Rate-Controlled Separations: Phillip C. W Anka T
Rate-Controlled Separations: Phillip C. W Anka T
SEPARATIONS
PHILLIP C. W ANKA T
Schoo/ of Chemica/ Engineering, Purdue University,
West Lafayette, Indiana, USA
A Catalogue record for this book is available from the British Library
Library of Congress Cataloging-in-Publication Data available
® Printed on acid-free text paper, manufactured with ANSI/NISO Z39.48-
1992 and ANIS/NISO Z39.48-1984 (Permanence of Paper)
Part III
MEMBRANES
NOMENCLATURE CHAPTERS 12 AND 13
Cb bulk concentration
cg gel concentration
em . concentration in membrane
Cp penneate concentration
Cw wall concentration
623
624
d,D diameter, cm or m
D Diffusivity, cm zIs
energy, joules
I current, amps
L length, cm or m
MW molecular weight
n total moles solute, Section 12.5.
high pressure
low pressure
permeate pressure
partial pressure
charge,coloumbs
r radial direction
R radius, cm
R retention
R gas constant
RO inherent retention = 1 - c;,/cw
~ pore radius, cm
t time, s or days
tm membrane thickness, cm
T absolute temperature, K
T f , T; freezing temperature of solution and pure solvent, K
v voltage, volts
v volume, cm3 or L
VP vapor pressure
w width, cm
x axial distance, cm
Y' y/h
gas mole fraction on high pressure side
Greek Letters
n selectivity, KsolvlK.olute in RO
£ porosity
a cut = Fp/Fin
v kinematic viscosity = Wp
~ efficiency of current utilization, Eq. (12-67)
7t Pi,3.141592654 ....
P density, g/cm 3
A
A membrane is a physical barrier between two fluids. In over 99% of the cases
of current industrial interest the membrane is made from a polymer. This poly-
mer is cast or spun or extruded to form a continuous film without holes in the
desired geometry. The simplest geometry used for membrane separators is the
stirred cell shown in Figure 12-1. The fluid enters into the upper chamber. Part
of the feed stream is transferred through the membrane to the lower chamber.
With a perfect membrane only species A will transfer through the membrane
while species B will remain in the upper chamber. Either species may be the
desired product. The stirrer is used to improve mass transfer rates and prevent
the buildup of species B at the membrane surface.
630
631
~
~
,- I
~
-
Membran e
t +
Cp
III
High fluxes are usually very desirable since the transfer area required
will be lessened. Thus we want a high driving force. In the cases of gas per-
meation, reverse osmosis (used for purifying water) and ultrafiltration (used for
concentrating large molecules) this means a high pressure drop across the
membrane. The membrane must be mechanically strong to withstand these
high pressures. Mechanical strength requires a thick membrane. However,
632
a Feed Phase
Thin
Skin
Porous
Support
Layer
Permeate
Eq.(12-1) shows that thick membranes will have low fluxes and hence large
are$ will be required to achieve the desired transfer rate. A fundamental prob-
lem which stopped the commercial use of membranes for many years was how
to make a membrane which was thin enough to have high fluxes while at the
same time it was mechanically strong.
The solution to this problem was found in the late 1950's and early
1960's by Sidney Loeb and S. Sourirajan (Loeb and Sourirajan, 1960, 1963).
Their solution was to make a membrane which was asymmetric or anisotropic.
This type of membrane is illustrated in Figure 12-2. The membrane was cast to
have a very dense thin layer or skin on one side. This layer was typically 0.1 to
1.0 microns thick. The thin skin does the actual separation of the species.
Thus in Eq.(12-1) the thickness term is very small. The thin skin is supported
by a much thicker porous layer. This layer provides the required mechanical
strength, but it is so porous that no separation occurs in this layer. These mem-
branes can be operated at pressure drops in excess of 1000 psig. The Loeb-
Sourirajan membrane was made from cellulose acetate for desalination of
water by reverse osmosis. Methods for casting these membranes are reviewed
by Kesting (1985) and Soltanieh and Gill (1981). The use of asymmetric mem-
branes has been extended to a variety of other polymers and to other membrane
separation methods. The history of membrane separations is discussed by
Lonsdale (1982). Not all membrane separators use asymmetric membranes.
The exceptions will be discussed in the individual sections.
The most common structures are shown in Figure 12-3 (Kesting, 1985;
634
b c
(-C- y- )
\ C=N n
(-S02~r@-o~~?)
H ~ cf H n
f
) -CF2,C-) n
(-(CF2 CF2,
to-poj
O-(C~ yF )y-OC~C~S02F g
CF3
CH 3 n
Pusch and Walch. 1982; Soltanieh and Gill. 1981; Ward et al., 1985). Addi-
tional examples of polymers used for membranes are discussed by Kesting
(1985). Lloyd and Meluch (1985). Finken (1985). Pusch and Walch (1982), and
Ward et al., (1985). For example. Figure 12-3 shows only one of the possible
polyamides (Kesting, 1985; Pusch and Walch, 1982). The cellulose (Figure
12-3a) and cellulose acetate membranes were the firstcommercialIy successful
membranes. Although still commonly used, considerable research has gone
into developing membranes with better chemical and thermal resistances.
Polysulfone (Figure 12-3c) and Nafion (Figure 12-3f) have outstanding resis-
tances to chemicals. Polystyrene cross-linked with divinylbenzene (Figures 9-1
a and b) also has excellent chemical resistance and is used as a backing for
composite membranes and as the basis for electrodialysis membranes.
With a high flux rate species B which does not transfer through the mem-
brane may build up to a high concentration at the surface of thel membrane.
This is called concentration polarization and is illustrated in Figure 12-4.
Since the transfer rate of species B through the membrane is low, the permeate
concentration is low, cp «Cb' Species A (which might be the solvent) has a
high flux through the membrane. Thus. there must be a convective flow
towards the membrane. This convective flow carries B to the membrane sur-
face. Since the flux of B is low. the concentration of B will build up at the
membrane wall. Thus, a concentration gradient is produced. This buildup of B
I
+ Back Diffusion
Convective Flow
I
---:--- cb
-~·I
y=8
Membrane
Figure 12-4. Build-up of retained species at membrane wall (concentration
polarization).
636
is counteracted by diffusion into the bulk fluid. If diffusion rates are high as is
true in most gas systems, then the concentration buildup will be modest and
concentration polarization will not be a problem.
The general design requirements for any membrane system can now be listed.
1. Thin active layer of membrane.
2. High permeability for species A and low permeability for
species B.
3. Stable membrane with long service life.
4. Mechanically strong.
5. Large surface area of membrane in small volume.
6. Concentration polarization eliminated or at least controlled.
7. Easy to clean if necessary.
8. Inexpensive to build.
9. Low operating cost
This list of desirable design conditions is general. Items specific to each of the
membrane devices will be discussed later. A historical perspective on mem-
brane modules with many early advertisements is given by Kremen (1986).
Currently, spiral wound and hollow fiber systems have most of the market.
The simplest geometry for a membrane separator is the stirred cell shown
in Figure 12-1. Stirred cells are commonly used in laboratories, but the mem-
brane surface area is small. Thus for large scale commercial applications other
637
a b
Porous Tube
O
------------r-.
I \
( \
I I
I I
\ I
- - -
\ I
- - - - - - ~ ...'
Membrane Coating
(Inside or Outside)
----..:: Spacers
Reten~
c
... :~
Permeate
Permeate channel
Feed--(]lilililililililllll: ·
L Hollow fibers
Distribution
Figure 12-5. Geometries of commercial membrane systems. a. Plale-and-
frame. b. Tubular. c. Spiral wound. d. Hollow fiber.
geometries are required. The most commonly used commercial systems are
shown in Figure 12-5. The plate-and-frame system is similar to a plate-and-
frame filter press with a series of flat membrane sheets. The area per volume
ratio is significantly higher than stirred cells but lower than spiral wound or
hollow fiber systems. The sheets can be quite close together to reduce concen-
tration polarization. The major advantage of plate-and-frame systems is they
638
can be taken apart for cleaning if necessary. Their major application has been
in the food industry where fouling is a major problem. The plate-and-frame
geometry is also used for electrodialysis.
The hollow fiber system shown in Figures 12-5d and 12-6 is also one of
the two most popular designs. The fibers are extruded from polymers such as
Nylon or polysulfone. A typical reverse osmosis hollow fiber would have a
inner diameter of 42 microns, an outer diameter of 85 microns, and a 0.1 to 1.0
micron skin on the outer surface (Applegate, 1984). For gas permeation the
fibers are made somewhat larger to reduce the pressure drop inside the tubes.
For ultrafiltration the fibers are much larger (500 to 1100 micron i.d.) and the
skin is on the inside of the fiber. These numbers indicate that a considerable
amount of engineering design can be done with the fibers. The fibers can have
fairly thick walls and with flow radially inward large pressure gradients can be
639
- 1126mm------·---1
-+ Solution outlet
tS9-m;-
diameter
Case
Hollow-liber membrane
Tube.eat
handled. The major advantage of the hollow fiber systems is that a huge
number of fibers (4.5x106 ) can be packed inside a 25.4 cm (10 inch) diameter
cartridge. Thus the area/volume ratio is very high (about 5000 ft2 per ft3) and
costs are kept down. A major disadvantage of hollow fibers is particulate
solids can permanently clog the tiny fibers. Thus dirty fluids must be
thoroughly cleaned before being sent to the hollow fiber separator. Details of
hollow fiber systems are given by Orofino (1977) and Caracciolo et al., (1977).
All of the basic systems are usually built in a few basic modular sizes. A
"module" is a complete unit such as a hollow fiber separator capable of pro-
cessing a given amount of fluid with a given separation. The details of a hol-
low fiber module used for ultrafiltration are shown in Figure 12-6 (Warashina
et al., 1985). A single module is often not sufficient for the required flow rate
and/or separation. Thus modules are cascaded in a variety of ways to produce
the desired separation. One advantage of modular construction is the modules
640
$ - -~
a b
OOOOOOOIOUI.
FIN
YIN
C d
~'I~iQt~ ~ ~ -- i-- ~
! ~
-- ---
e f
Figure 12-7. Idealized flow patterns for flow inside module. a. Completely
mixed. b. Mixed on feed (high pressure side). c. Mixed on per-
meate (low pressure) side. d. Cross-flow. e. Co-current. f.
countercurrent.
The flow patterns inside the module have a large effect on the outlet con-
centrations and flow rates of the high and low pressure products. Several ideal-
ized flow patterns are shown in Figure 12-7. The completely mixed case, Fig-
ure 12-7a, is the easiest to analyze but the least advantageous for separation.
The system which is mixed on the feed side, Figure 12-7b, is close to the flow
pattern which occurs for most hollow fiber systems for gas permeation and
reverse osmosis. Figure 12-7c shows the idealized flow pattern of a hollow
fiber system for ultrafiltration where the feed is inside the fibers. The solution
is well-mixed on the permeate side. Plate-and-frame and spiral wound systems
can have a modification of the cross-flow system, Figure 12-7d, or they could
641
l:
~ ~
fA
~B d41r'
fA
1
e
~~
~ Recycle
~ C A
~
Figure 12-8. Cascades for membrane separators. a. Parallel. b. Series. c.
Parallel-series. d. Recycle. e. Permeate in series. f. Counter-
current fractionation. g. Batch. Species A is more permeable
and species B is less permeable.
operating costs will be high and the use of this countercurrent cascade in a
membrane process will be unusual. In the batch operation shown in Figure
12-8g recovery in a single pass is insufficient, and operation is continued until
enough of the more permeable species has been recovered. Batch systems are
common in applications such as food processing where frequent cleaning is
required. Modifications of these cascades used for gas permeation are shown
in Figure 12-11.
thus this process introduces new concepts. Following this, pervaporation which
involves vaporizing a liquid at the membrane surface will be considered.
Finally, the experimental liquid-membrane systems will be briefly discussed
(12-2)
where h is the volumetric flux, Fp,A is the steady state volumetric transfer rate
of species A through the membrane, A is the membrane area, PAis the permea-
bility of the membrane for species A. the driving force L\p is the change in the
644
partial pressure of the species across the membrane, and lm is the thickness of
the membrane skin. Obviously, consistent units must be used. The usual units
for permeability are cm3 [S1P)crn/cm2 s (cm Hg). Since the partial pressure is
the mole fraction times the total pressure, PA = YAPtot, Eq. (12-2) can also be
written as
(12-4)
For the completely mixed system in Figure 12-7a the external mass bal-
ance at steady state is
(12-5)
FinYin = FoutYout + Fpyp
where F is the molar flow rate and y is the mole fraction. This mass balance is
often written as
(12-6)
Yin =(1- 0) Yout + 0 yp
where the cut 0 is the permeate flow rate divided by the feed flow rate.
(12-7)
645
The cut is usually one of the design parameters. With a specified, we can solve
Eq. (12-6) for Yom
Yin - a yp (12-8)
Yout = 1- a
(I2-9a)
(12-9b)
F p (1 - yp ) = P B1m
ApB [PH(1- Yom) - pdI - yp) ]
where PA and PB are the molar densities at the permeate conditions, and A is
the area for mass transfer. Dividing Eq. (12-9a) by (12-9b), we obtain
yp PA
A Yout - [~ 1 yp
(12-10)
-"--- = ClAB -,..- I:'H
----~---''-----
1- yp PB PL
(1 - Yout) - - (1 - Yp)
PH
,..
Yp PA (12-11)
--=ClAB -A-
I- Yp PB 1 _ Yin - a yp _ PL (1 _ Y )
I-a PH p
646
(12-12)
where
a= [_9 + PL
1-9 PH
1[a PB~A
AB -1]
(12-13a)
PA
b=(I-aAB)- A [
PL
-+--+-- ---
PH
9
1-9
Yin
1-9
1 1
1-9
(12-13b)
PB
A
1
A
PA (12-13c)
c=aAB -A- Yin
[ 1-9
PB
and then
The root which gives yp between zero and one is used. For ideal gases
PA/PB = 1 and the equations simplify. Alternate manipulations of these equa-
tions are shown in Example 12-1.
unit 10-10 [CC (STP) cm]/[cm 2 scm Hg] is defined as 1 Barrer. The gas is
perfectly mixed on both sides of the membrane. The high pressure feed
gas is 90 mole % C~ and 10 mole % CO2 , The permeate pressure is 1.1
atm, and the cut e = 0.5. Determine the membrane selectivity, Yp,eo" the
CO2 flux and the Cf4 flux if Lm = 1 J..UTI.
Solution
..
SeIecUVlty, Pea, 15
a.co..~ = - - = - - = 31.25. The permeate mole frac-
Pea. 0.48
tion Yp = Yp,eo, can be determined from Eqs. (12-12) to (12.14). From
Eq. (12-13),
a= r 0.5
L1- 0.5
+ Q]
20
(31.25 - 1) = 31.91
~
c (31.25) [~:! ] ~ 6.25
Where we have assumed PAIPB = 1.0. Then from Eq. (12-14)
Since O:s; yp :s; 1, use the minus sign. yp = 0.183. Now using mass bal-
ance Eq. (12-8)
Yout
= 0.1 - 0.5 (0.183)
0.5
=0.0168
648
Jeu. - [ 0.481~~0- 10
1(76) [20 (0.9832) - (1.1) (0.817)1 - 6.85 x 10-'
Notes: l.) Most of the CO2 is removed, yet permeate gas is still
mainly C~. This is important since driving force for CO2
transfer remains positive.
3.) JCll. > J co• despite much higher CO2 permeability. This
occurs because of the much higher driving force for C~
6.) Note that transfer of CO2 is "uphill" in the sense that CO2
mole fraction is higher on the permeate side. The pressure
drop allows this direction of mass transfer since
Pin,co. > Pp,co •.
,. 140
,.. 15 150 ..
E%
U E
III
120
'C02
EU ~C02 ......
~ ~
r
-, 100
,
C E
....
~ lOr tH2 100 III U
u' u%
Q
i!
=
~. 8 80
iIII
'u
-I
\ U = il-;' 60
§~ ~~ r:
1U
5 \
\
50 e.!!
i =
tEEu
Q, U
40
I!
c
~~H'
C Co CH.
0
'iii .!!!. ::.
=
= H. 02 JiI.::: 2 .. E
.~!
20
C OJ. H4 0 Cu
3.5 4.0 4.5
III ..
%CII 0
0'2.5 3.0 u
C
0 2.5 3.0 3.5 4.0 4.5
Lennard-Jones collision diameter (i) !
'-'
Lennard-Jones cOllision diameter (i)
a b
CO CH.
,
He H. O. N. C02
10,000
~-~---
Silicone
1,000 rubber /
.-- - 'V
!
Natural f
rubber I
.... ~ I.
)... '\/ L
Poly->(" \.
c
Co
styrene~ \ -- /
/I
:::. Cellulose _"- \ 1/
acetate '\ -/
"J
0.1
o 2.5 3.0 3.5 4.0 4.5
Lennard-Jones collision diameter (i)
Silicone Rubber
Defect
}-Thin Skin
Polysulfone
Support Layer
easiest to measure the ratio Pltm and to not separate the tenns. Other commer-
cial applications are discussed by Schell (1985) and Stem (1986).
