You are on page 1of 36

POROSITY

José Ignacio Calvo


Universidad de Valladolid, Valladolid, Spain

Aldo Bottino
Universitá di Genova, Genova, Italy

Pedro Prádanos, Laura Palacio, and Antonio Hernández


Universidad de Valladolid, Valladolid, Spain

1 INTRODUCTION

When searching within the vast literature covering the wide range of separation processes
in which membranes have found application, it is unquestionable that the membranes
involved in such a variety of processes will differ significantly in structure and con-
stitution, and consequently in their functional behavior. Therefore, it is necessary to
characterize membranes such that they can be selected and/or adjusted to each foreseen
process, seeking the best adapted filter or the way of optimizing the required parameters
during their production. It must be taken into account that any small change in the param-
eters of formation of a membrane can change its structure. These structural changes can
have a significant impact on the membrane’s performance in the processes being designed.
According to the previous paragraph, a simple but complete definition of the term
characterization can be given when applied to a synthetic membrane (note that this
could also be used for natural membranes, but those are not within the scope of this
article). Therefore, we can define characterization of a membrane as “the most complete
knowledge of its constitution, structure, and functional behavior, obtained through the
combined and critical use of adequate methods and techniques.”
The main goal being pursued in the characterization of a membrane is the prediction
of its performance when used in a given process. This makes it difficult to achieve a good
characterization of a membrane, as it can be used in very different processes in which
transport mechanisms are even unalike. For example, it is well known that the earliest
applications of synthetic membranes, and still the most extended, used the membrane as
a porous filter or sieve. For this type of membranes, the main factor to be taken into
account, for a correct selection, is related to the size of their pores, especially the average

Encyclopedia of Membrane Science and Technology. Edited by Eric M.V. Hoek and Volodymyr V. Tarabara.
Copyright © 2013 John Wiley & Sons, Inc.

1
2 POROSITY

pore size and pore size distribution, which can determine, in the first instance, the size
of the molecules to be retained.
There are many other applications subsequently developed, such as gas permeation,
where the membranes used are dense. In these cases, other characteristics of the mem-
brane are much more important than its porosity (e.g., parameters such as chemical
affinity for certain species and diffusion coefficients to various gases or mixtures). In
this sense, it would be convenient to start by differentiating clearly between porous and
dense membranes. It is normally considered that a pore has a size that can be observed
under the microscope. Thus, the boundary between dense and porous membranes would
be located in the detection limit of the most powerful microscopes. When this distinction
was proposed (1), the most potent microscopes were electronic ones, which, in the best
conditions, reached a resolving power close to 2 nm (2). Although we have more pow-
erful microscopes [those based on atomic force microscopy (AFM) or scanning probe
microscopy (SPM); or the new-generation electronic microscopes] that are capable of
reaching the atomic range, it seems reasonable to maintain the boundary between pore
and molecular gap in the values previously mentioned.
On the other hand, one would think that after an appropriate structural and morpholog-
ical characterization, we could anticipate precisely and successfully the future behavior
and performance of a filter when applied in a given separation process. This view, as
most researchers know, is too simplistic, while reality is always more complex owing to
our imperfect knowledge of the exact mechanism leading to such behavior along with
the presence of many other factors (hidden variables) in the process that are not always
possible to assess. This does not imply that we would not be able to make an appropriate
decision on the membrane to be used in a certain process. In fact, this “know-how” will
come mostly from the knowledge of the largest possible number of parameters of the
membrane, paying special attention to those related to the membrane performance tested
under similar experimental conditions as found in the intended application.
Usually, the characterization parameters are divided into two groups: structural
parameters and functional ones (3), taking into account whether they provide information
on the structure or the performance of the analyzed filter. This article focuses on
those parameters related to the structure, of which the most important ones are the
following:

1. average or effective pore size;


2. surface (volumetric) density of pores, defined as the number thereof per unit surface
area (volume) of the membrane;
3. porosity by volume or total volume fraction of membrane that is occupied by the
pores or voids—in certain cases, it may also be desirable to characterize surface
porosity, defined parallel to volume porosity;
4. surface morphology, including the shape factor, which measures the degree of
deviation from a circular surface pore perimeter;
5. roughness, which highlights the differences in height that may be found on the
surface of a membrane;
6. tortuosity, because, in general, as the pores are not perfectly parallel, their length is
not equal to the membrane thickness (hardly evaluable on the other hand, especially
for asymmetric membranes);
POROSITY 3

7. statistical distribution of pore size (and actually of most of the parameters men-
tioned), because homoporous membranes are not found in the market, so these are
always more or less heteroporous.

Most techniques providing information on pore size distribution can also be used to
determine membrane porosity, which is given by the summation of the individual volume
(or surface) of all the pore sizes present in a distribution. Consequently, we will include
in this article all the methods giving information about such parameters (mainly pore
size distribution and porosity) and the resulting information will be discussed.
To determine the parameters that we have described, there are a number of experi-
mental techniques more or less complex, which allow us to evaluate various properties
of the membranes (4–12). Next, we will very briefly explain the most important of these
and then discuss them more extensively in the following sections:

1. microscopic techniques, covering scanning electron microscopy (SEM), transmis-


sion electron microscopy (TEM), and AFM, and the usual software aimed to derive
maximum advantage from such images;
2. solute retention test and modeling to convert this information into pore size distri-
butions;
3. dusty gas model, which uses gas permeation to evaluate membrane porosity;
4. picnometry and related methods, which are also able to determine membrane
porosity;
5. techniques based on the Kelvin equation, mainly gas adsorption desorption (GAD)
[for determination of pore size and the Brunauer–Emmett–Teller (BET)) area] but
also other techniques such as permporometry (PMP), thermoporometry (TMP), and
nuclear magnetic resonance (NMR);
6. Washburn-equation-based techniques: starting from bubble point, which lead to
liquid displacement porosimetry on both gas–liquid displacement porosimetry
(GLDP) and liquid–liquid displacement porosimetry (LLDP) interfaces. Finally,
mercury porosimetry as a nonwetting technique will be described.

In the following sections, all these techniques and corresponding features will be
discussed. This article concludes with a discussion section, where these features and the
actual complementarity of the techniques will be analyzed.

2 MICROSCOPIC TECHNIQUES

In this section, we consider the characterization methods based on microscopic techniques


that, along with the appropriate image analysis, lead to the evaluation of the membrane
parameters without any previous assumption on the pore geometry.
Microscopic techniques were used very early to characterize membrane filters. Visual
inspection of pore structure was an invaluable tool for in-depth knowledge of the filters
themselves. Nevertheless, as the developed filters entered the range of submicron pores,
that is, pores with sizes under the micrometer, optical microscopy was no longer useful
to achieve a real picture of the membrane structure because its resolution is limited by
the light diffraction pattern. Only the development of nonoptical microscopic techniques
4 POROSITY

allowed a substantial progress. Initially, electron microscopy (13) and subsequently, probe
microscopy (14) provided the impetus for microscopic characterization of membranes.
Currently, we can get information on membrane surfaces covering the entire range of
membrane filters.

2.1 Electron Microscopy: SEM and TEM


When a solid is reached by high energy electrons, there are many interactions between the
solid material and the electron beam. This interaction can not only be used to identify
the specimen and its elements but also to characterize physically the holes and pores
present at the surfaces.
TEM operates by flooding the sample with an electron beam, most commonly at
100–200 keV and detecting the image generated by the scattered electrons both elastically
and inelastically when passing through the sample. TEM operates in the magnification
range from 600× to 106 ×.
Similarly, a fine beam of medium energy electrons (5–50 keV) generates secondary
electrons that are used in the SEM technique. SEM devices can achieve magnifications
ranging from 20× to 105 ×, giving images with a great field depth, thus leading to
considerable information about the surface texture of the targeted material.
The main problems and drawbacks of microscopic observation by both transmission
and SEM are caused by the specific requirements in the preparation of a membrane
sample in order to avoid artifacts. The first step of the preparation is a careful drying of
the sample. Actually, to avoid collapse of the original structure, the freeze-dry technique
using liquid nitrogen or the critical-point drying method with carbon dioxide is usually
employed.
In order to observe cross sections by SEM, the dried membrane has to be firstly
fractured at liquid nitrogen temperature, and then fixed perpendicular to the sample
holder. Usually, samples are subsequently covered by a thin metallic layer (normally a
gold film of some tens of angstroms), increasing the production of secondary electrons
and thereby improving image contrast (15). As an example, in Figure 1, SEM images
of an AFC-80 membrane (polyamide NF membrane from paterson candy international
(PCI)) at different resolutions are shown.
For TEM observation, a more complicated procedure is required. The dried sample is
first embedded, if necessary, and then cut by a microtome. An embedding media without
any remarkable influence on the membrane must be chosen. The section must be thin
enough for electrons to penetrate (i.e., <50 nm). If the surface of the sample alone is to be
analyzed, a replica technique can be created by coating the membrane with a carbon film,
and then removing the membrane material by dissolving it, and analyzing this replica (16).
The maximum resolution of TEM is in the range from 0.3 to 0.5 nm, while SEM has
lower resolution. Scanning transmission electron microscopy (STEM) was investigated
in the 1970s, developing a field emission gun. Subsequently, a cold field electron
emission source was developed, allowing a direct visualization of light atoms such as
carbon and even hydrogen using TEM and a clean single-layer graphene substrate.
In both cases, high electron beam energies are required to reach maximum resolution.
As a result, the sample surface can be seriously damaged, making observation difficult,
especially when dealing with polymeric membranes. In response to this drawback, in the
early 1980s, the field emission scanning electron microscopy (FESEM), which nowadays
achieves very high resolution (up to 0.7 nm) even at low beam energy (accelerating
POROSITY 5

voltage of 1.5–4 kV), was developed and used to observe surface pores of ultrafiltration
membranes (17).

2.2 Probe Microscopy: STM and AFM


AFM is a relatively new characterization technique that presents very high possibilities of
development and application in the field of microscopic observation and characterization
of various surfaces.
A very small tip scans the surface and moves vertically according to its interaction
with the sample, similarly to what is done in scanning tunneling microscopy (STM).
AFM and STM differ in the method used to detect interactions. In STM, the tip is close
enough to the sample (both being electrically conducting) to allow a current to flow by
tunnel effect, and the sample or tip moves to keep this current constant. However, in
AFM, the tip is placed on a cantilever, whose deflection can be detected by the reflection
of a laser beam appropriately focused. This allows analysis of nonconducting materials,
making the method more convenient to study membrane materials (18–20).
Several operation procedures can be used in AFM:

1. Contact Mode AFM. It measures the sample topography by sliding the probe tip
across the sample surface. The tip–sample distance is maintained in the repulsive
range of the van der Waals forces.
2. Noncontact Mode AFM. The topography of the sample is measured by sensing the
van der Waals attractive forces between the surface and the probe tip held above
the surface. Of course, worse resolution than that obtained in the contact mode is
achieved. Nevertheless, sample damage risk is avoided or minimized.
3. Tapping or Intermittent Contact Mode AFM. This is a variation of the contact mode
and operationally it is similar to the noncontact mode. The cantilever is oscillated
at its resonant frequency with high amplitude (>100 nm), allowing it to touch the
sample during the oscillation. This procedure retains the high resolution achieved
in the contact mode, but minimizes surface damage by eliminating lateral friction
forces. Working in this manner, the tapping mode presents better features than the
contact and noncontact methods.

