 by Eq. (23) [15, 19, 22]. filtration of protein solutions can be determined from
the compressible cake filtration model [2, 9, 15].
1eEpsb ps psi (22)
In Fig. 10, the results of unstirred dead-end ultra-
filtration are plotted in the form of the reciprocal fil-
w 01b b
Rrs {(rsr)RΩ2} (23) tration rate (dq/dv) versus the filtrate volume v per
E(1b)
unit membrane area. The filtration rate around the
where R is the distance from the center of rotation to isoelectric point is much smaller than that at pH 7.
the bottom of the sediment, rs is the distance from the The solid lines indicate the theoretical predictions
center of rotation to the surface of the sediment, w 0 is based on the compression-permeability data obtained
the net solute volume of the entire sediment per unit in analytical ultracentrifugation. The experimental
cross-sectional area, and E and b are the empirical data are in relatively good agreement with the theory,
constants. Below psi, e is assumed to be constant, and indicating that the compressible cake filtration model
is equal to the porosity at pressure psi. accurately describes the ultrafiltration behavior.
Figure 9 represents the logarithmic plot of the Figure 11 shows the variations of the mass fraction
equilibrium thickness (Rrs) of the sediment against c of the solute across the cake calculated on the basis
the centrifugal acceleration RW 2. The plot shows a of the ultracentrifugation data. In the compressible
linear relationship in accordance with Eq. (23). It is cake resistance model, the solutes deposited on the
interesting to note that the sediment at the isoelectric membrane are treated as the cake, and it is attributed
point is much more compact than that at pH 7 to the hydraulic barrier. The cake tends to have a
because the BSA molecule carries no net charge at much more compact structure at the membrane in
the isoelectric point. Therefore, it would be expected comparison to a relatively loose condition at the sur-
that in protein ultrafiltration a compact filter cake face because the cake is compressible. Of consid-
forms around the isoelectric point. The relation be- erable practical interest is that this result is in
tween e and ps can be determined from the plot in the agreement with that obtained for cake filtration of par-

KONA No.21 (2003) 25


104
5 5. Measurement of Concentration Distribution
BSA
in Filter Cake of Ultrafiltration
s6103
p98 kPa A variety of theoretical models describing the foul-
4 pH 5.1 ing phenomenon during ultrafiltration quantitatively
pH 7.0
PREDICTED have been developed: the gel polarization model [24-
26]; the osmotic pressure model [27-29]; the bound-
3 ary layer model [30, 31]; and the cake filtration model
dθ/dv [s/cm]

[32-34]. Although a significant amount of research


has been conducted to better understand the dynam-
ics of ultrafiltration, the real mechanism of separation
2
remains incompletely understood.
The compressible cake filtration model has been
widely used to describe filtration behaviors of particu-
1 late suspensions [9]. Recently, some authors prefer to
use the compressible cake filtration model to explain
the mechanism of ultrafiltration [2, 15, 35], and it has
0 the potential for analyzing the membrane fouling dur-
0 0.5 1.0 1.5
ing ultrafiltration. The filtration f lux rate during dead-
v [cm]
end ultrafiltration of bovine serum albumin (BSA)
solutions could be well evaluated by use of the ultra-
Fig. 10 Evaluation of filtration rate based on ultracentrifugation centrifugation data on the basis of the compressible
data.
cake filtration model [19]. The model regards a stag-
nant filter cake on the membrane as the structural
assemblage of “particulate” solutes. In this model,
highly resistant filter cake of excluded protein mole-
0.4
cules accumulated in front of the membrane surface
BSA forms, and the filter cake acts as an additional hy-
s6103
p98 kPa
draulic resistance to f low in conjunction with that
0.3 θ2 h
pH 5.1 provided by the membrane, thereby reducing the fil-
θ6 pH 7.0
tration f lux rate. The model also takes the compress-
θ12 ibility of the filter cake into account. Therefore, an
c []

0.2
CAKE SURFACE, θ2 h understanding of the properties of the filter cake
formed by the accumulation of the protein molecules
θ6
0.1 on the membrane can serve as a basis for clarifying
θ12 the real mechanism of ultrafiltration.
Several approaches have been developed for mea-
0 suring the average value of the porosity in the filter
0 0.5 1.0 1.5
cake formed on the membrane surface in ultrafiltra-
x [mm]
tion of protein solutions. Nakao et al. [36] measured
the gel concentration by scraping a thin layer of
Fig. 11 Evaluation of distributions of solute concentration in filter deposited material from the surface of a tubular mem-
cake based on ultracentrifugation data.
brane in ovalbumin ultrafiltration. Iritani et al. [2]
used a batchwise filter device which had a sudden
reduction in its filtration area to determine the aver-
age porosity of the BSA cake. Also, in more recent
ticulate suspensions [23]. A much more compact fil- work by Nakakura et al. [35], the porosity was ob-
ter cake exhibiting a large resistance to f low may tained from the measurement of the electric conduc-
form around the isoelectric point than that which tivity within the filter cake during ultrafiltration of
forms at pH 7. Consequently, the filtration rate at the BSA solutions. It is, however, crucial to measure the
isoelectric point becomes lower than that at pH 7. porosity distribution since the filter cake formed on
the membrane during ultrafiltration is extremely thin

26 KONA No.21 (2003)


in general. The f lowing behaviors of filter cake have been
Relatively few investigators have measured concen- observed, as depicted in Fig. 12 schematically. It was
tration distributions within the protein layer accumu- clarified that the filter cake formed mainly on the
lated on the membrane during ultrafiltration. Vilker et lower part of the membrane surface due to the effect
al. [29] used an optical shadowgraph technique to of gravity during inclined ultrafiltration, as shown in
measure the concentration gradient within the polar- Fig. 12(a) [40]. The same phenomenon cannot be
ized layer on the membrane during unstirred dead- observed in downward dead-end ultrafiltration. Once
end ultrafiltration of BSA solutions. McDonogh et al. downward ultrafiltration was conducted after inclined
[37] reported the variation with time of concentration ultrafiltration, the filter cake was spread over the
in the polarized layer on the membrane during ul- membrane surface, as shown in Fig. 12(b) [41].
trafiltration of BSA and Dextran Blue solutions using In Fig. 13, the f lux decline behaviors in various fil-
the electronic diode array microscope. Gowman and tration modes are plotted in the form of the reciprocal
Ethier [38] used an automated, laser-based refrac- filtration rate 1/J (dq/dv) versus the cumulative fil-
tometric experimental technique for the measure- trate volume v collected per unit effective membrane
ment of concentration and concentration gradient area, which is well known as the Ruth plot [3] in the
in the concentration polarization layer during dead- ordinary cake filtration. For conventional downward
end ultrafiltration of the biopolymer hyaluronan. dead-end ultrafiltration (y0), the filtration rate de-
Fernández-Torres et al. [39] obtained the evolution of clines gradually due to the build-up of a BSA cake
concentration profiles during BSA ultrafiltration in an as filtration proceeds. The angle y represents the
unstirred cell using the technique of holographic angle between the filtrate f low and the direction of
interferometry. Their results demonstrated that a fil- gravity. The plot is virtually linear throughout the
ter cake was actually formed during BSA ultrafiltra- course of filtration in accordance with the Ruth filtra-
tion. Although these non-destructive methods have tion rate equation (1) [3] with negligible membrane
the enormous advantage, they are technically com- resistance compared with the filter cake resistance.
plex and expensive. In contrast, in inclined ultrafiltration (yp/3 rad), the
A potential method has been explored for measur- filtration rate becomes remarkably high because the
ing the concentration distributions in the filter cake filtrate passes through the upper part of the mem-
using the principle of the inclined ultrafiltration [40, brane which is less resistant. Furthermore, once
41], where the membrane was inclined and a large downward ultrafiltration was conducted after inclined
amount of filter cake was formed. The concentration ultrafiltration, the filtration rate dramatically
distributions in the filter cake were measured after decreased because of the remarkable increase of the
downward ultrafiltration was performed in the wake total filtration resistance. Surprisingly, the value of 1/J
of inclined ultrafiltration. This destructive method is in this case lies on the extension line in ordinary
simple, and inexpensive. downward ultrafiltration, as shown in the figure. This