A first cut at design of a gas penneator can be made once the penneabili-
ties, pressure drop, thickness of the active layer, and area per volume of the
spiral wound or hollow fiber system are known. To detennine the amounts and
compositions of both high and low pressme products the flow patterns inside
the separator and the product cut must be specified. These details are discussed
in Chapter 13.
The single stage systems shown in Figures 12-5 to 12-7 are often not
optimal and modifications of the cascades shown in Figure 12-8 are often used.
Three cascades used commercially for gas penneation are shown in Figure
12-11 (Spillman, 1989). The two-stage process with recycle shown in Figure
12-11a is used to improve recovery of the slower penneating species at the cost
of additional compression and membrane area. If the concentration of the per-
meating species is very high, the "premembrane" shown in Figure 12-11b can
be useful for bulk removal. According to Eq. (12-3) a low penneate pressure
will decrease membrane area or increase product recovered; however, if
recompression is required a low penneate pressme increases recompression
costs. One compromise is the two stage process shown in Figure 12-11c where
the penneate pressures are different. Since gas compression is done in stages,
the two penneate pressures are chosen to match the compression scheme.
Membrane
~~~~~~--~8 E
Feed
Membrane
Corr4:lress
Permeate
b
Bulk Removal Ptrification
Permeate
c 60% Hydrogen
20% Nitrogen _ _ _ _ ....-,
20% Hydrogen
20% Inerts r-
,--_ _ _ - , 4;~%N:~r~r~:n
2,000 psi
1,000 psi
gas composition and flow rates, temperature and pressure, and the required
outlet gas composition for both high and low pressure gases. The manufacturer
then designs and builds a system which he guareentees will meet the required
specifications. Unfortunately, this arrangement is not as full-proof to the pur-
chaser as it sounds. If the actual operating conditions are different than
specified, the guareentee may be voided. In addition, the supplier will tend to
over-design to be sure he meets the specifications. Finally, tum-key units tend
to be more expensive than systems where the buyer does the design. It is best
if the buyer knows as much as possible about the separator even when he is
buying a tum-key system. This allows the buyer to write a more advantageous
contract and to do a better job bargaining. Additional factors which can be
important in design are discussed in Chapter 13.
The economics of gas permeation separation compared to competing
processes is discussed in detail by Spillman (1989). He concludes:
definition is useless unless we know what osmosis is. Thus we need to digress
slightly into the thennodynamics of membranes.
(12-16)
Jll.solvent = Jl2,solvent
Note that the equilibrium condition is on solvent only. Since solute does not
pass through the membrane, solute ~s not in equilibrium. Thermodynamic
arguments can now be used to derive equations which allow calculation of the
osmotic pressure (for example, see Reid, 1966). Assuming that the liquid is
incompressible and that activities can be estimated from vapor pressure meas-
urements, the result is
ltVsolvent
VPpure solvent
= TIT In [ VP .
1 (12-17)
soluuoo
where vsolvent is the partial molar volume of the solvent Equation (12-17) is
usually quite accurate, but is not always convenient to use since it requires
656
7t in atmospheres
(12-18a)
7t=cRT
(12-18b)
7t=ac
(12-18c)
The osmotic pressure of solutions is not widely tabulated and vapor pres-
sure data may not be available. Fortunately, the osmotic pressure can be
related to freezing point depression which is easy to measure and is widly tabu-
lated (e.g. in Handbook of Chemistry and Physics). The osmotic pressure is
related to freezing point depression by (Reid, 1966),
(12-19)
K.olv
Jso1v = --(.1p - .11t)
(12-20a)
1m
where K.olv is the permeability of the membrane to the solvent (usually water),
.1p is the pressure drop across the membrane, and .11t is the osmotic pressure
difference across the membrane. Note that the pressure driving force is
reduced by the osmotic pressure difference. As the solution becomes more
658
I
2. To find productivity at any other concentration i ! I
---
-~I--
.. __
and pressure, multiply data in Table I by the ~T-~
J~_
._.
l.-
appropriate factor
..J.--
---
2.0
--
II:
I
~
·. V V
~ ~ -~
..
N()CI leeO c.-- l--
.-- -
l-- V
f-
--
>-
I-
....--V
:;
t .\,'!l~: ....
.- ... I
---
.. V
...
\O,~ ...-::.
~
o k:
V k-
~
~k
o ..
II: . r:: j:; . '
..- i-"'""
1.0
p. .
Q.
..
~ :. I
~~
009911\
L.;..rl ... 'l:,
00 . ;::;...,.-- V
~., .... . " ,. . .. 11\
I--
:.;- ~ ~
·
..
--
...
L--I-
,. t'
'!)o,oo~j...--
• I • •
..
"-j • 1" "
~.
V~ r-
k ~ l-- ttO~
>-:: ...
...
.. "1'
· t~ •• , t
V
o
400 500 600 700 800
PRESSURE, psig
(b)
-= =
_. --,+ . . . ' ....... -,,:, ---'-,. .
~.:: :-::-: HI. Reference is 800 psig,' 30·/~ recovery
~
,
. ... . .. land 30,000 ppm NaCI feed. i-- h---,' .~.~. --·'~--lf~· I~
~I'-=:~: I~2. To find salt passage at any other can- I-+Lt::t=--.:We-:-- •::=;
~ 2.0 1---., -;w .J centration and recovery, multiply data - '. - 40 r'jt:-t:-:::.:. ..
t; ~= ~ ~ in Table 1 by the appropriate factor. ~f-W+. ~:.. Jyt::~ --:-.
~ :::.... ::-.: h ++T ::tixf-:;:;=:-:- _ . ::::.f:-'-"' -+j- ~,Itil: :-:.I.::-::;;II::r:::
-./
~ t:::""-R::- !+H=t-:H-'FH+t~ ~¥ - ;:;- : +,(--= .:: j/': 3oo4=lr :-":::
~_ ~ -+-4-+1-+-+-+
r=;=1---t+.:...J.
. ,.. ~. fEf,,+t. ;-~f=I '.~" -I--~.-. V_'
:::. .
_
:
Q. 1--. !=h.
Real membranes are partially permeable. Thus there will be some solute
flux through the membrane. Neglecting concentration polarization, the flux of
solute is
(12-21)
where Ksolutc is the permeability of the membrane to the solute, and c p is the
permeate concentration. Conservation of mass requires that
(12-22)
which relates the two flux equations. Experimentally the rejection coefficient
(12-23)
(12-26)
As M increases the solute flux increases since any solution which leaks through
the membrane is more concentrated.
(12-27)
661
Note that this predicts that solvent flux decreases as M increases. Combining
Eqs. (12-22), (12-23), (12-25) and (12-27), we obtain an expression for the
rejection coefficient when there is concentration polarization.
(12-28)
Solution
A. Define. The problem is clearly defined in the problem statement.
B. and C. Explore and Plan. Although the definition is clear, the path
to solution is not. We obviously need to determine osmotic pressure.
Equation (12-17) is accurate. Vapor pressure data (perry and Green,
1984), p. 3-73) at 30°C is:
From the pure water flux we can find Ksolv/t.n from Eq. (12-20a). Then
for part a J so1v is obtained from Eq. (12-20a) when M = 1.0 and from
Eq. (12-27) when M = 1.2. For the real membrane with M = I, cp can
be found from Eq. (12-23). Then Jso1v is determined from Eq. (12-20a).
We can also use the measured value of R to find a = K.olvlK.olute from
Eq. (12-24a). This gives us K.olute/tn,. This value is needed when M =
1.2. Combining Eqs. (12-22), (12-25) and (12-27) we can solve for Cp.
Once Cp is known Eq. (12-27) is used to find Jso1v .
c
x = 0.5 (27.2) = 544.5c, or a = 544.5
Ksolv
When M = 1.2, J solv =- - [~p - M 1t (Cb)]
fro
J so1v :;;:; (60) [49 - (1.2) (27.2)] =979.9 L/m2 day
Pan b. Real Membrane. For M = 1 we observe c p = 0.005. Rearrang-
ing Eq. (12-23a), we obtain
(12-20b)
Note that solvent flux of the real membrane is greater than that of the
ideal membrane since ~1t is decreased.
Ksolv R (12-24c)
a = Ksolute = -:-----:---:-:---=:-:-
(~p - ~1t) (1 - R)
For M = 1.2 we need to solve Eqs. (12-22), (12-2S) and (12-27) simul-
taneously. The resulting equation is
(12-29)
K.olv ( A
- - LlP-
M 1t (» Ksolute
Cb + - -
t.n t.n
= 60 [49 - (1.2) (27.2)] + 163.3 = 1143.2
and
K.olv
- - - M Cb = - (163.32) (1.2) (O.OS) =- 9.8
1m
E. Check. Can use Eq. (12-22) as one check. For Part b with M = 1.2
(the most difficult problem),
1. Note that the units on the gas constant R controls the units of 1t.
2. The value K.olv/lm determined from data can be used under dif-
ferent conditions.
7. If we do not have perfect mixing, Cb and c;, will both vary. Thus
1t (Cb) and 1t (c;,) vary which means the flux varies as we go down
the membrane. Mass balances for other situations are illustrated in
Section 13.3.
(12-30)
(12-31)
(12-32a)
Since D/o = k, the mass transfer coefficient, this result is often written as
M=-=exp
CW
Cb
-S01V
- [J 1
k
(12-32b)
3. The diffusivity increases. This shows that large solutes will have
worse concentration polarization.
The spiral wound and hollow fiber configurations are commonly used for
RO. The common membrane polymers include cellulose acetate, polyamides
(see Figure 12-3), and a composite membrane developed by FilmTec. The
composite membrane is a crosslinked polyamide barrier layer supported by a
microporous polysulfone coating on a JX>lyester web. This complex membrane
has high flux, high salt rejection, a pH range from 4 to 11, and a wide range of
temperature stability. Other membranes are listed in Figure 12-3 and Table
12-2.
The skin of the RO membranes have pores which are several Angstroms
or larger in diameter. These pores are larger than the hydrated ions, yet the
ions are excluded. The separation is clearly not a sieving effect. The current
physical model of the separation effect can be summarized as follows (Kesting,
1985). The water molecules are preferentially attracted to the hydrophilic
membrane surface. An ordered water sheath of thickness t (- ~) exists at the
membrane surface and repels ions. This water sheath pours into pores. If the
pore is less than 2t (- 10~) in size the ordered water sheath will be continuous
and ions will be excluded. In the occasional large pores (imperfections) with
diameter > 2t, the ordered water sheath does not fill the pore. In the center of
these pores ordinary water plus ions pass. The observed rejection coefficient is
the sum of the almost pure water from the small pores and the contaminated
water from the larger pores. These membranes are often called "solution-
diffusion" membranes.
0\
0\
00
Table 12-2. Rejection coefficients and fluxes for a variety of RO membranes (Cadotte et al, 1988; Pusch
and Walch, 1982). HF = hollow fiber, PF =plate and frame, SW = Spiral wound.
0-
0-
\0
670
Permeate
Product
Storage
F
Pressure Bypass
Control
Valve Waste Water
to Disposal
penneators with bypass lines is shown in Figure 12-13. The bypass lines are
used to mix in raw water if less pure water is required. The water which does
not pass through the membranes is more concentrated in the dissolved solids.
Usually, this stream is a waste stream which must be disposed of. The product
water is sent to a storage tank where additional chemicals may be added. For
example, if the water is for drinking chlorine or ozone will be added here.
Chlorine is detrimental to the membranes and thus is added after the mem-
branes unless it is required to stop biological growth. When treating highly
fouling materials such as food products, a cleaning cycle must be added. This
requires shut-down of the equipment, and it requires quite a bit of hardware in
addition to that shown in Figure 12-13. Details of cleaning cycles and the
chemicals used for cleaning are given by Eisenberg and Middlebrooks (1986).
Unfortunately, this can greatly increase the costs.
Membranes have a finite lifetime which depends upon the operating con-
ditions. Membrane replacement is usually treated as an operating cost. In
addition, there is usually a long tenn flux decay caused by membrane compac-
tion and fouling of the membrane. To take this into account the initial flux of
the membrane is usually not used in the design calculations. Instead the flux
after a few months of operation is used since the additional decay after this
time is quite slow. Obviously, the membrane must be protected from chemical
upsets since the membranes can be destroyed quite quickly.
the economics of the particular case. The economics of RO are often favor-
able, and RO is used very extensively for treating brackish water, seawater, and
tap water. One of the largest RO units is at Ras Abu Janjur, Bahrain, which
produces 12.7 million gallons of potable water per day from brackish well
water. The system consists of seven trains of hollow-fiber RO units with 2100
membrane modules per train. RO applications and economics are discussed by
Belfort (1984), Cheremisinoff and LaMendola (1983), Applegate (1984), and
Eisenberg and Middlebrooks (1986).
Reverse osmosis has also been used for separation of organics from
water. Although this is not yet a common industrial separation, it may be
important in particular cases. For example, dilute £-caprolactam solutions are
concentrated commercially using RO. If the organic recovered dissolves poly-
mers or if the module needs to be steam sterilized, ceramic membranes can be
used (Hsieh, 1988). Data is presented by Eisenberg and Middlebrooks (1986),
Leeper (1986), and Sourirajan (1977).
Ultrafiltration (UP) is used for retaining larger solutes than in RO, but many of
the operating characteristics are very similar to RO. The osmotic pressure in
UP is invariably quite low since the molar concentration of the high molecular
weight molecules separated by UP is quite low even when the weight concen-
trations are high. Because a large osmotic pressure does not have to be over-
come, much lower operating pressures are used in UP than in RO. Typical
operating pressures are in the range from 10 to 100 psig.
The membranes used in UP are usually asymmetric polymers used in one
of the geometries shown in Figure 12-5. A variety of polymers such a cellulose
acetate, aromatic polyamide, polysulfone, polyvinyl chloride, polyacrylonitrile
and polycarbonate have been used to cast membranes (see Figure 12-3).
Polysulfone has a wide temperature tolerance (to 93° C) and a wide pH range
(0.5 to 13) and is probably the most commonly used membrane material
(Applegate, 1984). The membranes are formed so that the dense skin is
microscopically porous. Thus the separation mechanism can be considered (in
a somewhat over-simplified form) to be sieving of the molecules. The mem-
branes can be tailor made to fix the desired pore size distribution and thus pro-
vide for different molecular weight cutoffs. The molecular weight cutoff is
loosely defined as the molecular weight below which species begin to pass
through the membrane. The values for molecular weight cutoffs can range
from 1000 to 80,000. Values for a variety of membranes are listed in Table
12-3. Molecular weight cutoffs should be employed with discretion since siev-
0\
--.I
~
Table 12-3. Molecular weight cutoffs and fluxes ofUF systems. (pusch and Walch, 1982; and
Warashina et ai., 1985). HF:;: hollow fiber, PF:;: plate and frame, SW:;: spiral wound
0'1
-.I
VI
676
08
.: 06
>
c
Q)
~ 04
02
, Nominal molecular welCJht
exclusion of membrane
Molecular welCJht
Figure 12-14. Solute retention on Amicon Diaflo membranes (porter, 1979).
Reprinted with permission from P.A. Schweitzer (Ed.), Hand-
book of Separation Techniques for Chemical Engineers (1979).
Copyright 1979, McGraw-Hill.
ing depends on size and shape of the hydrated molecule, and molecular weight
is only a rough guide to size and shape. In addition, there is considerable reten-
tion of molecules of lower molecular weight. Roughly, a sharp cut-off UP
membrane will have R = 0 when the molecular weight is 1(2 the molecular
weight which has R = 0.9. Diffuse cutoff membranes have a much broader
range of fractionation. Typical retention data are shown in Figure 12-14
(porter, 1979). A comparison of Tables 12-2 and 12-3 is interesting. Note that
the UP membranes with their much larger pores have considerably larger fluxes
at lower pressures than the RO membranes. Details of UP membranes are dis-
cussed by Blatt (1976), Cooper(1979), Kesting (1985), Lloyd (1985), Porter
(1979), and Pusch and Walch (1982).
For the large molecules separated by UP the molar concentration will be quite
low even if the weight concentration is high. Thus, the osmotic pressure differ-
ence ~1t will be small and is usually ignored. The resulting solvent flux equa-
tion is,
Kso1v
J I01v = - - ~p
(12-33)
tm
For small and intennediate macromolecules (5000 or 100,000 daltons) it may
be necessary to include the osmotic pressure tenn. This is discussed further in
Chapter 13.
For sieve type membranes flow of both solvent and solute is in pores.
Let R represent the fraction of solvent flux carried by pores which exclude the
solute. Then l-R of the solvent flux is in pores which do not exclude solute.
With no concentration polarization the liquid passing through these pores will
have a solute concentration Cb. This gives a solute flux of
(12-34)
In addition, Eq. (12-22) must remain valid since it represents a mass balance.
Combining Eqs. (12-22), (12-33) and (12-34) and solving for R, we obtain
Eq.(12-13). Thus in sieve type membranes the solute rejection can also be con-
sidered as the fraction of solvent flux in pores which reject solute molecules.
(12-35)
J I01uIe =MCb (1 - R) J I01v
678
y=8
cp (12-36)
R
app
= 1- -
Cb
=1- M (1 - R)
l\p
J lalV = --.-!;....-- (12-37)
_tm_+_ls
Kaalv Kg
679
where tg is the thickness of the gel layer and Kg is the permeability of the gel
layer. Since 19 can vary throughout a run, Eq. (12-37) is usually not useful for
predictions. Often, the gel layer will have far more resistance to flow than the
membrane and the gel will control the solvent flux. In these cases it does not
matter what type of membrane is used as long as it retains the solute and causes
a gel to form. The value of Kg can be estimated from the Carman-Kozeny
equation (Blatt et ai, 1970).