× – × –

FIGURE 1 SEM images of an AFC-80 membrane.


6 POROSITY

The techniques that have been discussed give an account of the sample topography. In
addition, other properties of the surfaces can be obtained, by analyzing the different forces
between the sample and the tip. For example, the phase contrast provides information
about differences in surface adhesion and viscoelasticity. Both magnetic force microscopy
(MFM) and electric force microscopy (EFM) measure the magnetic (or electric) force
gradient distribution above the sample surface. Surface potential microscopy measures
differences in local surface potential across the sample surface. Force modulation mea-
sures relative elasticity/stiffness of surface features, and lateral force microscopy analyzes
the frictional force between the probe tip and the sample surface. Finally, electrochem-
ical microscopy measures the surface structure and properties of conducting materials
immersed in electrolyte solutions with or without controlling the electric potential. In
many of these techniques, an appropriate consideration of the measured forces is neces-
sary to eliminate the contribution of the topographical images.
In Figure 2, an AFM tapping mode and a force modulation image of a cross-sectional
view for a C04 (polycarbonate track-etched microfiltration membrane from cyclopore)
membrane are shown. They were obtained with a Nanoscope IIIA from Digital Instru-
ments. In Figure 3, the pore size distribution of the AFC-80 as obtained from image
analysis of AFM tapping images is shown. Finally, Figure 4 shows a 3D AFM image of
a NF90 membrane from Dow Filmtec®, acquired by tapping mode.

2.3 Computerized Image Analysis


Image analysis can be carried out by means of a variety of software packages, some of
which are supplied by the main optical or electronic microscopes manufacturers (Jeol,
Leica, Karl Zeiss, Nikkon, etc.) to complement their devices.
In all cases, each image is first digitized by assigning to each pixel a gray level
ranging from 0 (black) to 255 (white). Then, a clearfield equalization is made to each
image field to eliminate parasite changes in gray levels owing to uneven illumination.
Obviously, a perfect clearfield equalization should require a blank image with a perfectly
flat sample of the same material equally treated and acquired in the same way. In fact,
this is impossible and even inconvenient as far as uneven illumination is concerned,

(a) (b)

μm μm

FIGURE 2 AFM (a) tapping and (b) force modulation images of sections of a C04 microfiltration
membrane soaked in paraffin. (Please refer to the online version for the color representation of the
figure.)
POROSITY 7

10
9
8
7
f (dimensionless) 6
5
4
3
2
1
0
0 1 2 3 4 5 6 7 8
dp (nm)

FIGURE 3 Pore size distribution of the AFC-80 from AFM tapping images.
6.000 nm

10

20

30

nm

FIGURE 4 3D AFM tapping mode image of a NF90 membrane. (Please refer to the online
version for the color representation of the figure.)

owing to the roughness of the sample itself. What can be done is to use what is called
pseudoclearfield equalization by dividing the original image into a convenient number
of rectangles. Then we can assign to all pixels the intensity such that a 95% of the
original pixels have a lower intensity. Finally, these rectangles are placed together by
linear interpolation from rectangle to rectangle and subtracted from the original picture.
8 POROSITY

Once illumination effects are eliminated, the image gray spectrum is compressed to
get the maximum contrast and definition. Then the images are redefined according to an
assigned gray threshold level under which every pixel is assigned to 1 and the rest to
0. The resulting binary picture is improved by scrapping isolated pixels, in such a way
that all the remaining 1s in the matrix are assumed to belong to a pore. Finally, the pore
borders are smoothed in order to reduce the influence of the finite size of pixels and low
definition.
Of course, a correct selection of threshold gray level is fundamental to perform a
correct analysis of accurately assigned pores. Customarily, the gray spectrum is analyzed
and the threshold placed centered in the peak-to-peak valley of the almost bimodal
distributions obtained. Unfortunately, sometimes the spectra are so flat that this technique
is only of relative help to make a correct threshold election (21). In any case, eye
inspection facilitates the process of selection of several reasonable threshold candidates
whose outcomes are conveniently averaged.
Some parameters can be obtained directly from the SEM pictures (22, 23), such as the
surface pore density or number of pores per surface unit, NT , and the porosity or porous
surface fraction, . In order to get the pore size distribution, other parameters can be
selected to study their statistical distributions, namely, pore area Ap and pore perimeter
Pp , along with two indirect parameters that can be defined as follows (23):
The equivalent or Feret pore diameter is

Ap
dp = 2 (1)
π

and the pore shape factor is


Ap
sp = 4π (2)
Pp2

According to these definitions, the equivalent pore diameter is the diameter that a
pore of area Ap should have if it had a circular section in the surface, while the pore
shape factor is the ratio between the actual pore area and the corresponding area of a
circle with the same perimeter. Actually sp should be unity if the pore sections on the
membrane surface were perfectly circular.
Consequently, the area and perimeter of each pore are measured immediately from the
microphotographs, while diameter and shape factor are calculated by using Equations 1
and 2. Then we obtain four size distributions for each membrane: pore area, perimeter,
diameter, and shape factor.
In the case of AFM pictures, we also have information on height, which permits the
study of pore entrances in greater detail. In this case, it is possible to use flat projections of
the surface at different heights along with the level profiles at several judiciously chosen
lines. The simultaneous use of images and profiles greatly facilitates the identification of
the entrance of individual pores (however, the diameter deep in the membrane may not
be determined directly by surface AFM owing to convolution between the tip shape and
the pore).
Some other parameters are specific to AFM and are usually implemented by the online
and/or offline analysis software included in the apparatus. In particular, roughness can be
analyzed. This analysis is based on determining the heights of the tip over a baseline or
POROSITY 9

reference level, Z . A statistical treatment of such heights leads to definition of the median
value, Zmed , the mean, Zm , or the maximum peak-to-valley height in the profile, Rt .
Moreover, some typical roughness parameters are defined, such as the average
roughness, Ra ,
1
n
Ra = |Zi − Zm | (3)
n
i =1

where n is the total number of points in the image matrix. Moreover, the mean square
roughness, Rms , is usually evaluated from the Fourier transform of the surface profile.
The roughness versus scanned area pattern is characteristic of a given material and
preparation, and defines the fractal dimension, dfr , which is evaluated as

dfr = 3 − α

where α is the so-called roughness exponent that can be calculated as the slope of
roughness versus scan size in a double log plot (24).
This fractal dimension experimentally well represents how roughness increases with
scan size with accurate fittings to experimental data (25). Some preliminary results on
the fractal dimension of these membranes without further correlation have been already
published by us (26). Surfaces can have fractal dimensions ranging from 2 to 3. A fractal
dimension close to 2 means a surface that is very planar, whereas a fractal dimension
close to 3 indicates a surface that almost fills the third dimension, resembling a volume
rather than as an actual surface.
Figure 5 shows roughness versus scan size for a polysulfone support polysulfone (PSf)-
GTS and activated composite membranes (ACM) consisting in an aromatic polyamide
deposited on the support without any extractant (25).

100 d = 2.33
Ra (nm)

10 d = 2.73

PSf-GTS
ACM-0 PSf-GTS
1
0.01 0.1 1 10 100 1000
Scan size (μm)

FIGURE 5 Roughness versus scan size of PSf-GTS and the corresponding ACM membrane.
Source: Reprinted from Reference (25) with kind permission from Springer Science and Business
Media.
10 POROSITY

3 PORE SIZE DISTRIBUTION BY SOLUTE RETENTION

The fitting of true retention, R, versus flow, JV , provides an estimation of the pore
radius that depends on the solute radius, rs . This permits an evaluation of the pore size
distribution (27–30). The basis of such calculation can be summarized as follows.
In order to evaluate the pore size distribution of a partially retaining membrane, we
should place ourselves under such conditions that retention could be attributed to a pure
sieving mechanism. We assume that for each solute, there is a fraction of pores that
totally retain the solute, while it passes freely through the rest of the pores. Then the
mass balance for each solute is
JV csp = JVt csm (4)

where JV is the total volumetric flux and JVt is the volumetric flux transmitted through the
nonrejecting (transmitting) fraction of pores. The concentrations csm and csp correspond
to the solute on the feed side of the membrane and in the permeated solution, respectively.
On the other hand, the ratio of the transmitted volumetric flux, JVt , and pure water flux,
Jwt , passing through the transmitting fraction of pores is

JVt η
= m (5)
Jwt η0

where ηm and η0 are the solution and solvent viscosities, respectively. For low csm , this
ratio can be approximated by 1 in such a way that Equation 4 can be rewritten as

Jwt = JV (1 − R) (6)

Using
csp
R =1− (7)
csm

and approximating
JV = Jwt + Jwr ≡ Jw (8)

Equation 6 provides
Jwr = JV R (9)

Therefore, Jwt and Jwr (pure water flux passing through the retentive pores) can be
evaluated once JV for each R is known, according to Equations 6 and 9. The ratio Jwt /Jw
versus the solute molecular weight or size gives the accumulated fraction of flux passing
through the pores of a size bigger than that of the solute. Thus, if many solutes with
different sizes are used, the cumulative pore size distribution should be obtained (29, 30).
To reproduce the experimental data, different two-parameter sigmoid curves with hor-
izontal asymptotes at Jwt /Jw = 1 and 0 have been used.

Jwt
= F (rp ; {ai }) (10)
Jw

where {ai } is a set of parameters to be evaluated by fitting Equation 10 to experimental


results. Usually, a two-parameter distribution is sufficient (parameters a and b).
POROSITY 11

0.07
<rp> = 0.37 nm
0.06

0.05
d(N/NT)/drp (m–1)

0.04

0.03

0.02

0.01

0.00
0.30 0.32 0.34 0.36 0.38 0.40 0.42 0.44
rP (nm)

FIGURE 6 Pore size distribution of an HL nanofiltration membrane (GE Water and ProcessTech-
nologies) as obtained from retention experiments for a series of PEGs.

The derivative, f (rp ) = d[F (rp )]/drp = d(Jwt /Jw )/drp , is the probability density func-
tion and provides the flux distribution through differently sized pores. Because the flow is
proportional to the fourth power of the pore radius, rp4 (according to the Hagen–Poiseuille
equation), the pore size distribution, except by a normalization constant, could be obtained
as    
Jw
d NNT 1 d Jwt 1
= 4 = 4 f (rp ; a, b) (11)
drp rp drp rp

As an example, Figure 6 shows the pore distribution obtained from the fitting of
retention versus flow measurements for a series of polyethylene glycol (PEG) through a
commercial nanofiltration membrane.

