You are on page 1of 7

Ind. Eng. Chem. Res.

1996, 35, 3179-3185 3179

Hydrodynamic Permeability of Hydrogels Stabilized within Porous


Membranes
Vivek Kapur,† John C. Charkoudian,‡ Stephen B. Kessler,§ and John L. Anderson*,|
Colloids, Polymers and Surfaces Program, Department of Chemical Engineering, Carnegie Mellon University,
Pittsburgh, Pennsylvania 15213, Millipore Corporation, 80 Ashby Road, Bedford, Massachusetts 01730, and
HemaSure Inc., 140 Locke Drive, Marlborough, Massachusetts 01752

The purpose of this work was to demonstrate that cross-linked polymer gels can be stabilized
against mechanical and osmotic forces by confining them in a microporous support. The
hydrodynamic (Darcy) permeability was measured for neutral and charged polyacrylamide (PA)
gels synthesized in semirigid membranes having a hydraulic mean pore diameter of 0.5 µm and
a porosity of 67%. The permeability was determined by measuring the flow rate of aqueous
solutions as a function of pressure drop across the membranes. The membrane-supported gels
were stable and yielded a constant permeability when the pressure drop was increased to 300
bar/cm. No swelling/deswelling was observed with the charged gels (0.3-0.4 equiv/L) when the
ionic strength was varied between 0.01 and 1.0 M, and the permeability was essentially
independent of ionic strength. The permeability of the neutral gel varied as φ-3.3 where φ is the
polymer volume fraction, whereas literature data for bulk PA gels shows the dependence to be
φ-1.4. The permeability of the membrane-supported neutral PA gel was greater than the
literature values for the bulk gel at low φ but comparable to the bulk gels when φ > 0.08.

Introduction where v is the superficial solvent velocity through the


gel, η is the solvent viscosity, and P is the pressure. The
The promise of gels as viable separation devices for above equation assumes the gel is isotropic. The
large-scale processes rests on the ability to stabilize hydrodynamic screening length, xk, is a measure of the
them in configurations that take advantage of their
mean spacing between the fibers (polymer chains)
space-filling capacity yet maintain their integrity. Poly-
forming the gel network. Darcy’s law is valid at
meric gels can be selective media based on molecular
size (Gehrke and Cussler, 1986; Tong and Anderson, distances greater than xk from boundaries where a no-
1996) and ion exchange (Boschetti et al., 1995; Gehrke slip velocity condition is imposed; within distances of
et al., 1989). The basic engineering problem is to protect about xk, Brinkman’s equation (Brinkman, 1947; How-
gels against mechanical and osmotic forces. An example ells, 1974) describes the velocity field.
of this challenge is chromatography. Highly cross- Measurements of k for gels are valuable for two
linked gels have been used for the packing of chromato- reasons. First, k depends on the the microstructure of
graphic columns; however, the large pressure gradients the gel, and hence its measurement gives insight into
required to achieve flows through beds of particles less the microstructure; and second, k along with solute
than 0.05 mm diameter can cause deformation and diffusion coefficients must be known to model molecular
collapse of the gel particles. transport within gels (Johnson et al., 1996; Tong and
An approach to overcome the stability problem is to Anderson, 1996). However, experimental determination
fabricate gels within pores of a rigid matrix such as of k for bulk gels is difficult because of their compress-
silica particles (Boschetti, 1994) or a rigid polymer ibility. Often a plot of superficial velocity versus applied
membrane. If the pores are much larger than the pressure across a thin film of gel is nonlinear because
effective mesh dimension of the gel, then the gel will the gel compresses and k decreases as pressure in-
control molecular and viscous transport within the pores creases.
(Kim and Anderson, 1991). Such “gel in a shell”
composite structures offer mechanical strength without There are two objectives of this work: first, to
sacrificing the functionality of the gel (Boschetti et al., demonstrate that a cross-linked gel can be stabilized
1995); that is, the gel network can be optimized with with respect to mechanical (pressure) and osmotic forces
respect to chemistry, charge, and cross-linking. by synthesizing it in a porous membrane; second, to
An important property of gels is their hydrodynamic compare the hydrodynamic permeability of membrane-
(Darcy) permeability, k, which is defined by Darcy’s law: supported gels with bulk gels. We report experimental
results for the permeability of cross-linked polyacryl-
k amide (PA) and derivatized (charged) PA gels that were
v)- ∇P (1) formed within the pores of poly(vinylidene fluoride)
η
membranes. The data are broken into two sets, one for
the neutral gels and the other for the charged gels. With
* Author to whom correspondence should be addressed: the neutral gels, we are interested in the variation of
Phone: (412)268-6986. Fax: (412)268-7139. E-mail: jlacheme permeability with polymer volume fraction of the gel
+@andrew.cmu.edu.
† Present address: E. I. duPont de Nemours and Co.,
and how our results compare with literature data for
Experimental Station, Wilmington, DE 19880.
the permeability of bulk gels (Tokita, 1993; White, 1960)
‡ Millipore Corporation. as well as the stability (i.e., the constancy of the
§ HemaSure Inc. permeability) at large pressure gradients. With the
| Carnegie Mellon University. charged gels, our focus is on the ability of the porous
S0888-5885(96)00015-2 CCC: $12.00 © 1996 American Chemical Society
3180 Ind. Eng. Chem. Res., Vol. 35, No. 9, 1996