(12-38)
Note that the gel permeability will be very sensitive to the particle diameter ~
in the gel layer. If e=O.5 and the viscosity is 1.0 cp, then Kg = 3xlO-ll for one
micron particles and Kg = 2x10- 16 for 30 Angstrom particles. Thus a one
micron layer of these gels with a pressure drop of 100 psi across the gel would
have solvent fluxes of 200 and 0.0013 cm 3 /cm 2 sec, respectively. Obviously, a
gel layer can have a major effect on the solvent flux.
(12-37)
Since cg and Cb are constants, the solvent flux is the variable. This is different
from Eq. (12-32) where Cw was unknown. The solvent flux is set by the rate at
which the solute can diffuse back from the gel layer to the bulk fluid. Increas-
ing the pressure drop will cause a brief temporary increase in the solvent flux.
However, during this period of increased flux transfer of solute to the gel layer
is greater than the rate of transfer of solute back into the fluid. The gel layer
must then become thicker. According to Eq. (12-37), J I01v will then decrease.
This decrease in J I01v continues until the gel layer is thick enough so that the
solvent flux has obtained the value predicted by Eq. (12-29). In other words,
680
the engineer has no control over the flux rate except by increasing the dif-
fusivity (increase temperature) or decreasing ~ (stir or use thin channels). For
high molecular weight polymers diffusivities are quite low and the solvent flux
is low if gels form. The mass transfer coefficient k can be estimated for many
flow geometries. This analysis is done in Chapter 13.
Experiment a: Run pure water and obtain Jsolv = 7000 L/m2 day with
~p= 5 atm.
Solution
a. From the pure water data we can find Ksolv/1m from Eq. (12-33).
~. n3 L
k = In (Cg/Cb) = In (0.14/0.02) = 119.74 m2 day
Mter a short period of time, tg increases and J so1v is given by Eq. (12-
39). Since k, c g and Cb are unchanged from part a, J so1v = 233. Solving
for tg/kg
Note that in both parts the resistance of the gel, fglKg, is significantly
greater than the membrane resistance, tmlKm. This is commonly the
case when proteins are ultrafiltered.
Although simple and very appealing, there are significant problems with
the gel layer model (Fane, 1986; Porter, 1979; Wijmans et aI., 1984). The con-
centration cg found by extrapolation may depend upon equipment geometry.
The measured cg may be too low and independent experiments show no gel-
ling. In addition, c g may be estimated as > 100% which is physically impossi-
ble. The mass balance, Eq. (12-31) does not include convection in the axial, x,
direction and the boundary condition at y = 5 is hypothetical. However, the
model is useful for correlating data. Modifications of this model are discussed
in Chapter 13.
The presence of a gel layer may have conflicting effects on the rejection
of solutes. If the gel layer is very viscous, but remains fluid so that molecules
682
diffuse in the layer, solute molecules will continue to pass through the mem-
brane. Since Cw>Cb, the fluid passing through pores which do not reject solute
will be concentrated. This will cause the apparent rejection coefficient to drop.
This situation is similar to concentration polarization without a gel layer. The
second case occurs when the gel layer has a solid-like nature and solute
molecules are essentially immobile. Now the rejection coefficient can increase
since the gel layer traps the solute molecules and very few of them can diffuse
into the pores.
The fonnation of a gel layer also has a large effect on the separation of
mixtures containing several solutes of widely different size. Suppose we wish
to separate a low molecular weight compound from a very high molecular
weight polymer. When run separately, the low molecular weight compound
has R=O while the polymer has R=l. What happens when the mixture is
ultrafiltered? Our first guess is probably to use superposition and assume each
molecule is unaffected by the presence of the other molecule. If no gel layer is
fonned, this assumption is probably close to true. Then the small molecules
have R=O and the polymer has R=1. The separation is easy. If a gel layer
fonns, the gel serves as a membrane which has different retention characteris-
tics than the original membrane. The gel layer may retain the small molecules.
If this happens both solutes may have retentions near 1.0 and separation will
not be obtained. Thus gel layers change the fundamental nature of UF.
Once fonned, the gel layer may not come off easily. In this case the
membrane is said to be fouled. Fouling often occurs when the solute is
adsorbed to the membrane or when charges OR the membrane attract colloidal
particles of opposite charge. The entire UF installation is often controlled by
the presence of gel layers and fouling. Fouling can also occur from precipita-
tion of salts, and is discussed in detail by Fane (1986) and Potts et al. (1981).
The batch system shown in Figure 12-8g is easily analyzed for simple
cases. A mass balance on the entire system (tank + pipes + UP module) is
dV (12-40)
-=-AJ
dt 1
SOY
Where V(t) is the volume of liquid in the tank + pipes + UP module. If osmotic
pressure is negligible and no gel forms, Jso1v is given by Eq. (12-33). Then
(12-42)
If a gel forms then J. olv is a constant given by Eq. (12-39). Now Eq.
(12-40) becomes
dV (12-43)
- = -A k In(c ICb)
dt g
Since R = 1, the total moles of solute in the system, n, must remain constant
(12-44)
n = c(t) V(t) = constant
dV (12-45)
""(it = - A k In (cg VIn)
+ 1
(3)(3!)
[(In V)3 - (In V )3] + ...
o = -A kt In (cg/n)
(12-46)
(12-47)
685
dV
dt
= _ A K.olv l\p
1m
[1 - [al\pMn 1~V 1 (12-48)
with the initial condition V = V0 when t = 0. The solution to Eq. (12-48) is (see
Problem 12-C4),
V _ V _ a Mn In aMn
vo - - -
l\p
[ V _ a Mn
1= _ A
K.olv
l\
P t (12-49)
o l\p 1m
l\p
Note that when a = 0, Eq. (12-49) reduces to Eq. (12-42). Equation (12-49) is
also valid for batch operation of RO when R = 1.
Ultrafiltration and reverse osmosis systems are often used together. For
example, cheese whey is the left over liquid from making cheese. The whey is
approximately 1% protein, 5% lactose (milk: sugar), 1% salts, and the
remainder is water. The whey has a very high Biological Oxygen Demand
(BOD) and is expensive to dispose of properly. One processing route which
utilizes all of the components of whey is shown in Figure 12-16. The protein is
recovered by UP. This protein concentrate is quite valuable and can either be
dried and sold or it can be added wet to cottage cheese or ice cream. The per-
meate from the UP system is sent to RO where the sugar and salts are concen-
trated by removing water. This is desirable since it makes both the downstream
processing steps smaller and cheaper. The concentrated solution can then be
fermented with yeast to produce a dilute ethanol solution. The ethanol is
recovered by distillation or other separation scheme. The waste liquor contain-
ing about 3% salts is sent to evaporators to recover the salts which can be used
as fertilizer. Note that the UP system is first. The UP system serves to protect
the RO system from particulates and protein which can cause fouling and gel
formation. Both UP and RO systems would be cleaned at least once a day.
F
r====t-- Permeate
(Water)
Storage
Fermentation
Tanks ''----95wt%
Ethanol
Disti Ilation
Water Vapor
Evaporators
L---=r- Stea m
Conc.
Salt
Solution
Figure 12-16. Simplified flow diagram for whey processing.
processing fruit juices (Stana, 1977). In concentration of fruit juices the pro-
cessor wants to recover all of the pulp, the sugar, and the volatile flavor
ingredients. The final product may be from 20 to 25% pulp and 20 to 25%
sugars. This product is very viscous and has a very high osmotic pressure.
(Open a can of frozen orange juice and let it melt. This is the desired final pro-
duct) A high recovery of sugar is desired; thus, the permeate leaving the RO
system should be close to zero concentration. A 50 wt% solution of sugar has a
very high osmotic pressure (see Table 12-1), and any leakage through the
membrane will cause significant contamination of the waste stream. Thus a
single RO module will be difficult to use and the membrane arrangement used
in Figure 12-16 for whey processing is not a good scheme. Instead, the pro-
cessing scheme shown in Figure 12-17 can be used (Stana, 1977). Note that
687
Concentrate
40-50%
Pulp
Raw
Juice
UF
the UF system is again placed first to protect the RO systems. Since the UF
membrane does not retain the sugar, the sugar concentration is the same in the
permeate and the high pressure products. Thus, although 7t is significant, d7t is
approximately zero and the usual UF equations can be used. The first RO sys-
tem produces the desired low concentration waste water, but the sugar solution
is concentrated to only 25%. This keeps M reasonable. The second RO sys-
tem produces the desired sugar solution, but the permeate is a 10 to 15% solu-
tion which is recycled to the first RO system. This keeps d7t modest in the
second RO system also. The pulp and sugar are then mixed to produce the
desired final product.
12.6. DIALYSIS
be in the opposite direction from the solvent flux. Dialysis also usually does
Rot use asymmetric membranes, but instead uses homogeneous membranes
without skins.
The major commercial use of dialysis is hemodialysis for removal of
waste products such as urea and creatine from patients who have had kidney
failure. This application is commonly known as the artificial kidney even
though hemodialysis is not a good mimic of the way the kidney functions (Col-
ton and Lowrie, 1981; Ward et ai, 1985, Lysaght et ai, 1986). Other commer-
cial applications include the recovery of caustic in the manufacture of rayon,
salt removal in pharmaceuticals manufacturing, and recovery of spent acid in
the metal industry. Except in medical applications, the engineer is more likely
to be involved with RO, UP, or electrodialysis than with dialysis. Both plate-
and-frame and hollow fiber dialyzers are available (Lacey, 1979; Klein et ai,
1987).
As small solutes pass through the microporous membrane, concentration
gradients will develop on both sides of the membrane. This is illustrated in
Figure 12-18. The flux of the permeable solutes through the membrane can be
written as
Ksolute (12-50)
J solute = (c m h - cm '!)
lm .
where Cb and Cd are the bulk fluid concentrations on the feed and dialysate
sides, and k~lute is the total or overall mass transfer coefficient. If the solvent is
689
Flux of small
--l-+- molecules
- - - c b Small molecules
Dialysate
Side
Feed Side
- - - - - Macromolecules
Membrane
Figure 12-18. Concentration gradients in dialysis.
stagnant, the overall mass transfer coefficient can be found from a sum of resis-
tances.
1 1m 1 1 (12-52)
--- +--+--
k;olute - Ksolute k~olute k~olute
where k!olule and k~lute are the mass transfer coefficients on the feed and
dialysate sides of the membrane. The individual mass transfer coefficients can
be estimated from standard correlations such as those in Chapter 13. The three
terms on the right hand side of Eq. (12-52) are often the same order of magni-
tude.
Solvent flux (see Eq. (12-27» can be in either direction in Figure 12-18.
If the pressures on both sides of the membrane are equal, the solvent will flow
from the dialysate to the feed side because of osmosis. This osmotic flow can
be drastically increased by the concentration polarization of macromolecules,
and in turn will tend to reduce the concentration polarization. Unfortunately,
the osmotic flow also decreases the flux of the small solutes since the convec-
tive flow is in the opposite direction to the solute diffusive flux. The solvent
690
flow can be stopped or reversed by applying a positive pressure on the feed side
of the membrane. Dialyzers are often operated with essentially no solvent flux,
or with a small flux of solvent into the dialysate.
Any of the flow patterns shown in Figure 12-7 can be used in a dialyzer.
Hemodialyzers are usually countercurrent If the volumetric flow rates of
liquids on the feed <IF and dialysate sides Qd and the overall mass transfer
coefficient kT can be assumed to be constant, a variety of steady-state flow case
can be solved (Michaels, 1966). The correct average concentration difference
in Eq. (12-51) is the logarithmic mean of the inlet and outlet concentration.
(12-53)
where 1 and 2 are the two ends of the dialyses. For example, if countercurrent
flow is used,
(12-54a)
(12-54b)
Solving Eqs. (12-51), (12-53) and (12-54) with the mass balance
(12-55)
(12-57)
691
In the derivation of Eq. (12-57) the volumetric flow rates Qp and Qd were
assumed to be constant. At equilibrium, cP,out = Cd,in' Thus the denominator on
the LHS of Eq. (12-57) is the maximum possible concentration change in the
feed stream while the numerator is the actual concentration change. The LHS
of Eq. (12-57) is thus the fraction of maximum solute removal which is
attained.
Results for other geometries are also easily obtained Michaels (1966).
For parallel flow of feed and dialysate the result is,
=
1 + Qp/Qd
If the dialysate is completely mixed and the feed is in plug flow, the
result is
(12-59)
pass amino acids and salts which it is desirable to retain. This latter problem is
solved by using a dialysate fluid which is a physiological electrolyte solution
containing the appropriate concentrations of the salts. Since salt concentrations
are the same on both sides of the membrane, there will be no transfer of the
salts. In hemodialysis a positive pressure is applied to the feed (blood) side in
order to remove some water (which can be considered the solvent). Equipment
used in medical applications must satisfy a large number of constraints such as
preventing blood clotting. Details of hemodialysis are discussed by Colton and
Lowrie (1981), Lysaght et al (1986), and Ward et al (1985). Other membrane
separations used in artificial organs are discussed by Lysaght et al (1986).
In the usual dialysis process exclusion of large solutes is based on steric
effects. In ion exchange dialysis anion exchange membranes are used. The
membranes are the same as those used in electrodialysis (see the next section).
These membranes will exclude cations (ions with positive charge), but the
exclusion of H+ cations is poor (see Chapter 9 for a discussion of Ion
Exchange). Thus, if a solution of acid and metal salts is on one side of the
membrane and water or a dilute acid solution is on the other side of the mem-
brane, the anions (ions of negative charge) readily transfer through the mem-
brane. To keep electroneutrality cations must also transfer. Since the only
cation which can transfer through the membrane is H+, the acid is separated
from the metal ion. This is illustrated in Figure 12-19a. Ion exchange dialysis
is used commercially to recover metal ions in the plating industry. The systems
use plate-and-frame arrangements similar to those used in electrodialysis
except no current is required. Water and feed solutions are alternated in the
different compartments. Other geometries could also be used.
Donnan dialysis is another variant using ion exchange membranes. The
stack utilizes all cation or anion exchange membranes. For purposes of this
explanation we will assume that all of the membranes are cation exchange
membranes since our purpose is to concentrate a valuable cation. A dilute
solution containing the valuable cation (for example Cu++) is circulated in the
odd-numbered compartments. In the even-numbered compartments a concen-
trated solution of cheap acid is circulated. One cell is illustrated in Figure 12-
19b. The W ions will transfer through the membranes due to the concentration
driving force. To keep electroneutrality, either the anion must transfer through
693
Acid
* +
! CuS0 4 acid
l
Metal
Salt
Anion conc.
l
+ dilute CUS04
H+
H2 S0 4
Metal
Salt Acid
t \ Cation
\
Exchange
~ Membrane
* Anion
Exchange
Membrane
the presence of the electrical field the cations will migrate towards the cathode
and anions towards the anode. The cations can pass through the cation
exchange membranes but not through the anion exchange membranes. The
anions can pass through the anion exchange membranes but not through the
cation exchange membranes. Because of the alternation of the cation and anion
exchange membranes the result is to concentrate ions in the odd numbered
compartments and to dilute the ions in the even numbered compartments in
Figure 12-20. In this way a desalted water and a concentrated brine are
formed. In commercial systems from 100 to 600 unit cells (pairs of mem-
branes) are in a stack. Reviews of electrodialysis are given by Applegate
(1984), Hwang and Kammermeyer (1975), Klein et al (1985), Komgold
(1984), Lacey (1979), Mintz (1963), Solt (1976), and Spiegler (1984).
used in special cases. Ion exchange membranes are discussed in detail else-
where (Kesting, 1985; Flett, 1983; Klein et al, 1987; Komgold, 1984; Lacey,
1979; Pusch and Walch, 1982; Solt, 1976; Strathmann, 1985). Lists of sup-
pliers and membrane properties are given by Komgold (1984), Spiegler (1984),
and Strathmann (1985).
(12-60a)
(12-60b)
(12-60c)
The anode can be made from the same materials as the cathode or from stain-
less steel. The anode and cathode compartments are separated from the stack
by membranes. These compartments are vented to remove the gases fonned.
Since the stack is in series, the current flow through each membrane and
each compartment must be the same. The reason for using a large number of
unit cells is to use this current many times and thus produce more desalted
water. The current is carried by the ions. The fraction of current carried by an
ionic species is the "transference number". This fraction can be different for
the different ions since current is the product of the charge times the velocity
times the number of ions transferred. Small ions such as W will carry more
current than large ions such as Cl- since the H+ have a higher velocity. In a
solution such as KCI where the ions are approximately the same size and thus
have the same velocity, the transference numbers will be approximately equal.