4 GAS PERMEABILITY

Gas permeability measurements on dried membrane can be carried out in order to obtain
a rapid evaluation of mean pore size and effective porosity of the membrane. These
measurements are very simple to perform and do not require instrumentation that is too
sophisticated and expensive. The membrane is mounted into a proper testing cell, and the
gas flux through the membrane is measured as a function of the pressure. In many cases,
simple and cheap soap bubble flowmeters can be used to evaluate the gas flow rate. Also,
great computing capabilities are not required because a simplified model derived from
the “dusty gas” model (DMG) (31, 32), which describes the gas transport through the
membrane in terms of Poiseuille flow and Knudsen diffusion, can be very satisfactorily
used to elaborate flux–pressure data (33–44).
12 POROSITY

On the basis of this model, a straightforward correlation between the permeance P


(i.e., the ratio between the flux and the transmembrane pressure difference) and the mean
pressure Pm between the feed and permeate side of the membrane holds:

P = α + βPm (12)

The slope β and the intercept α can be easily determined from a linear fitting of flux
and pressure experimental data.
These two parameters are related to the mean pore radius rp and the effective porosity
(i.e., the ratio between the actual porosity, , and the effective pore length, L) through
the following equations:

  
16 β 8RT 0.5
rp = μ (13)
3 α πM
 8RT μ
= β (14)
L rp2

where R is the gas constant and T , M , and μ are, respectively, the temperature, the
molecular weight, and the viscosity of the tested gas.
A modified DGM model that takes into account the gas compressibility as well as the
variation of the gas temperature that occurs in certain membrane separation process such
as membrane distillation has been recently proposed (45).

5 POROSITY MEASUREMENT TECHNIQUES

Among the structural parameters in membranes characterization, we now focus on the


porosity. This is a very important parameter in membrane science and technology. It
plays a key role not only in characterizing other porous materials, such as powders, but
also in heat transfer or in the chemistry of catalytic materials. This parameter can be
obtained indirectly (46) from measurements made by using different techniques (e.g.,
liquid permeability determinations). Nevertheless, in this section, the most important
specific techniques for porosity determinations will be discussed.
In general, porosity measures the membrane percentage occupied by pores over the
whole surface (surface porosity) or solid volume (volumetric porosity). In real situations,
depending on the method employed, only the passing-through pores are measured (per-
colation porosity) or the presence of incomplete pores or cavities inside the bulk material
or on its surface (total porosity) are taken into account.
The volumetric porosity () is given by the ratio between the empty volume (Ve ) and
the total volume (Vt ) of the sample:

Ve V − Vs V
= = t =1− s (15)
Vt Vt Vt

where Vs is the solid volume of the sample.


POROSITY 13

5.1 Apparent Density Method


This method is simply an estimate of the porosity (2) and never can be considered as an
exact measuring method, but as a fast way to obtain a rough idea of the actual value of the
sample. In this method, the sample volume is obtained from estimations or measurements
of its mass (ms ) and the density of the material from which the membrane is made (ρs ).

Vs m ρapp
=1− =1− s =1− (16)
Vt ρs Vt ρs

where ρapp is the apparent density. Obviously, this method gives only an average value
of porosity and does not serve for asymmetric membranes, where strong differences in
porosity and apparent density are found between the skin layer and support.

5.2 Pycnometric Method


If the membrane is immersed in a liquid of known density ρl , it will fill the space inside
the pores (Ve ). This volume can be obtained by successive weighting: the dry sample (m1 ),
the pycnometer totally filled with the liquid (e.g., water) (m2 ), and the same pycnometer
now containing the sample and also fully filled with the liquid (m3 ):

m1 + m2 − m3
=1− (17)
ρl Vt

5.3 Mercury Porosimetry


This method can be considered a variation of the ones described (in fact, it can be
considered a simple pycnometry), where the liquid introduced into the pores is mercury
in this case, although it is significantly more accurate from the experimental point of
view. The sample is placed in a cell, called a penetrometer, that has been previously
calibrated so that its total volume is known (Vp ). A large enough quantity of mercury
is introduced in the penetrometer and is forced to enter inside the membrane pores by
applying increasingly high pressures. The pressure and volume of mercury intruding
(Vint ) into the sample (which is ultimately the value of void volume of membrane) are
recorded. The total sample volume is previously computed by measuring the dimensions
of the sample or by other methods such as from the bulk density.

Vint
= (18)
Vp − ρHg (mp+Hg+s − mp − ms )

where mp+Hg+s is the mass of the penetrometer with the sample and mercury inside, mp
is the mass of the penetrometer, and ms is the mass of the dry sample.

6 KELVIN EQUATION-BASED TECHNIQUES

The adsorption phenomena have already been known since early times. In 1777, Scheele
(47) showed that air increases the coal volume in a reversible phenomenon, because if
the carbon is heated, the air is evacuated. In a most general phenomenon, several works
14 POROSITY

demonstrated that the quantity of adsorbed gas is proportional to the solid surface, so
that it could be considered as a surface phenomenon that takes place at the gas–solid
interface, where the gas molecules attach to the surface (physi- or chemisorption). These
systems are studied by the adsorption isotherm, where the amount of gas adsorbed is
evaluated as a function of the equilibrium pressure (given as relative pressure, pr , to
saturation pressure) at constant temperature.
In the study of porous solids, the pores are usually classified according to their size,
as follows: micropores (pores smaller than 2 nm), mesopores (in the range from 2 to
50 nm), and macroporoes (pores larger than 50 nm). These sizes determine the pore-filling
characteristics (monolayer adsorption, capillary condensation, etc.).
The shape of the adsorption isotherm is related to the internal structure of the adsorbate
such that, according to the International Union of Pure and Applied Chemists (IUPAC)
recommendations (48, 49) we can distinguish six main groups (Fig. 7):

Type I This corresponds to the so-called Langmuir isotherm (50) and appears when
only micropores are present in the sample.
Types II and III They appear in nonporous solid or in solids that only have macropores
(multilayer adsorption process). The two types differ from the relative weakness of

Type l Type lV

B
Micropores Mesopores

Capillary
Condensation

Type ll Type V
Amount adsorbed

Weak
substrate
B
Non-porous

Type lll Type Vl

Weak
substrate
Layering

Relative pressure (pi /p0)

FIGURE 7 Types of physisorption isotherms. Source: Adapted from Reference (48) © IUPAC
1985.
POROSITY 15

the solid–gas interaction (the case of the isotherm III). The Type II isotherm has
an inflection, designated by B , where the adsorbed gas quantity corresponds to the
monolayer capacity.
Types IV and V These processes have a hysteresis between the adsorption and desorp-
tion curves (51) corresponding to mesoporous solids, for which the Kelvin equation
is completely applicable. The difference between Types IV and V relies on weak
interactions present in the latter. IUPAC, in these cases, proposes a new classifi-
cation, in which the hysteresis loop is categorized into four types, ranging from
spheroidal particles to platelike particles.
Type VI This displays a stepwise multilayer adsorption due to the phase transition
of the molecular layer adsorbed or due to the adsorption on different faces of the
crystalline solid.

The types of adsorption isotherms described are only ideal cases; in practice, there
are many intermediate possibilities and these types cannot be distinctly seen very often.

6.1 Gas Adsorption–Desorption


Gas adsorption–desorption measurements are widely used for porous materials charac-
terization, mainly for the determination of the surface area and the pore size distribution.
This technique, using nitrogen as adsorbate, is very useful to determine the surface
area and mesopore size distribution. For the micropore size distribution, more specific
molecules are necessary in order to obtain reliable results.
There is a wide spectrum of commercially available devices for the determination of
the adsorption isotherms with gases or vapors. Considering the method used to determine
the amount of gas adsorbed, they can be divided into three main groups:

1. Volumetric Techniques. These techniques measure the volume and pressure of gas
at equilibrium (static or continuously).
2. Chromatographic Techniques. These use measuring techniques based on chro-
matography.
3. Gravimetric Techniques. These techniques determine the changes of the solid mass
with a microbalance.

6.1.1 Determination of Surface Area. The total membrane area where there may be
adsorption can be characterized by using the BET method, which is generally considered
a standard procedure. This method is based on an adsorption kinetics model (52).
Its main assumptions are as follows:

1. In all layers, except the first one, the molar enthalpy of adsorption corresponds to
the enthalpy of condensation, L.
2. In all layers except the first one, condensation–evaporation conditions are equal.
3. At the saturation pressure (i.e., pr = 1), all gases are condensed on the solid surface,
tending to an infinite number of adsorption layers.

On the basis of these assumptions, we obtain a relatively simple expression:


p 1 C −1 p
= + (19)
V (p0 − p) Vm C Vm C p0
16 POROSITY

where Vm is the total volume adsorbed in the monolayer per mass unit, V is the volume
adsorbed per mass unit, and C is a constant that is dependent on the isotherm shape.
As seen, if we represent p/V (p0 − p) versus pr , a linear relation is obtained, allowing
evaluating C and Vm from the slope and the intercept. In practice, this model fits only
relative pressures in the range of 0.05–0.3. Finally, taking into account the molecular
cross-sectional area Am (16.2 Å2 for nitrogen molecule), the surface area should be

Vm
A(BET) = NA Am (20)
Vg

where Vg is the gas molar volume at STP conditions, and NA is the Avogadro’s number.

6.1.2 Determination of Pore Size. In order to determine the pore size distribution, it
is necessary to develop the analysis of adsorption–desorption isotherms, distinguishing
whether they correspond to micropores or mesopores. For the case of macropores, the
isotherms might be analyzed close to the saturation pressure (pr = 1), where the conden-
sation on the pore walls starts; but this is not actually useful for exact measurements.

6.1.2.1 Mesoporous Analysis. The key mechanism in the mesopores adsorption is cap-
illary condensation. In this case, the adsorbed molecules behave as a liquid (except
in the first layer), and the interphase between gas and adsorbed molecules is in fact a
vapor–liquid equilibrium. This equilibrium is determined by the curvature of a meniscus;
so the method used has to take this into consideration.
Mesopore size calculations are usually made with the aid of the Kelvin equation,
which, in its simplest form, can be written as

2γ Vl
rk = (21)
RT ln(pr )

where rk is the radius of the equivalent hemispherical meniscus (Kelvin radius), γ is the
surface tension, and V1 is the molar volume of the liquid condensate.
In the case of cylindrical pores, the pore radius is rp = rk + t, where t is the correction
made in rk to take into account the thickness of a layer already adsorbed on the pore
walls. The correct calculation of t is a critical point to relate the relative pressure with
pore size; strictly it should be measured for a totally flat surface of the same material. As
this is normally not possible, a phenomenological correlation is used, which, in the case
of mesopores, is the Halsey’s correlation (53, 54). This correlation is based on a number
of measurements made with flat materials, with the thickness given in angstroms:
⎡ ⎤1/3
5 ⎦
t = 0.354 ⎣   (22)
p0
ln p

For a parallel-sided slit, the slit width, dp , is given by dp = rk + 2t.