membrane to stabilize the gel when the ionic strength


of the solution changes.

Experimental Section
A summary of materials and methods is given below.
More details are available elsewhere (Kapur, 1995).
Synthesis of Gel-Pore Composite Membranes.
The neutral PA gel membranes were made at the
Millipore Corporation (Bedford, MA). A poly(vinylidene
fluoride) (PVDF) membrane (Durapore, Millipore Corp.)
of nominal pore diameter 0.22 µm was used as the
support. These membranes can be obtained in two Figure 1. Apparatus for measuring Darcy permeability of the
forms: with a thin polyacrylate coating to make them membrane: km ) ηvm/(-∆P/L) where L is the membrane thickness
hydrophilic (Steuck, 1986) or without any coating (0.125 mm), η is the viscosity of the aqueous solution, and vm )
(hydrophobic). The hydrophilic membranes were used Q/Am.
for synthesis of the neutral PA gels. The membrane was
first soaked for 2 min in a degassed aqueous solution of vp(wm+g - wm)
φ) (2)
acrylamide monomer plus the cross-linking agent (N,N- πRm2L
methylenebisacrylamide) and photo-initiator (trade-
name Darocur 2959 ) 4-(2-hydroxyethoxy)phenyl-(2-
where Rm is the radius (2.35 cm), L the thickness (0.125
propyl) ketone, Ciba Geigy, Inc.). The membrane was
mm),  the porosity of the membrane (0.67), and vp the
then removed from the solution and placed between two
partial specific volume of PA (0.7 cm3/gm (Munk et al.,
thin polyethylene sheets. Excess solution was removed
1980)). The second method was based on determination
by gentle application of a rubberized roll bar. Polym-
of the nitrogen content of the membrane (Desert Ana-
erization was accomplished by transporting the sand-
lytics, Tucson, AZ). The nitrogen determination was
wich through a Fusion Systems UV (H bulb) unit at 12
converted into PA content per unit volume of membrane,
ft/min. After UV curing, the membrane was immersed
from which φ was calculated (Kapur, 1995). The values
in deionized/filtered water (MilliQ purified). The mem-
of φ by these two methods were in good agreement in
branes were stored in buffered solutions at pH 5.9 or
all cases. The nitrogen-based analysis was probably
pH 8.5. The cross-link density (CL), which is defined
more accurate, so its result was used for φ except for
as the ratio of (mass of bisacrylamide)/(mass of acryl-
the two lowest values of φ for which the nitrogen
amide), was 6% for all the data reported here. The UV
analysis was not available.
energy was too low to expect any significant covalent
The polymer volume fraction of the charged PA gels
binding of the PA to either the PVDF surface or the thin
was determined by the gravimetric method only. Equa-
polyacrylate coating.
tion 2 was used, and the specific volume was assumed
The charged membrane-supported gels were made at
to be the value for PA (0.7 gm/cm3). From representa-
BioSepra Inc. (Marlborough, MA). Hydrophobic Du-
tive samples of the membrane-supported gels, we found
rapore membranes (i.e., no coating on the PVDF) of 0.22
that φ ) 0.12 ( 0.01 for the Q form and φ ) 0.16 ( 0.01
µm nominal pore diameter were used as the support.
for the S form.
Gels based on monomers carrying either positive (Q,
Measurement of Hydrodynamic Permeability.
quaternary amine) or negative (S, sulfonate) groups
The membrane hydraulic permeability (km) is defined
were synthesized using methods similar to those de-
by
scribed by Girot and Boschetti (1993, 1995), which result
in substantially homogeneous, clear gels of low CL. The
two referenced patents pertain to the formation of such
ηvm
km ) (3)
gels within porous mineral oxide particles. To form the (-∆P/L)
same gels within microporous PVDF membranes, the
following procedure was used. The internal surfaces of where vm is the water flow rate (Q) divided by the
the membranes were treated using methods similar to available membrane area (Am ) 5.07 cm2). The ap-
those described in the above referenced patents. Work- paratus is sketched in Figure 1. The membrane was
ing inside a nitrogen-purged glovebag, each membrane mounted on a porous stainless steel screen that provided
sample was placed inside an individual polyethylene negligible resistance to flow. The low-pressure reservoir
pouch. Solution containing monomers, cross-linkers, was a block of Plexiglas with holes drilled at a converg-
and polymerization catalyst was applied to each mem- ing angle. The total volume of the low-pressure reser-
brane using a microliter syringe. Each pouch was then voir was 0.35 cm3.
sealed, and the glovebag was heated to 85 °C for 1.5 h Pressurized nitrogen was used to force water through
to effect polymerization. The resulting gel-filled mem- the membrane. The water was obtained from a MilliQ2
branes were washed in deionized water and stored in system. Buffer solutions of sodium dihydrogen phos-
appropriate buffers. The charge density based on the phate/sodium hydroxide and sodium chloride (pH ) 5.9)
pore volume was 0.4 mequiv/cm3 for the Q form and 0.3 or sodium borate/hydrochloric acid and sodium chloride
mequiv/cm3 for the S form. (pH ) 8.5) were prefiltered in-line with 0.2 and 0.1 µm
The volume fraction (φ) of polymer (monomer plus pore size polyester Nuclepore membranes. Pressure
cross-linking agent) in the neutral PA gel membranes was measured with Validyne DP-15 transducers. The
was determined in two ways. The first method was solvent flow rate was measured by following the move-
based on the weight of the dry membrane before ment of a bubble in a precision-bore capillary (radius )
synthesis of the gel (wm) and the weight of the dry 0.275 mm) connected to the low-pressure side of the
membrane after the gel was formed (wm+g). The gel Plexiglas cell. The hydrodynamic resistance of the
volume fraction was computed from membrane without gel in the pores was more than 100
Ind. Eng. Chem. Res., Vol. 35, No. 9, 1996 3181
Pore volume and pore-size distribution of sample
hydrophobic membranes were determined by mercury
porosimetry (Micrometrics Corporation, Norcross, GA).
From these data we found that  ) 0.67 and the mean
pore diameter equaled 0.4 µm.  was calculated from
the measured specific pore volume (mL/g of solid) and
the density of PDVF (1.77 gm/mL). The mean pore
diameter was defined as the diameter at 50% volume
intrusion.
The Darcy permeability of the bare membranes (kmo)
was about 2-3 × 10-11 cm2. The Kozeny equation
(Happel and Brenner, 1973) relates the permeability to
the hydraulic radius (rh ) pore volume/surface area) and
porosity () of the porous medium:

rh2
kmo ) (5)
K*
where K* is the Kozeny constant, which equals 2 for
parallel capillary-type pores of circular cross section and
Figure 2. Sample data for three gel-filled membranes (neutral
PA) with different polymer volume fractions (φ) in the pores. km
25/6 for consolidated structures (for example, packed
is the slope of the straight lines. beds) with  less than 0.5. The value of K* increases
as  increases and is about 5 for  ≈ 2/3. Taking the
times the sum of the resistances of all other components equivalent pore diameter to be 4rh and using the value
in the system. The membrane permeability was deter- 2 × 10-11 cm2 for the Darcy permeability, we estimate
mined by measuring Q versus ∆P, correcting η for the the hydraulic mean pore diameter to be 0.5 µm, which
temperature during the experiment (20-25, ( 0.5 °C falls between the values obtained by electron microscopy
during an experiment), and then using eq 3 to compute (0.6 µm) (Allegrezza and Kazan, 1993) and the poro-
km. Sample data for three gel-filled membranes are simetry data (0.4 µm).
plotted in Figure 2. Conversion of measurements of km to the permeability
With some of the gel-filled membranes at low φ there of the gel itself (k) requires knowledge of /τ where τ is
was a very slight curvature to the plot of vm versus ∆P. the “tortuosity factor”:
The values of km reported here are the extrapolations
to zero pressure. The reduction in km in going from the
(τ)
-1
k) km (6)
lowest to the highest pressure (-∆Pmax ≈ 3 bar) was
always less than 7% of the extrapolated zero-pressure
value of km. The ratio /τ can be experimentally determined by
Since cross-linked gels are elastic media, it is impor- measuring the diffusion rate (M) of a solute through a
tant to recognize elastic relaxation in making pressure- bare Durapore membrane as a function of concentration
flow measurements. Scherer (1989) modeled the time difference:
course of compression of a elastic thin film that is
supported by a plane on the low-pressure side. If time M ) kD(-∆C) (7a)
zero is defined when the pressure drop is first impressed
across the film, then the time required before the stress  DoAo
kD0 ) lim kD ) (7b)
in the film becomes essentially linear is given by ωf∞ τ L
L2η where Do is the diffusion coefficient of the solute, Ao is
τe ) (4) the membrane area for these diffusion experiments (2.9
Mk
cm2), and ω is the rotation rate (rad/s) of the stirring
where M is the uniaxial elastic modulus of the film. For bars. A diaphragm diffusion cell with rotating stirring
polymeric gels with φ ) 0.1, the osmotic modulus is of bars oriented parallel to the membrane surface on both
order 1 bar and increases as φ2.3 (Sellen, 1987). Taking sides was used. Extrapolation of the experimental kD
this value for the elastic modulus, a bulk gel film of to infinite stirring speed, as shown in Figure 3, allows
thickness comparable to our membranes (≈0.1 mm) the determination of the intrinsic membrane coefficient
would have τe ≈ 100 s; steady-state flow would only be kDo (Malone and Anderson, 1977). The two solutes were
expected for times longer than this. The times for a flow ribonuclease A and bovine serum albumin. Because
measurement at one pressure were about 103 s. The both solutes are 2 orders of magnitude smaller than the
system was allowed to equilibrate for 10-60 min before mean pore size of the membrane, Do essentially equals
the flow rate was recorded. Moreover, two repeatable the diffusion coefficient in solution.
measurments were taken for each increment in pres- The mean value of /τ for the two solutes is 0.29.
sure. Given that the porosity of the bare membranes was 0.67,
Characterization of Bare Durapore Membranes. we have τ ) 2.3, which is consistent with the value
The Durapore membranes had a nominal pore diameter derived from the Kozeny constant (K*/2 ) 2.5).
of 0.22 µm. We measured the porosity of sample If Darcy’s law (eq 1) is valid in the gel phase of the
membranes by gravimetric methods and found  ) 0.67 membrane, then the pressure obeys Laplace’s equation
( 0.02. The nominal thickness of the membranes was in the gel-filled pore spaces because ∇‚v ) 0. Note that
0.125 mm; this value was confirmed by our measure- the solute (protein) concentration within the gel-free
ments. These characterizations were done with hydro- pore spaces of the bare membrane also obeys Laplace’s
philic membranes. equation during diffusion. The mathematical analogy
3182 Ind. Eng. Chem. Res., Vol. 35, No. 9, 1996