696
cations onions
I I
I Well Mixed 1
I Zone I
1 I
1 1 1 1
18 81 18 81
1 1
Ion Concentration
I I
I I
Cation Anion
Exchange Exchange
Membrane Membrane
(12-62)
E=J2nRt
698
(12-65)
where R(CF) is the resistance per cell determined at the feed concentration. For
example, if we recover a product water which is one half as concentrated as the
feed this result is
where n is the number of unit cells, 11, is the efficiency due to the semipennea-
bility of the membrane (co-ions transfer through a membrane that they should
be excluded from), l1w is the efficiency related to water transfer through the
membrane, and 11m is the efficiency since some of the current invariably leaks
through the manifold holding the membranes. All of the efficiencies are less
than 1.0 and the overall electrical efficiency is often about 0.9. However, n
often varies from 100 to 600. Thus, ~ will usually be significantly greater than
1.0. An important exception to this occurs as the feed concentration becomes
high (3 to 5 mol/liter). As this happens 11, and hence Elec. Eff. approach zero
since the coions are not strongly excluded from the membrane. At these high
concentrations ~ becomes small and energy requirements become too large for
electrodialysis to be economical.
The membrane area A for the cation and anion membrane can be
estimated from
(12-68)
(12-69)
A=nAm
The area Am is set by the dimensions of the plate-and-frame module which are
standardized by the manufacturer. Substituting in
(12-70)
(12-71)
n=
1115 11m l1w
Solution
a Rearranging Eqs. (12-66) and (12-67),
- g [ 1 equiv ]_ eq
CF - Cdil - (1.2 L) 58.45 g - 0.0205 L
= 0.920
Assuming that cell resistance is inversely proportional to concentration
Eq. (12-65) is valid. Rearranging,
Since there are 300 Ljmin, this is 1412 joules/L = 337.5 callL.
3.) If the cell design and size or spacer design are changed,
R and Elec. Eff. may not be the same for laboratory and
commercial units. Also, if the high concentrations are
very different 115 may be different in the two units.
703
12.8. PERVAPORATION
VP Vacuum
2
---p
liquid side and dry on the vapor side, the diffusivity varies markedly across the
membrane. The evaporation at the vapor side may have a major effect on the
observed selectivity. For the separation of trace organics from water, the rela-
tive volatility of the trace organic is often very high even though the boiling
point of the organic is higher than water's boiling point. This occurs because
these are very nonideal solutions with high activities. Finally, because of eva-
poration the process is nonisothermal. The temperature change affects solubil-
ity, diffusivity, and evaporation. Currently, a complete model including all
these effects is not available.
(12-72)
Since flux is proportional to dcJdx, the steady state form of Fick's second law
becomes,
(12-73)
(12-74)
'1 = '1,1 at x=O , Ci = Ci.2 at x=tm
(12-75)
D(c) =Do exp (ac)
705
The parameters Do and a are functions of temperature and the properties of the
polymer.
(12-76)
where S* is the Henry's law constant and is a form of solubility. Then the flux
is
(12-77)
Note that Eq. (12-77) does not fit the standard form given in Eq. (12-1). If we
force fit pervaporation into the form of Eq. (12-1) using ~P as the driving force,
we obtain a "permeability" P of
J1m
P= - = -
Do
[exp(aS "Pl)-exp(aS "P2)]
(12-78)
~P Mp
This result shows that the permeability is a function of the partial pressures, in
addition to the diffusivity and the solubility. Since both solubilty and dif-
fusivity usually increase as the temperature increases, the permeability and per-
meation rate increase with temperature increases. The selectivity between dif-
ferent species may decrease as temperature increases, although for commercial
alcohol-water pervaporation there is little decrease in selectivity (Sander and
Soukup, 1988).
(12-79)
Pz = YiP
The flux and concentration profiles are assumed to be unchanged by the pres-
ence of the other diffusing components. Unfortunately, superposition is seldom
valid. The presence of a second component affects the permeation rate of the
706
30
25
.s
u
20
...."
c:
-2 15
"<-
"
0-
Il
til
o 10
o
Weight fraction of benzene in liquid
o = permeation 40° e • =equilibrium 40 0 e
t:l. = permeation 5Qoe • = equilibrium 50° C
o = permeation 60 0 e • = equilibrium 60 0 e
If selectivity and flux are known from laboratory data, we can design the
707
-
z-Oyp (12-80)
Xout - I-a
(12-81)
(12-82)
708
Equations (12-81) and (12-82) apply to local values at the same location of the
membrane. For a perfectly mixed system
(12-83)
yP =Y , Xout = X
(12-84)
yP = 1 + (ClAB
, - 1) Xout
The global character of this expression for perfectly mixed systems greatly
simplifies the remaining analysis.
(12-85)
This equation can obviously be solved for yp with the quadratic formula (see
Example 12-5). Equation (12-85) is linear in z and 9; thus, if yp and 9 are
known
(12-86)
Cl~ (Cl~ 1) z Yp - Yp
9= (12-87)
Z- -
, 2
(ClAB - 1) (Yp - yp)
These equations are obviously valid and easy to use if Cl~ is constant.
However, Figure 12-23 shows that Cl~ is a function of liquid mole fraction x.
The equations are still valid, but must be used in a trial-and-error fashion. For
709
1. Guess Xout.guess
The value of the cut is often controlled by the energy balance. For the system
shown in Figure 12-24 the energy balance is,
(12-89)
which is
(12-90a)
or
(12-90b)
The high temperature, Tin, is limited by the stability of the membrane. The low
temperature, Tout = T p' is limited by the need to have a vapor on the penneate
side. Thus, if T p is decreased a very low pressure may be required. Since
latent heats are significantly greater than specific heats, the amount of energy
710
Recycle
Permeate Recycle
Product I Product 2
required to vaporize the permeate will not be available in the feed unless the
permeate rate is low. For removal of trace organics permeate rates will be low
and sufficient energy is usually available in the feed. For breaking azeotropes
conentrations are usually significantly higher and heat effects become
important. A recycle stream and/or interstage heaters (see Figure 12-25) are
used when necessary to provide sufficient sensible heat for the vaporization.
Additional heat input into the pervaporation unit is provided by the hot distil-
late streams. The value of e per pass is low, but e..r for the entire unit shown in
Figure 12-25 can be high. It is desireable to keep the temperature drop quite
modest (5 to lO°C) since fluxes are significantly higher at higher temperatures
(Sander and Soukup, 1988). Thus a large number of stages or high recycle rate
is desirable. The use of these equations for a one-pass system is illustrated in
Example 12-5. Recycle systems are left to Problem 12-C3.
often has much larger selectivities but lower fluxes than RO. For trace oganics
the high activity of the organic gives high relative volatilities and thus both
high selectivities and high fluxes. In addition, pervaporation will not be limited
by osmotic pressure.
(1) Includes 2 wt% sodium citrate. (2) Includes 2 wt% benzoic acid.
-...J
W
714
Solution
We can find the cut a from Eq. (12-90b). First, we need consistent
units.
c
p.B
= 0.625 caV °C [
g
cal
1000 kca1
1[74.12
gmole
g 1= 0.046 kcal
gmole °C
Note that z = 0.1 refers to water mole fraction since a' = a~. Then yp
is also water mole fraction.
yp = 10.45 ± 4.35
10.5 0581 mo Ie fracllon
=. . water
l00lb/hr
Area = Feed Rate/Flux = 2 = 500 ft2
0.2Ib/hr ft
A liquid membrane has a layer of liquid which serves as the separation medium
instead of the solid polymer used in the other membrane methods discussed
717
a b ~ Feed Phase
o 00 0
Feed Phase o 00 0 Recelvlng
..
0 0 00 0
o 0 0 Phase
0
Liquid 0
Membrane
o
Liquid Exploded
View
Membrane
Receiving Phase
previously. Two types of liquid membrane systems have been extensively stu-
died. In supported liquid membranes the liquid is held in a porous matrix
which serves to support the liquid (Danesi, 1984-85; Noble et ai, 1986; Ward,
1970; Noble and Way, 1987; Way et ai, 1982, 1985). The equipment and
operating procedures for supported liquid membranes are very similar to those
for the other membrane separations discussed in this chapter. In emulsion
liquid membranes a double emulsion is formed with the receiving liquid encap-
sulated by an immiscible material which is also immiscible with the outer fluid
(Cahn and Li, 1976; Li, 1971; Marr and Kopp, 1982; Noble et ai, 1986; Way et
ai, 1982). The emulsion liquid membrane systems are operated in a manner
very similar to liquid-liquid extraction. Liquid membrane systems have been
an area of considerable research, but only waste water treatment and gas sen-
sors for ion selective electrodes are commercial (Noble and Way, 1987). In the
future these separation techniques may become important as commercial
separation devices.
the feed phase and the receiving phase. The feed and receiving phases can be
essentially the same liquid and in many cases will be an aqueous solution. The
liquid membrane serves to separate the feed and receiving phases. In order to
increase the capacity of the receiving phase, it will often contain a chemical
which will react with the diffusing solute. The supported liquid membrane sys-
tems can use one of the geometries shown in Figure 12-5. One of the major
problems with this type of liquid membrane is bleeding which is loss of the
liquid membrane as it dissolves into the feed and receiving phases. Supported
liquid membrane systems have apparently not been commercialized yet.
The emulsion liquid membrane system is shown schematically in Figure
12-26b. The system is a double emulsion with the receiving phase distributed
within the liquid membrane drops which are dispersed within the feed phase.
The emulsion type systems are fairly simple to make and have a very large sur-
face area. Normal extraction equipment is used for the contacting. Mter the
extraction, the liquid membrane usually must be broken to recover the solute.
The receiving phase usually contains a chemical to react with the solute to
increase the capacity. Emulsion type systems have been extensively designed
and piloted for a variety of commercial processes such as copper recovery,
phenol removal from waste water, and hydrocarbon separations.
Both types of liquid membrane systems can use facilitated transport
(Cussler, 1971; Goddard, 1977; Noble et ai, 1989). Facilitated transport is
shown schematically in Figure 12-26c. A carrier contained within the liquid
membrane reacts with the solute and then diffuses across the membrane. In the
receiving phase the carrier-solute complex is broken and carrier diffuses into
the receiving phase. The carrier then diffuses back across the membrane to
pick up another "load". This mechanism is similar to the carrying of oxygen in
blood by hemoglobin. In the second mechanism shown in Figure 12-26c a
second molecule of A is back-transported across the membrane. The back-
transported molecule supplies the energy for the process (Cussler, 1971).
Examples of this process are using W to supply energy to separate cations.
Facilitated transport is of interest since it can provide much faster mass transfer
rates with higher selectivities than the liquid membrane alone, and because
facilitated transport appears to be important in transfer across cell membranes
in living systems. In terms of a solution-diffusion membrane, facilitated tran-
719
sport involves increasing the solubility and/or the diffusion rate of the solute.
Noble et al (1989) list a large number of systems which have been developed
for facilitated transport. Note that these facilitated transport chemical systems
can often be used to advantage in other separation processes such as absorption
or extraction, and these alternatives may be cheaper than liquid membranes.
At the end of this chapter you should be able to meet the following objectives.
4. Describe gelling and fouling, and estimate the effects on flux and
rejection. Explain why flux is determined by back-diffusion when
a gel layer exists.
REFERENCES
Beaver, E.R., P.V. Bhat, and D.S. Sarcia, "Integration of membranes with other
air separations technologies," AIChE Symp. Ser .• 84 (261), 113 (1988).
Belter, P.A., E.L. Cussler, and W.-S. Hu, Bioseparations. Downstream Pro-
cessingfor Biotechnology, Wiley-Intersciences, 1988, Chapt 9.
Blatt, W.F., A. Dravid, A.S. Michaels, and L. Nelson, "Solute polarizatio,' and
cake formation in membrane ultrafiltration: Causes, consequences and COl trol
techniques," in J.E. Flinn (Ed.), Membrane Science and Technology, Plenum
Press, NY, 1970,47-97.
Caracciolo, V.P., N.W. Rosenblatt and V J. Tomsic, "Du Pont's hollow fiber
membranes," in S. Sourirajan (Ed.), Reverse Osmosis and Synthetic Mem-
branes, Ottawa, Canada, National Research Council of Canada, 1977, Chapt.
16.
721
Chern, R.T., WJ. Koros, H.B. Hopfenberg, and V.T. Stannett, "Material selec-
tion for membrane-based gas separations," in D.R. Lloyd (Ed.), Materials Sci-
ence of Synthetic Membranes, Am. Chern. Soc., Washington, DC, 1985, Chapt.
2.
Colton, c.K. and E.G. Lowrie, "Hemodialysis: Physical principles and techni-
cal considerations," The Kidney, Vol. 2, 2nd ed., Philadelphia, W.B. Saunders
Co. 1981,2425-2489.
Cussler, EL., "Membranes which pump," AIChE Journal, 17, 1300 (1971).
Flett, D.S. (Ed), Ion Exchange Membranes, Ellis Horwood Ltd., Chichester,
England. 1983.
Finken, H., "Asymmetric membranes for gas separations," in D.R. Lloyd (Ed.),
Materials Science of Synthetic Membranes, Amer. Chern. Soc., Washington,
D.C., 1985, Chapt. 11.
Gienger, J.K. and R.J. Ray, "Membrane-based hybrid processes," AIChE Symp.
Ser.,84 (261), 168 (1988).
Henis, J. M.S. and M.K. Tripodi, "The developing technology of gas separating
membranes," Science, 220 (4592), 11 (April 1, 1983).
Klein, E., R.A. Ward, and R.E. Lacey, "Membrane processes-Dialysis and elec-
trodialysis", in R.W. Rousseau (Ed.), Handbook of Separation Process Tech-
nology, Wiley, New York, 1987, Chapt. 21.
Li, N.N. and RB. Long, "Permeation through plastic films," AIChE Journal,
15, 73 (1969).
Lloyd, D.R and T.B. Meluch, "Selection and evaluation of membrane materi-
als for liquid separations," in D.R. Lloyd (Ed.), Materials Science of Synthetic
Membranes, Amer. Chern. Soc., Washington, DC, 1985, Chapt. 3.
724
Lysaght, MJ., D.R. Boggs, and M.H. Taim isto , "Membranes in artificial
organs", in M.B. Chenoweth (Ed), Synthetic Membranes, MMI Press, Mid-
land, MI, 1986,100-117.
Noble, RD., C.A. Koval, and JJ. Pellegrino, "Facilitated transport membrane
systems," Chern. Engr. Prog., 85 (3), 58 (March 1989).
Noble, R.D. and J.D. Way (Eds.), Liquid Membranes, Theory and Applications,
ACS Symp. Ser., No. 347, ACS, Washington, D.C., 1987.
Noble, RD., J.D. Way and A.L. Bunge, "Liquid membranes," in Y. Marcus
(Ed.), Ion Exchange and Solvent Extractions, vol. 10, Marcel Dekker, NY.
Potts, D.E., R.C. Ahlert and S.S. Wang, "A critical review of fouling of reverse
osmosis membranes," Desalination, 36,235 (1981).
Stannett, V.T., W J. Koros, D.R. Paul, H.K. Lonsdale, and RW. Baker,
"Recent advances in membrane science and technology," Adv. Polymer Sci ..
32.69 (1979).
Stem, S.A. and H.L. Frisch, "The selective permeation of gases through poly-
mers," Ann. Rev. Mater. Sci., 11, 523 (1981).
Ward, R.A., P.W. Feldhoff and E. Klein, "Membrane materials for therapeutic
727
Way, J.D., RD. Noble and B.R Bateman, "Selection of supports for immobil-
ized liquid membranes," in Lloyd, D.R (Ed.), Materials Science of Synthetic
Membranes, ACS, Washington, DC, 1985, Chapt 6.
Way, J.D., RD. Noble, T.M. Flynn and E.D. Sloan, "Liquid membrane tran-
sport: A survey," J. Memb. Sci., 12, 247 (1982).
HOMEWORK
A. Discussion Problems
A2. What are the advantages and disadvantages of each module shown in
Figure 12-5?
A4. In gas permeation too high a selectivity can be detrimental and may
actually limit the removal of the permeating gas from the feed. Explain
this phenomenon.
A8. Figures 12-4 and 12-15 show no concentration polarization on the per-
meate side. Figures 12-18 and 12-21 show mass transfer on both sides
of the membrane. Explain these differences.
B. Generation of Alternatives
B 1. The RM membrane is a clever and commercially successful solution to
the problem of how to make essentially defect free gas permeation
membranes. Brainstorm at least 5 other solutions to this problem.
C. Derivations
Cl. Show that Eq. (12-32a) is a solution to Eq. (12-31) and the appropriate
boundary conditions.
C2. Show that Eq. (12-39) is a solution to Eq. (12-31) and the appropriate
boundary conditions.
a. Calculate a
D. Single-Answer Problems
D2. Data for a hollow fiber RO module from Dow Chemical is given in
Table 12-2. Details of how the test were done are not complete.
a. If there was no concentration polarization, calculate Ksolv/tM for this
system.
b. IfM = 2 during the experiment, determine K.olv/tM'
You can use the van't Hoff equation.
b) This membrane had RO = 0.96. What was the salt permeability for
731
a In the stirred cell find Jso1v and Jsolule if .1p = 5 atm and Cb =20g/L.
b. In the hollow fiber system find J so1v , RaP!' and J so1uIe if .1p = 5 atm
and Cb = 2Og/L. Other conditions are the same as in previous
experiment
Dl1. We are separating oxygen and nitrogen by gas permeation with a sil-
icone rubber membrane. The membrane module is perfectly mixed.