A number of different computational procedures have been proposed to obtain the
mesoporous size distribution from adsorption data: Dollimore and Heal assume cylindrical
pores (55, 56).
POROSITY 17

6.1.2.2 Microporous Analysis. In the range of micropores, the Kelvin equation is no


longer valid. The process of micropore filling leads to a distortion of the isotherm shape
in the monolayer range, and in the absence of capillary condensation, the next layers
are filled, giving a multilayer adsorption. In this case, several approximations have been
proposed. The simplest one is based on a reformulation of the Kelvin equation, to take
into account the thickness of the adsorbed layer. However, these approximations give
nonconsistent results (57).
By using the thermodynamics of the adsorption, more sophisticated descriptions
can be found as, for example, the Dubinin–Radushkevich equation (58–60) or the
Horväth–Kawazoe method (61, 62). A model leading to reasonably good results is the
micropores method (MP method) proposed by several authors (63), which is based on the
t-plot described by Lippens and de Boer (64). These authors developed a method adapted
for the analysis of the microporous structure, consisting in graphing the adsorption
isotherm in terms of the thickness of the adsorbed layer. In this case, the t-values can
be calculated by using the empirical relation proposed by Harkins and Jura (65):
⎡ ⎤1/2
13.99
t =⎣  ⎦ (23)
p0
0.034 + ln p

6.1.3 Determination of Porosity by Gas Penetration. In a way similar to that used in


pycnometries, a method using gas penetration has been proposed for porosity determi-
nation (66, 67). The evaluation is made from measurements of the gas introduced in a
holder (in several selected situations) at constant flow (number of moles increase linearly
with time), and collecting data of pressure (p) versus time (t). The gas used is helium,
which is quite small in size and it does not adsorb on the holder walls and on the pore
walls at the temperature conditions of work (77 K, as the holder is placed in a Dewar
containing liquid nitrogen). In a reference holder, nitrogen gas is introduced, and the
saturation pressure is measured and used to normalize the sample pressure, in order to
take into account the atmospheric pressure variations.
Working at low pressure (between 7 and 16 kPa), helium can be considered as behaving
as a perfect gas. Then we can consider that the following equation holds

k2
p = k1 + (24)
Vg

where k1 and k2 are unknown constants, and Vg is the volume occupied by the gas. By
measuring in three different situations (first the empty holder, then the holder containing
a nonporous sample of known volume, and finally with the sample), the volume of the
sample can be determined. From this volume and the geometric volume of the sample,
porosity can be evaluated.

6.2 Other Techniques Based on Capillary Condensation


6.2.1 Permporometry. The permporometric technique is based on the capillary
condensation of a vapor inside the membrane pores, leaving other pores open to allow
18 POROSITY

a noncondensable gas to pass through them. According to the Kelvin equation, which
models capillary condensation, by controlling the relative vapor pressure of the
condensable gas, we can change the size of the pores that open to the flow of the
noncondensable gas (Kelvin radius, rk ). If flux measurements are started when the
relative vapor pressure of the condensable gas is unity, all the pores in the membrane
should be closed (i.e., filled with condensed liquid), avoiding any diffusive flux of the
noncondensable gas through the membrane. When relative vapor pressure is slightly
below 1, the liquid contained in the biggest pores starts to vaporize, opening these
pores. The measured flow of the noncondensable gas can be easily translated in terms of
pore number once the appropriate gas transport model is taken into account. By steadily
decreasing the relative pressure until all the pores are opened, both the differential and
integral pore size distributions can be obtained.
In this kind of determination, as occurs in the case of the mesopores analysis through
adsorption–desorption measurements, it must be taken into account that the size available
to flux is not the real size of the pore. So it is necessary to add the thickness of the
adsorbate layer, t, which is a function of pressure, as shown earlier. A good approximation
to determine t is again the Halsey correlation.
Once a geometric model for the pore and a model for the gas transport through the
pores are supposed, this method allows us to determine the absolute (differential or
integral) distribution of number of pores active to flux for the membrane. The model to
be used should depend on the working conditions (type of gas, temperature, pore sizes,
etc.). Nevertheless, in most situations where this technique was used, a diffusive Knudsen
model seems to be adequate.
Since this technique is based on the Kelvin equation, it is applicable only to mesopores,
whose strict limits are conditioned by the gas and working pressure. However, these limits
can be considered, as previously mentioned, as ranging from 2 to 50 nm.
The method has been used and compared with other ultrafiltration membranes charac-
terization methods, with very good results (3). Recently, it has been demonstrated that the
assumption of a bundle of nonintersecting pores leads to erroneous pore size distributions
that underestimate significantly the correct average pore size in PMP (68). Mourhatch
and coworkers developed a pore network that takes into account the pore interconnectiv-
ity by using the invasion percolation model. In recent times, PMP has been used for the
first time to characterize flow-through micropore defects down to 0.7 nm in size in MFI
zeolite membranes. Helium was used as the nonadsorbing gas, and n-hexane or benzene
was used as the adsorbate (69).

6.2.2 Thermoporometry. TMP is based on the dependence of the melting or solidifica-


tion point of a substance on its surface curvature. In effect, when a fluid soaks a porous
material, the surface curvature of the fluid is determined by the pore geometry. Thus, a
determination of the resulting distribution of melting or solidification temperatures can
give information on the size distribution of the pores of the sample. As a first approxi-
mation, pore size can be related to freezing and melting temperatures empirically; such
a phenomenological correlation should be obtained by reference to well-known porous
test substances. However, it is also possible to obtain equations based on equilibrium
thermodynamics, relating those parameters.
When a liquid totally saturates a porous material at thermodynamic equilibrium, the
thermodynamics of phase equilibria (based on the Laplace and Gibbs–Duhem equations)
shows that the solid–liquid interface curvature (determined by the pore size) can be
POROSITY 19

related to the temperature of phase change (70):


 T0
V1
T = T0 − d(γsl Csl ) (25)
T SF
where T is the temperature for the phase change at a given curvature radius, T0 is this
temperature for a flat surface, vl is the molar volume of the liquid, SF is the molar
entropy change, γsl is the surface tension at the solid–liquid interface, and Csl is the
interfacial curvature. The curvature parameter is given by
a
Csl = (26)
rp − t

The radius of curvature of the solid–liquid interface is (rp − t). The geometrical radius
of the pore is rp , while t is the thickness of the condensation layer on the wall that does
not change its state of aggregation. The parameter a is 2 for an approximately spherical
surface (for both melting and solidification), whereas for a cylindrical pore a = 1 when
melting, and a = 2 for solidification.
According to both these equations, the decrease in phase equilibrium temperature,
T = T − T0 , can be related to pore radius. The following equation, derived from a
thermodynamical approach, can be tested for different porous materials and liquids:
A
rp (nm) = − +B (27)
T
It is worth mentioning that for melting and solidification, these constants A and B are
different. The differential of this expression is
A
drp (nm) = d( T ) (28)
( T )2
To change the temperature of a given porous material-soaking liquid system in T ,
a power y is spent. Then the differential volume of liquid changing its phase when
temperature changes in d( T ) is
y
dV =
 d( T ) (29)
Wa

where
 is a constant that depends on the apparatus and measurement conditions, and
Wa is the apparent energy of phase change.
We arrive at
dV ( T )2
= y
(30)
drp Wa

where
is a constant that depends on the sensitivity of the differential scanning calorime-
ter (DSC) used and the speed of temperature variation, the mass of the sample, and the
density of the liquid and the constant A. It is worth noting that the cooling or melt-
ing speed must be low enough (∼ 1–6 K/h) for the three phases to remain in constant
equilibrium and the temperature to be the same throughout the sample.
This method has been used to characterize porous materials, specifically ultrafiltration
membranes, giving good results for pore size distributions in the range of 2–30 nm
(71–73). A fairly detailed review of the technique has been published recently (74, 75).
20 POROSITY

6.2.3 NMR Porometry. Several approaches have been developed and used to charac-
terize the pore size distribution of membranes and other porous materials by means
of 1 H NMR (76–78). These approaches analyze the differences between the NMR
characteristics of a chemical species immersed into a bunch of pores depending on
whether it is in liquid or in solid phase.
Several studies have examined the phase transition of liquid confined within the pores
of a porous material by using the longitudinal (spin–lattice) and transverse (spin–spin)
relaxation times of the H nuclei. Relaxometry nevertheless requires an independent
knowledge of the surface-layer parameter that has to be determined independently by
nitrogen adsorption, for example. Nevertheless, another NMR method can be adopted
(79) to determine pore size distributions. This is NMR cryoporometry, which uses the
depression in the melting point of a material confined within the pores of a membrane.
The theoretical basis of cryoporometry is the Gibbs–Thomson equation. From this
equation, it is known that the difference between the bulk and the depressed melting
temperature is inversely proportional to a linear dimension of the liquid confined within
the pores. Nevertheless, the pore size distribution obtained by this method is negatively
affected by the differences in the magnetic susceptibility effects between the matrix and
liquid as a function of temperature. Allen et al. (80, 81) determined the pore size distri-
bution (PSD) of cyclohexane confined within porous materials, using a spin echo pulse
sequence (90◦ –τ –180◦ –τ –echo). They found that pore susceptibility differences could
be ignored when a spin echo pulse sequence was used. At a pulse separation time, τ ,
of approximately 10 ms, the signal intensity, I , from the liquid phase of cyclohexane
confined within the pores was obtained by them as a function of temperature. The cyclo-
hexane outside the sample behaves as bulk cyclohexane, and provides a reference point
for the change in melting temperature arising from the confinement.
On the basis of these considerations, cryoporometry has been successfully applied to
various porous materials, including ultrafiltration (UF) membranes with nonuniform and
asymmetric pore structures (82). In their work, Jeon et al. review extensively the basis
of NMR cryoporometry and its application to the evaluation of pore size distributions of
membranes.

7 FLUID PENETRATION TECHNIQUES

7.1 Bubble Point


In 1908, Bechhold (83) proposed a method to evaluate the pore size of microfilters from
viewing of first air bubbles arising from a water-filled membrane when air is forced to
flow through. He is considered to be the first person to have suggested the technique. This
technique is called bubble point method and serves mainly to discriminate the maximum
pore size present in a given filter. In that sense, it is obvious that the method is poorly
informative of the actual pore sizes of the membrane, but even so, this information can be
very useful, for example, for membrane integrity studies. Effectively, it is very common
in membrane manufacturing to run regular membrane integrity tests, to control the quality
of the production line. In any case, the bubble point method used as integrity test for
membranes was the starting point of the often called bubble-point-related techniques,
which actually give complete information on pore size distribution inside membrane
filters.
POROSITY 21

The fundamentals of the method have been extensively discussed (11, 84, 85), and
they are based on the Young–Laplace equation, which describes the pressure needed to
expel a fluid from a tube filled with another fluid. This equation when used for capillary
tubes (as membrane pores are usually modeled) is called the Washburn equation, and it
is the basis for all methods described in this article, along with the mercury intrusion
porometry. Usually, the Washburn equation, when applied to liquids that properly wet
the pore surface, is termed the Cantor equation, given by


p = (31)
rp

where p is the applied pressure, rp is the radius of the pore to be emptied, and γ
the surface tension of the liquid–fluid–membrane interface. In this equation, we have
supposed zero contact angle (θ = 0); otherwise the Washburn equation becomes (2).

2γ cos θ
p = (32)
rp

Several commercial devices have been developed on the basis of the bubble point
technique for evaluation of maximum pore size in membrane microfilters. Taking into
account the typical surface tensions found for normally used liquids, microfiltration (MF)
membranes need only some bars to detect their maximum pore size. Even tight ultrafil-
tration (UF) membranes presenting pores under 50 nm can be analyzed using such bubble
point devices under 10–15 bar, values that are reasonable. Some of such devices that
we can find in the market are Sartocheck® 4 (Sartorius), Flowstar XC (Palltronic®),
BpTester (PMI), or IntegrityTest II (Millipore) (86).
These equipment are normally used for the evaluation of the maximum pore size
present in a sample (87, 88). This analysis can be considered an integrity test (89, 90)
for detecting the presence of pinholes. The information is very useful in membrane
production lines, especially for virus membranes, where the quality requirements for
filters are very stringent (as expected, owing to the potential risk of any filter misuse or
damage (91)). There are also other integrity test systems not based on bubble point but on
other physical properties of the analyzed filters, some using gas mixtures (92) or others
that are based on permeabilities to several liquid combinations (93–95). A complete
review of the integrity test methods for virus membranes can be found at Guo et al. (96).

