Sei sulla pagina 1di 13

Desalination, 58 (1986) 199--211

199

Elsevier Science Publishers B.V., Amsterdam -- Printed in The Netherlands

NEW I N S I G H T S I N T O T H E S T R U C T U R E O F M I C R O P O R O U S
M E M B R A N E S O B T A I N E D U S I N G A NEW P O R E
EVALUATION METHOD
S

M.G. KATZ and G. BARUCH


Radiation Chemistry Department, Soreq Nuclear Research Centre, Yavne (Israel)
Tel. (08) 484418; Telex 03916

(Received October 30, 1985; in revised form January 9, 1986)

SUMMARY
T h e p o r e size d i s t r i b u t i o n o f the active pores o f Millipore MF-VS and
MF-VM m i c r o f i l t r a t i o n m e m b r a n e s , having n o m i n a l p o r e sizes o f 250 )k and
500 A, respectively, was investigated b y a new evaluation m e t h o d described
earlier. T h e average values o f t h e active p o r e sizes calculated f r o m t h e p o r e
size d i s t r i b u t i o n s o b t a i n e d b y this m e t h o d were f o u n d t o be systematically
smaller t h a n t h e n o m i n a l p o r e sizes i n d i c a t e d b y t h e m a n u f a c t u r e r . These
d i f f e r e n c e s are discussed in t e r m s o f t h e applicability and range o f validity
o f t h e gas p e r m e a b i l i t y m e t h o d w h i c h is f r e q u e n t l y used f o r evaluation o f
t h e average p o r e size o f m i c r o f i l t r a t i o n m e m b r a n e s .
active pores, p o r e size d i s t r i b u t i o n , m i c r o p o r o u s m e m b r a n e s :
measurement method.
Key words

SYMBOLS
A
D
de
F
k0
kl
m
P
P
P0

sample area (cm 2)


-- Kelvin p o r e d i a m e t e r = 2r (cm)
- - effective d i a m e t e r o f e q u i v a l e n t capillary [11] (cm)
v o l u m e t r i c gas f l o w rate (cma/s)
c o n s t a n t [11]
see 5 / k I b e l o w
m e a n h y d r a u l i c radius [11] (cm)
- - p e r m e a b i l i t y c o e f f i c i e n t = F ~ / A A p (cm/s)
m e a n absolute pressure ( d y n e / c m 2)
- - v a p o r pressure o f c o n d e n s e d liquid in the capillary ( d y n e / c m 2 )
v a p o r pressure o f liquid at a planar surface ( d y n e / c m 2 )
-

0011-9164/86/$03.50

1986 Elsevier Science Publishers B.V.

200
Ap
q2

pressure drop across the tested sample (dyne/cm 2)


-- tortuosity
R
gas constant (dyne c m / K mol)
r
Kelvin pore radius (cm)
T
absolute temperature (K)
average molecular velocity of gas (cm/s)
V
-- molar volume of condensed liquid (cm3/mol)
-- surface tension of condensed liquid (dyne/cm)
7
constant [11]
51kl
-- porosity
0
-- contact angle (degree)
-- molecular mean free path (cm)
-- viscosity (poise)
-

INTRODUCTION
The evaluation of the sizes of the active pores of selective microporous
membranes is of cardinal importance in the understanding of the selective
transport mechanism through such membranes. Active pores are defined
as those pores in the dense, active skin layer of a membrane in which the
selective transport process takes place. This study was undertaken in order
to evaluate the pore size distribution of the active pores o f Millipore microfiltration membranes using a new, recently developed technique [1--3] and
to compare the results with corresponding values obtained by the "gas
permeability m e t h o d " [4, 5]. The gas permeability m e t h o d is a frequently
used technique; however, by this m e t h o d , only the average pore size can be
obtained, rather than a pore size distribution. This m e t h o d also has certain
other limitations, discussed below, which must be understood in order to
assess correctly the significance of results obtained by it. The " w e t and dry
flow m e t h o d " [6] is the only known technique so far which is suitable in
principle for the evaluation of pore size distributions of active pores in
membranes. Unfortunately, this m e t h o d is practical mainly for the
characterization of macropores, i.e. pores in the micron size range down to
"~ 300 A. The characterization of mesopores (size range between 20--500 A)
which are the majority in most of the known microporous selective
membranes, is very difficult, if at all possible, by this experimental method.
In the preceding studies [1, 2], the feasibility of the m e t h o d presented in
this work was verified and the optimal conditions for performing the
measurement were investigated. The m e t h o d itself is based on the wellk n o w n p h e n o m e n o n of capillary condensation of liquids in micropores,
which is the basis of one of the most popular methods for characterization
of the pore size distribution in porous media [7]. It is k n o w n that the vapor
pressure of a liquid is dependent on the radius of curvature of its surface.