°
Inlet gas is 21 mole % 02' We desire a permeate product which is 27
mole % 2 , Membrane has a selectivity (lAB =P AIPB = 2.1. Pressure
ratio is PL/PH = 0.35. Treat the gases as ideal gases. What cut e must
be used?
flux is 0.416 x 10-5 m3 /m 2 sec. Assume the gel layer has a porosity of E
=0.5.
The basic concepts, definitions, and theories for a variety of membrane separa-
tors were presented in Chapter 12. In order to cover the entire area of mem-
brane separations while at the same time keeping that chapter a reasonable
length some of the more involved analyses were not included. These more
involved analyses will be discussed in this chapter. Chapter 12 is a prerequisite
for this material.
The bulk flow patterns in the membrane modules have a large effect on
separation. In Chapter 12 fluid on both sides of the membrane was assumed to
be perfectly mixed. In Section 13.3 perfectly mixed modules with concentra-
tion polarization are studied. Countercurrent flow is studied in Section 13.4.
733
734
The usual picture of concentration polarization was shown in Figure 12-4. The
solute mass balance was given by Eq. (12-31) when the rejection is 1.0 and cp
= O. When the rejection is not 1.0. the mass balance can be written as:
dc (13-1)
]solv cp +]solv c +D -
dy
=a
where the first tenn represents solute which passes through the membrane. The
boundary conditions are
(13-2a)
c=cw at y=O
(13-2b)
c = Cb at y= 0
(13-4)
k=D/o
This mass transfer coefficient k can be calculated from mass transfer correla-
tions for different flow regimes. This is the subject of Section 13.1.2. Substi-
tuting Eq. (13-4) into (13-3a), we obtain
JI01v is given by Eq. (12-27). Note that Equations (12-27) and (13-3b) are valid
at each point of the membrane since, in general, Cb, c w, cP ' J solv , and M vary
throughout the membrane module. For a mixed system Cb = COul is constant and
hence Cw, cP ' M, and Jsolv are constant. When RO = 1, simultaneous solution of
Eqs. (12-27) and (13-3b) gives J so1v and M. If RO < 1, we must simultaneously
solve Eqs. (12-22), (12-25), (12-27) and (13-3b). For high flux membranes
Eqs. (13-4) and (13-5) are both suspect.
Equations (13-3), (13-4) and the expressions for k are convenient to use
when Jsolv is constant and specified. If Jso1v is unknown, these equations must
be solved simultaneously with Eq. (12-27). Unfortunately, the polarization
modulus M is required to calculate J so1v in Eq. (12-27) and Jso1v is required to
determine M in Eq. (13-3). Thus an iterative trial-and-error solution is often
required Fortunately, if Jso1vlk «1.0 an approximate solution can be
developed (Rao and Sirkar, 1978). This solution is particularly simple for ideal
semi-impermeable membranes where RO = 1. Assume that the osmotic pres-
sure is a linear function of concentration following Eq. (12-18b) which is rea-
736
sonable for dilute solutions. When J101v/k « 1.0, the exponential tenn in Eqs.
(12-32a) and (13-3) can be expanded as
cw J 1 JIOIV
, (13-5)
M =-=exp - - - 1 +--
( so v )
Cb k k
(13-6)
Equations (13-5) and (13-6) can now be solved simultaneously for J so1v '
Ksolv
--(~P-acb)
J lmv = -1m- - - - - (13-7)
1 Ksolvacb
+ 1m k
When using this equation, the assumption that JI01v/k« 1.0 should be checked.
Equation (13-7) is useful when Cb is known. If Cb is not known, we need to
include mass balances (see Section 13.3.2.).
IfRO < 1 and/or (JlOlvlk) is not very small, the result from Eq. (13-7) can
be used as the first guess for J101v ' Then Eq. (13-3b) can be solved for M. Now
the solvent flux equation with RO < I, Eq. (12-27), can be solved simultane-
ously with Eqs. (12-22) and (12-25). This value of J solv can be used in Eq.
(13-3b) and the procedure can be continued until there is convergence.
Once M and JI01v have been detennined the required membrane area and
module length can be found. An external mass balance for the module is
(13-8)
J lmv A Cp + (F - J I01v A)cout =F Cp
In this equation F is the volumetric feed rate to the module (e.g. liter/hr) and
cout is the concentrated product on the feed side of the membrane. Since cout is
737
usually given, Eq. (13-5) can be solved for A. For tubular or hollow fiber sys-
tems,
(13-9)
A = (xLd)n
where n is the number of tubes of diameter d and length L.
. SC-2/3
k = UbJD (13-lOa)
(13-lOb)
where f is the Fanning friction factor which in turbulent flow can be estimated
from
(13-1Oc)
f = 0.0791 Re-1/4
(13-11)
k = 0.023 ~ ReO. 83 Sc 1/3
Fully developed turbulent flow on tubes will certainly occur for Re > 20,000
and in UP devices appears to occur at Re = 2000 (porter, 1979). These results
can be substituted into Eq. (13-3) to estimate the concentration polarization
modulus in turbulent flow. Note that concentration polarization in turbulent
738
flow does not depend upon the distance down the tube. For turbulent flow
between parallel sheets the same correlation can be used except the equivalent
diameter of the channel deq should be used
0.33 [ 2] 0.75
k=0.0443 ~;
[ ] 00: (13-13)
00
where d is the vessel diameter in cm, is the stirrer speed in radians/sec, v is
the kinematic viscosity in cm 2 /sec, and D is the diffusivity in cm 2 /sec. This
form is useful for laboratory stirred tanks, but the stirred tank configuration is
unlikely to be used on a large scale. An example calculation is part of Example
13-2.
Turbulent flow can be important in tubular (Figure 12-5b), stirred tanks
and plate-and-frame (Figure 12-5a) modules. Qualitatively, Eq. (l3-3b) shows
that M decreases if the mass transfer coefficient k increases. This can be
achieved with high velocities ~, high diffusivities, and low viscosities.
Operating at high temperature increases D and decreases)J.. Increasing Jsolv
increases concentration polarization. Thus high flux membranes have more
concentration polarization.
For laminar flow conditions the average mass transfer coefficient can be
estimated from (Blatt et al. 1970; Porter, 1979)
(13-14)
where Yw is the fluid shear rate at the membrane surface and L is the length of
the flow channel. The fluid shear rate in laminar flow in a round tube is
4~ (13-15)
Yw=T
739
3 Ub (13-17)
YW=-h-
(13-18)
These laminar flow correlations are used when the concentration polarization
layer is thin, which holds when the axial distance is much less than the entrance
length (see Example 13-2). These correlations be used when a gel does not
form to estimate Jso1v and Musing Eqs, (13-3b) and (13-7), or when a gel forms
(Section 13-2).
id tubes is being used. The average fluid velocity Ub = 1300 cm/sec and
oM> =50 atm. The cut S =0.10. Operation is at 18°C.
Find the average concentration polarization modulus M, the average
solvent flux Jaolv , Couto and cpo
Data:
at 18°C: DNaC! (0.4 gmoles/L) = 1.17 x 10-5 cm 2 /s
DNaC! (0.8 gmoles/L) = 1.19 x 10-5 cm2 /s
(Sherwood et aI. 1975, p. 37)
Solution
Eq. (13-3b) and R and hence S, from the solution to Eq. (12-29). We
can then check that Coo.ll is not changed significantly.
cc attn]
The factor [82.057 gmole K / [ 1.9872
gmole Cal]
K .
converts two different
values of the gas constant. The values ~ and Renc::rgy depend on the
units of .1~, v.olvent, and the desired pressure units for 7t.
Since osmotic pressure data for NaCI is readily available (perry and
Green, 1984, p. 17-23), the purpose of this calculation is illustrative.
Interpolating between concentrations and extrapolating to 18°C using
the tabulated data, we find 7t = 18.57 which is a 0.6% difference.
Assuming 7t is linear with concentration, a = 7t/c = 18.69/0.422 = 44.29
atm/gmole/L.
742
2. Calculate k.
A 0.422 gmole/L solution is about 2.4 wt% (same Table as Freez-
ing point depression). Linearly interpolating, p(WOe) - 1.0174 and p
(25°C) - 1.0140. Then doing another linear interpolation p(l8°C) -
1.0156 glee. [This is not the best way to estimate density, but since the
differences are small it will not cause much error].
Viscosity of a mixture can be estimated from (Reid et ai, 1977, p. 462)
where the Xi are mole fractions. Here component 1 is pure water and
component 2 is 25 wt% solution. From the same table which gives
freezing point depression:
3. Solvent flux.
J -
(1.24 x 10-5 ) (50 - (44.29) (0.422» = 0.000386 cm/s
101v - (1.24 x 10-5 ) (44.29) (0.422)
1+ 0.0418
4. Solute.
Ksolv RO 0.995
0.= = =- - - - - -----
Ksolute (AP- An) (l-RO) (70 - (44.29) (0.4» (1- 0.995)
0.= 3.806
Now the solution to Eq. (12-29) gives~. Constants for the quadratic
formula are
- - up- M 1t (»
Ksolv ( A
Cb +
Ksolute
1m 1m
~ = 0.0039 gmoles/L
Gel formation was discussed in Chapter 12 following Eqs. (12-37) and (12-39).
In this section more details of the calculation procedures will be given. The
formation of a gel depends mainly upon the solute type and concentration
although the membrane characteristics and hydrodynamics also affect gel for-
mation (Blatt et ai, 1970; Fane, 1986). Rigid chain, solvated macromolecules
such as polysaccharides can gel at concentrations well below 1 wt %. Flexible
chain, linear macromolecules often gel in the range of 2 to 5 wt%, while highly
structured spheroidal macromolecules such as proteins and nucleic acids gel in
the range from 10 to 30 wt %. Colloids can also form gels. Submicron pig-
ments or minerals gel in the range from 5 to 25 vol % while polymer lattices
require from 50 to 60 vol %.
For membranes with R=I, the solute mass balance was given in Eq. (12-
31) and the solution was given in Eq. (12-39),
cg (12-39)
J I01v = k In-
Cb
The mass transfer coefficient can be estimated from the mass transfer theories
discussed in Section 13.1.2. Experimental UP data can be used to easily esti-
mate k and cg • If J solv is determined at a series of bulk concentrations, Cb. a
plot of J solv versus In Cb will have a slope of -k and an intercept on the concen-
tration axis of In cg (see Problem 12-D4). The resulting k and cg values are
746
then useful for correlating data. Changes in conditions such as velocity or tem-
perature will change the mass transfer coefficient and can be correlated with
Eqs. (13-11), (13-13), (13-16) or (13-18).
Although very appealing the gel formation theory often does not agree
with experimental data (Fane, 1986; Le and Howell, 1984; Porter, 1979; Wij-
mans et ai, 1984; Zydney and Colton, 1986). For instance, the model does not
include any affect of the membrane, velocity, or feed concentration on cg , but
experimental results show such effects. Since cg is usually much less than the
solubility limit, it is difficult to determine fundamentally exactly what cg is
measuring. Solute-solute and solute-membrane interactions are important
experimentally, but are not included in the model. Experiments with colloids
show fluxes which can be one or two orders of magnitude greater than expected
(porter, 1979). (This is very helpful and helps explain why the earliest com-
mercial applications of UP were for processing colloids, but it is a major prob-
lem for the theory.) Finally, the theory predicts a limiting flux which is
independent of.1p while experiments often show a.1p dependence on the limit-
ing flux.
Many attempts have been made to explain these problems. For example,
in medium molecular weight applications the osmotic pressure can be impor-
tant and should be included in the model (Fane, 1986; Wijmans et ai, 1984).
Unfortunately, models with osmotic pressure or with gels have very similar
predictions and it is difficult to determine which is correcL For macro-
molecules> 100,000 daltons the osmotic pressure cannot have much influence.
747
Porter (1979) suggested that the colloid data can be explained by the tubular
pinch effect. This is the tendency of small particles flowing in small rubes to
migrate away from the tube walls. This effect will increase the mass transfer of
particles away from the wall and will increase the solvent flux. Although this
effect definitely exists, it does not appear to produce a large enough effect to
completely explain the data (Le and Howell, 1984). Other explanations for
high fluxes with particles are the particle diffusivity is much higher than
Brownian diffusivity because it is caused by particle-particle interaction
(Romero and Davis, 1988; Zydney and Colton, 1986). Other models which
include variable diffusivity or cake aging or adsorption or leaky pores all seem
to be able to explain only part of the phenomena. The mass balance, Eq. (12-
31), does not include axial convection of solute and the boundary condition c =
a
Cb at y = is hypothetical. Shen and Probstein (1977) and Trettin and Doshi
(1980) improved the model by including convection in the x direction (e.g. see
Eq. (13-48» and writing the boundary condition as c = Cb at Y = 00. They also
allowed for a concentration dependent diffusviity. Their results for the limiting
flux agreed better with experimental data for protein UP than Eq. (12-39). At
the current time gelling in UP must be considered only partially explained.
Future research will eventually unravel the complexities involved. The simple
gel formation model still serves as a simple physical picture and as an excellent
method for correlating data.
Fouling can occur in both RO and UP although it is often worse in UP
since dirtier streams are often processed. Fouling is a plugging or coating on
(external) or in (internal) the membrane which is partially irreversible. That is,
in normal operation the fouled membrane will stay fouled until a separate
cleaning step is employed. After clean-up part or all of the original flux will be
recovered. Fouling is complex because it can be caused by many different
phenomena such as precipitation, adsorption, electrostatic attraction, biological
growth, chemical reaction or polymerization.
In RO fouling can be classified as inorganic, particulate, or biological
(potts et aI, 1981). Inorganic fouling occurs when compounds such as CaC03 ,
Caso 4 , MgC03 , and silica or iron precipitate out onto the membrane. Since
the purpose of RO is to remove water and since the concentration of salts is
highest at the membrane wall, it should not be surprising that precipitation can
748
be a problem. The higher the water recovery the more likely precipitation is to
be a problem. Fortunately, precipitation can often be controlled by adding sul-
furic acid to adjust the pH in the range 4 to 6, and by adding compounds which
bind calcium (e.g. Calgon). Fouling from particulates can be caused by parti-
culates in the feed. This is relatively easy to control by prefiltering the feed. A
more serious problem is the presence of both inorganic and organic colloids.
Methods for solving these problems will be very specific for each case since
the chemistry will differ from case to case. Some membranes such as cellulose
acetate are susceptible to biological growth. This can be controlled by operat-
ing in the pH range from 4 to 6, removing dissolved oxygen, or chlorinating.
All of the above types of fouling can also occur in UF. In addition, since
macromolecules are often ultrafiltered, polymer precipitates or gels may form
which can be attached to the membrane by adsorption or electrostatic forces.
The polymer may also react to form a cross-linked gel layer. Generally speak-
ing, the higher the molecular weight of the solute the worse fouling will be.
Modules should be designed to have no stagnant regions since fouling is invari-
ably worse in these locations. Separate cleaning steps are often employed.
These include pulsing clean water, back-flowing clean water through the mem-
brane, mechanical cleaning with sponge balls or other methods, and chemical
cleaning such as a caustic wash. Fouling may also be reduced by changing the
membrane properties. For instance, electrostatic attraction can usually be
reduced by giving the membrane a negative charge since fouling colloids are
often negatively charged.
(13-19)
where Jso1v(t), Jinit , and Jasy are the fluxes at time t, initially and asymptotically.
The time t is often measured in days, and JlO1v(day 1) = Jinito The empirical
constant m is negative. An m = -0.1 corresponds to about a 45% decline in
flux in one year while an m = -0.03 corresponds to about a 16% flux decline in
one year. If the membrane is chemically or mechanically cleaned, much of the
original flux can be restored. The decline will then start again from this
749
restored flux. Every time the membrane is cleaned some flux is. pennanently
lost. Thus the membranes will have a finite life. This lifetime can vary from a
few months to several years depending on the service conditions. Membrane
replacement should be included as an operating cost. Membrane systems are
usually designed with an average flux. For a parallel cascade membranes can
be replace in a staggered pattern so that some membranes are always fresh and
others are near the end of their useful life.
In Chapter 12 the mass balances for completely mixed systems were con-
sidered. In this section we will first develop the external mass balances for
reverse osmosis and ultrafiltration with concentration polarization. The possi-
ble occurrence of a gel will be included in Section 13.3.2.
The balances will be done for the system shown schematically in Figure
13-1. The overall external mass balance is
(13-20)
Fin = Fout + Fp
(13-21)
Fin em = Fout Coot + Fpcp
750
~N,CIN
Figure 13-1. Schematic of completely mixed RO or UF with concentration
polarization.
The volumetric permeate flowrate can be determined from the flux as Fp = JA.
For a perfectly mixed container Coot = Cb' Equations (13-20) and (13-21) can
be solved to find Couto
(13-22)
which is Eq. (12-30). Note that these equations are essentially the same as Eqs.
(12-5) to (12-8) since the external balances are unaffected by what happens
inside the system. The permeate concentration is related to the wall concentra-
tion
(13-23)
(13-24)
The wall concentration can be removed from Eqs. (13-23) and (13-24) by sub-
stituting in Eq. (13-3b). The results are
(13-25)
751
= ------------
em
Cb (13-26)
8(I-RO) exp (Jsolvlk)
(1-8) + - - - - - - -
RO + (l-RO) exp (JlOlvlk)
Note that if RO =1.0, cp=O and Cotlt=cuJ(I-8). When RO =1, the flow patterns
and solvent flux are unimportant for determining concentrations.
