Sie sind auf Seite 1von 14

Langmuir 1993,9, 1801-1814

1801

Theoretical Interpretation of Adsorption Behavior of


Simple Fluids in Slit Pores
Perla B. Balbuenat and Keith E. Gubbins'
School of Chemical Engineering, Cornell University, Ithaca, New York 14853
Received December 30,1992. In Final Form: March 29, 1993
Nonlocal density functional theory is used to interpret and classify the adsorption behavior of simple
fluids in model materials having slit pores. A systematicstudy is reported for a wide range of the variables
involved temperature, pressure, pore width H, and the intermolecularparameter ratios td/tfi and u d / u ~ .
Adsorption isotherms, isosteric heats of adsorption, and phase diagrams are calculated. The isotherms
are related to those of the 1985 IUPAC classification;the range of variables corresponding to each of the
six isotherm types is determined, and the underlying factors leading to each of the types are elucidated.
In addition to the six types of the 1985classification,a seventhtype is identified,corresponding to capillary
evaporation. A similarstudy and classificationis reported for the heata of adsorptionand phase transitions
(capillary condensation and layering transitions) in pores. Since the materials studied do not exhibit
either heterogeneity or networking, the conditions leading to phase transitions are clearly seen. Where
possible,qualitative comparisons with experimentalobservations are made. The theoretical classification
reported here should provide a useful framework against which to interpret experimental data.
1. Introduction

The first classification of physical adsorption isotherms


for pure fluids was presented by Brunauer et a1.12 They
proposed five isotherm types, based on known experimental behavior. In 1985, the IUPAC Commission on
Colloid and Surface Chemistry3 proposed a modification
of this classification; in addition to the original five types
of Brunauer et al. they added a sixth type, the stepped
isotherm. These six types are shown schematically in
Figure 1. Type I (the Langmuir isotherm) is typical of
many microporous adsorbents (pore widths below 2 nm);
at relative pressures approaching unity the curve may reach
a limiting value or rise if large pores are present. Types
I1 and I11 are typical of nonporous materials with strong
(type 11) or weak (type 111) fluid-wall attractive forces.
Types IV and V occur for strong and weak fluid-wall forces,
respectively,when the material is mesoporous (porewidths
from 2 to 50 nm) and capillary condensation occurs; these
types exhibit hysteresis loops. Type VI occurs for some
materials with relatively strong fluid-wall forces, usually
when the temperature is near the melting point for the
adsorbed gas.
The interpretation of experimental adsorptionisotherms
is complicated in practice by uncertainties concerning the
morphology of the adsorbing material. Materials studied
are frequently heterogeneous,having not only an unknown
range of pore sizes but a range of pore shapes, active
adsorption sites, and blocked and networked pores. For
suchmaterialsthe measured isotherm is a weighted average
over the adsorption (and any phase transitions that occur)
due to these various effects. The interpretation has been
further clouded by the use of methods based on the Kelvin
equation,which is known to give large errors for micropores
and the smaller mesopores. The latter difficulty can be
largely overcome by the use of modern statistical mechanical theories, particularly density functional theory,
+ Present address: Department of Chemical Engineering, University of Texas, Austin, TX 78712-1062. On leave from INTEC,
Univereidad Nacional del Litoral, Santa Fe,Argentina.
(1) Brunauer,S.;Deming,L.S.;Deming,W. E.;Teller,E. J.Am. Chem.

SOC.1940,62, 1723.
(2)Brunauer, S. The Adsorption of Guses und Vupours; Oxford
University Press: London, 1945; pp 149-151.
(3) Sing, K. S. W.; Everett,D. H.; Haul, R. A. W.; Moecou, L.; Pierotti,
R. A.; RouquBrol,J.; Siemineieweka,T. Acre Appl. Chem. 1988,57,603.

In

MICROPORES

SUBSTRATE

WEAK

LAYER ING

Relative pressure, P/ Po

Figure 1. The six types of adsorption isotherm according to the


1986 IUPAC classification.

to analyze isotherm data: but the difficulties of accounting


for heterogeneity of various kinds, networking etc., is still
not resolved. If one neglectspore blockingand networking
and assumes the heterogeneity is due only to a distribution
of pore sizes and chemically heterogeneous sites on the
surface, one can approach the problem by writing the
adsorption rs in the form

where H is the pore width, ed is the attractive energy


between an adsorbed molecule and a chemically heterogeneous site on the surface, I'(H,ed) is the adsorption
isotherm for a material in which all pores are of width H
with energy tsf, as calculated by some accurate theory or
molecular simulation, and P(H,ed) is the probability
distribution for H and
for the real material, as
(4) Laetoekie, C.; Gubbine, K. E.; Quirke, N.Langmuir, in prese.