201

When a liquid is contained in a capillary tube, this dependence is given by


the Kelvin equation [8].
RTlnp/po

= 2")'V (cos O)/r

(1)

For the sake of simplicity, it is assumed that the contact angle 0 is zero.
Accordingly, the radius of curvature of the liquid surface is equal to the
radius of the capillary. It is evident from this equation t h a t the vapor pressure
of the liquid in a capillary increases as the radius of the capillary ihcreases.
By the classical m e t h o d [ 7 ] the investigated porous substrate is equilibrated
with known pressures of condensible vapors, and from the adsorption-desorption isotherms obtained, the pore size distribution profile of the
investigated substrate can be derived. Usually liquid nitrogen is employed
for these adsorption--desorption isotherms. This m e t h o d is n o t adequate
for evaluation of the active pore sizes of microporous membranes because
no distinction is made between the active pores in the dense layer of the
membrane and the background pores in the porous supporting layer.
The new m e t h o d employed in this study is based on this same principle,
with the difference that, after equilibrating the investigated membranes with
k n o w n pressures of the condensable vapors, instead of measuring the amounts
of vapors adsorbed or desorbed from the membrane during the equilibration
process, the permeability coefficient of an incondensable gas through the
membrane is measured. Since this permeability coefficient is related
exclusively to the active pores, its dependence on the pressure of the vapors,
in equilibrium with the investigated membrane, will yield the pore size
distribution of only the active pores.
EXPERIMENTAL
The measurement m e t h o d itself and the experimental set-up employed
in these measurements have been previously described [2]. The condensable
"gas" used was water vapor and the incondensable gas was air.
Two VS-type membranes (nominal pore size 2 5 0 A ) and one VM-type
membrane (nominal pore size 500 A) were investigated at 80 mm Hg absolute
pressure (the pore size distribution of one of the VS membranes was also
determined at 40 mm Hg).
In order to enhance the process of equilibrium between the water adsorbed
on the pores and the water vapor in the gas phase, the membranes were
treated with a dilute aqueous solution of glycerol and subsequently dried.
The experimental permeability coefficients P were determined by dividing
the permeating gas flow rates, normalized to atmospheric pressure F ' ~ ,
by sample area A and by the pressure drop Ap across the tested sample.
It should be n o t e d that the permeability coefficient thus obtained is not
normalized for sample thickness (as the length of the active pores is not
k n o w n precisely) and as a result its units are cm/s instead of the usual cm2/s.

202

o~

Sample VS(KI)2
p = 40 mm Hg
run no. 7
run no. 8
run no. 9

4.0
~

o
o

;unno:

Io

run no. 2

o.

3.0

2.0

Sample VS(gl) 1
p = 80 ram Hg
o

1.0

-~

Sample VM(gl) i

v
-~

Sample VS(gl)2
o

p = 80 mm H~

4.0

\ ~ \ \ ~

run no. 13

run
run
run
run
run

o
o
O

no.
no.
no.
no.
no.

12
14
16
15
17

p = 80 mm Hg
run no. 4
run no. 5

o
o

3.0

2.0

1.0 _

\\

'e,
l
I00

I
200

I'"~
300

3,

I
I00

I
200

I
300

D (A)

Fig. 1. The d e p e n d e n c e o f t h e p e r m e a b i l i t y t o air c o e f f i c i e n t P o n t h e size D o f t h e


largest p o r e s c o n t a i n i n g c o n d e n s e d w a t e r (or t h e smallest d r y p o r e s ) : C u m u l a t i v e p o r e
size d i s t r i b u t i o n profiles. In all t h e runs, t h e e q u i l i b r i u m t i m e s o f t h e t e s t e d s a m p l e s w i t h
t h e gas p h a s e w e r e o f 3 0 m i n o r m o r e . In r u n 15 a n d part o f r u n 16, t h e e q u i l i b r i u m
t i m e s w e r e limited t o 15 min, as a result o f w h i c h t h e p e r m e a b i l i t i e s o b t a i n e d ( s h o w n
along t h e b r o k e n line) w e r e l o w e r t h a n t h e e q u i l i b r i u m values.