For R O <1 Eqs. (13-25) and (13-26) are very useful if Jso1v is known.
Unfortunately, J so1v is usually not known. The solvent flux must now be deter-
mined from the flux Eq. (12-27). We will assume that the osmotic pressure
depends linearly on concentration so that Eq. (12-18b) is valid. Combining
Eqs. (12-27), (12-18b) and (13-23), we have
(13-27)
The wall concentration can be removed using Eq. (13-3b), and this result can
be solved simultaneously with Eq. (13-26). The equations are nonlinear and a
closed form solution cannot be obtained. If the exponential term is approxi-
mated as in Eq. (13-5), an algebraic solution can be obtained. The wall con-
centration becomes
(13-28)
which gives
(13-29)
(13-30)
752
These two equations can now be solved simultaneously. After some manipula-
tions the following quadratic equation results.
(13-31)
a' Jtolv + b' Jso1v + C' = 0
where
(13-32a)
(13-32b)
After determining k from the appropriate mass transfer correlation, J solv can be
found from the quadratic formula.
Once Jso1v is known, Cp and Cb can be calculated from Eqs. (13-25) and (13-26).
If gel formation is possible, calculate Cw from Eq. (13-28). If Cw < c g a gel
won't form and this solution is correct. If Cw > cg , a gel forms and this solution
is not correct (see Section 13.3.2). Also, this solution is based on Jso1vlk < < 1
and this restriction should be checked. If the assumption is invalid, we need to
solve Eqs. (13-3b), (13-26) and (13-27) simultaneously. Equation (13-33) can
be used for a first guess.
(13-34)
for R=1
753
(13-35)
J solv 1
-----=--=--------
w. COllC pol.
k(1-9)tro
(13-37)
(13-38)
754
where we assume that the solute is mobile in the gel and does not change RO.
Since cg is known, Cp and Cb can be calculated immediately.
(13-39)
which reduces to Eq. (12-39) when RO = 1.0. We can determine k from the
appropriate correlation and then calculate l solv • For laminar flow Eqs. (13-16)
and (13-18) are convenient for first estimates of the average value of k.
Find Cp, Cb and J I01v for feed concentrations of 0.04 and 0.20 wt frac.
albumin ifa = 0.5.
Solution
We will first assume a gel does not form and use Eqs. (13-32) and
(13-33) to find J 101v • Then we will find Ct, from Eq. (13-26) and Cw from
755
= 0.00594 cm/s
To estimate a in Eq. (12-18b) we will use Eq. (12-18a). Since the solu-
tion will be approximately doubled in concentration, use c = 8 wt%
albumin. Then for 1 liter have approximately: 960 g water = 53.3
gmoles water and 83.5 g albumin = 0.00119 gmoles albumin.
Thus
= 0.0285 atm
from Eq. (12-18a). We can use Cin as 0.08 wt. frac in Eq. (12-18b).
756
Then
1t 0.0285 atrn
a =- = 0 .08 wt fraC =0.356 atrn/wt frac.
c
= 0.00243 cm/s
where the positive sign is used to make J ao1v positive. To check the
assumption calculate
There will be some error in the calculation (see Notes) and a second
tria' would be required for accurate results.
Cb =
em = 1.9763 Cin
0.5 + 0.5 (0.008) exp (0.409)
0.992 + (0.008) exp (0.409)
757
(1.9763) em (1.409)
Cw :::: 0.992 + (0.008) (1.409) :::: 2.775 Cin
Note that since osmotic pressure is negligible, em does not affect the
calculation of J so1v and the other concentrations can be detennined as
functions of inlet concentration.
Thus, for the 20 wt% solution we need to use the solution when a gel
forms. With a gel formed we use Eqs. (13-37) to (13-39).
J so1v ::::
(0.992) (0.45) / (0.3964)
(0.00594) In [ 1 _ (0.008) (0.45) / (0.3964)
1:::: 0.00076 cm/s
Notes:
A. There are several approximations involved in this solution.
1-----1
~ L _____
:::::f1..J
~
Figure 13-2. Schematic of countercurrent gas penneator.
759
(13-40a)
"
-d [ql (l-Yl) ] = -d [<U(l-Y2)] = P~B [(l-Yl)PH - PL (l-Y2)] (13-40b)
In these equations ql and qz are the flow rates in moles/hour on the feed and
permeate sides of the membrane, and Yl and Y2 are the mole fractions on the
feed and permeate sides of the membrane, respectively. A is the membrane
area PA and PB are the permeabilities of the membrane. The second equation
p,
in each set comes from the flux Eq. (12-3). The molar density, is required to
convert the volumetric flux JAto a molar flux. These two mass balances can be
rearranged. Adding Eqs. (13-40a) and (13-40b) we obtain
(13-41)
(13-42)
A third equation is also required. This can be obtained from the overall
and component mass balances for the mass balance envelope shown in Figure
13-2.
(13-43a)
(13-43b)
760
Removing the unknown flowrate 'l2 and solving for Y2, we obtain
(13-43c)
Equations (13-41), (13-42), and (13-43c) are solved simultaneously. The boun-
dary conditions are:
(13-44a)
(13-44b)
In Eq. (13-44a) Y2.0 is the penneate mole fraction corresponding to Yout. This
penneate mole fraction can be obtained by taking the ratio of flux Eq. (12-3)
written for components A and B at the outlet. This result is
PL
Yout - PH Y2.0
Y2.0 PA (13-44c)
- - = aAB - --------
l-Y2.0 PB PL
l-Yout - - (I-Y2.o)
PH
Note that Eq. (13-44c) is essentially the same as Eq. (12-10) for a perfectly
mixed system.
"'~'
0.28 \':0:,
, '.,
.... ',
,
'\ ...... ',
0.27 \. .....
', '"
0.14
'\,".. ,, ,
~ '\, U
<l>
ClJ '~~
'"
"..... ,,
0.26 "\, ". ~
E
ClJ ,,', ".
a.
,"
\', .••.. c
<l>
" " 012
OJ)
c \ '..... ,
ClJ
" 0.25
OJ) ~ 0.10
>-
x ". c
0
'. \.'<-'." \ , 0
c \ ':,.... \ .;::;
0 u
.;::; 0.24 Countercurrent - - - Countercurrent
u
.' \ ".'\;" ,\ .::'" 008[
.::'" ---- Cross-flow " '';;.,'. \ - - - - Cross-flow
\
........ Cocurrent " 'i:. \ 0.06 ....•..... Cocurrent
0.23
,,
___ ._ Constant yp '. \ •...'\;." \ \ \ ----- Constantyp
_ •. _ Well mixed
",:\.\ ---- .. Well mixed
0.22
'\.\
, \\' '" 0.02
~-'\,~-~
0.21
I I I I
0
°T0 0.2 04 06 08 10
,~
Cut, !:I Cut
(a) (b)
Figure 13-3. Product concentration for air separation (ex =2.05, prJPH = 0.359, YF = 0.209). a. Permeate con-
centrations. b. Low pressure product concentrations. (Blaisdell and Kammermeyer, 1973).
Reprinted with permission from Chern. Eng. Sci., 28, 1249 (1973). Copyright 1973, Pergamon
--.J
Press. 0\
762
other than completely mixed and countercurrent are developed by Hwang and
Kammermeyer (1975).
In laminar flow both concentration and velocity boundary layers will form.
Since diffifusivities are low, the concentration boundary layer will be thinner at
the beginning of the tube. Both boundary layers become thicker as one
progresses down the tube. If the tube is long enough (or the channel is thin
enough) both boundary layers will eventualy fill the tube. Thus, one should
expect two solutions for laminar flow: an entrance region solution and a fully
developed far-downstream solution. The solutions have been generated for the
case where the velocity profile is fully developed immediately, but the concen-
tration profile is not. The solute was assumed to be completely rejected. Solu-
tions for both parallel plates and round tubes are available (Sherwood et ai,
1965.
The geometry for RO between two parallel sheets is shown in Figure
13-4a For laminar flow the velocity must first be calculated from the following
Navier-Stokes equations,
(13-45a)
(13-45b)
(13-45c)
(13-46a)
u=O at y=±h
763
a
t t t t t
yLx tv .. u th
th -
• tv w
• • •
b
2R
\
(13-46b)
v= Vw at y=±h
au (13-46c)
v=-=O
ay at y=O
-u= -
Uin
3 [ 1 -Vw
2
--x (1-y')
Urn h
1 - h- [ 2-7y'2_7y'4 J
2 [ 1 -Vw
420 v
1(13-47a)
and
v y' ,z vwy ,2 I(j (13-47b)
- = - (3 - y ) - - - (2 - 3y - y )
Vw 2 280 v
764
where Ujn is the average fluid velocity at the channel inlet, and y' = y{h. This
solution reduces to the usual solution for laminar flow between flat plates when
vw =0.
These solutions for velocity can be inserted in the steady state mass bal-
ance for solute. When rejection is perfect, R = I, this balance is,
a [uc - D ax
ax ac] + ay
a [vc - D ac] = 0
ay
(13-48)
b) at the wall the flux of solute to the wall equals the backward difussion of salt
c Vw
ac
= D ay at y= ± h
(13-49b)
(13-49c)
~=O at y=O
oy
Sherwood and his coworkers also obtained asymptotic solutions for cer-
tain regions. The solutions are written in terms of the dimensionless variable
(13-50)
765
l:::,. a =0.0677
o a =0.27
o a =0.50
10.0
M-I
o.I 1..-----'L-...J'---'-J'-'---'--...L-L-'-'---'----L-.<-L-.l----1.----1.--'--I...J..-"----"-,--'-'
~ =8L/3a 2
Figure 13-5. Solutions for laminar Bow between parallel plates (Sherwood
et ai, 1965). Reprinted with permission from Ind. Eng. Chem.
Fundam .. 4, 113 (1965). Copyright 1965, Amer. Chern. Soc.
(13-51)
M = 1 + 1.536(~)1!3
For larger values of ~ the approximate solution for the entrance region is
(13-52)
M = 6 + ~ - 5 exp(--.J'fj3)
Very far downstream ( large ~ ) the boundary layer will completely fill the
tube. Once this occurs M will be constant.
(13-53)
A similar but more complicated analysis can be done for laminar flow in
round tubes. The geometry is shown in Figure 13-4b. The Navier-Stokes
equations were solved by Yuan and Finkelstein (1956). Sherwood et ai. (1965)
used this velocity solution in the solute mass balance. The approximate solu-
tion for the entrance region for v! x R/(4llm D2) s 0.02 is
(13-54)
M= v3XR2
w + 6 - 5 exp [v3
- ( w XR 2 )112 1 (13-55)
4 Uin D 12 llm D
Note that the concentration polarization depends upon the axial distance x.
Far downstream the boundary layer will completely fill the tube. Once
this occurs M will be constant. This far-downstream solution was obtained
numerically and was presented in graphical form. This result is shown in Fig-
ure 13-6 (Sherwood. et ai, 1965). To determine if the entrance region solu-
tions, Eqs. (13-54) or (13-55), or the far-downstream solution (Figure 13-6)
should be used. calculate M by both methods. The smaller M should be used.
Although usually quite accurate, these results for both turbulent and lam-
inar flow are approximate for a variety of reasons. The no-slip boundary condi-
tion at the wall is not strictly true for a porous solid since there can be lateral
fluid movement in the pores (Fane, 1986). Fluid properties P,1l and v and
767
100
80
60 -
40
0
20
M-I 10
8
6
..b
vwR
Figure 13-6. "Far-downstream" solution for round tubes. (Sherwood et ai,
1965). Reprinted with permission from Ind. Eng. Chern. Fun-
dam .. 4, 113 (1965). Copyright 1965, Arner. Chern. Soc.
are D = 5.2 x 10-7 cm 2 /sec, P = 1.01 glee, J.L = 0.0095 cm 2/sec, CL = 0.99
gH 2 0/ml (solvent concentration). The operation is inside round hollow
fibers which are 0.01 cm in radius. Bulk v~locity at the inletis 3.2
cm/sec. Flux rate is 1.8 x 10-4 glsec cm 2 • The totaIlength of each hol-
low fiber is 50 cm. Assume RO = 1.0 and Ppc:rmeate = 1.0.
If the hollow fibers are increased to 500 cm long, estimate the average
concentration polarization.
Solution
vw _- - - -_ 1.8 x 10-4 -
flux _ 1. 80 x 10-4 Cm/S
Pw 1.00
_D_ =
Vw R
1
5. 2x 0- 7
(1.8 x 10 ) (0.01)
=0.2888
0.421
M=0.421+6-5exp [- [- 3 - ]'h] =2.98
Since entrance region result is smaller, use that value of M.
769
v3
w
R
X
= (0.421) -250 = 4.21
4lljn D2 25
As a check on the solution for L = 50 we can use Eqs. (13-27) and (13-
3b).
Note that 0 = D/k = 5.2 x 10-7 /1.555 x 10-4 = 3.34 x 10-3 cm < R.
If we apply this correlation when L = 500,
k =0.00007218, M =12.108
In this section two models for ttansport inside the membrane will be developed.
The structure of irreversible thermodynamics will be used for these models and
will be briefly presented first Then the solution-diffusion model appropriate
for RO membranes will be developed in a simplified form. Finally, a frictional
model appropriate for UP membranes which have small pores will be dis-
cussed.
(13-56)
Ji = Flux i = Ln (Force on i) + L ~j (Force on j)
Thus the flux of species i is proportional both to the force on that species and to
the force on other species. This is usually written as
J.=T··X+
1 ~ 1
~ T··X
~ '-'iJ J
(13-57)
jU,.l)
and
(13-59)
Equation (13-59) states that a species will move in the direction of the force
applied on it Equation (13-58) requires that ~j =0 if Lii = O. Thus, if a com-
ponent will not move through a membrane due to forces acting directly on the
component then it will not pass through the membrane at all. Thennodynami-
cally, ~j can be either positive or negative since Eq. (13-58) only limits the
absolute magnitude of the coupling coefficients L ij . In practice all known Lij
are ~ O.
One other condition on the phenomenological" coefficients is Onsager's
reciprocal relationship.
(13-60)
~j = Lji
This states that coupling of two species is the same regardless of which is
experiencing the force and which flux we are considering. This relationship is
based on microscopic reversibility.
(13-61)
where Ili is the chemical potential of species i and Yi is the external force on
component i. The external force in ion exchange membranes is an electrical
force. In porous membranes where separation is based on sieving the Yi are
frictional forces between the component and the membrane.
772
(13-63)
(13-65)
(13-66)
where we have assumed that Vl is constant so that we could do the last integra-
tion in Eq. (13-64). Note that &t is from one membrane interface to the other.
That is, from c'm to C"m in Figure 13-7.
It will be our purpose to first determine the appropriate terms for the
external forces, to then arrange the equations in a form which allows experi-
mental determination of the coefficients Ln and ~j' and finally to predict the
773
=
Assume that there is no coupling of fluxes (Lij 0), and that external forces are
unimportant (Yi = 0). This assumes that convective flow of water or carrier gas
through pores or pinholes is negligible. This is appropriate for solution-
diffusion membranes used in reverse osmosis and gas permeation. Combining
Eqs. (13-57), (13-61), and (13-66), we obtain
(13-67)
where we have approximated grad J..li as ~Itm. J' is the mass flux in units such
as g/cm 2 s, and J'solv = J solv Psolv. This result agrees with Eq. (12-20a) if we
relate
_ Kao1v Psolv
L 11- (13-68)
VI
The minus sign in Eq. (13-67) comes in because the +x direction shown in Fig-
ure 13-7 is in the direction of osmotic flow for the irreversible thermodynamic
argument. The solvent density appears because mass and volumetric fluxes are
MEMBRANE
SUPPORT
e"2m
Assume that sorption and desorption at the membrane surface are rapid
so that diffusion controls. Then Lu can be related to the concentration within
the membrane Cim and the mobility IDjm .
(13-69)
The mobility is the velocity per unit force on one mole of particles. In dilute
solutions the mobility can be related to the diffusivity.
(13-70)
mim =D 1m /RT
where R is the gas constant and D 1m is the solvent diffusivity within the mem-
brane. Since the diffusivity and the solvent concentration are often approxi-
mately constant, Lll is often approximately constant. Now the solvent flux is
(13-71)
]'solv =
Remember that .rut refers to the osmotic pressure difference across the mem-
brane. The solvent permeability can be estimated by comparing Eqs. (13-71) to
(13-67) and (13-68)
(13-72)
(13-73)
where J 2 is a mass flux which is the same as in Chapter 12. Since the last term
is usually negligible, we obtain
(13-74)
The phenomenological coefficient Lzl can again be related to the mobility and
for dilute solutions to the diffusivity.
D'}m (13-75)
~2 = m:an C']m = - - '2m
RT
Since c'}m can change significantly in the membrane, L22 may not be constant.
For dilute solutions
[ ~l
dC:an T,p
RT (13-76)
(13-77)
D 2m K z D 2m Kz (13-79)
Jsolute = ---(czw-cZp) = - (M CZb - cp )
t.n t.n
This result agrees with Eq. (12-25) and shows that
(13-80)
Note that this result also agrees with Eq. (12-15) used for gas permeation which
is also a solution-diffusion process.