0743-7463/93/2409-1001$04.00/00 1993 American Chemical Society

Balbuena and Cubbins

1802 Langmuir, Vol. 9,No. 7, 1993

determined experimentally. The difficulty with even this


simplified approach is that we do not yet have reliable
methods for determining the probability distribution
function P(H,ed)or even the simpler pore size distribution
P(H)except for some rather easily characterized materials.
In view of this rather unsatisfactorysituation, we believe
it is useful to analyze the behavior, I'(H,e,f), for single
pores of simple geometry. Once a sound understanding
of this simpler case is achieved, the additional effects due
to chemical heterogeneity, networking, etc. can be evaluated. Accordingly, in this paper we use density functional
theory to determine the effect of the molecular and state
variables on adsorption isotherms, heats of adsorption,
and phase transitions for simple fluids adsorbed in pores
of slit geometry. The variables involved are temperature
and pressure, pore width (H),and the ratios of the
intermolecular potential parameters, e d e e and o,f/uft,
where e and u are parameters in the Lennard-Jones
potential and sf and ff subscripts indicate values for the
solid-fluid and fluid-fluid interactions,respectively. Since
heterogeneity is absent in our model, the relation between
adsorption type, phase transitions, etc. and the underlying
molecular and pore properties can be clearly seen. We
first determine the range of parameter space (kTleff,H/uft,
e,f/eft, and uduft) corresponding to each of the adsorption
types of the W A C classification; in doing this we also
introduce a new type, VII, which corresponds to capillary
evaporation (drying). This is followed by a similar analysis
of heats of adsorption and phase transitions (layering
transitions and capillary condensation) in terms of these
sameparameters. Where possible we relate these findings
in a qualitative way to experimental results.
2. Theory

2.1. Model. The system consists of a single slit pore


having two semi-infiiite parallel walls separated by a
distance H. The pore is open and immersed in a very
large reservoir containinga single-componentfluid at fixed
chemical potential p and temperature T, the totalvolume
of the system being V. The fluid inside the pores feels the
presence of the solid surfaces as an external potential, and
on reaching equilibrium its chemical potential equals the
bulk chemical potential. For the fluid-fluid intermolecular
pair potential energy we use the cut and shifted Lennardpotential, given by
Jones (U)

~ , ~ ( ifr r, <
) r,
if r > rc
(2)
where r, = 2 . 5 is
~ the cutoff radius and uuis the full LJ
potential,
u&) = uU&)
=O

-~

uU,ff= 4e[(uff/r)12- ( u , / ~ ) ~ I
(3)
The advantage of using the cut and shifted potential is
that comparisons ofthe theoretical resulta with molecular
simulations can readily be made. Where comparisons of
the theoretical results with experimental data are made,
the potential parameters (eff, uft) used should be those
fitted to the cut and shifted potential, rather than those
for the full U potential. When theoretical calculations
are made for the adsorptionisotherm with the two potential
models of eqs 2 and 3 using the samepotential parameters,
the isotherm for the full U model is shifted to lower
pressures than for the cut and shifted U and exhibits a
higher adsorption on pore filling. The results for the two
models are compared for a typical case in Figure 2, using
the theory described below. The shift of the capillary
condensation to lower pressures in the case of the full U

2.0r

r:
1.5 .

'.O

0
10-1

10-2

10-3

IO0

P/ Po

FYgum 2. Adsorption isotherms for fluids in a slit pore of width

H*= 6 at Tz = 0.8, d e n = 0.3,a d a n = 0.9462. Resulta for both


the full LJ (rc* = -) and the cut and shifted LJ potential (re*
= 2.6) are shown. Vertical lines are capillary condensation; the
approximate extent of thermodynamichyekreeis is ale0 shown.

model is similar to the shift in the condensationtransition


found for bulk fluids for the two modelsa6 For fluids in
pores this shift becomes smaller for smaller pores and vice
versa.
For the solid-fluid interaction the full U model is used.
We neglect the lateral solid structureof the wall and obtain
the external potential due to the solid by integrating the
LJ potential between one fluid molecule and each of the
molecules of the solid over the lateral solid structure.6 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,