203

RESULTS

AND DISCUSSION

The results obtained in this study are summarized graphically in Figs. 1


and 2. The desorption isotherms shown in Fig. 1 were obtained by plotting
the permeability coefficient P vs. the size D of the largest pores containing
condensed water (or the smallest pores which are already "dry"). These
plots constitute cumulative pore size distribution profiles of the investigated
membranes. The pore size frequency distributions shown in Fig. 2 were
derived from these cumulative distribution profiles by graphical derivation
and the bar graphs by the difference method, in the same fashion as described
earlier [ 1, 2 ].
The pore size distribution profiles obtained in this work indicate that no
significant difference is observed between VS-type and VM-type membranes
and that the most abundant pore sizes in both types are in the range of
8 0 - - 1 2 0 A . The average active pore sizes derived from these distribution
Sample VS (el)1
3.0 ~-

/'-",

i'. 0

run

no.

<D>

2
125

,[ l'i
"

3.0

/-",
/
/,

2.0

no.
3
<D> = 118 A

run

Sample VS(~l)2

1.0
//

c~
,o
I
o
x

Sample VS($i)2
5.0

4.0

3 0 -

/'L

"\,\

r u n no,

-T

....

-~

run

i0

no,

[~

' ,

/"

1.0

L,I

run

"',

<D> = 169 A

/-"--.. r~n.o. s

2.0

no.
9
<D> = 163 A

",

[/ [

'

l'0i"

],/]

[/' " "

ff

2.0

]
~

I,i00 I

I,

200

300
D (A)

Fig. 2. C o n t i n u e d o n p. 2 0 4 .

run no.

11

",

100

200

. . . . . I. . . . .
300

204
Sample VS(gl)2
..../
~,
,/
', ,

3"01-|

L'

Sample VM($1)I
F
|
L

run no. 6

~
,q
/I

/ /I

<0>o,ssA

run no. 12
<D> = 162 A

~',
I "-..

//,'

""]--

,
,
. . . . .

'o

3.o~.

2.0

"',

....
no. 5

..... [
~_

<D>=lS2A

/]

run
<D> = ,89 A

il

~'".L_

-......
I -'"

.--~.
'.0

~-I,/ l
I.Y

~.~.
I

I ........

~-

L/

I".

run no. ,4
<D> = 172 A

.,
F
3.0 ~-

. ~
/ I \\ I
l

--, . . . . .

i
~-[--'----~.'-'~
~

run no. 4
<D> = '54 A

run no. 'S


= 22 A

2.0

I.o
I'1

I
i00

I
200

I ......

r ....

I1

300

i
i00

~-~~ I'"" .....


200

7"--r"

300

D (A)
Fig. 2. Pore size f r e q u e n c y diagrams. T h e b a c k g r o u n d curves s h o w n in b r o k e n lines w e r e
derived graphically f r o m the cumulative distribution profiles s h o w n in Fig. I.

profiles are in the range of 120--190 A for the VS membranes and 160--190 A
for the VM membrane. These results are surprising considering the nominal
pore sizes of 250 A and 500 A given by the manufacturer for VS- and VMtype membranes, respectively [9].
Comparison of the pore distributions obtained for the VS membranes
in this work with the corresponding profiles determined earlier [2] indicates
that, in contrast to the present results, in the previous VS membrane the
most a b u n d a n t pores were f o u n d to be in the range of 180--200 A and the
average active pore size derived was in the range of 210--240 A which was
in reasonable agreement with the nominal value of 250 A. These findings
are summarized in Table I. On the other hand, it should be noted t h a t there
is fair agreement between the experimental and " n o m i n a l " dry permeability
data. Incidentally, in the VS 1 membrane, investigated earlier, where
reasonable agreement with the nominal pore size was determined, the dry
permeability was f o u n d to be lower, by almost an order of magnitude, than
the value suggested by the manufacturer and those evaluated in this study.
In the following discussion, an a t t e m p t is made to understand these
discrepancies. First, the disagreement between the data obtained for the
VS 1 membrane [2] and the data obtained in this study for the VS