Figure 3. Diffusion of proteins across a bare (no gel) membrane


to determine /τ. kD is the overall mass transfer coefficient (cm3/s)
defined by eq 7a, where kDo is mass transfer coefficient for
Durapore membrane (ω f ∞) defined by eq 7b, ω is the angular
velocity (rad/s) of the stirrers in the diffusion cell, and Do is the
diffusion coefficient (19 °C) of the protein in solution (cm2/s). /τ
is computed from the intercept of the best-fit straight lines and Figure 4. Hydrodynamic permeability (k) versus total polymer
eq 7b. The values are 0.31 for ribonuclease A and 0.26 for bovine volume fraction (φ) for PA gels. The symbols are from our
serum albumin; error estimates on each determination are ( 0.03. experiments (6% CL) with k determined from km using eq 7 and
/τ ) 0.29. Literature data for bulk PA gels are represented by
between pressure in the pressure-flow experiments with the lines: Tokita (1993), eq 9; White (1960), eq 10.
gel-filled membranes and protein concentration within
the bare membrane in the diffusion experiments means increasing CL implies that more cross-linking causes
that the tortuosity factor is the same for both processes. greater microscopic heterogeneity of the gel structure,
Thus, the hydrodynamic permeability of the gel itself as suggested by Weiss et al. (1981). This hypothesis is
(k) is related to the hydrodynamic permeability mea- also consistent with the measured increase in the
sured for the membrane (km) by eq 6 with /τ ) 0.29. partitioning of macromolecules with increasing CL
(Tong, 1995).
Results
Neutral PA Gels. The experimental values of k for White (1960) also measured the permeability of bulk
neutral PA gels are plotted against polymer volume PA gels. At CL ) 5%, he obtained data that are fit by
fraction of the gel in Figure 4. There was no measurable the following empirical expression over the range 0.04
effect of ionic strength (0.001 f 1.0 M) or pH (3.0 f < φ < 0.11:
9.0) (Kapur, 1995). The straight line is a power law fit
to these data, neglecting the values for φ < 0.04: k ) 1.015 × 10-16φ-1.083 (cm2) (10)