-(

tp&)lkT = A[ 2 -)lo
Usf
5 2

y:(

02
3A(0.61A

+z ) ~

where A = 2.rrps(cdkT)(usr)2(A)
and is the solid density.
The cross-parameters are calculated according to the
Lorentz-Berthelot rules, esf = ( ~ ~ e f t ) lad
/ ~=
; (uw + q ) / 2 .
The externalpotential involves severalinputs, two of which
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, ps = 114 nm3, A = 0.335 nm. The other two
variables are the relative strength of the solid-fluid to
fluid-fluid interactions, ed/eft, and the relative range of
the solid-fluid and fluid-fluid potentials, a,f/uff. Since eq
4 is the potential exertad by one wall, the external potential
for the slit geometry is

V*&m= 9,(d + 4#

-2)

(5)

The total adsorption per unit area, Fa*, is calculated


according to

where Fa* = I',uf?, p* = puf?, H* = Hiaft, and z* = daft.


Here FBis the number of molecules adsorbedper unit area
and p is the number density. Throughout this paper we
adopt the convention of defining dimensionless quantities
by using the fluid-fluid parameters, uff and eft.
(5) Powlee, J. G. Physica 1984,126A, 289.
(6) Steele, W . A. Surf. Sci. 1973,36, 317; The Interaction of Gosee

with Solid Surfaces; Pergamon: Oxford, 1974.

Langmuir, Vol. 9, No. 7, 1993 1803

Adsorption Behavior of Simple Fluids

The adsorption behavior depends on the independent


reduced variables T* = kBT/eft,H* (for pores), edetf, and
usst/uff. In most of our calculationswe have fixed the value
of uBf/uR= 0.9462, which corresponds to the U model for
methane on graphite. We vary the ratio e,f/etf in order to
change the value of A in eq 3. With the values we have
adopted for the solid density, the separation between
layers, A, and the ratio u8f/uff,the value of A is given by
31.18(eSf/eff)/!P. Changing the value of pa, the density of
the solid, represents a mathematically equivalent modification to changing the solid-fluid interaction parameter
tsf. Changing the b S f / b f f ratio has additional effects, since
it is raised to various powers, as shown by eq 4. This ratio
gives also the relative range of the potentials, which has
been shown7to be crucial in determining the order of the
wetting transitions. Moreover, we have found that it plays
an important role for solvation forces.8
2.2. Density Functional Theory. Several theories
have been used for inhomogeneous fluids, particularly
integralequation and density functionaltheories. We have
adopted the latter approach, since it is more tractable and
describes a wide range of surface-drivenphase transitions;
moreover, it provides results that are in good agreement
with molecular simulation for a wide range of condition^.^
Within this theory, the thermodynamic grand potential,
il, the free energy appropriate to the grand canonical
(T,V,p)ensemble, is a functionalof the one-particle density
distribution, p(r). The equilibrium density profile is
obtained by minimizing this functional. 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 samevalue for the free energy. We adopt
the nonlocal mean field version of this theory due to
Tarazona.'OJ1 The grand potential energy functional Q[p(r)] is the sum of the intrinsic Helmholtz free energy
functional F[p(r)l and two other terms corresponding to
the contributions of the bulk chemical potential p and the
external potential V,a(r),
il[p(r)l = F[p(r)l- J d r p ( W - Vext(r))

(7)

where p(r) is the fluid number density at point r. The


Helmholtz free energy is expanded about a WCA reference
system of molecules with purely repulsive forqes, and this
is replaced by the free energy of a fluid of hard spheres
of diameter d in the usual ~ a y . The
~ J perturbation
~
term
involves the attractive potential ua+,t(r- rl).

where Fh[p(r)l is the free energy functional for an


inhomogeneous hard sphere fluid, pW,r') is the pair
distribution function, and uatt is given by

- uLJVC)
uLJ(r)- uLJ(rc)

-eff
uatt=

(0

r < r,
rm< r < rc
r > re

(9)

thereby neglecting correlations due to attractive forces,


so that
Tarazona's model expresses the hard sphere free energy
as a contribution of an ideal gas and an excess part. The
ideal term is exactly a functional of the local density p(r),
while the excess part is considered a functional of a
smoothed density, p , which is defined as

p(r) = Jdr'w(lr - $1; p(r))p(r')