(a) (b)

SOLUTION
SOLUTION

MEMBRANE

FILTER CAKE
FILTRATE FILTER CAKE
MEMBRANE
FILTRATE

Fig. 12 Schematic diagram of flowing filter cake: (a) inclined ultrafiltration, and (b) downward ultrafiltration.

KONA No.21 (2003) 27


105 brane surface for the different values of the filtrate
8
volume v collected until inclined ultrafiltration was
BSA ψ0 switched to the downward mode. The profiles in this
s5103 ψπ/3 rad → ψ0
pH 7.0 v10 cm
figure may be considered to be similar to those in
6 p98 kPa ψπ/3 rad → ψ0 ordinary downward ultrafiltration, as expected from
v20 cm
the results of f lux decline behaviors shown in
1/J [s/cm]

Fig. 13. The filter cake which was thicker than 1 cm


was formed at the filtrate volume of 20 cm. It is very
4
difficult to obtain such a thick cake as this from down-
ward dead-end ultrafiltration. The solid lines on the
figure indicate the theoretical predictions based on a
2 compressible cake filtration model. In the compress-
ible cake filtration model, the concentration distri-
butions of the solutes deposited on the membrane
treated as the filter cake act as the hydraulic barrier.
0
0 5 10 15 20 25 The pressure gradient on the liquid passing through
v [cm] the solute varies, and the porosity in the cake is not
constant [42]. The figure distinctly demonstrates that
Fig. 13 Relation between reciprocal filtration rate and filtrate vol- there exists a concentration distribution within the fil-
ume per unit membrane area in various filtration modes. ter cake formed on the membrane. The concentration
c has a maximum at the membrane surface (x0) and
decreases markedly with the increase of x, which
shows the same trend as that observed for the cake
filtration of particulate suspensions [9, 18]. It was ver-
result supports that the structure of the filter cake ified that the solutes forming the cake are compact
formed on the membrane such as the concentration and dry at the membrane whereas the surface layer is
distributions is almost the same between both opera- in a wet and soupy condition. It was also found that
tions. the thickness of the filter cake formed on the mem-
In Fig. 14, the mass fraction c of BSA on the mem- brane increases as filtration proceeds.
brane measured after changing the angle y from p/3
rad to zero is plotted against distance x from the mem-
6. Role of Solution Environment in Properties
of Filter Cake
0.5 Cake filtration of suspensions of solid particles in
BSA aqueous solutions is frequently encountered in widely
s5103 diversified fields. In cake filtration, the nature of the
0.4 pH 7.0
p98 kPa cake formed upon the medium surface during filtra-
ψπ/3 rad → ψ0 tion will govern the filtration rate of the remaining
v10 cm
0.3 v20 cm suspension. For fine particle suspensions, colloidal
PREDICTED forces control the nature of the filter cake. Colloidal
c []

forces arise from interaction between the suspended


0.2 particles. The two main colloidal forces are the attrac-
tive van der Waals forces which originate from f luc-
tuating dipoles as a result of the motions of outer
0.1
s5103 electrons on the interacting particles and the repul-
sive electrostatic forces due to the presence of like
charges on the particles and a dielectric medium.
0
0 0.5 1.0 1.5 Whilst for a given system the van der Waals forces
x [cm] are essentially constant, the electrostatic forces will
vary with the surface charge of the suspended parti-
Fig. 14 Distribution of solute concentration in filter cake for dif- cles, which varies with the solution environment.
ferent filtrate volumes. Therefore, the filtration behaviors of the colloids are

28 KONA No.21 (2003)


affected significantly by the solution properties, in- the charge. Since the most loose filter cake forms
cluding pH and electrolyte strength. Considerable around the isoelectric pH, the filter cake is most per-
work has been published on the effects of the solution meable. Thus a av is much smaller near the isoelectric
environment on the filtration behaviors [43-52]. point.
Figure 15 shows the average specific filtration re- It is interesting to note that the results in protein
sistance a av and the average porosity eav of the filter ultrafiltration had a distinctly different behavior. In
cake formed in particulate microfiltration at different protein ultrafiltration of BSA solution, the filter cake
values of pH under otherwise identical filtration con- is in its most compact state around the isoelectric
ditions [53]. The particles used in the experiments point [2, 54], as shown in Fig. 16. Since the BSA mol-
was titanium dioxide of the rutile form with an origi- ecules are hydrophilic colloids, their stability in the
nal mean specific surface area size of 0.47 µm and solution would appear to be influenced not only by
with an isoelectric point of 8.1. It should be noted that the presence of a surface charge on the protein but
the resistance a av is a measure of the filtrability of the also by hydration of the surface layers of the protein.
suspensions. A larger a av causes a smaller filtration The BSA molecules, because of hydrated layers sur-
rate. It can be seen that a av goes through a minimum rounding them, are not destabilized by such consider-
around the isoelectric point. Thus, the largest filtra- ations as depression of the electrical double layer.
tion rate can be obtained when the particle carries no Thus, the BSA molecules have water bound to them
charge. Tarleton and Wakeman [49] reported similar even around the isoelectric point. The hydrophilic
results for crossf low microfiltration of anatase sus- BSA molecules maintain a dispersed state in the solu-
pension. However, this result is in sharp contrast to tion due to hydration of the surface layers of the pro-
that obtained in protein ultrafiltration [2, 54]. It is tein even around the isoelectric point. When a BSA
clear that eav is much larger near the isoelectric pH. molecule acquires a charge, the filter cake becomes
Since the titanium dioxide particles are hydrophobic loose and wet due to electrostatic repulsion between
colloids, they are destabilized around the isoelectric the charged BSA molecules. This contrasts to the
point where the van der Waals attraction is more dom- compact filter cake around the isoelectric point. The
inant. Consequently, the particles will tend to come average specific filtration resistance a av has a definite
together, i.e. to f locculate, and the very porous f locs maximum around the isoelectric point since a com-
are then formed. Thus, it is speculated that the filter pact filter cake provides a large hydraulic f low resis-
cake formed from such porous f locs has often loose tance.
and wet structures. On the other hand, the filter cake Charge effects are weakened in the presence of
becomes compact and dry when the particle carries solids. In Fig. 17, the average porosity eav of the filter