k ) 4.35 × 10-18φ-3.34 (cm2) (8) These permeabilities are an order of magnitude lower
than those measured by Tokita. A possible explanation
The exponent on φ is larger than the value 1.5 derived
is that the hydraulic permeability of bulk gels is difficult
from the theory of deGennes (1979) for entangled
polymer networks. Data for the sedimentation of linear to measure accurately due to their compressibility.
polymers in solution above the semidilute concentration Since White’s measurements were performed at 20-fold
yield an exponent near 1.5 (Mijnlieff and Jaspers, 1971; higher pressures than Tokita’s, greater compression of
Nystrom and Roots, 1980; Ethier, 1986). the gel is to be expected, which would result in lower
Tokita (Tokita, 1993; Tokita and Tanaka, 1991) apparent permeabilities. It is our opinion that Tokita’s
measured the Darcy permeability of bulk, neutral PA results are the most representative of bulk, neutral PA
gels that were synthesized as films 1 mm thick. Tokita’s gels.
data for 2.2% cross-linked gels are represented by the In Figure 4, we compare our data for the membrane-
following empirical expression (Tong, 1995): supported gels (CL ) 6%) with the data of Tokita (1993)
and White (1960). It appears that the membrane-
k ) 2.64 × 10-16φ-1.42 (CL ) 2.2%) (9a) supported gels were more permeable than the bulk gels
at low φ, but the differences were small when φ g 0.08.
Tokita also measured the effect of cross-link density at In general though, the differences between Tokita’s data
φ ≈ 0.035. The following empirical expression fits his and ours are less than the differences in permeability
data over the range 0.7-24.0% cross-link density (Tong, cited for bulk PA gels by different research groups.
1995):
Charged PA Gels. Figures 5 and 6 display data for
k ) 2.48 × 10-14CL0.295 (φ ≈ 0.035) (9b) the charged gels. Table 1 summarizes the results. The
reproducibility between membranes with the same
where CL ) % cross-link density (100 × mass of cross- nominal gel composition is excellent; the standard
linker/monomer). The increase in permeability with deviation of the measurements is less than 10% of the
Ind. Eng. Chem. Res., Vol. 35, No. 9, 1996 3183

Figure 5. Flowrate data for membranes filled with charged PA Figure 6. Flow versus solution ionic strength (M ) molarity) of
gels at 0.05 M ionic strength. The symbols represent different charged PA gels. The order of the experiments is given by the order
membranes. The straight lines are linear best fits of all the data of the symbols in the legend (top ) first). The ionic strength was
to eq 3. (a) Q gel: km ) 0.122 × 10-14 cm2; (b) S gel: km ) 0.0435 computed using the standard formula 1/2∑izi2Ci where zi and Ci
× 10-14 cm2. are the equivalent valence and concentration of ions of type i.
Increases in ionic strength were achieved by adding sodium
Table 1. Results for Hydrodynamic Permeability of chloride to the buffered solution.
Membrane-Supported, Charged PA-Based Gelsa
Figure 4 for comparison with the neutral gels. The
gel type IS (M) k × 1014 (cm2)
permeability of the charged gels was approximately the
Q 0.05 0.420 ( 0.035b same as the neutral gels for the given values of φ.
Q 0.01 0.342 Figure 6 is interesting in that there is only a small
0.10 0.452
1.00 0.679 effect of ionic strength on the permeability. The small
S 0.05 0.150 ( 0.013c increase in k in going from 0.01 to 1.0 M can be
S 0.01 0.152 explained in two ways. First, electroviscous effects that
1.00 0.169 retard flow are damped by increasing the ionic strength.
a Q form: quaternary amine, pH ) 8.5, charge density ) 0.4 Second, the effective solvent quality for a charged
mequiv/cm3 gel, φ ) 0.12 ( 0.01. S form: sulfonate, pH ) 5.9, polymer decreases at higher ionic strength; a decrease
charge density ) 0.3 mequiv/cm3 gel, φ ) 0.16 ( 0.01. IS means in solvent quality can result in an increase in the
ionic strength of the buffered solution, defined as 1/2∑izi2Ci where permeability of entangled polymer chains (Mijnlieff and
zi and Ci are the equivalent valence and concentration of ions of Jaspers, 1971). At higher ionic strengths the charged
type i. b Six different membranes prepared in the same manner.
groups on the PA chains are less repelled so that
The ( value is the standard deviation of the data. c Four different
membranes prepared in the same manner. The ( value is the Brownian fluctuations in local gel volume fraction are
standard deviation of the data. greater, leading to a higher permeability.
Figure 6 shows that the charged gels in the pores
mean for both positive (Q) and negative (S) gels. No were stable with respect to large osmotic stresses. We
hysteresis in the pressure-flowrate measurements was found that bulk samples of these gels collapsed to less
observed when the pressure was first increased and than 20% of their total volume (φ increased by a factor
then decreased. The mean values of permeabilty for the of 5 or more) when the ionic strength was increased from
Q and S gels at 1.0 M ionic strength are plotted in 0.01 to 1.0 M. It appears that confining the gel in a
3184 Ind. Eng. Chem. Res., Vol. 35, No. 9, 1996

