Sie sind auf Seite 1von 10

AST 22(10)_64 31/01/05 1:31 pm Page 773

773

Review

Physisorption Hysteresis Loops and the Characterization of Nanoporous


Materials
Kenneth S.W. Sing1 and Ruth T. Williams2∗ (1) School of Chemistry, University of Exeter, Stocker Road,
Exeter EX4 4QD, UK. (2) Chemistry Department, The Open University, Walton Hall, Milton Keynes MK7 6AA, UK.

(Received 8 June 2004; accepted 15 June 2004)

ABSTRACT: The classification of adsorption hysteresis loops recommended


by the IUPAC in 1984 was based on experimental observations and the applica-
tion of classical principles of pore filling (notably the use of the Kelvin equation
for mesopore analysis). Recent molecular simulation and density functional
(DFT) studies of the physisorption of gases by model pore structures have
greatly improved our understanding of the mechanisms of hysteresis and it is
therefore timely to revisit the IUPAC recommendations. In this review, we con-
clude that there is no immediate need to change the IUPAC classification of
physisorption isotherms and hysteresis loops. However, in the light of recent
advances, we are able to offer a revised checklist for the analysis of nitrogen
isotherms on nanoporous solids: this includes a carefully regulated application
of DFT in place of a classical procedure such as the well-known
Barrett–Joyner–Halenda (BJH) method.

HISTORICAL INTRODUCTION

From the standpoint of classical thermodynamics, the physisorption (i.e. physical adsorption) of a
non-reactive vapour by a porous solid is a reversible process. The general form of the physisorp-
tion isotherm for a given vapour–solid system at a temperature T is:

(
n = f P / P0 ) T , system
(1)

where n is the amount adsorbed at P/P0, the relative pressure. Provided that equilibrium is attained,
it follows that the value of n should depend only on P/P0 and not be affected by any previous
changes of pressure. In accordance with equation (1), physisorption isotherms of non-reactive
gases on stable, non-porous surfaces are generally completely reversible and therefore the corre-
sponding adsorption and desorption pathways are identical (Gregg and Sing 1982). In contrast,
physisorption isotherms given by many porous materials are not reversible over particular ranges
of P/P0. This phenomenon is termed adsorption hysteresis, but its exact form and extent are
dependent on the system and the operational temperature.
Hysteresis in the adsorption of water vapour by silica gel was reported by van Bemmelen (1896),
but the first detailed theory of adsorption hysteresis was put forward by Zsigmondy (1911),

*
Author to whom all correspondence should be addressed. E-mail: r.t.williams@open.ac.uk.
AST 22(10)_64 31/01/05 1:31 pm Page 774

774 K.S.W. Sing and R.T. Williams/Adsorption Science & Technology Vol. 22 No. 10 2004

who along with many others [see Brunauer (1944)] suggested the likelihood of its close connection
with capillary condensation. As is well known, the equilibrium vapour pressure above a curved
liquid surface is dependent on the mean radius of curvature, rK, of its meniscus. This dependence
is given by the Kelvin equation, which is usually expressed in the approximate form:

(
ln P / P 0 ) = −2 γ ν1 / rK RT (2)

where γ is the surface tension and νl the molar volume of the liquid condensate; R and T are
the gas constant and temperature, respectively. When equation (2) is applied to vapour sorption,
it is assumed that the meniscus curvature is directly controlled by the pore size and shape.
However, it was realised at an early stage that γ and νl are unlikely to remain constant at high
meniscus curvature and that the Kelvin equation is no longer applicable when the pore size is
reduced to molecular dimensions. Indeed, by the 1930s, it was accepted by most investigators that
there is always some form of adsorption on the pore walls before the onset of capillary condensa-
tion (Foster 1932; Brunauer 1944).
In the simplest case of adsorption in uniform cylindrical pores, the meniscus curvature is
usually assumed to have radial symmetry. The meniscus shape is therefore hemi-spherical if the
contact angle, θ, is zero. Then the pore radius, rp, is:

rp = rK + t (3)

where t is the thickness of the pre-adsorbed layer. However, if there is a finite contact angle, θ,
between the adsorbed layer and the condensate, the relation becomes:

rp = rK cos θ + t (4)