CZp
[
g salt
cm 3
1 = J'
J, [~l
[g water 1
c
Ip
[g water
cm 3
1 (13-81)
101v cmz s
(13-82)
Assuming that cZw » CZp and defining Kl = Clm I Cl p as the water distribution
coefficient in the membrane, we obtain
(13-83)
777
From Eq. (13-83) we can estimate the properties required to obtain a given con-
centration ratio. Note that the concentration ratio in RO membranes is propor-
tional to (Ap - Ax). More separation is achieved by increasing Ap.
Solution
l
~= 1(8.095 cc/gmole ::.:: 0.000752 atm-l
RT 82.057 (cc atm)
gmoleK
1
(293.16 K)
CZw c2b
-=-=100
c2p c2p
and
The solute flux is due to the sum of contributions from the convective
flow and diffusion.
(13-84)
Jz = c2m U + J2djf
where u is the center-of-mass velocity of the pore fluid. The term c2rn u
represents -~1 dJlddx. The diffusive flux can be written as LzZX 2 • Lz2 can
again be related to the mobility, and X z is the force from Eq. (13-61). Thus,
aJl2] dC2m
J2di[ = m2 c 2rn [- [ aC2m p,T ~ + Y2
1 (13-85)
where Y2 is the frictional force on the solute due to friction against the mem-
brane. This external force can be related to the solute velocity within the pores
as
(13-86)
where f23 is a friction factor between the solute (2) and the membrane (3). The
mobility mz can also be related to a friction factor except now it is the friction
between the solute and solvent
(13-87)
mz = l/fz1 = D21 / RT
where (2) is solute and (1) is solvent. Combining these equations and solving
for J 2, we obtain
(13-88)
(13-89)
780
As u approaches zero this gives us diffusion with a drag factor applied. The
solute flux can also be related to the solvent flow as
(13-90)
J 2 = E U c2p
(13-91)
which allows one to predict the permeate concentration once the solvent velo-
city is known. For high negative velocities Eq. (13-91) simplifies to,
The terms .11t' and .11t" are the osmotic pressure differences at the two
membrane-liquid interfaces.
~1t" = c'~2m J-
1t [ 1t(c2p)
(13-95b)
781
c2m 1
[ddxP+ Y E(MWh (13-97)
u = -.!!. 2
E
where Y2 is the force per mole of pore fluid given by Eq. (13-86), and
C7m/e(MWh is the molar concentration of solute in the pore fluid. The product
of these is the force per volume of the pore fluid. Combining Eqs. (13-86),
(13-92), and (13-97), we obtain
u =- H
E
[.11mpA. +
m fn J2
E(MW)2
1 (13-98)
The solute flux J2 is given by Eq. (13-90). If this equation is substituted into
Eq. (13-98) and the result is solved for u, we obtain
u =- H
E
[ H
1 + (MW)
~23
2E
c2p
j.11mpmA. (13-99)
Note that u is negative because of the chosen direction for x. Solution of Eqs.
(13-91) and (13-99) simultaneously allows prediction of the solvent velocity
and the permeate concentration through the membrane. (See Problem 13-C3).
In Eq. (13-91) C2w is considered to be known since it can be calculated from the
concentration polarization analysis. The solute flux can then be determined
from Eq. (13-90). The solvent flux can be found from Eq. (13-81) which after
substituting in Eq. (13-90), is
(13-100)
J'lOlv = EU Cl p
782
At the end of this chapter you should be able to meet the following objectives.
1. Predict the polarization modulus for both laminar and turbulent flow
when no gel layer is formed using mass transfer correlations.
2. Predict the mass transfer coefficient and the solvent flux for UP when a
gel forms. Determine whether or not a gel does form. Discuss the problems
with the gel formation theory.
3. Explain the mass balances for a completely mixed RO systems with con-
centration polarization. Estimate the solvent flux for this system.
4. Explain the mass balances for countercurrent gas permeation and explain
qualitatively why countercurrent operation is superior to other flow patterns.
5. Use the exact solutions for laminar flow to predict the concentration
polarization modulus in tubes or between flat plates.
REFERENCES
Berman, A.S., "Laminar flow in channels with porous walls," J. Appl. Phys.,
24, 1232 (1953).
Blatt, W.P., A. Dravid, A.S. Michaels, and L. Nelson, "Solute polarization and
cake formation in membrane ultrafiltration: Causes, consequences and control
techniques," in J.E. Flinn (Ed.), Membrane Science and Technology, Plenum
Press, NY, 1970,47-97.
Le, M.S. and I.A. Howell, "Alternative model for ultrafiltration," Chern. Eng.
Res. Des., 62,373 (1984).
Potts, DE., RC. Ahlert and S.S. Wang, "A critical review of fouling of reverse
osmosis membranes," Desalination. 36. 235 (1981).
Rao, G. and K.K. Sirkar, "Explicit flux expressions in tubular reverse osmosis
desalination," Desalination. 27. 99 (1978).
Shen, J.J.S. and RF. Probstein, "On the prediction of limiting flux in laminar
ultrafiltration of macromolecular solutions," Ind. Eng. Chern. Fundarn., 16,459
(1977).
Sherwood, T.K., P.L.T. Brian, RE. Fisher and L. Dresner, "Salt concentration
at phase boundaries in desalination by reverse osmosis," Ind. Eng. Chern. Fun-
darn .• 4. 113 (1965).
Soltanieh, M. and W.N. Gill, "Review of reverse osmosis membranes and tran-
sport models," Chern. Eng. Commun .• 12. 279 (1981).
Yuan, S.W. and A.B. Finkelstein, "Laminar pipe flow with injection and suc-
tion through a porous wall," Transactions ASME. 78.719 (1956).
Zydney, A.L. and C.K. Colton, "A concentration polarization model for the
filtrate flux in cross-flow microfiltration of particulate suspensions," Chern.
Eng. Commun., 47, 1 (1986).
HOMEWORK
A. Discussion Problems
A2. Why are there multiple solutions for concentration polarization in lam-
inar flow?
A3. Theoretical analyses of turbulent flow are much more complicated than
laminar flow analysis. However, the concentration polarization analysis
for turbulent flow is simpler than the exact analysis for laminar flow.
Explain this paradox.
A5. How can you determine if the solution obtained with Eqs. (13-3b) and
(13-16) or (13-18) is reasonable?
786
A6. On one page develop your key relations chart for this chapter. This will
be a challenge because you will have room for only the more important
equations.
C. Derivations
C1. Derive the ratio of k/.1p for both laminar and turbulent flow. Compare
these results.
C3. If a constant b < < 1, then exp (b) - l+b. Using this approximation,
simplify Eq. (13-91) when (- u lmAf21IRn < <1. Then solve Eqs. (13-
91) and (13-99) simultaneously to determine c2p' Note u < O. Note that
this assumption of small b will often not be valid.
C4. Prove that the lowest driving force for a countercurrent gas permeation
module equals the driving force for a completely mixed module.
D. Single-Answer Problems
D7. An experimental gas permeation system with flat plates is being tested.
Rejection, R O = 0.96. The flow path is 100 cm long and the plates are
0.02 cm apart. The high pressure side is at ten atmospheres and the low
pressure side is at one atmosphere. The total gas flux through the mem-
brane is J = 1.0 x 10-3 cc (SzTP). The inlet gas velocity is 100 cm/s.
cm s
Operation is at 25°C and the system is an ideal gas. Find the average
concentration polarization modulus.
D8. We wish to use a cellulose acetate membrane to purify tap water for an
electronics planL The tap water contains 120 ppm IDS (total dissolved
salts) and we wish a product which is 1 ppm or less . .1p = 110 atm.
D9. Merten (1966) reports values for transfer of sucrose through a cello-
phane membrane at 25°C. At infinite dilution, E = 0.67 and
b. ~
Note: Watch your units.
Part IV
distillate rate
L column length. m
p pressure
p permeate rate
R gas constant
793
794
T temperature, K
To surroundings temperature, K
Ts steam temperature
Greek
1t osmotic pressure
Throughout this book we have been discussing mainly known results - we now
move into the unknown area of selection and sequencing of separations. This
is an area which has been extensively studied for distillation and to a lesser
extent for extractive and azeotropic distillation - all equilibrium staged
processes. Very little research has been done for the rate controlled separations
which are the subject of this book. In this chapter we will discuss what has
been done, and extend by analogy the results obtained for equilibrium staged
processes to rate controlled processes. The results obtained are logical, but are
not proven; thus, they should be used carefully.
795
796
However, there are a few abstract ideas which are useful in looking at the
separation problem.
IOr----------------------------------------------,----~
Vitamin 8-1: •
\\ined gold •
Gold from
Penicillin _
sea water
o
• Bromme from sea water
•
Sulfur from
stack gas (cost)
-,
-4 0 12 14
-log (.:on,~ntration. mole fraction)
Figure 14-1. Relation between pure product cost and raw material concen-
tration (Sherwood et ai, 1975). Reprinted with permission
from T.K. Sherwood, RL. Pigford and C.R. Wilke, Mass
Transfer, McGraw-Hill, 1975. Copyright 1975, McGraw-Hill.
798
Chemical
Mechamcal DiffusIOnal Modifications
Electrostatic Separator
T
Emulsion Separator
Heterogeneous
Expression Homogeneous
Filtration Mass Spectrometer
Flotation
High-gradient Magnetic
Impingement Separator
~ Pressure Diffusion
Thermal Diffusion
Ultracentrifuge
Magnetic Electrophoresis
Sedimentation
Scrubber
Non-Equilibrium Equilibrium
E IEF
ITP
Sink-float
!
Pervaporation Molecular
Molecular Sieve Clathration Azeotropic
Gas permeation Distillation
Sublim/Desublim Crystallization Extractive
RO
Precipitation Flash
UF
Zone Melting Reactive
Gas Diffusion
Ion Exchange Steam
Chromatography
Ion Exc'~s'on Vacuum
Affinity
Leaching Evaporation
Capillary
Washing Foam Frac!.
Electrochromatography
Drying Solids Stripping
GLC
Liquid-Liquid
GSC
Extraction
Ion Ex.
Dual Temp. Exchange
LLC
Liquid Membrane
HPLC
LSC 3-phase
PC Dialysis
SEC
TLC
Figure 14-2 has over 70 separation methods listed, and it does not
include all methods. In addition most methods have many variants which are
not listed, and many methods can be done with a variety of solvents, adsor-
bents, additives, or membranes. The result of all these variations is there are
literally thousands of separation methods available if one includes all the varia-
tions. The engineer's job is to select methods which will solve the problem
economically.
The separations listed in Table 14-1 are also marked as to whether they
use an energy (E) or mass (M) separating agent. Separations using an energy
separating agent use heat (e.g. crystallization) or pumping energy (e.g. gas per-
meation) to effect the separation. Separations using a mass separating agent
add an additional species such as a solvent or purge gas to the mixture to be
separated. Note that the insoluble membrane, adsorbent or packing material is
not considered to be a mass separating agent. However, the solvent or carrier
gas in chromatography and the purge gas or displacer in adsorption is a mass
separating agent The disadvantage of adding a mass separating agent is that
usually the mass separating agent has to be removed before the product can be
used. Thus, a separation scheme using an energy separation agent is usually
required in addition to the mass separating agent processes. Separations such
as precipitation or adsorption can use either an energy or a mass separating
agent depending on how they are operated. Separations such as electrochroma-
tography use both energy and mass separating agents. Gradient chromatogra-
phy systems have two or more mass separating agents.
(14-1)
where Xi,1 is the mole fraction of component i in phase 1. Phases 1 and 2 are
assumed to be in equilibrium. For distillation Cli,j is the relative volatility. For
00
Table 14-1. Property Differences for Separations otv
(King, 1980; Null, 1987; Rudd et al., 1973)
Property
Differences Molecules Macromolecule Particles
GLC(M)
flash dist. (E)
vapor distillation (E)
pressure absorption (M)
(reI. volatility) sublimation/desublimation (E)
stripping (M)
evaporation (E)
high vacuum distillation (E)
solubility crystallization
absorption/stripping (M)
leaching (M)
precipitation (E or M)
00
ow
00
Property ~
Differences Molecules Macromolecule Particles
~ isotachophoresis (E) ~
f- LC (M) ~
adsorption (M or E)
~ foam fractionation (M) ~
high gradient
magnetic mag. separation magnetic separator
electrical electrostatic
conductivity separator
~ k~ (14-2)
~.=-=-,
J AJ k.J
The value of the separation factor required before the separation method
is of interest depends on several factors. The ease of multi staging is very
important. In this context multistaging means that the separation acheived can
be multiplied many times and the separation agent can be reused. A packed
bed chromatograph and a packed bed distillation column are both multistage
systems even though no physical stages are present Equilibrium processes
such as distillation, adsorption and chromatography can easily be operated with
many stages, and the separation agent can be reused many times with low addi-
tional capital investment. Thus the separation acheived with a low separation
808
studied. Although energy use may be high, the energy is usually supplied by
low pressure steam which is often very cheap. The net result is that if distilla-
tion can be used it should be considered as one of the alternatives. Guidelines
for when distillation cannot be used are discussed in Section 14.5.
Solution
The minimum reversible work required for the complete separation of a mix-
ture of ideal gases is (Keller. 1982)
(14-5)
If the feed, PF. and product, pp. pressures are equal this result simplifies to
c
(14-6)
Wmin,T,p =-RT 1: (Yi,F In Yi,P)
i=l
(14-7)
Wmin,T,p = -RT ~ [Xi.F In (Yi Xi.F)]
1=1
}----..- I at T,pp
2 at T, Pp
Feed _
PF ,T, Yi,F
1------< } - - - . n at T, Pp
\Membranes
(14-9)
W rev =RT In (Pfinal / Pinitial)
Since Plinal = Pp and Pi,initial = Pi in Eq. (14-8), the reversible work for one gas is
(14-10)
W~v,i = RT In (Pp / Yi,F PF)
The total reversible (and hence minimum) work for the separation is the sum of
Wrev,i for all components. The result is Eq. (14-5). More detailed develop-
ments which include partial separation and nonisothermal operation are given
by King (1980).
Solution
W'nun,T ,p Wmin,prod
xE,F (cal/gmole feed) (cal/gmole EB prod)
0 0
0.1 247.52 2475.2
0.2 381.01 1905.1
0.3 465.12 1550.4
0.4 512.44 1281.1
0.5 527.77 1055.5
0.6 512.44 854.1
0.7 465.12 664.5
0.8 381.01 476.3
0.9 247.52 275.0
1.0 0 0
815
Note that the minimum work required to separate a 0.0 and a 1.0
mole fraction feed are both zero since these feeds are already pure.
The minimum work required for separation per mole of feed is
symmetric around a feed mole fraction of 0.5. For nonidealliquids
this will not be true. The third column in the table is of consider-
able interest since it shows the minimum energy required per mole
of ethylbenzene product. As expected, this value decreaBl!s as the
feed becomes more concentrated in ethylbenzene.
Weq=Q [ T-T
TO 1Ep (14-12)
816
.-
Q)
Q)
0.
E 6
-
0
u
o~
>, c-
uo
c'';::;
.!!:! ~ 4
.5,? ro
:t;c.
Q)Q)
VI
E
~
E 2
'xro
~
Figure 144. Maximum energy efficiency for simple distillation (Ho and
Keller, 1987). Reprinted with permission from Y.A. Liu, H.A.
McGee and W.R. Epperly (Eds.), Recent Developments in
Chemical Process and Plant Design, Wiley-Interscience, 1987.
Copyright 1987, Wiley-Interscience.
where Q is the heat supplied at absolute temperature T and the surroundings are
at To. The efficiency Ep is the power cycle efficiency compared to a Carnot
cycle. From this definition of equivalent work Null (1980) derives approximate
equations for the energy requirements of several commercial separation
processes. The following development is based on Null's (1980) development.
(14-13)
where ~ and Qc are the reOOiler and condenser heat loads, respectively; Dis
the distillate product rate; A. is the latent heat of vaporization of the distillate
product; and RD = LID is the external reflux ratio. Combining Eqs. (14-12) and
817
",0
"
~Q
.,
>
~ ..,
c
"-
0'
Q)
~ N
T s' F (el
Figure 14-5. Equivalent work requirement for normal distillation. To =
37.8°C (Null, 1980). Reprinted with permission from Chern.
Eng. Prog., 76(8),42 (1980). Copyright 1980, American Insti-
tute of Chemical Engineers.
(14-13), we can determine the equivalent work required for simple distillation.
(14-14)
where Ts is the temperature of the stearn used to supply the heat. Note that Ts
> Treboiler since some temperature difference is necessary for heat transfer. To
is the temperature of the cooling water used. The predicted equivalent work
values normalized by DA. are given in Figure 14-5 (Null, 1980). Null (1980)
also presents results for somewhat more complicated distillation columns.
818
Although the relative volatility does not appear explicitly in Eq. (14-14) or Fig-
ure 14-5, a. is an important parameter in determining Weq since a. controls the
required external reflux ratio. The value of Weq for distillation can now be
compared to the results for other processes.