7.2 Gas–Liquid Displacement Porosimetry


Departing from the Bechold’s original idea, Erbe (97) proposed to continue increasing
the pressure from the bubble point so that pores of successively decreasing sizes were
gradually emptied. Thus, the measurement of gas flowing through the pores that are
already opened would give us information about the distribution of pore sizes. The
bubble point should correspond to the largest pore size present in the membrane, while
successive flows of air should account for smaller pore sizes present in the sample. To
assess the contribution of these pores to the total distribution, a transport model through
them is necessary to allow the evaluation of the increase of air permeability when smaller
and smaller pores are opened. Taking into account the usual applied pressures and the
pore sizes involved, the Hagen–Poiseuille law for convective flow inside capillaries can
22 POROSITY

be applied for such a purpose. Later on, commercial equipment appeared, which included
automated data acquisition and further calculations. The accuracy of the method led it to
reach the status of a recommended standard (98, 99).
Experimentally, the technique consists of filling the membrane with a liquid that fills
all the pores. Since the contact angle depends on the interaction between the membrane
and the liquid, according to the membrane to be analyzed, different liquids should be
used. In fact, the real membrane character (hydrophobic or hydrophilic) being unknown,
then the best choice is that of a standard liquid having both hydrophobic and hydrophilic
radicals (e.g., hydrocarbon components with a polar functional group). This is the case
of the alcohols or some halogen compounds. Porometer manufacturers usually supply
these liquids (Porofil®, Fluorinert®, Porefil®, Porewick®, Galwick®, or similar combi-
nations), which guarantee fast and complete wetting for almost all materials, and show
low surface tension, low vapor pressures, and low reactivity (85).
The wet sample is then subjected to an increasing pressure applied by a gas source.
As the gas pressure increases, it eventually reaches a point where it can overcome the
surface tension of the liquid in the larger pores, which is subsequently expelled. A still
higher pressure rise would discharge the liquid from the smaller pores, in accordance
with the Washburn equation. With the opening of the pores of different sizes, the volume
of gas flow JV increases successively until all the pores are empty. The so-called wet flow
curve is obtained by representing the pressure applied versus the flow of gas through the
sample when the liquid is being expelled. If air continues to be pushed after the sample
is completely dried (there is no fluid remaining within the pores), a dry flow curve is
obtained. A comparison of both curves, using appropriate transport models, allows the
calculation of the contribution of each pore size to the total permeability along with the
pore size population.
As already stated, the Hagen–Poiseuille model is normally used for accounting for the
fluid transport inside capillary tubes. But for gaseous fluids, the Knudsen model could
also be applied, and the pertinence of one or another model depends on the relative size
of the mean free path of the gas molecules and the capillary diameter. Proper transitions
between the Knudsen and Poseuille models for analyzing GLDP results must be taken
into account to get more precise results (85).
It is also important to note another source of error in the interpretation of porosimet-
ric data. In both convective (Hagen–Poiseuille) and diffusive (Knudsen) flow regimes,
the model uses an actual pore length to calculate pore contribution to total flow or per-
meabilty. This length is usually identified with the membrane thickness for symmetric
membranes, while for asymmetric ones, this should correspond to the thickness of the
active layer. Both cases suppose cylindrical pores normal to membrane surface with no
tortuosity. But neither is the actual thickness easy to evaluate (especially for asymmetric
membranes), nor can the assumption of pore length equal to filter thickness without tor-
tuosity be assured. In fact, a recent work has improved data calculations by determining
the tortuosity factor for each pore class in the distribution (100).
The range of applicability of this method depends on the characteristics of the liquid
in which the membrane is immersed, due to to both the surface tension of gas–liquid
interface and the contact angle between the liquid and the membrane material. The usual
organic liquids (names already mentioned) allow the analysis of pore sizes below 0.1 μm
at an applied pressure of about 10 bars.
The GLDP technique has been applied to a range of porous materials, not only mem-
branes but also many other porous materials. The upper range of application of the
POROSITY 23

technique can be extended to pores up to 200 μm, allowing the application of the tech-
nique to many sieving filters and also woven and nonwoven textiles.
More or less on the basis of the first air–liquid porometers from Coulter, acquired
a long time ago by Beckmann and now no longer used, more equipment appeared to
cover the gap; these were led by the Coulter porometer II. Among those, it can be
commented that the PMI device, one of the first in the market, claims to lead down
to almost nanometer range. Other companies manufacturing air–liquid porometers are
Quantachrome or Porometer (a company originated at Benelux Scientific).
Using such equipment, many groups continue evaluating pore size characteristics of
membrane filters or comparing their results with those coming from different techniques
(101). In literature, the technique is sometimes also termed capillary flow porometry
(102) or liquid extrusion porometry (103), but they all refer to the same procedure.
Most equipment use a data smoothing algorithm, which, along a very steep increase of
pressure (typically more than 200 pressure steps are used in a complete analysis), lead
to very soft almost Gaussian distributions. Also remarkable is the reproducibility of the
method. As an example, Figure 8 shows the porosimetric runs for several coupons of a
polytetrafluoroethylene (PTFE) membrane used for osmotic membrane distillation. The
resulting pore size distributions for all six coupons are shown in Figure 9, with a very
low dispersion.
Finally, it can be commented that the technique can be also used to evaluate the
internal fouling patterns in typical foulant filtration experiments (104).

7.3 Liquid–Liquid Displacement Porosimetry (LLDP)


This method, first introduced by Erbe (97, 105) in 1933 and then refined by Grabar and
Nikitine (106) in 1936, has the same principle as that of gas–liquid porosimetry, but the

PTFE, 0.2 μm
12
First sample
Second sample
10
Third sample
Fourth sample
8 Fifth sample
Air flow (mL/min)

Sixth sample

0
0.5 1.0 1.5 2.0 2.5
Pressure (bar)

FIGURE 8 Porosimetric run using a Coulter porometer II for six different samples of a PTFE
membrane of nominal pore size 0.2 μm. (Please refer to the online version for the color represen-
tation of the figure.)
24 POROSITY

PTFE, 0.2 μm
2.5
First sample
Second sample
2.0 Third sample
Fourth sample
Diff. permeability (%)

Fifth sample
1.5 Sixth sample

1.0

0.5

0.0
0.08 0.12 0.16 0.20 0.24 0.28 0.32
dp (μm)

FIGURE 9 Pore size distributions from the porosimetric analysis shown in Figure 8 (PTFE
membrane of nominal pore size 0.2 μm, used for osmotic membrane distillation). (Please refer to
the online version for the color representation of the figure.)

gas used to perform the porosimetric measurement is replaced by a liquid (the so-called
displacing liquid), which is not miscible with that (the so-called wetting liquid) filling
the membrane pores. The method is well described in the Kesting’s book (84).
The measurements can be carried out by following the same procedure employed
for GLDP, that is, by gradually increasing the pressure of the displacing liquid and
measuring the flow rate through the membrane pore as a function of the applied pressure
(107–115) as well as through an alternative and more accurate route that consists of
increasing stepwise the flow rate of the displacing liquid through a precise volumetric
pump and measuring for each step the corresponding pressure (116–131). A syringe
pump has been proved very useful for LLDP measurements because it provides very
stable fluxes without fluctuations and need to use any sort of dampening (125–131).
Pairs of liquids with very low interfacial tension (Table 1) are very suitable to test
membrane with pore sizes within the nanometric range, as these pores can be occupied
by the displacing liquid at very low pressures that do not alter the membrane structure.
For example, according to the Cantor equation, for water–isobutanol system (γ = 1.7
mN/m) and complete membrane wetting (θ = 0), the wetting liquid that fills the pore
with a radius of 0.034 μm will be displaced by the displacing liquid at a pressure of
1 bar, that is, a value at which the membrane compaction that represents an important
source of error in LLDP measurements (133) would be negligible. A great advantage of
this method lies on the fact that the membrane is tested under wet state, that is, in a
condition very close to that occurring in important membrane separation processes, such
as dialysis, micro-ultra-nanofiltration, etc. The method is clearly sensitive to interactions
between the membrane and the liquid pairs. Swelling phenomena during testing may
considerably alter the permeation through the pore, thus leading to unreliable results.
POROSITY 25

TABLE 1 List of Different Immiscible Wetting Liquid-Displacing Liquid Pairs Obtained


from the Phase Demixing of a Two- or Three-Component Mixture, Which Can be Selected
for LLDP Measures According to the Membrane Pore Size

Liquid Pair Composition, v/v Interfacial Tension, mN/m Temperature,◦ C References

Water/isobutanol/ 25/15/7 0.35 18.7/24 106


methanol
Water/isobutanol/ 4/5/1 0.80 20/23 106
methanol
Water/isobutanol 1/1 2.2 18.7/21 106
Water/isobutanol 1/1 2.0 25 132
Water/pentanol 1/1 4.80 25 132
Water/octanol 1/1 8.50 25 132

Various more or less automated devices have been developed in different research labs
to perform the measurements, but only a very limited number of commercial devices are
available (134). A typical example of porosimetric run using an LLDP-automated device
is shown in Figure 10, and the corresponding pore size distribution for a polysulfone
commercial membrane from Sartorius is shown in Figure 11.
To summarize the advantages of LLDP, mostly common to GLDP, we can conclude
that

1. this technique only accounts for pores open to flux;


2. the membrane is tested in wet state, that is, in conditions close to those employed
during a micro and/or ultrafiltration process;

Sartorius PSF membrane


3

2
Flux (mL/min)

0
0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2
Pressure (bar)

FIGURE 10 Porosimetric run using a liquid–liquid displacement porosimetry (LLDP) device


for a PSf membrane from Sartorius. (Please refer to the online version for the color representation
of the figure.)
26 POROSITY

Sartorius PSF membrane


20

15
Diff. permeability (%)

10

0
0 10 20 30 40 50
Rp (nm)

FIGURE 11 Pore size distribution obtained from liquid–liquid displacement porosimetry


(LLDP) analysis shown in Figure 10. (Please refer to the online version for the color representation
of the figure.)

3. this technique covers most of the range of pore sizes usually found in synthetic
membranes in conjunction with GLDP. LLDP is able to analyze pores under the
nanometer range (129);
4. the results are fully reproducible;
5. the structural information acquired can be used to get an estimation of MWCO
(135), which is very useful for membrane manufacturers.