(11)
where w(lr - r'l) is a weighting function chosen to give a
good description of the hard sphere direct pair correlation
function in the uniform fluid over a wide range of densities.
In Tarazona's theory, this is carried out by expandingthe
uniform fluid hard sphere direct correlation function
c(r1,rz) given by the density functional theory in powers
of the number density and matching the coefficients in
this expansion to the ones in the expansion of the PercusYevick expression for the hard sphere c(r1,rz). Terms up
to second order in the density are retained. The inputs
to the model are the intermolecular potentials and an
equation of state for the excess Helmholtz free energy for
hard spheres. The equivalent hard sphere diameter, d, is
calculated as a function of temperature, as suggestad by
Lu et
The explicit form is the one that approximates
the B a r k e ~ H e n d e r s o ddiameter
~
dlqf = ( a l P + q ) / ( a r 3 T * + a&
(12)
where the constants a1 a 4 were obtained by requiring
good agreement between theory and simulation at low
temperature^.'^ Equation 7 is solved numerically for the
density profile, given the conditions of bulk density,
temperature, and separation between walls. A simple
iteration scheme is used. The Carnahan-starling e x p r e
sionI6is used for the hard sphere excess free energy. This
theory has been s u c d u l l yapplied to calculateadmrption
properties of hard sphere (HS) and W fluids near hard
walls and U ~ a l l s . ~It~has
J ~ also been used to study
layeringtransitions18and prewettingefor particular values
of the intermolecular potential parameters and to predict
the bulk freezing transiti~n,'~
but a parametrized form for
the density had to be imposed. In the presence of a
structuredwall, at sufficientlylow temperatures, the fluid
might experience a fluid-solid transition, the ordering
being influencedby the wdordering. Mederoe, Tarazona,
and NavascuW' have applied the same model to understand phase transitions of submonolayer fie. Again, a
parametrized form for the density is used. In thie paper
we take the walls to be structureless, and no parametrized
form is used for the density profile; hence, no solid p
k
are predicted. However, surface transitions among fluid
phases are found with the model; the most important for
the description of adsorption behavior are wetting, layering
transitions, and capillary condensation.

...

where rmis the value of the U potential at the minimum.


The attractive term is treated in mean field approximation,

3. Adsorption Isotherm Behavior


In this section we use the theory to explore the effecte
of P ,H*, cat/ctt, and usfluffon the form of the adsorption

(7) Teletzke, G. F.; Scriven, L. E.; Davis, H. T. J. Chem. Phys. 1983,


78, 1431.
(8)Balbuena, P. B.; Berry, D.; Gubbms, K. E. J. Phys.Chem. 1993,
97, 937.
(9) Evans, R. In Inhomogeneous Fluids; Henderson, D., Ed.; Dekker:
New York, 1991; Chapter 5.
(10) Tarazona, P. Phys.Rev. A 1988,31, 2672.
(11) Tarazona, P.; Marini Bettolo Marconi, U.; Evans, R. Mol. Phys.
1987, 60, 673.
(12) Haneen,J. P.;McDonald, I. R. Theory of Simple Liquids, 2nd ed.;
Academic Press: London, 1986, pp 184-192.

(13) Lu, B. Q.; Evans, R.; Telo da Gama, M. M. Mol. Phys. 1986,66,
1319.
(14) Barker, J. A.; Henderson, D. J. Chem. Phys. 1967,47,4714.
(16) Telo da Gama, M. M. Private communication.
(16) Carnahan, N. F.; Starling, K. E. J. Chem. Phys. 1969,51,636.
(17) Peterson,B. K.; Gubbins,K. E.; Heffelfimger,G. 5.; Marini Bettolo
Marconi, U.; van Swol, F. J. Chem. Phys. 1988,88,6487.
(18) Ball, P. C.; Evans, R. J. Chem. Phys.1988,89,4412.
(19) Tarazona, P. Mol. Phys. 1984,52, 81.
(20) Mederoe, L.; Tarazona, P.; Navasculs, G. Phys.Rev. B 1987,96,
3376.