porous matrix minimized such osmotic swelling/deswell- 13 to the data of Tokita (1993) at 2.2% CL (see eq 9a)
ing. Furthermore, the recovery of the same value of k gives af ) 5.7 Å, which is comparable to the value (6.5
when the ionic strength was first increased from 0.01 Å) determined from measurements of the partitioning
to 1.0 M and then reduced to 0.01 M (Figure 6a) of globular proteins in bulk gels (Tong and Anderson,
demonstrates the stability of this system. 1995). Theoretical models for hydrodynamic perme-
ability do not explicitly account for cross-link density.
Discussion Tokita’s data (see eq 9b) show an increase in k as CL
increases. This trend has also been noted by Weiss et
Hydrodynamic Theories of Permeability. Be- al. (1981). The random network models represented by
cause polymeric gels have a chain-like structure, they eq 11 would require af to increase as CL increases if it
are often modeled as a network of straight cylinders of is applied to gels at different cross-link density. A
radius af. The Darcy permeability is expressed in physical interpretation of this increase in af can be
general as a function of the volume fraction of chains constructed by imagining that an increased cross-link
(φ): density results in “bundling” of the polymer chains, that
is, microsegregation of the polymer.
k ) af2f(φ) (11) Bulk versus Membrane-Supported Gels. Because
the pore size (≈0.5 µm) of our membranes was much
where f(φ) depends on the spatial arrangement of the greater than the hydrodynamic screening length of the
chains. Several models are based on a regular periodic gel (xk < 4 nm), the microstructure of the gel should
array of cylinders. Sangani and Acrivos (1982) consid- be the same in the pores as in a bulk gel. Figure 4
ered a regular square array of cylinders with flow suggests that there might be a difference between
oriented perpendicular to the axes of the cylinders at membrane-supported neutral gels and bulk gels at low
an angle of 45° with respect to the alignment of the φ; however, Tokita’s data could be in error at low φ
array. They numerically solved the Stokes equations because of some compression of the bulk gel when
for this alignment and computed the drag on each subjected to pressure.
cylinder. Tsay and Weinbaum (1991) converted the One might speculate that there were defects in the
drag per cylinder to a Darcy permeability and obtained low φ gels synthesized in the porous membranes, which
the following empirical expression from the calculations would explain the higher permeabilities. We believe
of Sangani and Acrivos: this was not the case. Measurements of diffusion and