1015
8 1.00
BSA
1012
12 0.7 s6103
p98 kPa
RUTILE 6
10 s0.02
cs0.2 mol/m3
p196 kPa 0.95
αav [m/kg]

8
εav []
αav [m/kg]

4
εav []

6 0.6

4
2
0.90
2

0 0.5 0
2 4 6 8 10 12 2 4 6 8 10 12
pH pH

Fig. 15 Effect of pH on average specific filtration resistance and Fig. 16 Effect of pH on average specific filtration resistance and
average porosity of filter cake formed in microfiltration of average porosity of filter cake formed in ultrafiltration of
titanium dioxide suspensions. BSA solutions.

KONA No.21 (2003) 29


cake formed in particulate microfiltration is plotted
7. Properties of Filter Cake Composed of
against absolute values of zeta potentials. This figure
Mixtures
shows a close relation between eav and the zeta poten-
tial. It is surprising that the plots can be represented Especially when two types of particles which may
by a unique curve irrespective of the sign of the zeta be membrane foulants in combination are present in
potential. The cake porosity increases with decreas- the feed f luid, the situation becomes more complex.
ing zeta potential. The porosity remains constant in The filtration behaviors can be strongly affected by
the range of the zeta potential below ca. 20 mV. As the the nature of the interaction between the dissimilar
zeta potential decreases, the electrostatic repulsion components.
decreases and thus the porous f loc forms. Therefore, The variations of the filtration rate with time and
the average porosity of the filter cake increases. The the properties of the filter cake were examined for
repulsive force that stems from the overlapping of the microfiltration of binary mixtures of fine particles
electric double layers, depends on the presence or [55]. The average porosity in the filter cake of the
absence of ionized salt in suspension. The repulsive binary particulate mixtures was explained well by the
force is less effective at higher ion concentrations and mixture packing model in which small particles fill
cannot counteract the van der Waals attraction, and so the voids between large particles. Further, the aver-
the solution f locculates. Therefore, it is conceivable age specific filtration resistance of the binary mix-
that the filtration rate is markedly inf luenced not only tures was well evaluated on the basis of the additive
by the solution pH but also by the solution ionic law of the effective specific surface area. However, in
strength. The result for the NaCl concentration of 200 most of published works on filtration of binary partic-
mol/m3 at pH 10.5 is shown in the figure. The magni- ulate mixtures [56, 57], the effects of the physico-
tude of the zeta potential decreased remarkably as a chemical factors such as the pH and salt concentra-
result of the addition of NaCl compared with that for tion on the filtration performance have received little
the NaCl concentration of 0.2 mol/m3 at the same pH. or no attention.
The average porosity eav of the filter cake is aug- Properties of a filter cake formed in dead-end
mented markedly by the addition of salt. This is microfiltration were examined using binary particu-
because the addition of salt destabilizes the suspen- late mixtures of titanium dioxide and silicon dioxide
sion by reducing the double layer repulsion between having the different values of the isoelectric point for
rutile particles. It should be emphasized that the plot each material [58]. The original surface mean diame-
for the NaCl concentration of 200 mol/m3 lies on the ter of silicon dioxide is 0.92 µm. The silicon dioxide
unique curve. particle has negative charge because it contains only
dissociable group of one type, the silanol groups
SiOH. In Fig. 18, the average specific filtration resis-
tance a av and the average porosity eav of the filter
0.7
cake at pH 4.5 are shown against the mixing ratio
pH 10.5 (st/s) of particles, where st is the mass fraction of tita-
cs200 mol/m3
nium dioxide particles in the bulk suspension, and
the total mass fraction s of particles is kept constant.
It is immediately obvious that a av tends to become
smaller in a mixed system of two components in com-
εav []

0.6 parison with a single component system. A smaller


specific filtration resistance was generally caused by a
larger cake porosity. At this pH value, a titanium diox-
ide has a positive electrical charge, whilst a silicon
RUTILE
s0.02 dioxide is electronegative. For the single component
cs0.2 mol/m3 pH 10.5 suspension, the particles are likely to be well dis-
p196 kPa cs0.2 mol/m3
persed by repulsive electrostatic forces, and form
0.5
0 20 40 60 cakes of higher resistance during filtration. However,
|ζ| [mV] the particles in mixed suspension come together due
to heterocoagulation associated with a coulombic
Fig. 17 Relation between average porosity of filter cake and attractive force and the London-van der Waals force,
absolute value of zeta potential. and the very porous f locs are formed prior to deposi-

30 KONA No.21 (2003)


1012 1015
10 0.7 10 1.00
TiO2 /SiO2 BSA/LYSOZYME
s0.02, pH 4.5 s6103
8 cs0.2 mol/m3, p196 kPa 8 pH 6.9
p98 kPa
0.95
6 6
αav [m/kg]

αav [m/kg]
εav []

εav []
0.6

4 4
0.90

2 2

0 0.5 0 0.85
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
s t /s [] s b /s []

Fig. 18 Effect of mixing ratio of particles on average specific fil- Fig. 19 Effect of mixing ratio of proteins on average specific fil-
tration resistance and average porosity in microfiltration tration resistance and average porosity in ultrafiltration of
of binary particulate mixtures at pH 4.5. binary protein mixtures at pH 6.9.

tion. Thus, the coagulated colloids form highly per- average porosity of the cake decreases markedly by
meable filter cakes, minimizing decreases in the fil- adding the other protein to the single protein solution
tration rate. For reference, the calculations based on because of a higher attraction between solutes. The
the additive law on the specific volume and the effec- decrease of the porosity eav roughly corresponds to
tive specific surface area are shown as the dotted the increase of the resistance a av. The resistance
lines [56]. It is obvious that the effect of the surface shows a definite maximum at the mixing ratio of
charge of the particles on the cake properties should about 0.6 because the most compact cake forms at
be considered. that mixing ratio. Since the protein molecules are
It is of significance to compare the results with hydrophilic colloids, their stability in the solution is
those obtained in dead-end ultrafiltration of the mixed inf luenced not only by the presence of a surface
protein solutions. It was observed that the results charge on the protein but also by hydration of the sur-
obtained with protein ultrafiltration were in sharp con- face layers of the protein. Even in the pH range where
trast to those with particulate microfiltration [17]. the two protein molecules have opposite electrical
The properties of the cake formed on the retentive charges, protein molecules are well dispersed in the
membranes in dead-end ultrafiltration of binary pro- solution due to hydration of the protein. Conse-
tein mixtures have been studied by using a mixture quently, the mixed protein cake becomes compact
of the two proteins BSA and egg white lysozyme. because of the attractive force associated with oppo-
The molecular weight and the isoelectric point of sitely charged solutes.
lysozyme are 14,300 Da and 11.0, respectively. In In microfiltration of binary particulate mixtures, the
Fig. 19, the properties of the filter cake at pH 6.9 are effects of adding electrolyte on a av and eav at two dif-
shown against the mixture ratio of the proteins, ferent pH values are shown in Fig. 20. The addition
where sb is the mass fraction of BSA in the solution, of salts leads to a decrease of the zeta potential
and the total mass fraction s of solutes is kept con- because of a less extensive electrical double layer. It
stant. Of particular importance is the surprising result is observed that at pH 4.5 a av increases and eav
that the porosity eav shows a distinct minimum. In the decreases slightly, but discernibly with the increase
case of the single solute solutions, the average poros- of the salt concentration. The result is explained by
ity is relatively large, since only an electrostatic repul- assuming that the coulombic attractive force between
sive force acts between the solutes. In contrast, in the the particles decreases due to a decrease of the zeta
mixed protein solutions, the BSA and lysozyme mole- potential, and as a result the effect of heterocoagula-
cules are oppositely charged at pH 6.9. Therefore, the tion is weakened. At pH 9.6, it is found that a av

