You are on page 1of 8

Fundamentals of Adsorption

Proc. IVth Int. Conf. on Fundamentals of Adsorption, Kyoto, May 17-22, 1992
Copyright 0 1993 International Adsorption Society

Theoretical Interpretation and Classification of Adsorption


Isotherms for Simple Fluids

Perla B. Balbuaal, Christian Lastoskie', Keith E. Gubbins' and Nicholas QuirkeZ


1) School of Chemical Engineering, Cornell University, Ithaca, NY 14853, USA
2) BP Research Centre, Middlesex TW16 7LN, UK

ABSTRACT
We apply nonlocal density functional theory to the interpretation and classification of adsorp-
tion isotherms in terms of the underlying molecular properties and pore geometry. We present
results for single walls and slits (parallel walls) varying pore width, temperature, and the pa-
rameter ratios ,,€ €,/, (welldepth) and u,,/c,, (molecular diameters) for the solid-fluid and
fluid-fluid potentials. In a second study we present an improved method for the determination
of pore sise distribution. The theoretical adsorption isotherms for single pores are correlated
an a function of the pressure and pore width. The pore-size distribution is then found by
comparing this correlation to the experimental adsorption data for nitrogen in porous carbons.
INTRODUCTION
The &st classification of adsorption isotherms for pure fluids was presented by Brunauer et
al.[l]. The scheme was based on van der Waals theory of physical adsorption. The IUPAC
Commission on Colloid and Surface Chemistry [2] extended Brunauer's classification by adding

w WEAK

CAPlLURY CON 4
DEWSIllON

Relatlve pressure, P/Po

Figure 1. IUPAC classification of adsorption isotherms.


27
28 P. B. Balbuena, C. Lastoskie, K. E. Gubbins and N. Quirke

class VI, the stepped isotherm, as seen in Figure 1. Langmuir [3]had previously described
Class I, which is known as the Langmuir isotherm , and class I1 corresponding to multilayer
adsorption. Brunauer et. al. [4],interpreted these two classes with the BET equation, and
extended this equation to cover the first five types of isotherms of Figure 1. Both the Langmuir
and the BET equations are extensively used together with the Frenkel-Helsey-Hill (FHH) [5-71
equation to correlate and interpret experimental adsorption data. A comparison between the
BET, the FHH and the Langmuir (case VI) equations was made recently by Everett et. al.
[8,9], and their validity was checked for some conditions. Although it is possible to extract
some information about the physical significance of the fitted parameters, in most cases the
conclusions are valid only within certain ranges.
In this paper we use a molecular based theory that allows us to relate the different classes of
isotherms to the fluid-fluid and fluid-wall interaction energy parameters, the pore size and the
temperature. Our calculations are based on the single pore model, and neglect interconnection
of the pores. As discussed by several authors [lo-121,such interconnections are believed to be
important in determining the shapes of hysteresis loops, and we plan to indude these effects in
future work.
Nitrogen adsorption measurements have been routinely used for determining the pore size
distribution (PSD) of porous carbons. To calculate the PSD, it is necessary to assume a model
for pore filling which relates the pore width to the condensation pressure. A commonly employed
model from classical thermodynamics is the Kelvin equation:

where p is the pressure at which condensation occurs in a slit of width H, po is the saturation
pressure, R is the ideal gas constant, and pl and 7 are the liquid density and surface tension
of nitrogen at absolute temperature T. The Kelvin equation is accurate in the limit of large
pores and low temperatures, but fails when applied to narrow pores, as it does not take into
account the wetting of the porous surface prior to condensation. For micropores, adsorption is
enhanced due to increased fluid-fluid attractive forces and stronger solid-fluid attractive forces
from the overlap of the wall potentials. Hence, PSD analysis based upon the Kelvin equation
underestimates the sizes of the micropores present in the material.
In this work, we adopt a nonlocal mean field density functional theory model of adsorption. A
local form of this theory was originally proposed by Seaton, Walton and Quirke [13].To solve
for the PSD, we express the experimental isotherm N(p) rn