[x ]
filtration of proteins through the same gel-filled mem-
2.377
π branes that were studied here, when taken together
f(φ) ) 0.0572 -2 (12a) with the hydrodynamic permeability data, show con-
φ
clusively that the gel substantially filled the porous
Drummond and Tahir (1984) also solved the Stokes membranes (Kapur, 1995; Kapur et al., 1996). In
equations for flow past regular arrays of parallel aggregate, these experimental results prove that the
cylinders. Their results are represented by the following membrane-supported gels were essentially defect free.
equation: One possible explanation for the different permeabili-
ties for the bulk and membrane-supported neutral gels
1 at low φ is that the method of synthesis was different.
f(φ) ) [-ln φ - 1.476 + 2φ - 1.774φ2 + 4.076φ3]
8φ In the membrane-supported neutral gels, the polymer-
(12b) ization was photo-initiated whereas the bulk gels stud-
ied by Tokita (1993) were chemically initiated. Righetti
The difference between the above two expressions is 6% et al. (1981) have found that gels synthesized with
or less for 0.01 < φ < 0.15. Durlofsky and Brady (1987) different initiators and promoters can have different
and Phillips et al. (1989,1990) numerically solved the properties.
Stokes equations for flow through networks of chains
of spheres (af ) sphere radius) arranged in periodic Summary
configurations. Their calculated permeabilities are The hydrodynamic permeability of the membrane-
comparable to those predicted from eq 12. supported gels was reproducible between membranes
Jackson and James (1986) reviewed experimental for a given cross-link density and polymer volume
data for flow through fibrous media over a large varia- fraction. Compression of the gel, which would have
tion of af and found that the following expression, which been indicated by a lack of proportionality between
is semi-empirical in nature, fits much of the data: water flow rate and applied pressure, was negligible up
to pressure gradients as great as 300 bar/cm. The
3 membrane-supported charged gels were resistant to
f(φ) ) - [ln(φ) + 0.931] (13)
20φ osmotic deswelling/swelling, and their permeability
varied little when the ionic strength was changed from
Spielman and Goren (1968) modeled a network of 0.01 to 1.0 M. The permeabilities of the positive (Q)
straight cylinders by considering it a Brinkman fluid. and negative (S) gels were the same as for the neutral
The result for a random orientation of the cylinders is gels at the same polymer volume fraction.