205
TABLE I
PORE SIZES AND DRY PERMEABILITIES OF AIR IN MILLIPORE MICROFILTRATION MEMBRANES
Pore Size (/~)

VSI*

VS(gl)I

VS(gl)2

VM(gl)I

Nominal
(given by the manufacturer)

250

250

250

500

Average
(determined by the present method)

210--240

120--125

155--194

162--189

Most abundant
(determined by the present method)

180---200

80--120

80--120

80-120

Dry permeability
(cm/s)
Nominal
(based on manufacturer's data at
atmospheric pressure)
at 80 mm Hg
at 40 mm Hg

3.65

3.65

3.65

4.87

0.49
0.67

3.0
--

4.4
5.0

4.9
--

* v s 1 represents the untreated VS membrane investigated earlier [ 2 ].

m e m b r a n e s , will be addressed. T h e n , t h e d i f f e r e n c e b e t w e e n o u r d a t a a n d
t h e d a t a p r o v i d e d b y t h e m a n u f a c t u r e r will be discussed.

Results o f this study vs. those obtained earlier [2] for the VS membranes
I t was f o u n d in t h e c o u r s e o f o u r e x p e r i m e n t s t h a t bringing t h e m e m b r a n e s
i n t o e q u i l i b r i u m w i t h w a t e r v a p o r was n o t as simple a n d s t r a i g h t f o r w a r d a
p r o c e s s as it s e e m e d a f t e r initial e x p e r i m e n t s [ 2 ] . In t h a t s t u d y , t h e o p t i m a l
c o n d i t i o n s f o r e v a l u a t i n g t h e p o r e size d i s t r i b u t i o n s b y this m e t h o d w e r e
investigated a n d a single s a m p l e o f V S - t y p e m e m b r a n e was u s e d t h r o u g h o u t
t h e s t u d y . We f o u n d t h a t t h e s a m p l e c o u l d b e b r o u g h t i n t o an e q u i l i b r i u m
s t a t e w i t h given partial pressures o f w a t e r v a p o r t h r o u g h a d s o r p t i o n (i.e.
w e t t i n g b y e q u i l i b r a t i o n w i t h h u m i d gas) a l m o s t as well as b y d e s o r b t i o n
(i.e. drying).
H o w e v e r in later studies, w h e n n e w VS s a m p l e s w e r e i n t r o d u c e d f o r
m e a s u r e m e n t , it was f o u n d t h a t t h e y c o u l d n o t be w e t t e d b y e q u i l i b r a t i o n
w i t h w a t e r v a p o r - s a t u r a t e d air. T h e n e w m e m b r a n e s , as s u p p l i e d b y t h e
m a n u f a c t u r e r , c o u l d be w e t t e d o n l y w h e n b r o u g h t i n t o c o n t a c t w i t h liquid
w a t e r , t h r o u g h capillary s o r p t i o n . T w o o b s e r v a t i o n s w e r e m a d e at this stage:
(a) T o bring t h e s e m e m b r a n e s t o d e s o r p t i o n e q u i l i b r i u m , i.e. equilibrating
t h e w a t e r - s o a k e d m e m b r a n e s w i t h a given partial pressure o f w a t e r v a p o r ,
was also v e r y difficult, if n o t i m p o s s i b l e , a n d t h e results o b t a i n e d w e r e n o t
reproducible.