Balbuena and Gubbina

1804 Langmuir, Vol. 9, No. 7,1993

/ I

f
,

pre-wetting

0.9

transition

0.2

0.4

0.6

0.8

I .o

06

08

10

P/PO

0
P/ PO

Figure 3. (a) Classificationof adsorption isothermsfor a single


surface for Q/UE = 0.9462. (b) Schematic vapor-liquid phase
diagram,showing the prewetting line, wetting temperature, and
surface critical temperature, T,c(these temperatures, and also
the prewetting line, will depend on the value of Ea/w). (c)
Schematicadsorption isotherms for various temperatures for a
value of $en of around 0.10, where prewetting can occur;
isotherms corresponding to partial wetting, complete wetting,
and a prewetting transition are shown.

isotherms. The range of values of these variables that


correspond to the various isotherm types (Figure 1) is
determined. We first consider the case of a single planar
surface (nonporousmaterials) and then consider pores of
slit geometry. A brief preliminary discussion of some of
these results has been reported previously.21
3.1. Single Planar (Nonporoua) Surfaces. For a
= m), only types 11,111,and VI adsorption
single surface (H*
isotherms of Figure 1are found. The ranges of T* and
est/cff corresponding to each of these classes are shown in
Figure 3 for usduff = 0.946, the value corresponding to
methane on graphite. The dashed lines in the figure
dividing the types correspond to qualitative changes in
the adsorption behavior (they are not phase transitions)
and are approximate and are accurate to about 0.02 in
e,f/cnand 0.05 in T*. The subscript f on some of the classes
in Figure 3 refers to the fact that the adsorption remains
PO; such behavior is observed below the
finite as P
wetting temperature, where there is only partial wetting.
For values of e d e f f below about 0.18, only class I11is found.
For higher values of this ratio class 11is found at the higher
temperatures, while layering transitions (class VI) occur
at temperatures below about 0.8. The wettingtemperature
as a function of esf/eff is also shown. This was calculated
in the usual way22by solving Young's equation, rs = ra1
+ yc cos 0, to find the temperature for which cos B = 1;
here the y's are the surface tensions for the interfaces
indicated and B is the contact angle. The liquid-gas surface
tensions were interpolated from values calculated by Lu
et aZ.,13and the solid-gas and solid-liquid surface tensions
were calculated from the theory. The bulk phase diagram
shown in Figure 3b illustrates the different wetting regimes
schematically, for some fiied value of the ratios e d e f f and
u,f/uff. The prewetting line (shown dashed) represents a
coexistence between thin and thick adsorbed films and

(21) Balbuena, P. B.; Gubbine, K. E. Fluid Phase Equilib. 1992, 76,


21.
(22) Tarazona, P.; Evans, R. Mol. Phys. 1983,48,799.

02

0 4

P/P"

Figure 4. Adsorption isotherms calculated from the theory for


a single wall with
= 0.9462: (a, top) for
= 0.10;(b,
bottom) for CJJCR = 0.20.

ends at the surface critical temperature Tuc,whose value


depends on the values of e,fIeff and ussfluff. Prewetting
transitions were found in our calculations at temperatures
above the wetting value, for a small range of cudenvalues
near 0.10-0.12, for reduced temperatures around 1.101.14. Schematicadsorption isotherms for the single surface
for c,f/eR values in the range where prewetting occurs are
shown in Figure 3c. Several theoretically calculated
isotherms are shown in Figure 4 for two different fluidwall strengths. The weaker fluid-wall strength used in
the results of Figure 4a results in prewetting transitions
over an intermediate temperature range. The isotherm
for T* = 0.7 in Figure 4b shows layering, but this
temperature is above the critical temperature for layering
transitions; at somewhat lower temperatures the steps
shown become discontinuities, i.e. first-order layering
transitions.
3.2. Slit Pores. We present results at three different
and 1.4,corresponding
temperatures, T* = kT/etf= 0.5,0.8,
to a low, intermediate, and supercritical regime. The
critical temperature of the bulk fluid has been estimated
by Powles6 to be 1.12. Thus, we can expect phase
transitions to occur in pores at the temperatures T* = 0.5
and 0.8, but not at the supercritical temperature of T* =
1.4. The ranges of H*and e d e g corresponding to each of
the adsorption types of the IUPAC classification scheme
are shown in Figure 5, for a uduff value of 0.9462
(corresponding to methane on graphite) for the three
temperatures. The divisions between the different types
of isotherms in Figure 5 are not always clear-cut; we expect
the divisions shown to be accurate to about 0.02 in e d e
and to 1unit in H*.We note that classes I, N,and V are