KONA No.21 (2003) 31


1012 tration resistance a av obtained in dead-end ultrafiltra-
10 0.7
tion of titanium dioxide suspensions containing BSA
TiO2 /SiO2
s0.02, s t /s0.5
molecules using 10 kDa membranes are illustrated
8 p196 kPa for various pH values against the mass ratio (sb/s) of
BSA to the net colloids. At pH 4.2 and 5.1, the value of
a av increases with an increase of the mass fraction of
6
αav [m/kg]

BSA. This is because the specific resistance a av of

εav []
0.6
pH 4.5 BSA is significantly larger than that of titanium diox-
4 pH 9.6 ide. At pH 6.0, however, in the range of quite small
BSA fractions a av tends to become smaller than that
of the cake composed of titanium dioxide alone (sb/s
2
0). The result is perhaps surprising when one con-
siders the high specific resistance of BSA. At pH 6.0
0 0.5 the positive charge of the titanium dioxide particle is
0 100 200
cs [mol/m3] cancelled by the adsorption of negatively charged
BSA molecules. The f locculation of titanium dioxide
Fig. 20 Effect of NaCl concentration on average specific filtration therefore occurs due to a marked decrease of the zeta
resistance and average porosity. potential of the complex of titanium dioxide particles
and BSA molecules. This brings about the specific
behavior of a av at pH 6.0.
The average specific filtration resistance a av for dif-
decreases and eav increases markedly with increasing ferent concentrations of sodium chloride is plotted in
NaCl concentration. At this pH, the repulsive force Fig. 22 against the mass ratio of BSA to the net col-
that stems from the overlapping of the electric double loids. The value of a av decreases markedly by the
layers, is reduced because of the charge-shielding addition of an appropriate volume of sodium chloride
due to the presence of salts, and consequently parti- and BSA. This is due to charge-shielding of both par-
cles tend to aggregate [59]. Thus, the cakes formed ticles and proteins caused by the presence of the
are more open and have a lower resistance to f low. salts, thus reducing the electrostatic repulsion be-
In Fig. 21, the profiles of the average specific fil- tween titanium dioxide and BSA.

10 16 10 14

10 15 10 13

10 14 10 12
αav [m/kg]

αav [m/kg]

10 13 10 11
TiO2 /BSA TiO2 /BSA
s0.01 s0.01
p98 kPa pH 4.2
10 12 pH 4.2 10 10 cs0
pH 5.1 cs300 mol/m3
pH 6.0

10 11 10 9
0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.01 0.02 0.03 0.04 0.05
s b /s [] s b /s []

Fig. 21 Effect of mixing ratio of proteins on average specific fil- Fig. 22 Effect of salt addition on average specific filtration resis-
tration resistance in ultrafiltration of particle/protein tance in ultrafiltration of particle/protein mixtures.
mixtures.

32 KONA No.21 (2003)


where cl and sl are the mass fractions of lysozyme in
8. Fractionation of Mixtures by Filtration
the filtrate and bulk feed solution, respectively. Filtra-
Solute/solute interaction plays an important role in tion of binary protein mixtures with this membrane
both the filtration rate and the solute rejection when resulted in nearly complete retention of BSA, but the
two or more proteins are ultrafiltered [60-62]. It is variation in the lysozyme rejection with the filtrate
generally accepted that solutes whose molecular di- volume. In the incipient stages of filtration, the mem-
mensions are clearly small enough to permeate the brane exhibits a high lysozyme rejection which may
membrane are substantially retained by that same be attributed to adsorption of the lysozyme solutes on
membrane when larger solutes are present [63]. the hydrophobic surface of the membrane [65]. The
In Fig. 23, the solute rejection behavior following rejection Robs,l is high during ultrafiltration in the
ultrafiltration of binary BSA/lysozyme mixtures con- absence of NaCl (cs0). The BSA molecule is nega-
taining equal amounts of each protein is shown in the tively charged at pH 7, while the lysozyme molecule
form of the apparent rejection Robs,l of lysozyme ver- has a net positive charge at this pH value. The filter
sus the cumulative filtrate volume v per unit effective cake formed on the membrane surface consists of the
membrane area collected in the filtration time q [64]. admixture of two proteins very well packed to form a
Ultrafiltration experiments were carried out with the compact layer because the coulombic, attractive force
upward filtration mode, in which the filtrate f low was associated with oppositely charged solutes, resulting
opposite to the direction of gravity, using hydropho- in high rejection of lysozyme. It can be seen that the
bic, sorptive polysulfone membranes with a molecular lysozyme rejection decreases with the increase of the
weight cut off of 30,000 Da, making it essentially NaCl concentration. In the case of a charge-stabilized
impermeable to BSA, but permeable to lysozyme. colloid such as protein, solute interactions depend on
Solution pH was adjusted to 7, which is pH between the magnitude of the surface charge and/or on the
the isoelectric points of both proteins, and the NaCl extent of the electrical double layer, and this depends
concentration cs was varied. The apparent rejection on the total electrolyte concentration. When salts are
Robs,l of lysozyme can be defined by added, this leads to a less extensive electrical double
layer. Such charge-shielding between the protein mol-
cl
Robs,l 1 (24) ecules resulting from the existence of salts would
sl
reduce electrostatic attraction between BSA and
lysozyme molecules. Thus, the solute rejection at first
declines to a minimum value. However, as filtration
proceeds, the compressible filter cake of retained
BSA/LYSOZYME BSA solutes provides a barrier to transport of the
s b5104, s l5104, pH 7 smaller lysozyme solutes [66], and hence the rejec-
1.0 p98 kPa, UPWARD
tion of the lysozyme solutes rises dramatically.
In Fig. 24, the f lux behavior of the experiment
described in Fig. 23 is shown in the form of the reci-
0.8
procal filtration rate (dq/dv) versus v. This plot is well
known as the Ruth plot [3] in the cake filtration of
Robs,l []

particulate suspensions. In principle, the value of


0.6
dq/dv is directly proportional to the hydraulic resis-
tance due to solute accumulation on or in the mem-
0.4 brane. It would be expected that a rather dense, finely
porous filter cake of retained BSA/lysozyme mixtures
cs0
forms on the membrane surface during the early
0.2 cs150 mol/m3 stages of filtration without added salt. Since the
cs300
hydraulic resistance of this filter cake is remarkably
cs600
large, the value of dq/dv rises rapidly to a near-
0
0 2 4 6 8 10 asymptotic value with increasing filtrate volume.
v [cm] However, increase in the thickness of the filter cake is
suppressed because the filter cake is exfoliated con-
Fig. 23 Effect of salt addition on apparent lysozyme rejection at tinuously by the effect of gravity with the progress of
pH 7. filtration, similar to upward ultrafiltration of single