[ ]
The important result is that our experiments dem-
x2 5x K1(x) onstrate that stable, essentially defect-free and non-
x2 ) 4φ (14)
3 6 K0(x)
+
compressible gels can be fabricated within microporous
supports.
where x ) af/xk and Kn is the modified Bessel function
of the second kind. Values of k computed from eq 14 Acknowledgment
are about 25% higher than from eq 13 over the range This research was supported by NSF Grant CTS-
0.01 < φ < 0.15. Note that the algebraic form of eq 14 9122573 and by BioSepra, Inc. We greatly appreciate
is less convenient than eq 11. A least-squares fit of eq the assistance of Dr. J. Y. Koo of BioSepra, Inc., who
Ind. Eng. Chem. Res., Vol. 35, No. 9, 1996 3185
synthesized the membrane-supported charged gels. Munk, P.; Aminabhavi, T. M.; Williams, P.; Hoffman, D. E. Some
J.L.A. wishes to express his gratitude to Eli Ruckenstein solution properties of polyacrylamide. Macromolecules 1980, 13,
for his encouragement, advice, and friendship over the 871.
Nystrom, B; Roots, J. Molecular transport in semidilute macro-
years. molecular solutions. J. Macromol. Sci.-Rev. Macromol.Chem.
1980, C19, 35.
Literature Cited Phillips, R. J.; Deen, W. M.; Brady, J. F. Hindered transport of
spherical macromolecules in fibrous membranes and gels.
Allegrezza, A. E.; Kazan, G. Image analysis of GV and DV AIChE J. 1989, 35, 1761.
membranes. Internal Report; 1993, Millipore Corporation: Phillips, R. J.; Deen, W. M.; Brady, J. F. Hindered transport in
Bedford, MA, Dec 6, 1993. fibrous membranes and gels: Effect of solute size and fiber
Boschetti, E. Advanced sorbents for preparative protein separation configuration. J. Colloid Interface Sci. 1990, 139, 363.
purposes. J. Chromatogr. A 1994, 658, 207. Righetti, P. G.; Gelfi, C.; Bosisio, A. B. Polymerization kinetics of
Boschetti, E., Gurrier, L.; Girot, P.; Horvath, J. Preparative high- polyacrylamide gels. III. Effect of catalysts. Electrophoresis
performance liquid chromatographic separation of proteins with 1981, 2, 291.
HyperD ion-exchange supports. J. Chromatogr. B 1995, 664,
Sangani, A. S.; Acrivos, A. Slow flow past periodic arrays of
225.
cylinders with application to heat transfer. Int. J. Multiphase
Brinkman, H. C. A calculation of the viscous force exerted by a
Flow 1982, 8, 193.
flowing fluid on a dense swarm of particles. Appl. Sci. Res. 1947,
Scherer, G. W. Measurement of permeability. I. Theory. J. Non-
A1, 27.
Cryst. Solids 1989, 113, 107.
DeGennes, P. G. Scaling Concepts in Polymer Physics; Cornell
University Press: Ithaca, NY, 1979. Sellen, D. B. Laser light scattering study of polyacrylamide gels.
Drummond, J. E.; Tahir, M. I. Laminar viscous flow through J. Polym. Sci. 1987, 25, 699.
regular arrays of parallel solid cylinders. Int. J. Multiphase Flow Spielman, L.; Goren, S. L. Model for predicting pressure drop and
1984, 10, 515. filtration efficiency in fibrous media. Environ. Sci. Technol.
Durlofsky, L.; Brady, J. F. Analysis of the Brinkman equation as 1968, 2, 279.
a model for flow in porous media. Phys. Fluids 1987, 30, 3329. Steuck, M. J. U.S. Patent 5618533, 1986.
Ethier, C. R. The hydrodynamic resistance of hyaluronic acid: Tokita, M. Friction coefficient of polymer networks of gels and
Estimates from sedimentation studies. Biorheology 1986, 23, solvent. In Advances in Polymer Science 110. Responsive Gels:
99. Volume Transitions II; Dusek, K., Ed.; Springer-Verlag: Berlin,
Gehrke, S. H.; Cussler, E. L. Mass transfer in pH-sensitive 1993; pp 27-47.
hydrogels. Chem. Eng. Sci. 1989, 44, 559. Tokita, M.; Tanaka, T. Friction coefficient of polymer networks of
Gehrke, S. H.; Andrews, G. P.; Cussler, E. L. Chemical aspects of gels. J. Chem. Phys. 1991, 95, 4613.
gel extraction. Chem. Eng. Sci. 1986, 41, 2153. Tong, J. Partitioning and Diffusion of Macromolecules in Poly-
Girot, P.; Boschetti, E. U.S. Patent 5,268,097, 1993. acrylamide Gels. Ph.D. Thesis, Carnegie Mellon University,
Girot, P.; Boschetti, E. U.S. Patent 5,393,430, 1995. Pittsburgh, 1995.
Happel, J.; Brenner, H. Low Reynolds Number Hydrodynamics; Tong, J.; Anderson, J. L. Partitioning and diffusion of proteins and
Noordhoff: Leyden, 1973. linear polymers in polyacrylamide gels. Biophys. J. 1996, 70,
Howells, I. D. Drag due to the motion of a Newtonian fluid through 1505.
a sparse random array of small fixed rigid objects. J. Fluid Tsay, R.; Weinbaum, S. Viscous flow in a channel with periodic
Mech. 1974, 64, 449. cross-bridging fibers. Exact solutions and Brinkman approxima-
Jackson, G. W.; James, D. F. The permeability of fibrous porous tion. J. Fluid Mech. 1991, 226, 125.
media. Can. J. Chem. Eng. 1986, 64, 364. Weiss, N.; Van Vliet, T.; Silberberg, A. Influence of polymerization
Johnson, E. M.; Berk, D. A.; Jain, R. K.; Deen, W. M. Hindered initiation rate on permeability of aqueous polyacrylamide gels.
diffusion in agarose gels: Test of effective medium model. J. Polym. Sci.: Polym. Phys. Ed. 1981, 19, 1505.
Biophys. J. 1996, 70, 1017. White, M. L. The permeability of an acrylamide polymer gel. J.
Kapur, V. Transport in Polymer/Gel Modified Micropores. Ph.D. Phys. Chem. 1960, 64, 1563.
Thesis, Carnegie Mellon University, Pittsburgh, 1995.
Kapur, V.; Charkoudian, J. C.; Anderson, J. L. Transport of Received for review January 17, 1996
proteins through gell-filled porous membranes. J. Membr. Sci. Revised manuscript received June 18, 1996
1996, submitted for publication.
Accepted June 22, 1996X
Kim, J. T.; Anderson, J. L. Diffusion and flow through polymer-
lined micropores. Ind. Eng. Chem. Res. 1991, 29, 1008. IE960015Z
Malone, D. M.; Anderson, J. L. Diffusional boundary-layer resis-
tance for membranes with low porosity. AIChE J. 1977, 23, 17.
Mijnlieff, P. F.; Jaspers, W. J. M. Solvent permeability of dissolved
X Abstract published in Advance ACS Abstracts, August 15,
polymer material. Its direct determination from sedimentation
measurements. Trans. Faraday Soc. 1971, 67, 1837. 1996.

You might also like