Adsorption Behavior of Simple Fluids

Langmuir, Vol. 9, No. 7, 1993 1806

/ i
1-

12

H* = 7.5

3-

(VI

I
I

I CP/Pc3
0

0.2 Csf/@ff

a4

0.6

Figure 5. Classificationof adsorption isotherms for slit pores


for a d u =~ 0.9462: (a) !P = 0.5; (b) !P = 0.8; (c) !P = 1.4.The
linea in the f i i are approximateand refer to qualitativechanges
in the adsorption behavior.
specific to pores and do not occur on single surfaces.
Moreover, the classes referred to as 11,111,and VI in Figure
5differ from those for a single surface in that the adsorption
does not divergeto infinity as PapproachesPObut remains
finite because of the confinement of the fluid. All six of
the IUPAC types are found, and in addition we find a new
type, which we call VII, at the two lower temperatures. In
this type the fluid-wall forces are very weak and capillary
evaporation takes place, i.e. the gas-liquid transition in
the pore occurs at a pressure higher than the normal vapor
pressure, PIP" greater than 1. Isotherms showing negligible adsorption for all P < PO and finite adsorption at
P = PO have been reported experimentally, e.g., for Kr on
Na and on Na2023and for water on graphite;24however,
they correspond to class 111.
Several typical isotherms for the lowest temperature of
T* = 0.5 are shown in Figures 6 and 7. In Figure 6 the
effect of variation in the strength of the fluid-wall forces,
e,f/efi, at a fixed pore width of H* = 7.5 is shown. For very
weak walls type VI1 is found with capillary evaporation,
and as the wall strength increasesthe behavior passes first
to type V and then to VIf;in the latter case there is a single
layering transition from 0 to 1 layers at PIP of about
0.26,followed by capillary condensation at PIP" = 0.46.
In the figure the regions of thermodynamichysteresis are
also shown. These show metastable regions of the isotherms for which solutionsto the density functionaltheory
equations are found, even though they lie at higher
(adsorption)or lower (desorption)pressures than the value
for the true thermodynamic transitions; such solutions
correspond to localminimain the grand free energy,rather
than to the global minimum. The dashed lines show the
true thermodynamictransitions, calculated by evaluating
~~

(23)Pierotti, R.A.; Halsey, G.D.J. Plays. Chem. 1959,63,680.


(24) Avgul, N.N.;Berezin,G.I.; Kiselev, A. V.;Lygine, I. A. Zzu. Akud.
Nuuk SSR Otd. Khim. Nuuk 1961,2,205.

rY

I
I

2-

I
I

02

H* = 5 ( V I

I
I

06

04

0 8

I O

P/ P O

Figure 7. Typical isothermsfor !P = 0.6,ed/ee = 0.18, and a d / a ~


= 0.9462,showing the effect of varying pore width.
the grand free energy for each phase and determining the
point where they are equal. In Figure 7 we show the effect
of varying H*,keeping constant the fluid wall strength.
At this particular value of elJeff the behavior passes from
type I to V as the pore size increases. For somewhat
stronger fluid-wall strengths (see Figure Sa) the behavior
passes from type I to IV and then to VI. We note that in
the region marked type I in Figure 5a the isotherme
generally show a first-order phase transition from a gaslike
adsorbed phase to a liquidlike one; this occurs at low
pressures (see Figure 7). The pressure at which this
transition occurs depends strongly on both H*and e d / e ~ .
As H* decreases, or as e d e f t increases, the transition
pressure decreases. For H* = 2.5, !P = 0.6,and csf/ea=
0.3 (with uduff= 0.94621, for example, the transition occurs
at PIP 4 X 106, while for a pore of this size and
temperature when edeff= 0.4348 (the value for methane
on graphite, and close to the value for N2 on graphite4)the
transition is at a P I P value of 2.5 X IO-". When the
adsorption isotherm is plotted against PIP" on the usual
s d e of 0 to 1 such low pressure transitions are not visible,
and the isotherm has the usual Langmuir form. For H*
values below about 1.6,the adsorption drops rapidly to
zero for all relative pressures; the pores are now too small
to admit molecules. As H*increases, for e,$~valueslarger
than about 0.2, the behavior passes from type I to VIf
when the pore becomes large enoughto accommodatemore
than two layers of adsorbate.
For the intermediate temperature of !P = 0.8 the range
of edlefffor which type VI1 is found is smaller than at the

Balbuena and Gubbins

1806 Langmuir, Vol. 9, No. 7, 1993

O0.2
I
I
I

0-

"0

0.02

0.06

0.04

0.08

PO3/ kT

Figure 9. Adsorption isotherm for P = 1.4, udus = 0.9462,


and H* = 2.5. The behavior is type I", but for values of u/qt
slightly below 0.1 it changes to type 111".