7.4 Mercury Porosimetry


As stated in the introduction, we can include mercury porosimetry among the charac-
terization methods that are based on the Young–Laplace equation. The main difference
relies on the fact that in this case, it is mercury, a nonwetting liquid, that is forced to
enter the pores. Then, the pressure necessary to overcome the interfacial forces can be
related to the characteristic radius of each pore by using the Washburn equation, and the
corresponding contact angle, which usually varies in the range 112◦ –150◦ (6, 136).
The method was originally proposed by Washburn (136) and then developed by Ritter
and Drake (6), while Honold and Skau (137) were the first authors using it for analyz-
ing membrane materials. The method has proved to be reliable to determine pore size
distributions, pore shape characteristics, specific surface areas, and porosities (the last
mentioned has been commented on in Section 5.3).
By a stepped increase in applied pressure, mercury is forced to enter all the pores
present in the membrane. This includes, of course, the voids that appear between the
different sample parts (normally, the samples are in powder state or are broken into small
pieces before being placed into the measurement cell). The resulting curves showing the
intruded (or extruded in the consecutive depressurization step) volume versus the applied
pressure are called porograms, and their shape is strongly variable, depending on the
POROSITY 27

characteristics of the sample. But all porograms presents some common features that are
usually attributed to the presence of ink-bottle pores inside the porous samples:

1. There is always some hysteresis; that is, intrusion and extrusion pathways do not
match.
2. Also, after the completion of an intrusion–extrusion cycle, there is always a certain
portion of mercury that remains entrapped inside the sample (pore entrapment),
which prevents the curves from closing. In any case, this phenomenon usually
disappears after the second pressurization–depressurization cycle.

As stated, mercury porosimetry has been in use for a long time ago to characterize
porous materials, especially powders, ceramic materials (138), and zeolites. Mercury
porosimetry is a widely accepted technique for measuring pore volume and pore
size distributions in the macro-mesoporous range, and their applications have been
consolidated with the introduction of internationally recognized standards (ASTM
D4284, D2873, D4404, or DIN 66133) (139).
One of the advantages of this technique is that it covers, in a single device, a very
broad range of pore sizes. Usual mercury porosimeters have a range of application from
0.002 to 1000 μm. Nevertheless, to analyze smaller pores, intrusion pressures as high as
4500 bars would be necessary, with high risk of porous structure distortion.
Some mercury porometer suppliers are Carlo Erba, Fissons, Micromeritics, or Quan-
tachrome Co. In most of their devices, the amount of intruded mercury is determined
by the fall in the level of the interface between mercury and the compressing liquid.
Operationally, the sample is first evacuated and the penetrometer is then filled in the low
pressure port. Then a low pressure analysis is performed up to the maximum pressure
specified. Then the sample is moved to the high pressure port where a high pressure
analysis is performed up to the higher pressure achievable or selected.
From intruded mercury volume versus applied pressure data, the pore size distri-
bution can be obtained. The differential distribution of specific intruded volumes is
given by
p dV
Dv (dp ) = , (33)
dp dp

where p is the applied pressure and dp the corresponding pore diameter according to
the Washburn equation, while V is the specific volume (volume of mercury divided by
sample weight) and dV /dp is the derivative of specific intruded volume versus pres-
sure. Normally, data acquisition software uses appropriate derivation algorithms to get
smoothed curves.
Of course, a proper value of the contact angle should be used. For example, a value
of 130◦ for the contact angle seems to be valid in most materials and it is normally
implemented as default in most devices (136), along with the air–mercury surface tension,
γ = 0.474 N/m (140).
Many recent works can be found in literature regarding the use of mercury porosimetry
on characterization of membranes (101, 141, 142), ceramics (143), coal (144), and other
porous materials or nanoparticles (103, 145, 146).
28 POROSITY

8 CONCLUSIONS

In this article, we have briefly enumerated the main methods that are now available to
evaluate the pore size information of different types of membranes. When facing the
decision of what method we should use for analyzing a given membrane, there are some
questions we must take into account.
Those techniques related to permeation parameters (liquid displacement, solute reten-
tion test, PMP, etc.) allow determining the pore distribution for pores open to flux, and
then getting a mean pore size closer to the smaller pore diameter found in a cross section
of the membrane. These techniques are useful to characterize the skin or active layer,
but do not give any information about the inner structure of the membrane.
On the other hand, morphology-related techniques (microscopies, gas adsorption–
desorption, mercury porosimetry, etc.) give complete information about the whole pore
structure. Nevertheless, GAD or HgP need to assume a structural model for the pores,
and interpretation of the results becomes quite complex, especially for asymmetric mem-
branes. While microscopic techniques need combining surface and cross section images
of the membranes to know the actual pore structure, the latter are difficult to obtain
without artifacts.
In any case, all the techniques presented are useful as they contribute to obtain the
most complete picture of the actual pore structure of the membrane.
Therefore, if we should decide to use only one or two techniques, several questions
are to be considered, namely,

1. membrane type, material, or configuration, as some of these techniques are difficult


to adapt to different membrane configurations, or the material of the membrane can
suffer distortion or damage in during analysis;
2. moreover, it is important to decide whether we need to have the analyzed filter
ready for use after analysis, in which case, only nondestructive methods are to be
selected;
3. finally, there is always the issue about which of these techniques we have available
or can be acquired at a reasonable price.

Moreover, one must not forget the importance of knowing the actual membrane mate-
rial and its chemical structure (functional groups on the surface), which determine many
physical and chemical properties of the membrane. Spectroscopic techniques are there-
fore very useful and complementary to those techniques used to obtain the membrane
porosity and pore size distribution, giving a very complete picture of the membrane.
Finally, we must not forget that most of the methods described here can be used
for other porous materials, not only for membranes but also for catalyzers, minerals or
ceramic materials, powders, etc. Nevertheless, for those materials, especially powders,
if not consolidated, these techniques can lead to erroneous results. For example, those
techniques related to permeation parameters, such as GLDP, LLDP, or PMP, cannot be
applied to nonconsolidated powders.
Other interesting aspects of porosity characteristics of membranes have not been cov-
ered in this chapter, for example, the pore network models that have been used in past
decades to analyze porous materials or the fractal geometry that can be found in the sur-
face roughness analysis of membranes. These and other interesting points can be found
POROSITY 29

in many books and papers (147), but the limited space of this contribution does not allow
more extension.

REFERENCES

1. Mulder M. Basic Principles of Membrane Technology. Dordrecht, Netherlands: Kluwer; 1991.


2. Calvo JI. Caracterización de membranas de microfiltración. Aspectos estructurales y
funcionales [PhD thesis]. Valladolid, Spain: Universidad de Valladolid; 1995.
3. Cuperus P. Characterization of Ultrafiltration membranes. Pore structure and tope layer thick-
ness [PhD thesis]. Twente, Netherlands: Universiteit Twente; 1990.
4. Kamide K, Manabe S. Characterization technique of straight-through porous membranes. In:
Cooper AR, editor. Ultrafiltration Membranes and Applications. New York: Plenum Press;
1980.
5. Rasneur B. Porosimetry (characterization of porous membranes). Proceedings Summer School
on Membrane Sci. and Tech.; 1984; Cadarache, France.
6. Lowell S, Shields JE. Powder surface area and porosity. In: Scarlett B, editor. Powder Tech-
nology Series. New York: John Wiley & Sons, Inc.; 1987.
7. Palacio L. Caracterización estructural y superficial de membranas microporosas [PhD thesis].
Valladolid, Spain: Universidad de Valladolid; 1998.
8. Romanos GE, Stubos AK, Mitropoulos AC, Kanellopoulos NK. Pore structure characteriza-
tion of mesoporous inorganic membranes. Proceedings Euromembrane’97; Twente, Nether-
lands; 1997.
9. Bowen WR, Hilal N, Lovitt RW, Sharif AQ, Williams PM. Atomic force microscope stud-
ies of membranes: force measurements and imaging in electrolyte solution. Proceedings
Euromembrane’97; Twente, Netherlands; 1997.
10. Palacio L, Rodrı́guez ML, Prádanos P, Calvo JI, de Saja JA, Hernández A. Microscopic
study of a clean and fouled composite membrane. Proceedings Euromembrane’97; Twente,
Netherlands; 1997.
11. Hernández A, Calvo JI, Prádanos P, Palacio L. A multidisciplinary approach towards pore
size distributions of microporous and mesoporous membranes. In: Smith-Sørensen T, editor.
Surface Chemistry and Electrochemistry. New York: Marcel Dekker Inc.; 1998.
12. Bottino A, Capannelli G, Comite A, Ferrari F, Firpo R, Venzano S. Membrane technologies
for water treatment and agroindustrial sectors. C R Chimie 2009;12:882–888.
13. Riley RL, Gardner JO, Merten U. Cellulose acetate membranes: electron microscopy of
structure. Science 1964;143:801–803.
14. Binnig G, Quate CF, Gerber C. Atomic force microscope. Phys Rev Lett 1986;12:930–933.
15. Riedel C, Spohr R. Transmission properties of nuclear tranck filters. J Membr Sci 1980;
7:225–234.
16. Allen T. Volume 1, Particle Size Measurement. London: Chapman and Hall; 1997.
17. Kim KJ, Fane AG, Fell CJD, Suzuki T, Dickson MR. Quantitative microscopic study of
surface characteristics of UF membranes. J Membr Sci 1990;54:89–102.
18. Fritzsche AK, Arevalo AR, Moore MD, Weber CJ, Elings VB, Kjoller K, Wu CM. The
surface structure and morphology of polyvinylidene fluoride microfiltration membranes by
atomic force microscopy. J Membr Sci 1992;68:65–78.
19. Prádanos P, Rodriguez ML, Calvo JI, Hernández A, Tejerina F, de Saja JA. Structural
characterization of an UF membrane by gas adsorption-desorption and AFM measurements.
J Membr Sci 1996;117:291–302.
30 POROSITY