(14-15)
where Be is the bottoms product of melted crystals, At- is the latent heat of
freezing, and RM is the external reflux ratio for the melter. If the heat is sup-
plied to the me Iter at temperature TIM where TIM> TM• then the equivalent
work for the heated end of the crystallizer is
(14-17)
where ER is the Carnot cycle efficiency of the refrigeration system. The total
equivalent work for the melt crystallization process is the sum of Eqs. (14-16)
and (14-17).
W =B (R +1)'L[Ep(TSM-To)+(To-Te)] (14-18)
eq.cry e M "'f T1M T0 E
R
819
(14-20)
Since latent heats of freezing are typically about 1/5 of the latent heat of vapor-
ization, Eqs. (14-19) and (14-20) often show an energy advantage for crystalli-
zation. However, Fair (1987) notes, "The specific energy advantages of cry-
stallization have not been a factor in its selection for commercial use." These
results do show the potential of crystallization particularly if energy costs
escalate.
where Cp•sorb and Cp,B are the heat capacities of the sorbate and the bulk adsor-
bent in kJ/kgmole K and kJ/kg adsorbent K, respectively, and 'lavg is the aver-
age adsorbed phase loading at the end of the feed step. For a symmetric mass
transfer zone,
(14-22)
<lavg = <laat (L - 0.5 LMfZ) / L
820
'l'in Eq. (14-21) is a factor to account for the use of excess regeneration gas,
Tads is the average bed temperature at the end of the feed step, Trcg is the aver-
age temperature of the bed after the regeneration step, Mlads is the heat of
adsorption in kJ/kgmole, Nads is the average rate the adsorbate product is pro-
duced in kgmole/hr. Then; is the total regeneration heat required per kg mole
adsorbed.
(14-23)
where TRG is the supply temperature of the regeneration gas, TRG > Treg. This
result can be compared to Eq. (14-14) for distillation. Adsorption uses less
equivalent work when
(14-24)
Usually, MladJI > A., and hence C;/A »1. Adsorption typically adsorbs the least
volatile componenL Thus, NadJI/D corresponds to BID if either adsorption or
distillation can do a separation. Adsorption will be favored if BID is small
which means that a small amount of nonvolatile material is to be removed.
Adsorption is also favored if distillation is difficult and a large reflux ratio is
required. Null's (1980) numerical results are shown in Figure 14-6.
Reverse osmosis directly uses mechanical energy and the work required
is easily calculated. The mechanical work required is the volumetric flow rate
times the pressure increase. Although the entire feed stream is pressurized, part
of the mechanical work can be recovered from the high pressure waste stream.
Only the permeate product is completely degraded. Assuming that the permeate
pressure and feed pressures (before pressurizing) are the same, we obtain
(14-25)
where P is the volumetric flow rate of permeate product and ~p is the pressure
821
"...0
0
-
N
G)
110
>
-
co
u. co
(/)
..... -
c
.2
Q.
:!
N
0
(/) 0
".c
c(
110
(J
.c co
~
G) .,
>
0
..a N
co
0.1
0
0
Ix: 500 600
200 300 400
(93.3) (149) (204) (260) (316)
Ts ,F(C)
Figure 14-6. Comparison of equivalent work requirements for temperature
swing adsorption with regeneration at 316°C versus simple dis-
tillation (Null, 1980). Reprinted with pennission from Chern.
Eng. Prog .• 76(8), 42 (1980). Copyright 1980, American Insti-
tute of Chemical Engineers.
drop across the membrane. We have also assumed that the efficiency of energy
recovery from the high pressure waste stream is 100%. Obviously, ~p must be
greater than the osmotic pressure difference. From Eq. (12-27) ~p can be
related to the volumetric solvent flux Jso1v = PIA and the osmotic pressure
difference
Ap __
LJ.
PIA
~--'-~- + M 1t (Cb) - 1t (cp) (14-26)
K so1v 11m
where 1t is the osmotic pressure, A is the membrane area, and M is the concen-
tration polarization modulus. Combining these equations, we obtain
(14-27)
822
With reasonable values for fluxes and permeabilities the work calculated
from Eqs. (14-25) or (14-27) is very low when compared to the equilibrium
processes which involve a phase change (note that there is no latent heat term
in these equations). Thus, from an energy point of view reverse osmosis and
the other membrane separations are very appealing. However, RO is not a
complete separation process since the reject stream is a concentrated solute in
solvent, and cannot be pme solute. The concentration of the reject stream is
limited by the osmotic pressure which increases rapidly at high concentrations.
If a complete separation of solute and solvent is required then RO can be
advantageously coupled with another separation method in a hybrid process.
The appropriate strategy to use for process design and synthesis has been
extensively studied recently (Douglas, 1988; Rudd et al, 1973; Westerberg,
1982). We will follow Westerberg (1982) and present some guidelines without
examples. Detailed examples are given in these references, and in Blumberg
(1988) for liquid-liquid extraction.
will be made in the design that the use of complex and time-
consuming detailed calculation methods is unecessary. Douglas
(1988) discusses short-cut calculation methods for equilibrium
staged processes. Detailed calculation methods should be used in
the later stages of design.
5. When trying to prove that concepts will not wolk, use optimistic conditions.
Most initial design concepts will not work. During the screening
process the designer's job is to prove things will not work or can-
not be economic. This type of comparison is convincing only
when optimistic guesses are made (e.g. make the optimistic
guesses that there is no entrainment in a crystallization process or
that there is no concentration polarization in a membrane separa-
tion). If one were conservative in making the guesses, the failure
of the design might be due to the guess and one does not have a
clear reason for rejecting the design.
14.5 HEURISTICS
One problem with heuristics is that different heuristics often conflict with
each other. Then the proper choice of heuristic is unclear. All of the heuristics
should be preceeded with the words, .. IT all things are equal ... " Liu (1987) has
rank ordered heuristics for distillation and modified distillation systems.
In this section a large number of possible heuristics for the selection and
sequencing of separations will be listed and the reasoning for the heuristic will
be explained. These heuristics were developed either by generalizing the
heuristics developed for the equilibrium staged separations or from comments
in the literature. Except for the distillation heuristics, these heuristics have not
been rigorously tested and should always be checked with calculations.
825
20. In developing a flow sheet select the separation schemes first before
adjusting temperatures or flow rates.
This heuristic is stated by Rudd et al (1973) and Kelly (1987). Compared to
temperature and flow rate adjustment separations are much more expensive and
require more energy. Thus, the separation problems should receive priority.
8G. Favor separation methods which are known within the company.
The separation methods which the company is familiar with will require the
least design and development time. These methods will also be easiest to sell
to management. Heuristic #lG should not be forgotten when applying this
heuristic.
9G. For new products look for separations which can be put on stream quickly.
With new products the company which gets to the market first often captures a
large proportion of the market Thus. separation schemes which will take a
long time to develop are not favored for the first plant Known separation
processes will be favored (See heuristics 7G and 8G).
1OG. Consider hybrid plants when one separation method by itself is not ideal
for a given separation problem.
Hybrid plants use two or more separations for the same problem. An example
would be concentrating a stream by reverse osmosis before doing evaporation
even though evaporation by itself could do the separation. Hybrid plants can
utilize the strength of a separation in the concentration range where it is best,
and avoid weaknesses.
2D. If a mass separating agent is used remove it early and do not use a mass
separating agent to recover a mass separating agent.
827
This heuristic is commonly stated (Kelly, 1987; Liu, 1987; Nath and Motard,
1981; Rudd et aI, 1973). It probably does not strictly apply when solutes must
be dissolved in a solvent so that they can be processed. Rudd et al (1973) also
suggest considering the use of a chemical already present in the process even if
it is not optimum based on its properties alone.
3D. Prefer processes with only one and not two mass separating agents.
The use of an additional mass separating agent such as a gradient in chroma-
tography makes the downstream recovery more difficult However, Mazsaroff
and Regnier (1986) note that gradients are required for elution of
macromolecules when separation is due to surface forces.
6D. Avoid solids and if they are present remove them early.
Solids cause processing problems. Separations involving solids (e.g. adsorp-
tion and chromatography) are difficult to operate as countercurrent processes.
828
70. When a solid material needs to be processed consider putting the sample
into solution and removing insolubles to be a separation step.
This heuristic is explicitly stated for chromatography by McDonald and
Bidlingmeyer (1987), but can be extended to any process such as extraction,
crystallization, or adsorption where solids are first dissolved.
5C. If component compositions are similiar, favor 50/50 splits or equal size
parts.
This was also derived for distillation and serves to balance the distillation
column (Douglas, 1988; Kelly, 1987; Liu, 1987; Rudd et aI, 1973). Its use for
other separations is not clear.
3S. If the product is sold as a solid the last purification step should probably be
crystallization.
Crystallization can be used both to pnxluce crystals with the desired size distri-
bution and as a final polishing step.
b. An azeotrope forms.
c. lightgases
g. there are two sets of compounds with boiling point overlap. Note that
extraction is also commonly used in this case.
This set of heuristics is from Keller (1982).
6S. Consider displacement adsorption only when a.ds > 2 and (ldist < 1.2 to
1.3.
Keller (1982) notes that displacement adsorption processes are quite complex,
and the process should be considered as a last resort.
7S. In adsorption and chromatography consider methods with cheap pac kings
first
This heuristic is related to heuristic 5D. For adsorption this heuristic favors
activated carbon. For chromatography it favors plain silica, activated
alumina, or ion exchange chromatography systems.
831
lOS. In large scale liquid chromatography always optimize the solvent sta-
tionary phase system.
McDonald and Bidlingmeyer (1987) note that the selectivity ex is the most
important parameter, and increasing ex by optimization is well worth the
effort. However, Kopaciewicz and Regnier (1983) note that solvents need to
be compatible for chromatography columns in series or an extra separation
will be required between columns. Solvent selection is also very important in
absorption (Keller, 1982), azeotropic and extractive distillation (Van Winkle,
1967; Wankat, 1988), and extraction ( Blumberg, 1988; King, 1981).
As was noted at the beginning of this section many of these heuristics have
not been rigorously tested. In addition, an ordering of these heuristics for
nondistillation applications is needed. This implies that considerable research
is needed. In the meantime the heuristics must be used with caution.
Solution
Heuristic 5S suggests TSA may have high energy costs for break-
ing the azeotrope. An alternative such as PSA should be con-
sidered.
Heuristic 9S probably rules out chromatography.
Heuristic 12S would be applicable if solids were present in the
feed.
The net result are the flow sheets shown in Figure 14-7. Figure
l4-7a is a fairly standard distillation column followed by azeotro-
pic distillation. Reverse osmosis is optional as a preconcentrator
(Note: because of the unusual shape of the VLE curve this precon-
centration has little effect on the reflux ratio RD' Then according
to Eq. (14-13) 0& is not changed much by preconcentration.
Although suggested by the heuristics RO is probably not useful for
ethanol-water and is not shown on the other flow sheets. Precon-
centration by RO can be useful for other systems.) Figure l4-7b
uses distillation followed by liquid phase adsorption. This requires
a drain step, evaporation of liquid, and then thermal regeneration.
Since energy requirements will be quite high the vapor phase
adsorption alternative shown in Figure 14-7c should be considered.
The adsorption can be done as TSA or PSA. TSA is done com-
mercially, but from heuristic 5S we should also consider PSA.
This will require operating the distillation column under pressure.
This may be advantageous since a smaller diameter distillation
column can be used although equilibrium is less favorable.
a
Water + EtOH
Azeotropic
Distillation
EtOH
Water
b EtOH
Drain
Adsorption
F
c EtO
nT
PSA or TSA
Regeneration
Go.
F
835
1. Generate each possible neighbor using the two rules. This will
result in a lot of possibilities.
It should be evident that this method can involve a lot of work. Thus, one
needs to use short-cut estimation methods in the evaluation of flowsheets.
These heuristics are used in the order listed to generate the initial fiowsheet.
The following evolutionary rules are then applied:
Liu (1987) notes that rules 1,3, and 4 may be very valuable in developing less
expensive designs.
These evolutionary rules can be applied in a variety of ways. The stra-
tegy chosen by Nath and Motard (1981) makes this a global method. Evolu-
tionary rule 1 is done first. Then rules 2 to 5 are treated equally. When a
838
At the end of this chapter you should be able to meet the following objectives.
5. Use heuristics to develop an initial flowsheet Discuss the rationale for each
heuristic.
7. List and discuss the areas which need further research in selecting a separa-
tion method.
REFERENCES
Bravo, J.L., J.R. Fair, J.L. Humphrey, CL. Martin, A.E. Seibert and S. Joshi,
Fluid Mixture Separation Technologies for Cost Reduction and Process
Improvement, Noyes Publications, Park Ridge, N.J., 1986.
Evans, FL., Jr., Equipment Design Handbook for Refineries and Chemical
Plants, 2nd ed., Gulf Pub. Co., Houston, TX, 1980.
Garg, D.R. and J.P. Ausikaitis, "Molecular sieve dehydration cycle for high
water content streams," Chern. Eng. Prog .• 79 (4) 60 (1983).
Hendry, J.E., D.P. Rudd, and J.D. Seader, "Synthesis in the design of chemical
processes," AIChE Journal. 19, 1 (1973).
Ho, F.-G. and G.E. Keller, II, "Process integration", in Y.A. Liu, H.A. McGee,
and W.R. Epperly (Eds.), Recent Developments in Chemical Process and Plant
Design, Wiley-Interscience, New York, 1987, Chapt 4.
Jacobs, L.J., Jr., and W.R. Penney, "Phase Segregation", in R.W. Rousseau
840
King, CJ., Separation Processes, 2nd ed., McGraw-Hill, New York, 1980.
Perry, R.H. and D.W. Green (Eds.), Chemical Engineers' Handbook, 6th ed.,
McGraw-Hill, New York, 1984.
Robinson, C.S. and E.R. Gilliland, Elements of Fractional Distillation, 4th ed.,
McGraw-Hill, New York, 1950.
Rudd, D.F., G.J. Powers, and J.J. Siirola, Process Synthesis, Prentice-Hall,
Englewood Cliffs, NJ., 1973.
Seader, J.D., "Distillation," in RH. Perry and D.W. Green (Eds.), Perry's
Chemical Engineers' Handbook, 6th ed., McGraw-Hill, New York, 1984, Sec-
tion 13.
Seader, J.D. and AW. Westerberg, "A combined heuristic and evolutionary
strategy for synthesis of simple separation sequences", AIChE Journal, 23, 951
(1977).
Sherwood, T.K., RL. Pigford and C.R. Wilke, Mass Transfer, McGraw-Hill,
New York, 1975, Chapt. 1.
Tedder, D.W. and D.F. Rudd, "Parametric studies in industrial distillation. Part
I. Design Comparisons", AIChE Journal, 24,303 (1978).
Westerberg, AW., "Design research. Both theory and strategy", Chem. Eng.
Educ., 16, 12 (1982) and 16,62 (1982).
HOMEWORK
A Discussion Problems
a. Minimum work
843
b. Equivalent work
c. Heuristic
d. Evolutionary method
e. Local evolution
f. Global evolution
g. Separation factor
h. Diffusional separations
i. Mechanical separations
j. Hybrid process
k. Neighbor
j. Top-down!Bottom-up
A2. When one is developing new separation processes, distillation is often con-
sidered "the enemy". Explain why.
A3.Heuristics can be misleading. For example, heuristic 4D could be mislead-
ing when high pressure gas streams are processed. Explain why one prob-
ably wants to do the separation at high pressure.
A5.Explain the differences and similarities between local and global evolution.
A7.0n Figure 14-1 "Water from sea water", vitamin B-12, and "gold from sea
water" all falloff the empirical straight line. Hypothesize why each of
these three is different
844
AlO. What are "optimistic conditions" for comparison of the following separa-
tion processes?
a. zone melting
b. UP
c. ED
d. LC for protein purification
a. crystallization
b. adsorption
c. membrane separations
A12. Westerberg (1982) suggests that theories evolve from simple to complex.
Explain how one can do this for
a. Chromatography
b. Reverse osmosis
c. Solution crystallization
B. Generation of Alternatives
B1. Hybrid processs allow the designer to stretch his/her imagination. Mem-
845
B2. Suppose you work for a company whose business is selling processes.
Brainstorm and then pare down a list of separations which should be
explored further by R&D. Explain your final list of 5 items.
B3. Generate possible separations to recover acetic acid from a dilute aqueous
solution.
C. Derivations
C1.Develop a local evolutionary process which uses Nath and Motard's (1981)
evolutionary rules.
C4.Many of the heuristics in this chapter were developed from material in pre-
vious chapters. Generate two additional S heuristics based on information
in previous chapters.
D. Problems
Assume the melt product is essentially at the eutectic composition (13 mole
% p-xylene). Pressure is 101.3 kPa Determine the ideal equivalent work
(ep = eR = 1) for melt crystallization if a lOoC approach temperature is used
in the melter and the crystallizer. The ambient temperature To is 18°C.
Melter reflux ratio RM = 0.5. Use a feed rate of 100 kglhr as a basis.
(perry and Green, 1984, p. 3-123) Ar,M =26.045 calIg, Ar,p =38.526 cal/g
D3.Find the equivalent work in calI(kg product) required for the RO system in
=
Example 12-2b for M 1.2. Watch your units. Use a water density of 1
kg/liter.
Fl. Different chromatographic methods used for protein purification have dif-
ferent solvent requirements. Read Kopaciewicz and Regnier (1983) and
develop heurispcs for sequencing chromatography columns for protein
purification.