When capillary condensation occurs in slit-shaped pores, the meniscus curvature is assumed to
have cylindrical symmetry so the shape is then hemi-cylindrical if θ = 0. We therefore have:

w p = rK + 2 t (5)
where wp is now the effective slit width.
The relation between rp and rK is more complex if the pores are neither cylindrical nor slit-
shaped. For example, in the important case of the interstices between globular particles, capillary
condensation occurs in stages which are controlled by the particle-size distribution and the type
of particle packing (Haynes 1975). Furthermore, as argued by Derjaguin (1957) and Everett
(1958), equation (4) does not allow for the effect of the adsorption forces on the meniscus curva-
ture and θ is also likely to vary with the distance from the surface.
Returning to the question of the origin of physisorption hysteresis, we now see that there are
several ways in which capillary condensation may play a central role. First, there is the possibil-
ity (Zsigmondy 1911) that on going from adsorption to desorption, θ is reduced from a finite value
to zero. This may seem to be a reasonable hypothesis in view of the difference often found
between advancing and receding contact angles, but it would be expected to result in a reduction
in the size of the hysteresis loop when the adsorption/desorption cycle is repeated. Such changes
have been recorded [see Brunauer (1944)], but they can be attributed to either imperfect out-
gassing or modification of the pore structure [see Rouquerol et al. (1999)]. In fact, many hysteresis
loops are known to be completely reproducible over long periods of time (Rao 1941) and the
AST 22(10)_64 31/01/05 1:31 pm Page 775

Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials 775

permanence of the boundary curves has been demonstrated by systematic scanning experiments
(Rao 1941; Everett 1955, 1967). The extensive measurements of various loops over wide ranges
of temperature by Everett and his co-workers (Burgess et al. 1989) were discussed in terms of
hysteresis phase diagrams and critical temperatures.
It was suggested by Foster (1932) that hysteresis is due to a delay in meniscus formation.
According to this idea, capillary condensation equilibrium, as represented by the Kelvin equation,
is established over the desorption branch, while multilayer adsorption plays a more important role
along the adsorption branch. More recently, these principles have been refined with particular
reference to the regimes of stability, meta-stability and instability of the multilayer (Saam and
Cole 1975; Everett 1979).
An early attempt by Cohan (1938) to predict the width of hysteresis loops was based on the
notion that capillary condensation in an open-ended cylindrical pore entails the initial formation
of a cylindrical meniscus, while evaporation can take place from the hemi-spherical meniscus
formed at each end of the filled pore. The Cohan theory was generally accepted (Brunauer 1944;
de Boer 1958) until it was shown by Everett and Haynes (1972) that the mechanism is an over-
simplification since a long cylindrical film is less stable than the equivalent unduloid, which has
uniform mean curvature and contains the same amount of material. The unduloidal film surface
then allows bridging to occur spontaneously across the pore.
All the evidence already referred to leads us to believe that thermodynamic equilibrium is not
completely attainable over the adsorption branch of a loop. However, according to the ‘ink bottle’
theory of Kraemer (1931) and McBain (1935), equilibrium is not established along the desorption
branch if part of the condensate is trapped in wide cavities with narrow entrances. In this case, the
release of the retained condensate takes place only when the pore entrances become unblocked
when the value of P/P0 is reduced to an appropriate level. Furthermore, it has been tacitly assumed
by many investigators that pores of a given size can behave as a set of ‘independent domains’
(Everett 1955) and that the processes of filling and emptying do not depend on the presence or
absence of neighbouring pores. Scanning experiments have provided strong evidence (Everett
1967) that with many porous adsorbents this is not the case, and that the ease of removal of
condensate from an individual pore will depend on whether there is a clear channel linking the
pore to the outer surface. Indeed, it has become increasingly evident that most porous materials
possess complex networks of pores of different size and shape (Levitz 2002).
It might be thought that the dimensions of the pore entrances or constrictions within the
network could be assessed from the desorption branch of the loop (de Boer 1958). However, as
was first pointed out by Harris (1965), at 77 K many nitrogen isotherms exhibit very steep des-
orption curves at ∼ 0.42 < P/P0 < 0.50, which would seem to imply that a large proportion of
adsorbents possess pores in the narrow range of ∼ 1.7 < rp < 2 nm. In fact, the abrupt cut-off at
P/P0 = 0.42 corresponds to the lower closure point of a large number of nitrogen hysteresis loops.
Similar results were later observed with other adsorptives and it was concluded that the position
of each lower closure point corresponded to the lower limit of stability for the particular capillary
condensate at a given temperature [see Gregg and Sing (1982)].
It is consistent with these findings that well-defined hysteresis loops begin to appear in the multi-
layer/capillary condensation range when the pore size is increased to more than a few molecular
diameters (Brunauer 1944). However, other forms of sorption hysteresis have been observed at lower
P/P0 values (i.e. in the normal monolayer range of the isotherm). The explanations which have been
advanced for these phenomena include intercalation and expansion of layer structures and the ‘acti-
vated entry’ of molecules through narrow pore entrances [see Gregg and Sing (1982); Rouquerol et al.
(1999)]. These low-pressure forms of sorption hysteresis fall outside the scope of the present review.
AST 22(10)_64 31/01/05 1:31 pm Page 776