20. Calvo JI, Prádanos P, Hernández A, Bowen WR, Hilal N, Lovitt RW, Williams PM. Bulk
and surface characterization of composite UF membranes atomic force microscopy, gas
adsorption-desorption and liquid displacement techniques. J Membr Sci 1997;128:7–21.
21. Zeman L, Denault L. Characterization of microfiltration membranes by image analysis of
electron micrographs. Part. I. Method development. J Membr Sci 1992;71:221–231.
22. Swenson RA, Attle JR. Counting, measuring, and classifying with image analysis. Am Lab
1979;11:50–67.
23. Calvo JI, Hernández A, Caruana G, Martı́nez L. Pore size distributions in microporous mem-
bvranes. I. Surface study of track-etched filters by image analysis. J Colloid Interface Sci
1995;175:138–150.
24. Shilyaev PA, Pavlov DA. Fractal based approach to characterization of surface Geometry.
Proceedings of International Workshop “Scanning Probe Microscopy–2004”. N. Novgorod:
Institute for Physic of Microstructures, Russian Academy of Science; 2004.
25. Macanas J, Palacio L, Prádanos P, Hernández A, Muñoz M. Atomic force microscopy as a
suitable technique for surface characterization of activated membranes for metal ion facilitated
transport. Appl Phys A 2006;84:277–284.
26. Recio R, Palacio L, Prádanos P, Hernández A, Lozano AE, Marcos A, de la Campa JG,
de Abajo J. Permeability and selectivity of 6FDA-6FpDA gas membranes prepared from
different solvents. Desalination 2006;200:225–226.
27. Le MS, Howell JA. Alternative model for ultrafiltration. Chem Eng Res Des 1984;
62:373–380.
28. Nobrega R, de Balmann H, Aimar P, Sanchez V. Transfer of dextran through ultrafiltration
membranes: a study of rejection data analyzed by gel permeation chromatography. J Membr
Sci 1989;45:17–36.
29. Ochoa NA, Prádanos P, Palacio L, Pagliero C, Marchese J, Hernández A. Pore size distribu-
tion based on AFM Imaging and Retention of Multidisperse Polymer Solutes. Characterization
of Polyethersulfone UF membranes with dopes containing different PVP. J Membr Sci
2001;187:227–237.
30. Otero JA, Mazarrasa O, Villasante J, Silva V, Pradanos P, Calvo JI, Hernandez A. Three inde-
pendent ways to obtain information on pore size distributions of nanofiltration membranes. J
Membr Sci 2008;309:17–27.
31. Evans RB, Watson GM, Mason EA. Gaseous diffusion m porous media at uniform pressure.
J Chem Phys 1961;35:2076–2083.
32. Mason EA, Malinauskas AP. Gas Transport in Porous Media: The Dusty-gas Model . Ams-
terdam: Elsevier; 1983.
33. Schofield RW, Fane AG, Fell CJD. Gas and vapour transport through microporous mem-
branes. I Knudsen-Poiseuille transition. J Membr Sci 1990;53:159–171.
34. Datta R, Dechapanichkul S, Kim JS, Fang LY, Uehara H. A generalized model for the
transport of gases in porous, non-porous, and leaky membranes. I. Application to single
gases. J Membr Sci 1992;75:245–263.
35. Veldsink JW, Versteeg GF, van Swaaij WPM. An experimental study of diffusion and con-
vection of multicomponent gases through catalytic and noncatalytic membranes. J Membr
Sci 1994;92:275–291.
36. Beuscher U, Gooding CH. Characterisation of the porous support layer of composite gas
permeation. J Membr Sci 1997;132:213–227.
37. Wang D, Li K, Teo WK. Effect of temperature and pressure on gas permselection properties
in asymmetric membranes. J Membr Sci 1995;105:89–101.
38. Deshmukh SP, Li K. Effect of ethanol composition in water coagulation bath on morphology
of PVDF hollow fibre membranes. J Membr Sci 1998;150:78–85.
POROSITY 31

39. Changrong X, Huaqiang C, Hong W, Pinghua Y, Guangyao M, Dingkun P. Sol-gel synthesis


of yttria stabilized zirconia membranes through controlled hydrolysis of zirconium alkoxide.
J Membr Sci 1999;1621:181–188.
40. Li K, Kong JF, Wang D, Teo WK. Tailor-made asymmetric PVDF hollow fibers for soluble
gas removal. AIChE J 1999;45:1211–1219.
41. Fernández-Pineda C, Izquierdo-Gil MA, Garcı́a-Payo MC. Gas permeation and direct contact
membrane distillation experiments and their analysis using different models. J Membr Sci
2002;198:33–49.
42. Khayet M, Feng CY, Khulbe KC, Matsuura T. Preparation and characterization of polyvinyli-
dene fluoride hollow fiber membranes for ultrafiltration. Polymer 2002;43:3879–3890.
43. Pakizeh M, Omidkhah MR, Zarringhalam A. Study of mass transfer through new templated
silica membranes prepared by sol-gel method. Int J Hydrogen Energy 2007;32:2032–2042.
44. Calvo JI, Bottino A, Capannelli G, Hernández A. Pore size distribution of ceramic UF mem-
branes by liquid-liquid displacement porosimetry. J Membr Sci 2008;310:531–538.
45. Gao F, Chena XC, Yua G, Asumanaa C. Compressible gases transport through porous mem-
brane: a modified dusty gas model. J Membr Sci 2011;379:200–206.
46. Kulkarni SS, Funk EW, Li NN. Membranes. In: Ho WSW, Sirkar KK, editors. Membrane
Handbook . New York: Van Nostramd Reinhold; 1992.
47. Scheele CW. Chemical Observations on Air and Fire. London: J. Johnson; 1780. p 182–197.
48. Sing KSW. Reporting physisorption data for gas/solid systems with special reference to
the determination of surface area and porosity (Recommendations 1984). Pure Appl Chem
1985;57:603–619.
49. Rouquerol J, Avnir D, Fairbridge CW, Everett DH, Haynes JM, Pernicone N, Ramsay JDF,
Sing KSW, Unger KK. Recommendations for the characterization of porous solids (technical
report). Pure Appl Chem 1994;66:1739–1758.
50. Langmuir I. The adsorption of gases on planes surfaces of glass, mica and platinum. J Am
Chem Soc 1918;40:1361–1368.
51. Burgess CGV, Everett DH, Nuttall S. Adsorption hysteresis in porous materials. Pure Appl
Chem 1989;61:1845–1852.
52. Brunauer S, Emmett PH, Teller E. Adsorption of gasses in multimolecular layers. J Am Chem
Soc 1938;60:309–319.
53. Seifer J, Emig G. Mikrostrukturuntersuchungen an porósen Feststoffen durch Physisorption-
smessungen. Chem Eng Technol 1987;59:475–484.
54. Barrett EP, Joyner LG, Halenda PP. Determination of pore volume and area distribu-
tions in porous substances. I. Computations from nitrogen isotherms. J Am Chem Soc
1951;73:373–380.
55. Dollimore D, Heal GR. An improved method for the calculation of pore size distribution
from adsorption date. J Appl Chem 1964;14:109–114.
56. Gregg SJ, Sing KSW. Adsorption, Surface Area and Porosity. London: Academic Press;
1982.
57. Kaneko K. Determination of pore size and pore size distribution. 1. Adsorbents and catalysis.
J Membr Sci 1994;96:59–89.
58. Dubinin MM. The potential theory of adsorption of gases and vapors for adsorbents with
energetically nonuniform surfaces. Chem Rev 1960;60:235–266.
59. Dubinin MM. Fundamentals of the theory of adsorption in micropores of carbon adsor-
bents: characteristics of their adsorption properties and microporous structures. Carbon
1989;27:457–467.
60. Bhatia SK, Shethna HK. A modified pore filling with applications to determination of pore
size distributions. Langmuir 1994;10:3230–3243.
32 POROSITY

61. Horváth G, Kawazoe K. Method for the calculation of effective pore size distribution in
molecular sieve carbon. J Chem Eng Jpn 1983;16:470–475.
62. Gil A, Dı́az A, Montes M, Acosta DR. Characterization of the microporosity of pillores
clays by nitrogen adsorptions-application of the Horváth-Kawazoe approach. J Mater Sci
1994;29:4927–4932.
63. Mikhail RSh, Brunauer S, Bodor EE. Investigations of a complete pores structure analysis.
I. Analysis of micropores. J Colloid Interface Sci 1968;26:45–53.
64. Lippens BC, de Boer JH. Studies on pore systems in catalyts V. The t-method. J Catal
1965;4:319–323.
65. Harkins, WD, Jura GJ. The decrease of free surface energy as a basis for the development
of equations for adsorption isotherms; and the existence of two condensed phases in films on
solids. J. Chem. Phys 1944;12:112–113.
66. Felder RM, Huvard GS. Permeation, diffusion and sorption of gases and vapors. In: Fava RA,
editor. Methods of Experimental Physics. Volume 16, Part C, Polymers physical properties.
New York: Academic Press; 1980.
67. Palacio L, Prádanos P, Calvo JI, Hernández A. Porosity measurements by a gas penetra-
tion method and other techniques applied to membrane characterization. Thin Solid Films
1999;348:22–29.
68. Mourhatch R, Tsotsis TT, Sahimi M. Determination of the true pore size distribution by flow
permporometry experiments: an invasion percolation model. J Membr Sci 2011;367:55–62.
69. Korelskiy D, Grahn M, Mouzon J, Hedlund J. Characterization of flow-through micropores
in MFI membranes by permporometry. J Membr Sci 2012;417–418:183–192.
70. Brun M, Lallemand A, Quinson JF, Eyraud Ch. Changement d’état liquide-solide dans les
milieux poreux. I. Étude experimentale de la solidification de l’eau et du benzène. J Chim
Phys 1973;70(6):973–978.
71. Eyraud Ch, Bontemps M, Quinson JF, Chatelut F, Brun M, Rasneur B. Determination
of the pore-size distribution of an ultrafilter by-gas-liquid permporometry measurement-
comparison between flow porometry and condensate equilibrium porometry. Bull Soc Chim
Fr 1984;9–10:I237–I244.
72. Cuperus FP, Bargeman D, Smolders CA. Critical-points in the analysis of membrane pore
structures by thermoporometry. J Membr Sci 1992;66:45–53.
73. Broek AP, Teunis HA, Bergeman D, Sprengers ED, Smolders CA. Characterization of hol-
low fiber hemodialysis membranes-pore-size distribution and performance. J Membr Sci
1992;73:143–152.
74. Riikonen J, Salonen J, Lehto V-P. Utilising thermoporometry to obtain new insights into
nanostructured materials. Review part 1. J Therm Anal Calorim 2011;105:811–821.
75. Riikonen J, Salonen J, Lehto V-P. Utilising thermoporometry to obtain new insights into
nanostructured materials. Review part 2. J Therm Anal Calorim 2011;105:823–830.
76. Stapf S, Kimmich R. Molecular dynamics in confined monomolecular layers. A field-cycling
nuclear magnetic resonance relaxometry study of liquids in porous glass. J Chem Phys
1995;103:2247–2250.
77. Valckenborg R, Pel L, Kopinga K. Cryoporometry and relaxometry of water in silica-gels.
Magn Reson Imaging 2001;19:489–491.
78. Glaves CL, Smith DM. Membrane pore structure analysis via nmr spin lattice relaxation
experiments. J Membr Sci 1989;46:167–184.
79. Strange JH, Rahman M. Characterization of porous solids by NMR. Phys Rev Lett
1993;71:3589–3591.
80. Allen SG, Mallett MJD, Strange JH. Morphology of porous media studied by nuclear magnetic
resonance line shapes and spin-echo decays. J Chem Phys 2001;114:3258–3264.
POROSITY 33