206
(b) After numerous cycles of wetting and drying o u t of the membranes, t h e y
started to behave " p r o p e r l y " , and they could be brought into equilibrium
with the gas phase. The problem again was that reaching this situation with a
given membrane was a lengthy procedure which could take several days at
least and its use was impractical. Later, it was f o u n d that, upon treatment
of the membranes with dilute aqueous solutions of glycerol, they became
"well behaved" regarding the readiness to reach equilibrium with the gas
phase.
It was contemplated that, as a result of the glycerol solution treatment,
the active pore size distribution of the membranes could be altered due to
the possible presence of condensed glycerol in the pores. In order to verify
this, the dry permeabilities of several untreated VS membranes were also
measured by our m e t h o d at 80 mm Hg absolute pressure and found to be in
the range of 3.2--4.3 cm/s. The fair agreement between these values and the
dry permeabilities of the glycerol-treated membranes (shown in Table I)
can be taken as an indication that no liquid glycerol was present in the active
pores of the treated membranes, and the glycerol was totally adsorbed or
dissolved in the membrane material as a plasticizer. This glycerol content
was found, however, to very effectively enhance the readiness of the treated
membranes to reach equilibrium with water vapor.
One of the possible explanations for this observation may be that the
contact angle between the liquid water and the glycerol-swollen membrane
material is closer to zero than in the case of the untreated membranes.
Alternatively, a monolayer of glycerol could have been deposited on the
pore walls, drastically reducing the wetting angle, but w i t h o u t significant
effect on the pore dimensions.
The actual values of the contact angles between water and the membrane
material were not determined experimentally. On the investigated microporous membranes such a measurement was n o t practical due to the lack
of a constant and well-defined boundary between the liquid phase and the
membrane material. It is known, however, according to the Young--Laplace
equation [10], that the capillary-suction pressure of a liquid increases with
decreasing contact angle, reaching a maximal value at zero contact angle.
It is also reasonable to expect that the speed of capillary rise of a liquid is
proportional to the capiUary-suction pressure (provided all other parameters
and conditions are identical). Accordingly, the speed of capillary rise in
different porous substrates having the same pore sizes may be taken as an
indication of the relative values of the contact angles between the liquid
and the substrate. If, for example, the glycerol treatment of the membranes
lowers the contact angle between the water and the membrane material,
it is expected that the glycerol-treated membranes will absorb water faster
than untreated membranes of the same kind. The time it takes a drop of
water to soak through a dry membrane, referred to as the "soaking t i m e " ,
was measured. This soaking time was measured by applying a drop of water

207

on one side o f a dry, o p a q u e m e m b r a n e and watching the m e m b r a n e from


its o t h e r side. At the m o m e n t t h a t the c o n t o u r of the wet stain started t o
become visible from the ot he r side, the m e m b r a n e was considered to be
soaked. This m e a s u r e m e n t was p e r f o r m e d on glycerol-treated and unt reat ed
membranes, as well as on membranes which were soaked previously with
water or n-hexane and subsequently dried overnight in a vacuum oven. The
results obtained for Millipore VS- and VM-type microfiltration membranes
are summarized in Table II. The results confirm, as expect ed, that, as a result
o f the glycerol t r eat m e nt , the soaking time o f the membranes was reduced
considerably, suggesting that the cont act angle o f the water with the pore
walls was reduced. It can be seen f r o m the data obtained t hat the mere
t r e a t m e n t with water also reduced the soaking time o f the membranes
significantly b u t n o t as m uc h as the glycerol treatment. T he t r e a t m e n t with
n-hexane did n o t change the soaking time o f the membranes significantly.
If we accept these results as sufficient evidence for the assumption that the
c o n t a c t angle o f the water with the m e m b r a n e material is higher than zero
and its value is reduced significantly by the glycerol t r e a t m e n t , the discrepancy
between the pore sizes evaluated earlier f or unt reat ed VS membranes [2]
and the pore sizes obtained in this study is predicted by the terms o f the
Kelvin equation. This equation shows that the calculated pore sizes are
p ro p o r tio n al to the cosine of the c o n t a c t angle and incorrectly assigning a
zero value to the c o n t a c t angle in the u n t r e a t e d membranes led to exaggerated
pore size evaluations.
TABLE II
TIME (s) REQUIRED TO SOAK UNTREATED AND VARIOUSLY TREATED MILLIPORE MICROFILTRATION MEMBRANES WITH WATER (TYPES VS AND VM)
Membrane type
and treatment*

Application of water drop on


porous side

dense side

Untreated VS-type (as received)


VS-type with n-hexane
VS-type with water
VS-type with glycerol

9.12.1
8.40.8
2.40.4
1.760.13

13.61.3
11.31.1
4.50.6
2.420.45

Untreated VM-type (as received)


VM-type with n-hexane
VM-type with water
VM-type with glycerol

7.80.7
5.60.5
1.140.16
<0.5

16.01.1
11.50.6
4.10.3
0.980.15

*The m e m b r a n e s were s o a k e d w i t h n - h e x a n e , water and 1% weight glycerol in wateI


solution and subsequently dried overnight in a vacuum oven.