KONA No.21 (2003) 33


103 tration rate in upward ultrafiltration is slightly lower
6
than that in stirred ultrafiltration with stirring of 5.2
BSA/LYSOZYME rad/s, an effect of gravity acting on the filter cake in
s b5104
s l5104 upward ultrafiltration corresponds to that of very low
5
pH 7 shear stress.
p98 kPa
UPWARD Figure 26 shows the lysozyme rejection behavior
4 of the experiment described in Fig. 25. The figure
indicates that the lysozyme solutes pass almost com-
dθ/dv [s/cm]

pletely through the membrane in unstirred downward


3

103
2 2.5
BSA/LYSOZYME DOWNWARD
s b5104 UPWARD
cs0
s l5104 Ω5.2 rad/s
1 cs150 mol/m3 2.0
pH 7 Ω20.9
cs300
cs300 mol/m3 Ω52.4

dθ/dv [s/cm]
cs600
p98 kPa Ω104.7
0 1.5
0 2 4 6 8 10
v [cm]

1.0
Fig. 24 Effect of salt addition on filtration rate at pH 7.

0.5

protein solutions [40, 54]. The filtration rate increases


0
markedly on addition of NaCl because of the forma- 0 2 4 6 8 10
tion of a filter cake which is substantially free from v [cm]
lysozyme molecules. However, as filtration proceeds,
the filtration rate decreases gradually because the Fig. 25 Effect of hydrodynamics above membrane on filtration
lysozyme molecules are trapped in the pores of the fil- rate.
ter cake.
In Fig. 25, the reciprocal filtration rate (dq/dv) in
the unstirred downward and upward, and the stirred 1.0

downward ultrafiltration of binary BSA/lysozyme BSA/LYSOZYME


s b5104, s l5104
mixtures is shown as a function of v [67]. For these pH 7, cs300 mol/m3
0.8
experiments, ultrafiltration membranes with very low p98 kPa
DOWNWARD
adsorptivity for proteins were used. Solution envi- UPWARD
ronment was set at pH 7 and the NaCl concentration Ω5.2 rad/s
0.6
Ω20.9
Robs,l []

of 300 mol/m3 where electrostatic interactions be- Ω52.4


tween BSA and lysozyme weakened comparatively Ω104.7
as implied from Figs. 23 and 24. For conventional 0.4
unstirred downward ultrafiltration, the plot is virtually
linear throughout the course of filtration because of
the continuous formation of the filter cake on the 0.2

membrane in accordance with the compressible cake


filtration model [2, 9, 15]. For unstirred upward ultra-
0
filtration, the f lux decline is suppressed, similar to the 0 2 4 6 8 10
result of Fig. 24. The increase in the shear stress act- v [cm]
ing on the membrane surface by stirring leads to the
suppression of the cake deposition, and consequently Fig. 26 Effect of hydrodynamics above membrane on apparent
the marked increase in the filtration rate. Since the fil- lysozyme rejection.

34 KONA No.21 (2003)


ultrafiltration and are only a little retained by the to the propagation direction of the ultrasonic wave.
membrane in upward ultrafiltration except that the The oscillating frequency and the output power of the
high rejection appears in the incipient stages of filtra- ultrasonic irradiation were 25 kHz and 180 W, respec-
tion. It has been reported that in single protein solu- tively. It is evident that the ultrasonic fields can
tions the stirring resulted in increasing the solute contribute to the remarkable improvement in the
rejection [63]. It should be emphasized that also in filtration rate. As was shown in Figs. 25 and 26,
binary protein mixtures the higher stirring speed although the observed filtration rate becomes high by
causes the higher rejection of the solutes. This is increasing the shear stress on the membrane, the
because the concentration of lysozyme near the mem- transmission of the lysozyme molecules becomes
brane decreases with increasing shear stress acting less. In contrast, of particular importance is the sur-
on the membrane. prising observation that the ultrasonic irradiation
Figures 25 and 26 suggest that it is necessary to does not lower the lysozyme transmission regardless
control the hydrodynamics above the membrane in of involving a marked increase in the filtration rate.
order to obtain high lysozyme transmission simulta- One possible explanation for this result is that ultra-
neously with high filtration rate. While hard stirred sonic irradiation supplies vibrational energy to the fil-
ultrafiltration, which produces the high filtration rate ter cake to keep the solutes partly suspended, and
and solute rejection, is very effective for concentrat- therefore leaves more free channels for the filtrate
ing protein solutions, it is not suitable for fractionation f low.
of binary protein mixtures. It can be concluded from To clarify the effects of the ultrasonic irradiation,
the figures that unstirred upward or mildly stirred the lysozyme rejections observed in the cases where
ultrafiltration is more advantageous for the efficient the filtration rates with and without ultrasonic irradia-
fractionation of binary protein mixtures. tion are very similar are compared in Fig. 28. The
The effects of ultrasonic irradiation on the filtration figure shows that the filtration rate in the ultrasonic
rate and lysozyme rejection in upward ultrafiltration upward ultrafiltration irradiated under output power
of binary BSA/lysozyme mixtures using the ultra- of 180 W is substantially similar to that in downward
filtration membrane with very low adsorptivity for ultrafiltration stirred with a rotational speed of 20.9
proteins at pH 7 and high salt concentration are rad/s without ultrasonic irradiation. However, it is of
shown in Fig. 27 [67]. The whole of the filter was importance to note that there is a marked difference
kept immersed in the ultrasonic cleaning bath full of in the lysozyme rejection. The lysozyme transmission
water during the course of filtration. The filter was set in ultrasonic upward ultrafiltration is much higher
up so that the membrane surface was perpendicular than that in stirred ultrafiltration. Therefore, it is

103 102
2.0 1.0 10 1.0
BSA/LYSOZYME BSA/LYSOZYME
s b5104, s l5104 s b5104, s l5104
pH 7, cs300 mol/m3 0.8 8 pH 7, cs300 mol/m3 0.8
1.5 p98 kPa p98 kPa
dθ/dv [s/cm]

dθ/dv [s/cm]

0.6 6 0.6
Robs,l []

Robs,l []

1.0
, UPWARD
, UPWARD, ULTRASONIC 0.4 4 0.4
, Ω20.9 rad/s
, UPWARD, ULTRASONIC
0.5
0.2 2 0.2

0 0 0 0
0 2 4 6 8 10 0 2 4 6 8 10
v [cm] v [cm]

Fig. 27 Effect of ultrasonic irradiation on filtration rate and ap- Fig. 28 Comparison of apparent lysozyme rejection between
parent lysozyme rejection. cases with and without ultrasonic irradiation in which
similar filtration rate are observed.