I
I

IC

0 IIV)
0
0

I
I

0.2

06

04

08

I O

P/PO

Figure 8. Adsorption isotherms for T* = 0.8, u,dun = 0.9462:


(a, top) H* = 3; (b, bottom) H* = 10. For H* = 3 there is a
fiit-order phase transition at PIP" = 0.0035, which is not visible
on the scale of the plot shown here.

lower temperature, and the range for class V is correspondingly larger (Figure 5b). For values of d e f f larger
than about 0.2, on increasing H*the behavior first passes
from type I to IV, and at larger H* to type VIf. Typical
isotherms for !P = 0.8 are shown in Figure 8. At this
temperature when type I occursthe sharprise in adsorption
at very low relative pressures, correspondingto pore filling,
is sometimes continuous and is not always a first-order
transition; this is becauee for very smallpores the capillary
critical temperature lies below !P = 0.8. For somewhat
larger pores the capillary critical temperature is above 0.8
and class I with first-order transitions is found, while for
larger H*values class IV is found with capillary condensation; for such pores the transition usually occurs at PIP"
values of 0.03 or above. In the region of Figure 5b
corresponding to type VIfthe layering involves continuous
transitions at this temperature, rather than the sharp firstorder transitions seen at !P = 0.5 (see, for example, the
isotherm for eden = 0.2 in Figure 6, as compared to the
one for e d e f f = 0.435 in Figure 8b).
For the supercritical temperature of !P = 1.4 the vapor
pressure PO is undefined, and the IUPAC classification
scheme of Figure 1 is not strictly applicable. It is,
nevertheless, convenient to adopt an analogous scheme.
We therefore adopt the IUPAC schemewith the following
modifications: We plot against another dimensionless
(other chaica, e.g. PIP,, where P, is the
pressure, P a g l k ~
critical pressure for the bulk gas, could equally well be
used) and add the superscript sc to the type numbers to
remind ourselves that the temperature is supercritical. At
these high temperatures only three types are found, IW,

IIp, and IIIp; types IV, V, and VI, which involves firstorder transitions or are at temperatures close to such
transitions, cannot occur; in effect, classes IVt and Vf,
present at lower temperatures, shift to IIP and IIIP,
respectively. Some typical isotherms for micropores are
shown at this temperature in Figure 9.
The results shown in Figures 5-9 are for a fixed value
of a B f / aof~ 0.9462. Changing this ratio (equivalent to
changing the diameter of the adsorbate molecule, keeping
the substrate fixed) affects both the constant A in eq 3
and the terms (usf/z) raised to the powers 10,4, and 3. The
effect of increasing usduffis qualitatively rather similar to
decreasing edeaand is shown in Figure 10for some typical
cases at the intermediate temperature. For a carbon
surface, the size ratios shown, usduff = 0.90,0.9462, and
0.98, correspond to uff values of 0.425, 0.382, and 0.354
nm, respectively. These latter three values of uffareabout
the values for ethane, methane, and argon, respectively.
4. Heat of Adsorption

The isosteric heat of adsorption, qat,is the heat released


(per molecule) on transferring an infinitesimal amount of
the adsorbate from the coexisting bulk gas phase to the
adsorbed phase at some constant temperature, pressure
(and hence constant total adsorption Fa*),surface area A,
and pore width H,as defined, qnt is minus the enthalpy
change for such a transfer from the bulk to adsorbate
phases and so is positive since heat is released on
adsorption. It can be related to the entropies (SI,
internal
energies (LO,and volumes (V) of the two phases by8,26
Qat

- #"))
= T(,S"@

= LI(@- p)+ p(Vg) - lr(B)) (13)

where superscripts (a) and (8) refer to the values for the
adsorbed and bulk gas phases, respectively. In this
equation Sb),IJ(B),and v(B)are the values per mole for the
coexisting gas phase; fi(a),S(a),and v[B) are the partial
molal quantities for the adsorbed phase, e.g.
= (ab'/
a n ( a ) ) ~ p , ~where
p,
V is the total volume and n") is the
number of moles in the adsorbed phase at a particular T,
P, A, and H. The isosteric heat obeys the Clapeyron
equation
(25) Nicholson, D.; Pareonage, N. C. Computer Simulation and the
StatisticaZMechonics of Adsorption;AcndemicPreaa: London, 1982;pp
34-35.