776 K.S.W. Sing and R.T. Williams/Adsorption Science & Technology Vol. 22 No. 10 2004

The nomenclature of pores of different size has a long and chequered history, but as a result of
the pioneering work of Dubinin (1960, 1966) the term ‘micropore’ is now reserved for pores no
wider than ∼ 2 nm (Gregg and Sing 1982; Sing et al. 1985). The somewhat wider range of pores
of 2–50 nm, which are of particular significance in the context of capillary condensation, are
known as ‘mesopores’ (Sing et al. 1985). In recent years, it has become customary to refer
collectively to wide micropores (supermicropores) and narrow mesopores (i.e. no wider than a
few nm) as ‘nanopores’.
Several attempts have been made to classify hysteresis loops (Barrer et al. 1956; de Boer 1958).
The classification adopted by the IUPAC in 1984 [see Sing et al. (1985)] is shown in Figure 1. To a
large extent, this was based on the classical interpretation of hysteresis in terms of capillary conden-
sation and the extensive experimental work of Everett (1967), Burgess et al. (1989) and others [see
Gregg and Sing (1982)]. In view of the current interest in the application of molecular simulation and
density functional theory, this would seem an appropriate time to revisit the IUPAC classification.
The interpretation of physisorption hysteresis is important in the context of pore structure char-
acterization. Thus, nitrogen sorption measurements at 77 K are routinely used for characterizing
porous materials. The computational procedure proposed by Barrett, Joyner and Halenda (1951),
which is universally known as the BJH method, remains the most popular method for the

H1 H2
Amount adsorbed, n

Amount adsorbed, n

Relative pressure, P/P0 Relative pressure, P/P0

H3 H4
Amount adsorbed, n

Amount adsorbed, n

Relative pressure, P/P0 Relative pressure, P/P0

Figure 1. The four types of hysteresis loops identified by IUPAC.


AST 22(10)_64 31/01/05 1:31 pm Page 777

Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials 777

determination of the mesopore-size distribution, but the development of other user-friendly com-
mercial software for pore-size analysis has recently attracted a considerable amount of interest
(see Rouquerol et al. 1999, 2002).