where &in and are the widths of the smallest and largest pores present in the sorbent;
€Irna=
p(p, H)is the density of nitrogen in a slit of width H at pressure p ; and f(H) is the PSD of the
porous material. Mean field theory computations provide the individual pore isotherms p ( p , H),
which are then numerically fitted to the experimental adsorption measurements according to
equation (2) to yield the pore size distribution f ( H ) .
In our analysis we use the nonlocal version of mean field theory. In this respect our analysis
differs from that of Seaton et al., which employed the 1-s accurate local mean field theory.
While both versions correctly predict precondensate film growth and a transition to continuous
filling below a critical slit width, only the nonlocal theory reproduces the oscillations in the
fluid density profile near the solid-fluid interface as observed in simulations [14].
DESCRIPTION OF THE MODEL
We consider an idealized material in which the pores are all identical and non-connected and
are slits of width H. The pores are open and immersed in a very large reservoir containing
a single-component fluid at fixed chemical potential p and temperature T, the total volume
Interpretation and Classification of Isotherms 29

of the system being V. In such conditions, the fluid inside the pore feels the presence. of the
solid surfaces ae the action of an external potential and reaches an equilibrium situation where
its chemical potential equals the value of the bulk chemical potential. The grand canonical
ensemble provides an appropriate description of the thermodynamics. The theory we have
chosen describes the thermodynamic grand potential an a functional of the one-particle density
distribution. This functional takes its minimum value at equilibrium; this value is the grand
potential energy , and the density profile that gives this minimum value is the equilibrium
density profile. When more than one minimum exists, the one with the lower free energy is
the stable one. A phase transition occurs when two minima have the same value for the free
energy. Several density functional models have been used to study this problem [15]. Here
we use TarBllona’s non-local mean field theory [16,17]. The grand potential energy functional
n ( p ) is the sum of the intrinsic Helmholte free energy functional F ( p ) and two other terms
corresponding to the contributions of the bulk chemical potential p and the external potential
L(
r),

where p(r) is the fluid number density at point r. The Helmholtz free energy is expanded
about a reference system of hard-spheres of diameter d, and the perturbation term involves the
attractive potential &(lr - r’l). The attractive part is treated in mean field approximation,
thereby neglecting correlations due to attractive forces ,

Taraaona’s model expresses the hard sphere free energy as the sum of an ideal gas and an excess
part. The ideal term is a functional of the local density p(r), while the excess part is considered
a functional of a smoothed density, p(r) which is defined as: p(r)= J dr’ w(lr - r’l ;p (r))
p(r’), where the w((r - r‘l) are weighting functions chosen to give a good description of the
hard sphere direct pair correlation function for the uniform fluid over a wide range of densities.
This model has been shown to give very good agreement with simulation results for the density
profile and surface. tension of hard sphere (HS) and Lennard-Jones (LJ) potentials near hard
walls and LJ walls. In this paper, the walls are taken to be structureless. The inputs to the
model are: the intermolecular potentials, and an equation of state for the excess Helmholtz
free energy for the hard sphere fluid. The equivalent hard sphere diameter, d, is calculated as
a function of temperature, as suggested by Lu et. al [18]. The explicit form is the one that
approximates the Barker- Henderson diameter [19] , d/ v f f = (a1 T’ 02)/ +
( a3 T’ +a&
where the constants, a,,were chosen to give good agreement between theory and simulation
at low temperatures [18]. Equation (3) is solved numerically by minimizing n ( p ) to obtain
tho density profile, given the conditions of bulk density , temperature,and separation between
walls. A simple iteration scheme is used. The Carnahan-Starling expression is used for the
hard sphere excess free energy. The attractive part of the potential is represented by the WCA
form of the cut and shifted LJ, the cutoff value being 2.5. The external potential is obtained by
integrating the LJ potential between one fluid molecule and each of the molecules of the solid
over the lateral solid structure [20,21]. A sum is then performed over the planes of molecules
in the solid, the Separation between planes being A. This yields the ”10-4-3” potential,
2
hf(z)/kT = A[j(v,f/z)10 - ( # . f / ~ )-~(~,f)~/3A(O.61Az ) ~ ] + (5)