81. Allen SG, Stephenson PCL, Strange JH. Morphology of porous media studied by nuclear
magnetic resonance. J Chem Phys 1997;106:7802–7809.
82. Jeon JD, Kim SJ, Kwak SY. 1H nuclear magnetic resonance (NMR) cryoporometry as
a tool to determine the pore size distribution of ultrafiltration membranes. J Membr Sci
2008;309:233–238.
83. Bechhold H. The permeability of ultrafilters. Z Phys Chem 1908;64:328–342.
84. Kesting RE. Synthetic Polymeric Membranes: A Structural Perspective. 2nd ed. New York:
John Wiley & Sons, Inc.; 1985.
85. Hernández A, Calvo JI, Prádanos P, Tejerina F. Pore size distributions in microporous mem-
branes. A critical analysis of the bubble point extended method. J Membr Sci 1996;112:
1–12.
86. Calvo JI. Membrane characterization by porosimetric techniques. Lesson Presented at XXIV
Summer School on Membranes; 2007; Genoa, Italy.
87. Gadam S, Phillips M, Orlando S, Kuriyel R, Pearl S, Zydney A. A liquid porosimetry tech-
nique for correlating intrinsic protein sieving: applications in ultrafiltration processes. J Membr
Sci 1997;133:111–125.
88. Arkhangelskya E, Dueka A, Gitis V. Maximal pore size in UF membranes. J Membr Sci
2012;394–395:89–97.
89. Dosmar M, Wolber P, Bracht K, Troger H, Waibel P. A new in-place integrity test for
hydrophobic membrane filters. Filtr Sep 1993: 305–309.
90. Farahbakhsh K, Smith DW. Estimating air diffusion contribution to pressure decay during
membrane integrity tests. J Membr Sci 2004;237:203–212.
91. Peinemann K-V, Pereira Nunes S, editors. Membrane Technology. Volume 1, Membranes for
life sciences. Germany: Wiley-VCH Verlag GmbH; 2007.
92. Giglia S, Krishnan M. High sensitivity binary gas integrity test for membrane filters. J Membr
Sci 2008;323:60–66.
93. DiLeo A, Phillips M. Process for evaluating solute retention characteristics of membranes.
US patent 5.457.986; 1995.
94. DiLeo A, Phillips M. Integrity test for membranes. US patent 5.282.380; 1994.
95. Phillips M, DiLeo A. A validatable porosimetric technique for verifying the integrity of virus
retentive membranes. Biologicals 1996;24:243–253.
96. Guo H, Wyart Y, Perot J, Nauleau F, Moulin P. Low-pressure membrane integrity tests for
drinking water treatment: a review. Water Res 2010;44:41–57.
97. Erbe F. Blockierungsphenomene bei ultrafiltem. Kolloid Z 1932;59:32–44.
98. ASTM F316. Standard test method for pore size characteristics of membrane filters by bubble
poitn and mean flow pore test; 2003.
99. ASTM E1294. Standard test methods for pore size characteristics of membrane filters using
automated liquid porosimeter; 1999.
100. Shrestha A, Pellegrino J, Husson SM, Wickramasinghe SR. A modified porometry
approach towards characterization of MF membranes. J Membr Sci 2012;421–422:
145–153.
101. Gribble CM, Matthews GP, Laudone GM, Turner A, Ridgway CJ, Schoelkopf J, Gane PAC.
Porometry, porosimetry, image analysis and void network modeling in the study of the pore-
level properties of filters. Chem Eng Sci 2011;66:3701–3709.
102. Li D, Frey MW, Joo YL. Characterization of nanofibrous membranes with capillary flow
porometry. J Membr Sci 2006;286:104–114.
103. Manickam SS, McCutcheon JR. Characterization of polymeric nonwovens using
porosimetry, porometry and X-ray computed tomography. J Membr Sci 2012;407–408:
34 POROSITY

108–115.
104. Herrero C, Prádanos P, Calvo JI, Tejerina F, Hernández A. Flux decline in pro-
tein microfiltration: influence of operative parameters. J Colloid Interface Sci 1997;187:
344–351.
105. Erbe F. Die bestimmung der porenverteilung nach ihrer größe in filtern und ultrafiltern.
Kolloid Z 1933;63:277–285.
106. Grabar P, Nikitine S. Sur le diamètre des pores des membranes en collodion utilisées en
ultrafiltration. J Chim Phys 1936;33:721–741.
107. Hampl V, Spurný K. Analytical methods for determination of aerosols by means of
membrane ultrafilters. X. Comparison of methods used for the determination of the
mean radius and the pore distributions curve. Collect Czech Chem Commun 1967;32:
4181–4189.
108. Capannelli G, Vigo F, Munari S. Ultrafiltration membranes–characterization methods. J
Membr Sci 1983;15:289–313.
109. Kujawski W, Adamczak P, Narebska A. A fully automated system for the determination
of pore size distribution in microfiltration and ultrafiltration membranes. Sep Sci Technol
1989;24:495–506.
110. Wienk IM, Folkers B, van den Boomgaard T, Smolders CA. Critical factors in the determi-
nation of the pore size distribution of ultrafiltration membranes using the liquid displacement
method. Sep Sci Technol 1994;29:1433–1440.
111. McGuire KS, Lawson KW, Lloyd DR. Pore size distribution determination from liquid per-
meation through microporous membranes. J Membr Sci 1995;99:127–137.
112. Jakobs E, Koros WJ. Ceramic membrane characterization via the bubble point technique.
J Membr Sci 1997;124:149–159.
113. Gadam S, Phillips M, Orlando S, Kuriyel R, Pearl S, Zydney A. A liquid porosimetry tech-
nique for correlating intrinsic protein sieving: applications in ultrafiltration processes. J Membr
Sci 1997;133:111–125.
114. Lee Y, Jeong J, Youn IJ, Lee WH. Modified liquid displacement method for deter-
mination of pore size distribution in porous membranes. J Membr Sci 1997;130:
149–156.
115. Wu LQ, Huang P, Xu N, Shi J. Effects of sol properties and calcination on the performance
of titania tubular membranes. J Membr Sci 2000;173:263–273.
116. Capannelli G, Becchi I, Bottino A, Moretti P, Munari S. Computer driven porosimeter for
ultrafiltration membranes. In: Rodrı́guez-Reinoso F, Rouquerol J, Singh KSW, Unger KK,
editors. Characterization of Porous Solids. Amsterdam, The Netherlands: Elsevier Science,
Publishers B.V.; 1988. p 283–294.
117. Munari S, Bottino A, Capannelli G, Moretti P. Membrane morphology and transport proper-
ties. Desalination 1985;53:11–23.
118. Munari S, Bottino A, Moretti P, Capannelli G, Becchi I. Permoporometric study on ultrafil-
tration membranes. J Membr Sci 1989;41:69–86.
119. Gekas V, Tragadh G, Capannelli G, Bottino A. Correlation of direct porosimetric data and
performance of ultrafiltration membranes. Proc Biochem Int 1990;25:111–116.
120. Bottino A, Capannelli G, Petit-Bon P, Pegoraro M, Zoia G. Pore size and pore size distribution
in microfiltration membranes. Sep Sci Technol 1991;26:1315–1327.
121. Persson KM, Capannelli G, Bottino A, Trägårdh G. Porosity and protein adsorption of four
polymeric microfiltration membranes. J Membr Sci 1993;76:61–71.
122. Bottino A, Capannelli G, Grosso A, Monticelli O, Nicchia M. Porosimetric characterization
of inorganic membranes. Sep Sci Technol 1994;29:985–999.
POROSITY 35

123. Germic L, Ebert K, Bouma RHB, Borneman Z, Mulder MHV, Strathmann H. Characterization
of polyacrylonitrile ultrafiltration membranes. J Membr Sci 1997;132:131–145.
124. Lindau J, Jonsson A-S, Bottino A. Flux reduction of ultrafiltration membranes with different
cut-off due to adsorption of a low-molecular-weight hydrophobic solute-correlation between
fux decline and pore size. J Membr Sci 1998;149:11–20.
125. Calvo JI, Bottino A, Capannelli G, Hernandez A. Comparison of liquid–liquid displacement
porosimetry and scanning electron microscopy image analysis to characterise ultrafiltration
track-etched membranes. J Membr Sci 2004;239:189–197.
126. Sanz JM, Jardines D, Bottino A, Capannelli G, Hernández A, Calvo JI. Liquid–liquid porom-
etry for an accurate membrane characterization. Desalination 2006;200:195–197.
127. Almecija MC, Zapata JE, Martinez-Ferez A, Guadix A, Hernandez A, Calvo JI. Analy-
sis of cleaning protocols in ceramic membranes by liquid-liquid displacement porosimetry.
Desalination 2009;245:541–545.
128. Bottino A, Capannelli G, Comite A, Mangano R. Critical flux in submerged membrane
bioreactors for municipal wastewater treatment. Desalination 2009;245:748–753.
129. Sanz JM, Peinador R, Calvo JI, Hernández A, Bottino A, Capannelli G. Characteriza-
tion of UF membranes by liquid–liquid displacement porosimetry. Desalination 2009;245:
546–553.
130. Peinador RI, Calvo JI, Pradanos P, Palacio L, Hernandez A. Characterisation of poly-
meric UF membranes by liquid-liquid displacement porosimetry. J Membr Sci 2010;348:
238–244.
131. Peinador RI, Calvo JI, ToVinh K, Thom V, Pradanos P, Hernandez A. Liquid-liquid dis-
placement porosimetry for the characterization of virus retentive membranes. J Membr Sci
2011;372:366–372.
132. Fu J, Li B, Wang Z. Estimation of fluid-fluid interfacial tensions of multicomponent mixtures.
Chem Eng Sci 1986;41:2673–2679.
133. Morison KR. A comparison of liquid–liquid porosimetry equations for evaluation of pore
size distribution. J Membr Sci 2008;325:301–310.
134. PMI Porous Materials, Inc. Liquid-liquid porometer. Available at http://www.pmiapp.com/
products/liquid-liquid-porometer.html. Accessed Dec 20, 2012.
135. Calvo JI, Peinador R, Pradanos P, Palacio L, Bottino A, Capannelli G, Hernandez A. Liquid-
liquid displacement porometry to estimate the molecular weight cut-off of ultrafiltration
membranes. Desalination 2011;268:174–181.
136. Washburn EW. The Dynamics of Capillary Flow. Physical Review 1921;17(3):273–283.
137. Honold E, Skau EL. Application of mercury intrusion method for determination of pore-size
distribution of membranes filters. Science 1954;120:805–806.
138. Roĉek J, Uchytil P. Evaluation of selected methods for the characterisation of ceramic mem-
branes. J Membr Sci 1994;89:119–129.
139. León y León CA. New perspectives in mercury porosimetry. Adv Colloid Interface Sci
1998;76–77:341–372.
140. Rootare HM. A short literature review of mercury porosimetry as a method of measuring
pore-size distributions in porous materials, and a discussion of possible sources of errors in
this method. Aminco Laboratory News 1968;24(3):4A–4H.
141. Sarkar S, Bandyopadhyay S, Larbot A, Cerneaux S. New clay–alumina porous capillary
supports for filtration application. J Membr Sci 2012;392–393:130–136.
142. Luiten-Olieman MWJ, Winnubst L, Nijmeijer A, Wessling M, Benes NE. Porous stainless
steel hollow fiber membranes via dry–wet spinning. J Membr Sci 2011;370:124–130.
143. El-Kady AM, Ali AF. Fabrication and characterization of ZnO modified bioactive glass
nanoparticles. Ceram Int 2012;38:1195–1204.
36 POROSITY

144. Yao Y, Liu D. Comparison of low-field NMR and mercury intrusion porosimetry in charac-
terizing pore size distributions of coals. Fuel 2012;95:152–158.
145. Rigby SP, Fletcher RS, Riley SN. Characterisation of porous solids using integrated nitrogen
sorption and mercury porosimetry. Chem Eng Sci 2004;59:41–51.
146. Manso JM, Rodriguez A, Aragón A, Gonzalez JJ. The durability of masonry mortars made
with ladle furnace slag. Constr Build Mater 2011;25:3508–3519.
147. Sahimi M. Flow and Transport in Porous Media and Fractured Rock: From Classical Methods
to Modern Approaches. 2nd ed. Weinheim, Germany: Wiley-VCH; 2011.

You might also like