Hysteresis loops of types H1 and H2

Hysteresis loops of types H1 and H2 in Figure 1 are given by a variety of mesoporous oxide
xerogels and also different forms of porous glass. Such pore structures can be regarded as virtually
rigid and, since each isotherm has a well-defined plateau at high P/P0, it is possible to obtain an
unambiguous assessment of the mesopore volume (Rouquerol et al. 1999). Other notable features
of H1 and H2 loops are the upper and lower closure points associated with the boundary curves.
The interpretation of H1 and H2 loops has been greatly facilitated by the development in recent
years of a number of ordered silicate/aluminosilicate pore structures, the best known being
MCM-41 (Kresge et al. 1992). The uniform structure of MCM-41 consists of hexagonal arrays of
non-intersecting tubular nanopores of controlled diameter (typically 2.5, 3.4, 4.0 or 4.5 nm). The
considerable number of physisorption isotherms determined on these grades of MCM-41 over the
past 10 years (see Rouquerol et al. 1999) can be broadly divided into the three representative types
shown in Figure 2. Isotherms a and b are completely reversible, whereas isotherm c exhibits the
characteristic H1 hysteresis loop.

a b c
Amount adsorbed, n

Filled symbols correspond to desorption

0 0.2 0.4 0.6 0.8 1


Relative pressure, P/P0

Figure 2. Representative adsorption isotherms.


AST 22(10)_64 31/01/05 1:31 pm Page 778

778 K.S.W. Sing and R.T. Williams/Adsorption Science & Technology Vol. 22 No. 10 2004

Nitrogen isotherms determined at 77 K on 2.5-nm and 4.5-nm diameter samples of MCM-41


are respectively similar to a and c in Figure 2, whereas the nitrogen isotherm on the 4.0-nm sam-
ple is similar to b. This sharp pore-filling/emptying step is apparently located at the lower limit of
capillary condensation hysteresis (i.e. P/P0 ∼ 0.42 for nitrogen at 77 K), the condensate becoming
unstable at lower P/P0 values. The isotherms of argon and oxygen at 77 K determined on the
4.0-nm sample of MCM-41 sample (Branton et al. 1993, 1995) are both similar to c in Figure 2.
However, at 87 K, argon isotherms corresponding to a, b and c were reported for MCM-41
samples with respective pore diameters of 3.1, 3.6 and 4.0 nm (Neimark et al. 2003). These results
appear to confirm the dependence of the lower limit of hysteresis on the adsorptive and opera-
tional temperature.
Alongside the development of model pore structures, various attempts have been made to
interpret capillary condensation hysteresis in terms of the statistical mechanics of confined fluids
(Evans et al. 1986; Neimark et al. 2000; Seaton et al. 1989; Jorge and Seaton 2002; Pikunic et al.
2002). In particular, our understanding of the mechanisms involved in nanopore filling have been
greatly improved by the application of the Monte Carlo (MC) and molecular dynamics (MD)
methods of molecular simulation and the use of density functional theory (DFT). Recent work has
included the development of a new gauge-cell MC simulation method (Neimark and Vishnyakov
2000; Jorge and Seaton 2002) which involves the use of a reference gauge cell of limited capac-
ity and thus permits the level of density fluctuations to be controlled. This procedure allows one
to construct the complete phase diagram for a confined fluid in the form of a van der Waals loop
and thereby explore the limits of stability (referred to as spinodals) of the vapour-like and liquid-
like states.
Although still in its developmental stage, the gauge-cell approach has already provided a basis
for analyzing the different conditions for spontaneous and reversible transitions as revealed by
computer simulation and experimentation (Vishnyakov and Neimark 2001). It has also been pos-
sible to confirm the Everett–Haynes unduloidal film-bridging mechanism (Kornev et al. 2002;
Vishnyakov and Neimark 2003). By extending the application of the gauge-cell method, Jorge and
Seaton (2002) were able to make a direct comparison between the mechanisms of adsorption of
water and methane in graphitic nanopores. As expected, the pore-filling mechanisms were found
to be quite different: in contrast to methane, water does not form a dense monolayer on the inter-
nal carbon surface and the formation of a liquid-like bridge is therefore dependent on the
development of hydrogen-bonded clusters.
Very recently, Neimark et al. (2003) have used non-local density functional theory (NLDFT)
along with MC gauge-cell simulation to explain the changes in isotherm shape illustrated in Figure 2.
These and other fundamental studies (Lastoskie and Gubbins 2000) strongly suggest that the clas-
sical Kelvin–Cohan calculation of pore size begins to break down for pores as large as 10 nm.
We next consider the interpretation of type H2 hysteresis loops (see Figure 1) which were once
thought to be always associated with percolation effects in highly heterogeneous pore networks
(Gregg and Sing 1982; Sing et al. 1985). The complexity of capillary condensation in such hetero-
geneous pore structures is illustrated by the application of Monte Carlo simulation (Gelb and
Gubbins 1998) and percolation theory (Mason 1988; Seaton 1991). According to the latter
approach, the percolation threshold can be identified as the beginning of the steep desorption
branch of an H2 loop (Liu et al. 1992).
Very recently, a few highly regular pore structures have also been found to give H2 loops. One
example is SBA-16 — a templated, ordered silica structure, which has been shown by high-
resolution electron microscopy to consist of ∼ 10-nm spherical cavities arranged in a body-
centred array and connected through ∼ 2-nm openings (Sakamoto et al. 2000). Another ordered
AST 22(10)_64 31/01/05 1:31 pm Page 779

Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials 779

structure, which also gives typical argon and nitrogen H2 loops, is FDU-1 (Yu et al. 2000): this
has a regular periodic structure with cage-like pores of ∼ 15-nm diameter.
With particular reference to these cage-like structures, Neimark and his co-workers
(Ravikovitch and Neimark 2002a,b; Vishnyakov and Neimark 2003) are now engaged in a series
of systematic experimental and theoretical studies on pore-blocking effects. Special attention has
been given to the analysis of desorption scanning isotherms which have revealed some previously
unexplored features and have thrown new light on the mechanisms of desorption from ink-bottle
pore systems. Indeed, it now appears that the adsorption/desorption scanning behaviour of H1 and
H2 loops is quite different.
The application of NLDFT has also confirmed that if the cavities are sufficiently small (3 nm
< pore width < 6 nm), pore filling by nitrogen (at 77 K) can occur reversibly at equilibrium pres-
sures determined by the cavity diameters. Such reversible condensation/evaporation is not
observed in larger cavities. Instead, capillary condensation occurs spontaneously in the region of
the predicted spinodal point on the metastable adsorption branch. Desorption from these larger
cavities is determined by the window size unless this is smaller than the critical size (i.e. ca. 4 nm
for N2 adsorption at 77 K) corresponding to the limit of stability, in which case a cavitation process
of the liquid-like condensate leads to spontaneous desorption. It is evident that these findings are
broadly in agreement with earlier work (Rouquerol et al. 1999).
There is, however, one new aspect. Careful experimental work has indicated that the hysteresis
lower closure point for MCM-41 samples is ∼ 0.055P0 lower than the corresponding closure points
for SBA-16 and FDU-1. This appears to be in accord with the NLDFT predictions (Ravikovitch
and Neimark 2002a,b) and the results of Monte Carlo simulation (Vishnyakov and Neimark 2003).
We must therefore conclude that spontaneous evaporation/cavitation processes are to some extent
dependent on pore geometry. More work is required, however, to establish the exact relationship
between the limiting P/P0 value of the lower closure point of hysteresis and the pore shape.