where A= 2 r p, (c,f/kT) (A)and p, is the solid density. The cross-parameters are


calculated according to the Lorentz-Berthelot rules, E , ~ = ( E , , / E f f ) l l 2 ; u,f= (u,,+uff)/2. All
parameters are made dimensionless with the fluid-fluid LJ parameters. The external potential
30 P. B. Balbuena, C. Lastoskie, K. E. Gubbins and N. Quirke

involves several inputs. Two of them are characteristic of the surface itself: the solid density, and
the separation between layers. In all our calculations, we have used the values corresponding to
a graphite surface, p,=114 r ~ m - ~and
, A= 0.335 nm. The other two variables are the relative
strength of the solid- fluid to fluid-fluid interactions e . f / ~ f f and
, the relative molecular size
parameter of the solid-fluid and fluid-fluid potentials, u.j/u,f. The external potential for the
slit geometry is : Vczt(z,H)= q&,(z)+$,f(H-z), and the total adsorption per unit area r: is
calculated according to : = (1/2) J,"' p* (z*) dz'. The adsorption behavior depends on the
independent reduced variables T*=keT/Eff,ti'=H/ ujf (for pores), e . f / e f f and u s f l u f f .
RESULTS
We fix u . f / u f f= 0.9462, the value for Lennard-Jones methane on graphite. We have chosen to
vary the ratio ~ , f / e f as
f a way of changing the value of A in equation (8), but other variables
could be variedinstead, e.g. the solid density, the A parameter or the ratio of u.j/ujj. This last
ratio gives also the range of the potential, and determines the order of the wetting transitions;
we are presently investigating its effect on solvation forces [22].
Nonporous solids: single surfaces
Several previous etudies [23-251 have clearly established the relation between wetting and ad-
sorption. In this work, we have calculated the wetting conditions at fixed temperature. The
solid-gas and solid-liquid interfacial tensions rag and 7.l were determined as a function of e , f / e f f .
Young's equation relates these two interfacial tensions to the liquid-gas interfacial tension, 7fg,
and the angle B defined by the liquid and gas interfaces: 7ag=7.1t71g cosB. For the liquid-gas
interfacial tension values we have interpolated the values calculated by Lu et. al. [l8] for the
.
same fluid The wetting conditions are determined by the intersection of the two functions, rag
and 7,1+7lg, corresponding to cos 0=1. The wetting temperatures for a range of values of ~ . f / ~ f f
are shown in Figure 2-a, together with the regions where the different classes of isotherms are
found. The solid line in Figure 2-a shows the wetting temperatures and dashed lines (which
are approximate) divide the classes. For conditions of partial wetting,the adsorption reaches
a finite value at the saturation pressure; we refer to these classes as I I f , IZIj and V I f . At
temperatures above the wetting temperature , adsorption diverges at po. Accordingly, classes
I1 or VI are found for sufficiently strong substrates. For weak walls, the transition is from class
IZZf to class 111.
Porous solids: confined fluids
In this section we present results of adsorption of fluids in a slit-type geometry. Hence, there is
a new variable ,the slit width H , expressed in reduced units as ti'=H/ajf .The reduced bulk
critical temperature for this fluid is estimated by Powles [26] to be T,'= 1.119. We report our
results at three temperatures, one at a low temperature (T*=0.5) ,an intermediate (T*=0.8),
and a supercritical temperature (T*=1.4). In this way we can identify regions where each class
of isotherm can be found . In previous work [27] , we have extensively discussed the phase
transitions observed in pores in relation to the wetting transitions observed for a single planar
wall for the same intermolecular parameters. In Figure 2-b we show H*vs e . f / ~ f f, at T*=0.5.
At this temperature, adsorption is very small for sufficiently low values of the ratio e . f / ~ f .j
Capillary evaporation takes place at these conditions, i.e. the gas-liquid transition occurs at
P/Po > 1. This transition pressure decreases with the pore size, until a limit is reached where
the pore is so small that no molecule can fit into it. Experimental isotherms of this type have
been found for Kr on Na and Nu20 [28] ,and a20 on graphite [29]. For micropores, when the
relative wall strength is increased, class I , the Langmuir isotherm is obtained. In cases where
there is a phase transition to a liquid-like phase, this takes place at extremely low pressures.
There are many examples of experimental class I isotherms. They include Nz in zeolites and