KONA No.21 (2003) 35


believed that the ultrasonic irradiation is quite effec- cl mass fraction of lysozyme in filtrate
tive not only for concentration of proteins by ultra- cs NaCl concentration of solution (mol/m3)
filtration [68] but also for fractionation of the binary D inner diameter of inserted cylinder (m)
protein mixtures by ultrafiltration. Dh diameter of hole (m)
df average equivalent spherical diameter of
solute (m)
Conclusions
E empirical constant defined by Eq. (22) (Paβ)
As was seen from this review, an understanding of E0 empirical constant defined by Eq. (13)
the properties of the filter cake formed on the filter e local void ratio
medium or the membrane can serve as a basis for eav average void ratio
clarifying the real mechanism of cake filtration and eav,ω average void ratio from filter medium
membrane filtration. In recent years, experimental surface to distance w from filter medium
testing procedures have been newly developed such h height of filter chamber (m)
as filtration experiments in which a filter was sub- J filtration rate (m/s)
jected to a sudden reduction in its filtration area and Jω apparent solvent velocity relative to
ultracentrifugation experiments. The compressible solid at arbitrary position w (m/s)
cake filtration model which explicitly took the non- Kv Ruth coefficient of constant pressure
homogeneity and the compressibility of the filter filtration defined by Eq. (2) (m2/s)
cake into consideration was used to describe the L thickness of filter cake (m)
properties of the filter cake and the filtration behav- m ratio of wet to dry cake mass
iors. In protein ultrafiltration, the validity of the com- n empirical constant defined by Eq. (12)
pressible cake filtration model was verified by the p applied filtration pressure (Pa)
measurements of the concentration distributions in pm pressure loss through filter medium (Pa)
the filter cake using the principle of inclined filtration ps local solid compressive pressure (Pa)
where a large amount of filter cake is formed. More- psi pressure below which e remains constant (Pa)
over, it was shown that the solution properties (in par- R distance from center of rotation to bottom
ticular, the solution pH and the electrolyte strength) of sedimentation (m)
play an important role in the filtration behaviors of Robs,l apparent rejection of lysozyme defined
colloids. The filtration behaviors of mixtures can be by Eq. (24)
strongly affected by the nature of the interaction be- ri radial distance of sedimentation boundary
tween the dissimilar components. Factors inf luencing from center of rotation (m)
fractionation behaviors by filtration were clarified. rs distance from center of rotation to surface
The author believes that many existing separation of sediment (m)
problems would have been avoided by the application S sedimentation coefficient (s/rad2 )
of available scientific data although the random na- s mass fraction of solids in slurry, or mass
ture of most particulate dispersions has resulted in a fraction of solutes in bulk feed solution
difficult process problem. sb mass fraction of BSA in bulk feed solution
sl mass fraction of lysozyme in bulk feed solution
st mass fraction of titanium dioxide particles
Acknowledgements
in slurry
The author wishes to acknowledge the Ministry of u apparent liquid velocity relative to solid (m/s)
Education, Culture, Sports, Science and Technology, u1 filtration rate (m/s)
Japan for the financial support and all university part- v cumulative filtrate volume collected per
ners for the excellent cooperation. unit effective medium area (m3/m2)
vt cumulative filtrate volume collected per
unit effective medium area until
Nomenclature
cake surface reaches disk with hole (m3/m2)
Cc empirical constant defined by Eq. (13) v0 sedimentation velocity (m/s)
Ch ratio of volume of hydrous protein molecule w0 net mass of solid per unit effective
to volume of anhydrous protein molecule medium area (kg/m2)
c local mass fraction of solid or solute in x distance from medium surface (m)
filter cake α local specific filtration resistance (m/kg)