Hysteresis loops of types H3 and H4

The type H3 hysteresis loops depicted in Figure 1 are typically given by the adsorption of non-
polar gases by montmorillonite clays [see Barrer (1978, 1989)] and the aggregates of other platy
particles. It is evident that these H3 loops have characteristic desorption shoulders and lower
closure points, which in the case of N2 adsorption at 77 K are located in the region of 0.42P0. In
this one respect, H3, H2 and H4 loops are all similar. However, the H3 loop does not have a
plateau at high P/P0 values and therefore this form of isotherm should not be referred to as type
IV. Since there is no well-defined mesopore volume, caution should be exercised in the inter-
pretation of the uptake at high P/P0 (Rouquerol et al. 1999). As indicated in Figure 1, although
only the initial monolayer–multilayer section of the isotherm is reversible, the whole adsorption
branch of the H3 loop appears to exhibit the same shape as a type II isotherm (Sing et al. 1985).
In principle, any deviation from a true type II isotherm can be detected by the construction of a
comparison plot or αS-plot (Rouquerol et al. 1999). This pseudo-type II character is associated
with the metastability of the adsorbed multilayer (and delayed capillary condensation) and is
due to the low degree of pore curvature and non-rigidity of the aggregate structure. The shape of
the H3 loop is also linked with the non-rigid nature of the adsorbent and the location of the
characteristic shoulder is consistent with the destabilization of condensate at the limiting
P/P0 value.
The isotherms for the sorption of polar molecules (e.g. water, lower alcohols, pyridine) by
montmorillonite are of a quite different character. The hysteresis now extends over the complete
AST 22(10)_64 31/01/05 1:31 pm Page 780

780 K.S.W. Sing and R.T. Williams/Adsorption Science & Technology Vol. 22 No. 10 2004

range of P/P0 values and is the result of irreversible interlayer penetration (or intercalation) of the
polar molecules and the expansion and contraction of the clay particles (Barrer 1989).
Type H4 loops are given by many activated carbons and some other nanoporous adsorbents.
These isotherms are of a composite nature. In the least complicated case (as illustrated in
Figure 1), the initial region of reversible micropore filling is followed by multilayer physisorption
and capillary condensation. By applying an empirical method of analysis (e.g. the αS-plot), one
can often separate the isotherm into its constituent parts (Gregg and Sing 1982). It then becomes
clear that the H4 loop is equivalent to an H3 loop on the relatively small non-microporous part of
the adsorbent.

CONCLUSIONS AND RECOMMENDATIONS

Recent molecular simulation and density functional studies of the physisorption of gases by model
pore structures have greatly improved our understanding of adsorption hysteresis, but there
appears to be no immediate need to change the classifications of loops and isotherms that were
proposed in 1984 by the IUPAC [see Sing et al. (1985)]. Indeed, we consider the identification of
the type of isotherm and hysteresis loop to be a useful starting point in the characterization of a
nanoporous adsorbent. By taking account of the recent work, we are now in a position to offer
the following checklist as a revised version of that originally proposed by the IUPAC for the inter-
pretation of nitrogen isotherms at 77 K.
1. What is the isotherm type? If the isotherm is type IV or pseudo-type II, does it exhibit a
reproducible hysteresis loop? Is the isotherm reversible in the pre-loop region?
2. Construct an αS-plot or comparison plot, making use of standard isotherm data determined
on an appropriate non-porous solid. There is no significant microporosity if back-
extrapolation of the αS-plot gives a zero intercept,
3. If the isotherm is type IV and there is no detectable microporosity, the mesopore volume
can be evaluated from the plateau at high P/P0 values. If microporosity is detected, the
uptake at the plateau corresponds to the total capacity of micropores plus mesopores.
4. If the hysteresis loop is type H1, the desorption branch should be used in preference to
the adsorption branch to compute the effective, or apparent, mesopore-size distribution.
A computational procedure based on density functional theory is recommended for pore-
size analysis rather than the BJH–Kelvin method, but it must be kept in mind that the
validity of the derived pore-size distribution is dependent on the conformity of pore shape
and adsorbent structure to the chosen model system.
5. If the hysteresis loop is type H2, the desorption branch is likely to be dependent on network
and/or ‘ink bottle’ effects. It may be possible to arrive at a semi-quantitative evaluation of
the pore-size distribution from the adsorption branch, provided that allowance is made for
the metastability of the adsorbed multilayer.
6. If the hysteresis loop is type H3, the mesopore size distribution cannot be assessed reliably
from either branch of the loop. However, it may be possible to obtain useful information by
the application of an empirical method of isotherm analysis.
7. If the hysteresis loop is type H4, the effective micropore capacity may be obtained from an
appropriate αS-plot or comparison plot. If the pore shape and solid structure are known, the
density functional approach is appropriate for nanopore size analysis.
AST 22(10)_64 31/01/05 1:31 pm Page 781

Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials 781

8. The use of nitrogen adsorption alone cannot be expected to provide more than a semi-
quantitative evaluation of the distribution of narrow nanopores. A range of probe molecules
of different size is required to assess the available range of pore size.

REFERENCES

Barrer, R.M. (1978) Zeolites and Clay Minerals, Academic Press, London.
Barrer, R.M. (1989) Pure Appl. Chem. 61, 1903.
Barrer, R.M., McKenzie, N. and Reay, J.S.S. (1956) J. Colloid Sci. 11, 479.
Barrett, E.P., Joyner, L.G. and Halenda, P.P. (1951) J. Am. Chem. Soc. 73, 373.
Branton, P.J., Hall, P.G. and Sing, K.S.W. (1993) J. Chem. Soc., Chem. Commun. 1257.
Branton, P.J., Hall. P.G., Treguer, M. and Sing, K.S.W. (1995) J. Chem. Soc., Faraday Trans. 91, 2041.
Brunauer, S. (1944) The Adsorption of Gases and Vapours, Oxford University Press, Oxford.
Burgess, C.G.V., Everett, D.H. and Nuttall, S. (1989) Pure Appl. Chem. 61, 1845.
Coelingh, M.B. (1939) Kolloid-Z. 87, 251.
Cohan, L.H. (1938) J. Am. Chem. Soc. 60, 433.
de Boer, J.H. (1958) The Structure and Properties of Porous Materials, Everett, D.H., Stone, F.S., Eds,
Butterworths, London, p. 68.
Derjaguin, B.V. (1957) Proc. 2nd Int. Cong. Surface Activity II, Butterworths, London, p. 154.
Dubinin, M.M. (1960) Chem. Rev. 60, 235.
Dubinin, M.M. (1966) Chemistry and Physics of Carbon, Walker, P.L., Ed, Marcel Dekker, New York, p. 51.
Evans, R., Marconi, U.M.B. and Tarazona, P. (1986) J. Chem. Phys. 84, 2376.
Everett, D.H. (1955) Trans. Faraday Soc. 51, 1551.
Everett, D.H. (1958) The Structure and Properties of Porous Materials, Everett, D.H., Stone, F.S., Eds,
Butterworths, London, p. 95.
Everett, D.H. (1967) The Solid–Gas Interface, Flood, E.S., Ed, Edward Arnold, London, p. 1055.
Everett, D.H. (1979) Characterization of Porous Solids, Gregg, S.J., Sing, K.S.W., Stoeckli, H.F., Eds, S.C.I.,
London, p. 229.
Everett, D.H. and Haynes, J.M. (1972) J. Colloid Interface Sci. 38, 125.
Foster, A.G. (1932) Trans. Faraday Soc. 28, 645.
Gelb, L.D. and Gubbins, K.E. (1998) Langmuir 14, 2097.
Gregg, S.J. and Sing, K.S.W. (1982) Adsorption, Surface Area and Porosity, 2nd Edn, Academic Press,
London.
Harris, M.R. (1965) Chem. Ind. (London) 269.
Haynes, J.M. (1975) Colloid Science, Vol. 2, Chemical Society, London, p. 101.
Jorge, M. and Seaton, N.A. (2002) Mol. Phys. 100, 3803.
Kornev, K.G., Shingareva, I.K. and Neimark, A.V. (2002) Adv. Colloid Interface Sci. 96, 143.
Kraemer, E.O. (1931) A Treatise of Physical Chemistry, Taylor, H.S., Ed, Macmillan, New York, p. 1661.
Kresge, K.D., Leonowicz, M.E., Roth, W.J., Vartuli, J.C. and Beck, J.S. (1992) Nature (London) 359, 710.
Lastoskie, C.M. and Gubbins, K.E. (2000) Characterization of Porous Solids V, Unger, K.K., Kreysa, G.,
Baselt, J.P., Eds, Elsevier, Amsterdam, p. 41.
Levitz, P.E. (2002) Handbook of Porous Solids, Schuth, F., Sing, K.S.W., Weitkamp, J., Eds, Wiley–VCH,
Weinheim, p. 37.
Liu, H., Zhang, L. and Seaton, N.A. (1992) Chem. Eng. Sci. 47, 4393.
Mason G. (1988) Characterization of Porous Solids I, Unger, K.K., Rouquerol, J., Sing, K.S.W., Kral, H.,
Eds, Elsevier, Amsterdam, p. 323.
McBain, J.W. (1935) J. Am. Chem. Soc. 57, 699.
Neimark, A.V. and Vishnyakov, A. (2000) Phys. Rev. E 62, 4611.
AST 22(10)_64 31/01/05 1:31 pm Page 782