208
N o m i n a l p o r e sizes vs. the evaluated p o r e sizes
The considerable difference between the present results and the data
suggested by the m a n u f a c t u r e r regarding t h e pore sizes (see Table I), could
n o t be explained as being due to the higher c o n t a c t angle. Since comparison
o f our results with the data suggested by the m a n u f a c t u r e r regarding the dry
permeability coefficients of these membranes indicates good agreement,
these discrepancies can also n o t be explained by differences between various
p r o d u c t i o n lots. These differences are discussed below in terms of the
different principles on which the m eas ur em ent m et hods, e m p l o y e d for the
evaluation of the pore sizes, are based.
The m a n u f a c t u r e r does n o t specify t he m e t h o d e m p l o y e d for the evaluation
o f the nominal pore sizes. Since no i nf o rm at i on regarding the pore size
distribution is provided, it m a y be assumed t hat t he pore sizes were evaluated
by the gas permeability m e t h o d . By this m e t h o d the permeability coefficient
o f the m e m b r a n e to a given gas is evaluated at various values o f absolute
pressure and is treated in terms o f the Carman equation [ 1 1 ] .
4eSm
P -

em 2

3 k l q 2 ~+--koq2~? ~

(2)

This equation describes the permeability coefficient as com posed of two


terms: the c o n t r i b u t i o n o f the molecular (Knudsen) flow mechanism, which
is i n d e p e n d e n t o f the absolute pressure, and t he c o n t r i b u t i o n o f the laminar
flow mechanism, which is pr opor t i onal to the absolute pressure. F r o m the
linear plot o f P vs. ~ t he average pore radius value m m a y be derived.
One o f the basic assumptions o f this m e t h o d is t hat the evaluated pore
sizes are uniform, t hat is, the pores do n o t comprise a m i xt ure o f pores of
various sizes. It can be shown algebraically t hat by this m e t h o d the average
pore size o f an assembly of pores o f distributed sizes will always he biased
upwards [ 1 2 ] . The e x t e n t of the inaccuracy of the result obtained by the
permeability m e t h o d depends on the e x t e n t of divergence o f the pore sizes
in the evaluated membrane, i.e. t he average pore size obtained will be biased
upwards to a greater e x t e n t f or wider pore size distributions and to a lesser
e x t e n t for narrower pore size distributions.
It is k n o w n that the sizes of the active pores o f the Millipore microfiltration membranes which are t he subject o f this discussion are n o t uniform.
There is a lot of experimental evidence which indicates a rather wide pore
size distribution in these membranes. F r o m the bubble points given by the
m a n u f a c t u r e r [9] m a x i m u m values of 830 and 1 1 0 0 A can be evaluated
for VS- and VM-type membranes, respectively. The pore size distribution
profiles obtained by our m e t h o d also indicate t hat the pore sizes of the
investigated membranes were distributed over a rather wide range. Accordingly, in such membranes, the average pore size values obtained by the gas
permeability m e t h o d must he biased upwards and this m ay provide an

209
explanation for t he discrepancy between o ur results and the nominal pore
sizes.
The lack o f a difference between the pore size distribution profiles
obtained by o u r m e t h o d for VS- and VM-type membranes, m ay be considered
in the light o f several studies. Yasuda and Tsai [4] evaluated the average
pore sizes o f VS and VM membranes by the gas permeability m e t h o d and
obtained 340 and 4 7 0 A , respectively, which is a significantly smaller
difference ( ~ 40%) than the f a c t or 2 difference claimed by t he manufacturer.
Smolders et al. [5] evaluated the average pore size of VM membranes by the
same m e t h o d to be a b o u t 380- - 400 A which brings t he difference bet w een
the pore sizes o f the VS and VM membranes dow n to a b o u t 15%. These data
suggest that the nominal pore sizes of the VS and VM membranes given by
the m a n u f a c t u r e r c a n n o t be considered to be accurate. The evaluation o f t he
pore size distribution of these membranes by our m e t h o d suggests t h a t in
the pore size range 0--300 A there is no significant difference between the
pore size distributions of these two types o f membranes. It is possible t h a t
the larger pore sizes, which are inaccessible by our m e t h o d , are m ore a b u n d a n t
in the VM-type membranes than the VS membranes and this is the reason
b oth for the higher average pore sizes obtained for the VM-type m em branes
by the gas permeability m e t h o d and for the higher maximal pore sizes
determined in these membranes by t he bubble poi nt m e t h o d .
A d d i t i o n a l c o m m e n t s o n the gas p e r m e a b i l i t y m e t h o d
Besides that already discussed, there is anot her limitation o f the gas
permeability m e t h o d t hat deserves c o m m e n t . This m e t h o d is based on the
Carman equation, which predicts a linear relationship between the gas
permeability coefficient o f m i c r o p o r o u s membranes and the absolute pressure
at which this coefficient is determined. In reality, this linear relationship
is observed only dow n to certain minimal values of absolute pressure. Below
these values, the experimental permeability coefficients deviate upwards

0"02

0.038

0"03
0.030
O'02G

I~

0"8

I"E
p ..--~

~.-

$'a
m.'m.