36 KONA No.21 (2003)


a av average specific filtration resistance (m/kg) 11) Murase, T., Iritani, E., Cho, J. H. and Shirato, M.,
a1 empirical constant defined by “Determination of Filtration Characteristics Based upon
Eq. (12) (mn1s2n/kgn1) Filtration Tests under Step-up Pressure Conditions,” J.
Chem. Eng. Japan, 22, 373-378 (1989).
b empirical constant defined by Eq. (22)
12) Terzaghi, K. and Peck, R. B., “Soil Mechanics in Engi-
e local porosity
neering Practice,” Wiley, New York, 65 (1948).
eav average porosity 13) Tiller, F. M., “The Role of Porosity in Filtration. Part 2.
z zeta potential (V) Analytical Equations for Constant Rate Filtration,”
q filtration time (s) Chem. Eng. Progr., 51, 282-290 (1955).
m viscosity of filtrate (Pa・s) 14) Tanford, C., “Physical Chemistry of Macromolecules,”
r density of filtrate (kg/m3) Wiley, New York, 317 (1961).
rs true density of solid (kg/m3) 15) Iritani, E., Hattori, K. and Murase, T., “Analysis of Dead-
End Ultrafiltration Based on Ultracentrifugation Meth-
y angle between filtrate f low and direction
od,” J. Membrane Sci., 81, 1-13 (1993).
of gravity (rad)
16) Iritani, E., Mukai, Y. and Hagihara, E., “Measurements
W angular velocity (rad/s) and Evaluation of Concentration Distributions in Filter
w net solid volume per unit medium area Cake Formed in Dead-End Ultrafiltration of Protein
lying from medium up to an arbitrary Solutions,” Chem. Eng. Sci., 57, 53-62 (2002).
position in cake (m) 17) Iritani, E., Mukai, Y. and Murase, T., “Properties of Fil-
w0 net solid volume of entire cake per unit ter Cake in Dead-End Ultrafiltration of Binary Protein
medium area (m) Mixtures with Retentive Membranes,” Trans. IChemE,
Part A, Chem. Eng. Res. Des., 73, 551-558 (1995).
18) Tiller, F. M. and Cooper, H., “The Role of Porosity in Fil-
tration. Part V. Porosity Variation in Filter Cakes,”
References
AIChE J., 8, 445-449 (1962).
1) Fane, A. G., “Ultrafiltration: Facrtors Inf luencing Flux 19) Iritani, E., Hattori, K. and Murase, T., “Evaluation of
and Rejection,” Progress in Filtration and Separation, Dead-End Ultrafiltration Properties by Ultracentrifuga-
Wakeman, R. J. ed., Elsevier, Amsterdam, 4, 134 (1986). tion Method,” J. Chem. Eng. Japan, 27, 357-362 (1994).
2) Iritani, E., Nakatsuka, S., Aoki, H. and Murase, T., 20) Iritani, E., Hattori, K., Akatsuka, S. and Murase, T.,
“Effect of Solution Environment on Unstirred Dead-End “Sedimentation Behavior of Protein Solutions in Ultra-
Ultrafiltration Characteristics of Proteinaceous Solu- centrifugation Field,” J. Chem. Eng. Japan, 29, 352-358
tions,” J. Chem. Eng. Japan, 24, 177 (1991). (1996).
3) Ruth, B. F., “Studies in Filtration. III. Derivation of Gen- 21) Michaels, A. S. and Bolger, J. C., “Settling Rates and
eral Filtration Equations,” Ind. Eng. Chem., 27, 708-723 Sedimentation Volumes of Flocculated Kaolin Suspen-
(1935). sions,” Ind. Eng. Chem. Fundam., 1, 24-33 (1962).
4) Grace, H. P., “Resistance and Compressibility of Filter 22) Murase, T., Iwata, M., Adachi, T., Gmachowski, L. and
Cakes,” Chem. Eng. Progr., 49, 303-318 (1953). Shirato, M., “An Evaluation of Compression-Permeabil-
5) Shirato, M., Aragaki, T., Mori, R. and Sawamoto, K., ity Characteristic in the Intermediate Concentration
“Predictions of Constant Pressure and Constant Rate Range by Use of Centrifugal and Constant-Rate Com-
Filtration Based upon an Approximate Correction for pression Techniques,” J. Chem. Eng. Japan, 22, 378-384
Side Wall Friction in Compression Permeability Cell (1989).
Data,” J. Chem. Eng. Japan, 1, 86-90 (1968). 23) Shirato, M., Aragaki, T., Ichimura, K. and Ootsuji, N.,
6) Tiller, F. M., Haynes, S. and W-M. Lu, “The Role of “Porosity Variation in Filter Cake under Constant-Pres-
Porosity in Filtration: VII Effect of Side Wall Friction in sure Filtration,” J. Chem. Eng. Japan, 4, 172-177 (1971).
Compression-Permeability Cells,” AIChE J., 18, 13-20 24) Michaels, A. S., “New Separation Technique for the
(1972). CPI,” Chem. Eng. Progr., 64, 31-43 (1968).
7) Gale, R. S., “Recent Research on Sludge Dewatering,” 25) Porter, M. C., “Concentration Polarization with Mem-
Filtr. & Sep., 8, 531-538 (1971). brane Ultrafiltration,” Ind. End. Chem., Prod. Res. Dev.,
8) Murase, T., Iritani, E., Cho, J. H., Nakanomori, S. and 11, 234-248 (1972).
Shirato, M., “Determination of Filtration Characteristics 26) Trettin, D. R. and Doshi, M. R., “Ultrafiltration in an
due to Sudden Reduction in Filtration Area of Filter Unstirred Batch Cell,” Ind. End. Chem., Fundam., 19,
Cake Surface,” J. Chem. Eng. Japan, 20, 246-251 (1987). 189-194 (1980).
9) Shirato, M., Sambuichi, M., Kato, H. and Aragaki, T., 27) Kozinski, A. A. and Lightfoot, E. N., “Protein Ultrafiltra-
“Internal Flow Mechanism in Filter Cakes,” AIChE J., tion: A General Example of Boundary Layer Filtration,”
15, 405-409 (1969). AIChE J., 18, 1030-1040 (1972).
10) Shirato, M., Aragaki, T. and Iritani, E., “Analysis of Con- 28) Jonsson, G., “Boundary Layer Phenomena during Ultra-
stant Pressure Filtration of Power-Law Non-Newtonian filtration of Dextran and Whey Protein Solutions,”
Fluids,” J. Chem. Eng. Japan, 13, 61-66 (1980). Desalination, 51, 61-77 (1984).