782 K.S.W. Sing and R.T. Williams/Adsorption Science & Technology Vol. 22 No. 10 2004

Neimark, A.V., Ravikovitch, P.I. and Vishnyakov, A. (2000) Phys. Rev. E 62, 1493.
Neimark, A.V., Ravikovitch, P.I. and Vishnyakov, A. (2003) J. Phys. Cond. Mat. 15, 347.
Pikunic, J., Lastoskie, C.M. and Gubbins, K.E. (2002) Handbook of Porous Solids, Schuth, F., Sing, K.S.W.,
Weitkamp, J., Eds, Wiley–VCH, Weinheim, p. 182.
Rao, K.S. (1941) J. Phys. Chem. 45, 513.
Ravikovitch, P.I. and Neimark, A.V. (2002a) Langmuir 18, 1550.
Ravikovitch, P.I. and Neimark, A.V. (2002b) Langmuir 18, 9830.
Rouquerol, F., Rouquerol, J. and Sing, K. (1999) Adsorption by Powders and Porous Solids, Academic Press,
London.
Rouquerol, F., Rouquerol, J. and Sing, K.S.W. (2002) Handbook of Porous Solids, Schuth, F., Sing, K.S.W.,
Weitkamp, J., Eds, Wiley–VCH, Weinheim, p. 236.
Saam, W.F. and Cole, M.W. (1975) Phys. Rev. B 11, 1086.
Sakamoto, Y.H., Kaneda, M., Terasaki, O., Zhao, D.Y., Kim, J.M., Stucky, G., Shim, H.J. and Ryco, R.
(2000) Nature (London) 408, 449.
Seaton, N.A. (1991) Chem. Eng. Sci. 46, 1895.
Seaton, N.A., Walton, J.P.R.B. and Quirke, N. (1989) Carbon 27, 853.
Sing, K.S.W., Everett, D.H., Haul, R.A.W., Moscou, L., Pierotti, R.A., Rouquerol, J. and Siemieniewska, T.
(1985) Pure Appl. Chem. 57, 603.
van Bemmelen, J.M. (1896) Z. Anorg. Chem. 13, 239.
Vishnyakov, A. and Neimark, A.V. (2001) J. Phys. Chem. B 105, 7009.
Vishnyakov, A. and Neimark, A.V. (2003) Langmuir 19, 3240.
Yu, C., Yu, Y. and Zhao, D. (2000) J. Chem. Soc., Chem. Commun. 575.
Zsigmondy, A. (1911) Z. Anorg. Chem. 71, 356.

Das könnte Ihnen auch gefallen