-
aromatic hydrocarbons such as benzene on charcoal (301. For larger pore sizes, ti' > 5, and wall
strength stronger than ~ . f / e , f 0.2, layering transitions are observed in addition to capillary
condensation.
Interpretation and Classification of Isotherms 3I

At T*=0.8 (Fig. 2-c) capillary evaporation occurs only for very small values of c,f/cff. I n -
creasing this ratio, class V behavior is observed. The amount adsorbed per unit area is still

1.a 16

12
1.1

1'
0.0

0.7
3

0.6 0 I I
0 0.2 0.4 0.6 0 0.2 0.4 0.6
%I'ell % IEI f

16 1 1
16

t T *' = 0.8
12 - p v VIf
12

- B
6 - H o g

H'
e
6 - IV

3 - 3

0
0 0.1 0.2 O J 0.4 0.1 0.6 0 0.2 0.4 0.6
%IlE ff
I1

Figure 2. Isotherm classification:(a) Single walls. (b),(c),(d) Slit pores.

very small for low and moderate relative pressures, but at a pressure somewhat below Po the
fluid condenses inside the pore. The transition pressure decreases with the pore size, at fixed
c,!/cff. Examples of this class are n-hexane and n-octane on polytetrafluorethylene [31], and
water on polymethylmethacrylate [32]. For stronger walls, class I is found for micropores and
class IV for mmo and macropores. Class IV is found experimentally for a great many systems,
e.g. Ar and NZon porous glass [33].
At supercritical temperatures Po is undefined and the IUPAC classification of Figure 1 is not
applicable. Nevertheless, it is convenient to relate the observed isotherm behavior to the IUPAC
classification scheme. We therefore adopt the notation of Fig. 1, but with the superscript sc
to remind ourselves that we are referring to adsorption at supercritical temperatures. Figure
.
2-d shows the types of isotherm obtained for T*=1.4 At these higher temperatures there is
no condensation, and classes V and IV shift to 1117 and 117, the pore modification of classes
I11 and I1 of Fig. 1, reupectively.
Pore siae distribution analysis of porous carbons
To interpret experimental adsorption isotherms of nitrogen on porous carbon using mean field
32 P. 8. Balbuena, C. Lastoskie, K. E. Gubbins and N. Quirke
M

I 0

Figure 3. Capillary coexistence curve. Figure 4. Nonlocal theory isotherms


Solid line, Kelvin equation; dashed line, for nitrogen on carbon at 77 K. The
modified Kelvin equation; dotted line, ultramicropore isotherms are truncated
nonlocal theory; dash-dotted line, local to aid visualization.
theory. The symbols denote regimes of
continuous filling (triangles), capillary
condensation (open circles), and layering
transitions (filled circles). The inset
shows the ultramicropore local theory
results.

-
n

k
..
I

10 20 30 40 *-
- 0t
BP71 8,
s
v

'. ,
........
.... '. ......,,, 0

Figure 5 . Carbon isotherms, nitrogen Figure 6. Carbon PSDs. Solid lines,


adsorption at 77 K. Solid lines,nonlocal nonlocal theory; dashed lines, local
theory; dashed lines, local theory; theory. The dotted lines are the
symbols, experiment. extrapolation of the PSD beyond limits
imposed by the experimental sampling
(see text).
Interpretation and Classification of Isotherms 33