Fig. 3. Transition region for circular capillary, from Knudsen's data. (a) Observed curve;
(b) extrapolation from high pressure (reproduced from P.C. Carman- ref. 11 ).

210
from the predicted linear relationship, yielding a curve with a minimum at
certain absolute pressure values. This can be seen in a plot showing the
dependence of the experimental permeability coefficients on the absolute
pressure at low pressures, published by Carman [ 1 3 ] , given in Fig. 3. The
m i n i m u m of the experimental permeability coefficients corresponds to a
value of the absolute pressure at which the mean free path of the permeating
molecules has the same order of magnitude as the pore sizes of the porous
barrier layer (k and de, respectively, in Fig. 3). In the case of relatively large
pore sizes, this deviation from linearity does n o t have much significance,
but, as we evaluate smaller and smaller pore sizes, the absolute pressures
at which the deviation from linearity becomes significant become higher
and higher. Thus, for instance, for pore sizes in the range of 100 A, assuming
that the permeating gas is nitrogen (mean free path of 929 at atmospheric
pressure and 20C [14] the minimum of the experimental curve showing
the dependence between the permeability coefficient and the absolute
pressure is expected to be at about 9 atm. Naturally, the linear region of this
dependence is expected to be f o u n d at considerably higher pressures. The
absolute pressures at which these measurements are usually performed do
n o t exceed 100 atm and one should n o t be surprised therefore to find that
measurements are n o t in the linear region of the Carman equation. Allusion
to such a situation can be f o u n d in one of the recent studies o f Smolders
et al. [ 5] who f o u n d t h a t using gases with different mean free path values
yielded different average pore sizes for a Millipore VM membrane by the gas
permeability method. A possible explanation for these findings may be t h a t
the measurements were performed in the non-linear region of the Carman
equation. Since these deviations already contribute significantly in the case
of VM membranes with nominal pore sizes in the range of 500 A, considerably
larger deviations from linearity are to be expected with membranes of smaller
pore sizes.
ACKNOWLEDGEMENT
This study was supported by a grant from the National Council for
Research and Development, Jerusalem, Israel.
REFERENCES
1 M.G. Katz, Harnessing Theory for Practical Applications, World Filtration Congress
II, Vol. II, The Filtration Society, Uplands Press, Croydon, 1982, p. 508.
2 A. May-Maromand M.G. Katz, Proceedings of Europe--Japan Congresson Membranes,
Stresa, Italy, June 18--22, 1984, J. Membr. Sci. 27 (2) 1986.
3 M.G. Katz, Israel Patent Application No. 66,672, August 1982.
4 H. Yasuda and J.T. Tsai, J. Appl. Polym. Sci., 18 (1974) 805.
5 F.W. Altena, H.A.M. Knoef, H. Heskamp, D. Bargeman and C.A. Smolders, J. Membr.
Sci., 12 (1983) 313.

211
6

7
8
9
10
11
12
13
14

ASTM Designation F 316--80, 1980.


For example, F.A.L. Dullien, Porous Media, Fluid Transport and Pore Structure,
Academic Press, New York, 1979, p. 103.
W. Thompson (Lord Kelvin), Philos. Mag., 42 (1871) 448.
1978/1979 Millipore Catalogue & Purchasing Guide, Millipore Corp., Bedford,
Mass., 01730, U.S.A.
For example, S.J. Gregg and K.S.W. Sing, Adsorption, Surface Area and Porosity,
Academic Press, London, 1982, p. 175.
P.C. Carman, Flow of Gases through Porous Media, Butterworths, London, 1956,
p. 77.
Reference 11, p. 13 and 34.
P.C. Carman, Proc. Roy. Soc., 203A (1950) 55.
Handbook of Chemistry and Physics, 51st Edition, CRC Press, Cleveland, Ohio,
1970--1971, p. F--151.

Potrebbero piacerti anche