KONA No.21 (2003) 37


29) Vilker, V. L., Colton, C. K. and Smith, K. A., “Concentra- 44) Bacchin, P., Aimar, P. and Sanchez, V., “Inf luence of Sur-
tion Polarization in Protein Ultrafiltration,” AIChE J., face Interaction on Transfer during Colloid Ultrafiltra-
27, 632-645 (1981). tion,” J. Membrane Sci., 115, 49-63 (1996).
30) Wijmans, J. G., Nakao, S., van den Berg, J. W. A., 45) Kawalec-Pietrenko, B. and Ruszel-Lichodzijiewska, R.,
Troelstra, F. R. and Smolders, C. A., “Hydrodynamic “ The Inf luence of Physico-Chemical and Operating
Resistance of Concentration Polarization Boundar y Lay- Parameters on the Filtration of Precipitate of Hydrated
ers in Ultrafiltration,” J. Membrane Sci., 22, 117-135 Zr(IV) Oxide,” J. Chem. Tech. Biotechnol., 35A, 426-430
(1985). (1985).
31) van den Berg, G. B. and Smolders, C. A., “Flux Decline 46) McDonogh, R. M., Fell, C. J. D. and Fane, A. G., “Sur-
in Membrane Processes,” Filtr. & Sep., 25, 115-121 face Charge and Permeability in the Ultrafiltration of
(1988). Non-Flocculating Colloids,” J. Membrane Sci., 21, 285-
32) Reihanian, H., Robertson, C. R. and Michaels, A. S., 294 (1984).
“Mechanism of Polarization and Fouling of Ultrafiltra- 47) McDonogh, R. M., Welsch, K., Fane, A. G. and Fell, C. J.
tion Membranes by Proteins,” J. Membrane Sci., 16, D., “Incorporation of the Cake Pressure Profiles in the
237-258 (1983). Calculation of the Effect of Particle Charge on the Per-
33) Suki, A., Fane, A. G. and Fell, C. J. D., “Flux Decline in meability of Filter Cakes Obtained in the Filtration of
Protein Ultrafiltration,” J. Membrane Sci., 21, 269-283 Colloids and Particulates,” J. Membrane Sci., 72, 197-
(1984). 204 (1992).
34) Nakao, S., Wijmans, J. G. and Smolders, C. A., “Resis- 48) Schultz, J., Papirer, E. and Nardin, M., “Physicochemi-
tance to the Permeate Flux in Unstirred Ultrafiltration cal Aspects of the Filtration of Aqueous Suspensions of
of Dissolved Macromolecular Solutions,” J. Membrane Fibers and Cement. 4. Inf luence of the ζ Potential of the
Sci., 26, 165-178 (1986). Fibers on Filtration Efficiency,” Ind. Eng. Chem., Prod.
35) Nakakura, H., Yamashita, A., Sambuichi, M. and Osasa, Res. Dev., 22, 102-105 (1983).
K., “Electrical Conductivity Measurement of Filter Cake 49) Tarleton, E. S. and Wakeman, R. J., “Understanding
in Dead-End Ultrafiltration of Protein Solution,” J. Flux Decline in Crossf low Microfiltration: Prat II –
Chem. Eng. Japan, 30, 1020-1025 (1997). Effects of Process Parameters,” Trans. IChemE, Part A,
36) Nakao, S., Nomura, T. and Kimura, S., “Characteristics Chem. Eng. Res. Des., 72, 431-440 (1994).
of Macromolecular Gel Layer Formed on Ultrafiltration 50) Ueshima, K., Iizuka, H. and Higashitani, K., “Ordered
Tubular Membrane,” AIChE J., 25, 615-622 (1979). Structure of Monodispersed Particles Filtered at Con-
37) McDonogh, R. M., Bauser, H., Stroh, N. and stant Pressure,” J. Chem. Eng. Japan, 24, 647-652 (1991).
Grauschopf, U., “Experimental in situ Measurement of 51) Welsch, K., McDonogh, R. M., Fane, A. G. and Fell, C. J.
Concentration Polarisation during Ultra- and Micro-fil- D., “Calculation of Limiting Fluxes in the Ultrafiltration
tration of Bovine Serum Albumin and Dextran Blue of Colloids and Fine Particulates,” J. Membrane Sci., 99,
Solutions,” J. Membrane Sci., 104, 51-63 (1995). 229-239 (1995).
38) Gowman, L. M. and Ethier, C. R., “Concentration and 52) Willmer, S. A., Tarleton, E. S. and Holdich, R. G., “ The
Concentration Gradient Measurements in an Ultrafiltra- Importance of Cake Compressibility in Deadend Pres-
tion Concentration Polarization Layer, Part I: A Laser- sure Filtration,” Proc. 7th World Filtration Congress,
Based Refractometric Experimental Technique,” J. Vol. I, p. 27-31, Hungarian Chemical Society, Budapest,
Membrane Sci., 131, 95-105 (1997). Hungary (1996).
39) Fernández-Torres, M. J., Ruiz-Beviá, F., Fernández- 53) Iritani, E., Toyoda, Y. and Murase, T., “Effect of Solution
Sempere, J. and López-Leiva, M., “Visualization of the Environment on Dead-End Microfiltration Characteris-
UF Polarized Layer by Holographic Interferometry,” tics of Rutile Suspensions,” J. Chem. Eng. Japan, 30,
AIChE J., 44, 1765-1776 (1998). 614-619 (1997).
40) Iritani, E., Watanabe, T. and Murase, T., “Upward and 54) Iritani, E., Watanabe, T. and Murase, T., “Properties of
Inclined Ultrafiltration under Constant Pressure by Use Filter Cake in Dead-End Upward Ultrafiltration,” J.
of Dead-End Filter,” Kagaku Kogaku Ronbunshu, 17, Membrane Sci., 69, 87-97 (1992).
206-209 (1991). 55) Iritani, E., Nagaoka, H. and Murase, T., “Filtration Char-
41) Zeng, W. P., Iritani, E. and Murase, T., “Dynamic Behav- acteristics and Cake Properties in Microfiltration of
ior of BSA Cake Layer Deposited on Membrane in Binary Mixtures of Fine Particles,” Kagaku Kogaku
Dead-End Inclined Ultrafiltration,” Kagaku Kogaku Ron- Ronbunshu, 22, 648-654 (1996).
bunshu, 24, 158-160 (1998). 56) Shirato, M., Sambuichi, M. and Okamura, S., “Filtration
42) Tiller, F. M., “ The role of Porosity in Filtration, Numeri- Behavior of a Mixture of Two Slurries,” AIChE J., 9,
cal Methods for Constant Rate and Constant Pressure 599-605 (1963).
Filtration Based on Kozeny’s Law,” Chem. Eng. Progr., 57) Abe, E., Hirosue, H. and N. Yamada, “A Method for Pre-
49, 467-479 (1953). dicting the Filtration Characteristics of Body Filtration,”
43) Bacchin, P., Aimar, P. and Sanchez, V., “Model for Col- J. Soc. Powder Technol. Japan, 32, 550-556 (1995).
loidal Fouling of Membranes,” AIChE J., 41, 368-376 58) Iritani, E., Mukai, Y. and Toyoda, Y., “Properties of a Fil-
(1995). ter Cake Formed in Dead-End Microfiltration of Binary

38 KONA No.21 (2003)


Particulate Mixtures,” J. Chem. Eng. Japan, 35, 226-233 ration of Proteins by Ultrafiltration through Sorptive
(2002). and Non-Sorptive Membranes,” J. Membrane Sci., 69,
59) Chang, D. J., Hsu, F. C. and Hwang, S. J., “Steady State 189-211 (1992).
Permeate Flux of Cross-Flow Microfiltration,” J. Mem- 64) Iritani, E., Mukai, Y. and Murase, T., “Upward Dead-End
brane Sci., 98, 97-106 (1995). Ultrafiltration of Binary Protein Mixtures,” Sep. Sci.
60) Ingham, K. C., Busby, T. F., Sahlestrom, Y. and Castino, Technol., 30, 369-382 (1995).
F., “Separation of Macromolecules by Ultrafiltration: 65) Iritani, E., Tachi, S. and Murase, T., “Inf luence of Pro-
Inf luence of Protein Adsorption, Protein-Protein Inter- tein Adsorption on Flow Resistance of Microfiltration
actions, and Concentration Polarization,” Polymer Sci- Membrane,” Colloids Surfaces A: Physicochem. Eng.
ence and Technology, Ultrafiltration Membranes and Aspects, 89, 15-22 (1994).
Applications, Cooper, A. R. ed., Plenum Press, New 66) Iritani, E., Itano, Y. and Murase, T., “Dead-End Inclined
York, 13, 141 (1980). Ultrafiltration by Use of Dynamic Membranes,” Kagaku
61) Rodgers, V. G. J. and Sparks, R. E., “Reduction of Mem- Kogaku Ronbunshu, 19, 536-539 (1993).
brane Fouling in the Ultrafiltration of Binary Protein 67) Mukai, Y., Iritani, E. and Murase, T., “Fractionation
Mixtures,” AIChE J., 37, 1517-1528 (1991). Characteristics of Binary Protein Mixtures by Ultra-
62) Zhang, L. and Spencer, H. G., “Selective Separation of filtration,” Sep. Sci. Technol., 33, 169-185 (1998).
Proteins by Microfiltration with Formed-in-Place Mem- 68) Kokugan, T., Kaseno, T., Fujiwara, S. and Shimizu, M.,
branes,” Desalination, 90, 137-146 (1993). “Ultrasonic Effect on Ultrafiltration Properties of
63) Nakatsuka, S. and Michaels, A, S., “ Transport and Sepa- Ceramic Membrane,” Membrane, 20, 213-223 (1995).

Author’s short biography


Eiji Iritani
Dr. Eiji Iritani is Professor of the Department of Chemical Engineering at Nagoya
University. He received his PhD from Nagoya University in 1981. He has worked
extensively on most aspects of solid-liquid separation. These works now also
include membrane separation, water treatment, and colloid engineering.

KONA No.21 (2003) 39

You might also like