theory, we must select appropriate potential parameters. The fluid-fluid parameters are chosen
so that the saturation pressure and saturated liquid density of bulk nitrogen predicted by mean
field theory (using the full Lennard-Jones potential) are equal to the experimental values at the
nitrogen boiling point of 77.347 K. To assign the solid-fluid parameters, we fit the theoretical
adsorption isotherm on a nonporous surface (single wall) to match experimental nitrogen uptake
measurements on nonporous carbon. The t-curve of de Boer et al. 1341 provides our reference.
The parameter values obtained are crjf = 3.572 A; e f , / k = 93.98 K;craf = 3.494 A; and e , j / k
= 41.87 K (nonlocal theory) and 74.23 K (local theory).
Theoretical isotherms for nitrogen adsorption in carbon slit pores at 77 K were generated for
pore widths ranging from 6 A to 360 A. In Figure 3 we compare the capillary coexistence curves
for local and nonlocal theory. For reference, we also display the filling pressures predicted by
the Kelvin equation, and by a modified Kelvin equation where 61 in equation (1) is replaced
by R - 2t, and t is the film thickness at pressure P c / P o . The Kelvin-based models severely
overestimate the filling pressures of the micropores. Local theory offers an improved description
of pore filling compared to the classical thermodynamics methods, but it greatly underestimates
the filling pressures of the ultramicropores. Nonlocal theory, by contrast, provides isotherms
in quantitative agreement with simulation results [35]. The minimum in the filling pressure for
pore widths of N 7 A arises from the overlap of the attractive wall potentials, which enhances
adsorption. For pores narrower than 7 A, repulsive solid-fluid interactions dominate and the
filling pressure rises sharply. Several adsorption isotherms from nonlocal theory are shown in
Figure 4. It is interesting to note that nonlocal theory predicts layering transitions in the
filling of the supermicropores, and that the nonlocal isotherms exhibit steplike growth of the
multilayer film. These features are absent fiom the local theory description.
Using the nonlocal theory isotherms, we obtain the PSDs of porous carbons by numerically
fitting the integral of equation (2) to experimental adsorption data. To carry out the analysis
we must assume a functional form for the PSD. Although this choice of function for f ( H ) is
arbitrary, in practice it is found that a sufficiently large sample of uptake data (about twenty
points) constrains the shape of f ( H ) such that the numerical values of the PSD will be effectively
unique, regardless of the functional form chosen for the PSD. In this work, we select the trimodal
gamma distribution to model the PSD:

where a,,pi,^, are adjustable parameters which determine the size and shape of the distribution.
Figure 5 shows the local and nonlocal theory fits to the adsorption isotherms of two carbons,
BP71 and AC610. The local and nonlocal PSDs corresponding to these fits are shown in Figure
6, for BP71 and AC6lO respectively. To verify the uniqueness of the numerical values of the
PSDs, the fitting process waa repeated using the trimodal lognormal distribution; the results
were essentially indistinguishable from those presented in Figure 6.
As one might deduce from the shapes of the isotherms, the PSD of BP71 is largely mesoporous,
while AC610 is highly microporous. Each PSD is truncated at lower and upper bounds given
by the range of pressures sampled in the experimental isotherm, although the full range of pore
siees is employed in the fitting procedure. The two theories predict markedly different PSDs;
this is expected, as the nonlocal theory is a better model of adsorption in narrow pores, and
thus predicts more accurately the isotherm and PSD of the microporous carbon AC610.
CONCLUSIONS
The nonlocal density functional theory used here is believed to be quite accurate for slit pores
over the full range of conditions studied, so that the results obtained should provide a framework
for the behavior of simple fluids in single pores. While the present model can account for all
six of the IUPAC classes, we expect interconnections among pores to be a crucial factor in
34 P.B. Balbuena, C. Lastoskie, K. E. Gubbins and N. Quirke

accounting for the subclasses of behavior observed in the hysteresis loops of classes IV and V.
The pore s i x distribution analysis method presented in this work is based upon an improved
molecular description of adsorption, and thus it offers a more quantitatively accurate approach
to determining the PSDs of carbons than previous methods have afforded. In our continuing
studies, we shall apply the model to other fluid-sorbent systems, including silicas and aluminas.
ACKNOWLEDGMENTS
It is a pleasure to thank K.S.W. Sing for a helpful discussion. We thank the National Science
Foundation (grant no. CTS-8914907)and the Gas Research Institute for support of this work.
We thank S.M. Riddiford and J. Aukett for the nitrogen adsorption data for the porous carbons.
REFERENCES
[l] S. Brunauer, P.H. Emmett and E. Teller, J. Amer. Chem. SOC.,60(1938)309
[2]K.S.W. Sing et.al., IUPAC Commission on Colloid and Surface Chemistry including Catal-
ysis, Pure and Applied Chemistry, 57(1985)603
[3]I. Langmuir, J. Amer. Chem. SOC.,40(1918)1361
[4]S. Brunauer, L.S. Deming, W.E. Deming and E. Teller, J. Amer. Chem.Soc., 62(1940)1723
[5]J. Frenkel, in Kinetic Theory of Liquids, Oxford University Press, London, 1946.
[6]G. Halaey, J. Chem. Phys., 16(1948)931
[7]T.L. Hill, J. Chem. Phys., 17(1949)668
[8]D.H. Everett, Langmuir, 6(1990)1729
[9]C.G.V. Burgess, D.H. Everett and S. Nutall, Langmuir, 6(1990)1734
[lo]G.Mason, Proc. R. SOC.London A, 390(1983)47
[ll] M. Parlar and Y.C. Yortsos, J. of Colloid and Int. Science, 132(1989)425
[12]N. Seaton , Chemical Engineering Science, 46(1991)1895
[13]N.A. Seaton, J.P.R.B. Walton and N. Quirke, Carbon 27(1989)853
[14]Z. Tan and K.E. Gubbins, J. Phys. Chem., 94(1990)6061
[15]R. Evans in Inhomogeneous Fluids (Ed. D. Henderson), to be published by Dekker ,1992,
chapter 5
[MIP. Tarazona, Phys. Rev. A, 31(1985)2672
[17]P. Taramna, Phys. Rev. A ,32(1985)3148
[18]B.Q. Lu, R. Evans and M.M Telo da Gama, Mol. Phys., 55(1985)1319
[19]J.A. Barker and D. Henderson, J. Chem. Phys., 47(1967)4714
[20]W.A. Steele, Surf. Science, 36(1973)317
[21]Steele, W.A. in The interaction of gases with solid surfaces, Pergamon, Oxford, 1974,p.13
[22]P.B. Balbuena, D. Berry and K.E. Gubbins, (1992),to be published.
[23]D.E. Sullivan,J. Chem. Phys.,74(1981)2604
[24]G.F. Teleteke, L.E. Scriven and H.T. Davis, J. Chem. Phys., 78(1983)1431
[25]P. Tarazona and R. Evans, Mol. Phys., 48(1983)799
[26]J.G. Powles, Physica 126A(1984)289
[27]P.B. Balbuena and K.E. Gubbine, Fluid Phase Equilibria, (1992), in press.
[28]R.A. Pierotti and G.D. Halsey ,J. Phys. Chem., 63(1959) 680
[29]N.N. Avgul, G.I. Berezin , A.V. Kisilev and LA. Lygina, Iw. Akad.Nauk. SSSR Otd.
Khim. Nauk., 2(1961)205
[30]D.A. Cadenhead and D.H Everett, Conf. on Ind. Carbon and Graphite, Society of Chem.
Ind., 1958,p.272
[31]J.W. Whalen, J. Colloid Int. Sci., 28(1968)44t
[32]H.H.G. Jellinek, M.D. Luh and V. Nagarajan, V. Koll-Z.U. Polymere, 232(1969)758
1331 P.H. Emmett and M. Cines, J. Phys. Chem., 56(1947) 735
[34]J.H. de Boer, B.G. Linsen and T.3. Osinga, J. Catalysis , 4(1965)643
[35]B.K. Peterson, J.P.R.B. Walton and K.E. Gubbins, J.Chem.Soc.Far.Trans.2,82(1986)1789

You might also like