Adsorption Behavior of Simple Fluids

Langmuir, Vol. 9, No. 7,1993 1807


21

*
qcond
6

P/ P O

3'0

P/P0

Figure 10. Effectof variation in adusonthe adsorptionisotherm


for T+ = 0.8,u/f/tff
= 0.30 (a, top) H*= 2.5 (type I); (b,bottom)
H* = 7.5 (types w ,VI&

s@)
- #a)
E
!(d) T r.,w = p)- p")

(14)

(15)
Over small temperature intervals it is usually possible to
neglect the temperature dependence of qat, so that (15)
can then be integrated to give

+ constant

(16)
Equations 15and 16provide a convenient relation between
the isoeteric heat of adsorption and measurablequantities.
The assumptions implicit in (15) and (16) are usually
innocuous; in the work reported here we have checked the
ideal gas and W >> Va)approximations by including
higher order virial coefficient terms and f i d that the
approximations lead to errors of no more than a few
percent. An equation that is equivalent to (15) is2s

qst = fi)-l[U@) Pv(B)- u""' - n c a ) ( d ~ ) / d n ' a ' ) , ] (17)

where 2%) = P P ) / R Tis the compressibility factor of the


(26)W&,
1988,63,49.

. , ....

, , , ,,...I

16-4

10-3

, ,,

1'0-2

, ,,

10-1

100

P/ Po

whereP is the pressure of the bulk gas phase in equilibrium


with the adsorbed phase and the derivative on the lefthand side of the equation is taken with the adsorption, re,
kept constant. If we further assume that vCa) << V(g) and
that the ideal gas law holds for the coexisting gas phase,
so that P)
- Val = RT/P, eq 14 becomes

(In P)r, -4,JRT

G.B.;Panagiotopouloe,A. Z.; R o w h o n , J. S.Mol. Phys.

Figure 11. (a, top) Isosteric heat, qlt+ = qdca, and (b, bottom)
adsorption isotherms for apore of H* = 2.5,
= 0.4348, u / a a
= 0.9462, and !P = 0.8. In part a the heat of condensation,q d * ,
is also shown.
gas. Although (17) is convenient for use in molecular

simulations, it is easier to w e (16) in density functional


theory calculations, and we have therefore used it in thie
work.
At zero pressure there are no interactions among the
adsorbate molecules, and the bosteric heat of adsorption
qEto in this limit is given byS
qato= RT -

soH*

dz V&)

expt-V,,(z)/kT)/

( s o H * dz exp(-V,&)/kT))

(18)

We have carried out calculations of the isoeteric heat

based on eqs 18for zero pressure and 16for finite pressure,


for a reduced temperature of T* = 0.8 and for a size ratio
of ud/uu = 0.9462, corresponding to methane on graphite.
The results are shown as qlt* = q.t/eH, the reduced isosteric
heat, versus adsorption rE*
in Figures 11-16. In these
figures the heat of condensation for the bulk fluid, Qcond*
= PL- H&ert, where HL and HG are the enthalpy per
molecule for the coexisting bulk liquid and gas phases,
respectively, is also shown for comparison as a horizontal
line. Firsborder phase transitions(capillarycondensation,
layering transitions) are shown as dashed limes joining the
coexisting adsorbed phases. T h e calculated values of qBt*
are expected to be accurate to within about aO.15 unit.

Balbuena and Gubbins

1808 Langmuir, Vol. 9,No. 7, 1993


I

6
0

61

0.4

0.8

1.2

I .6

35

3.0

r3

0.5

1.0

1.5

2.0

2.5

3.0

I
I

1l

l
l

l
l
l

l
l
l

l
l
l
l

l
l
l
l

(I11

2.5 -

2.0 -

I
I
I
I

3.5

0.6
0.8
I .o
P/ Po
Figure 13. (a, top) Isostsric heat and (b,bottom) adsorption
isothermsfor H*= 10. Other conditionsare the same as in Figure

Also included in the figures are typical adsorption isotherms in the temperature region near to T+ = 0.8.
In Figures 11-13 we show results for esf/eg = 0.435,the
value for methane on graphite, for three different pore
widths, H* = 2.5,5,and 10. At H* = 2.5 (Figure 11) the
isotherms are of type I, with continuous fding at this
temperature, which is above the capillary critical temperature. The values of the isosteric heat lie well above
the bulk heat of condensation, a result of the very strong
fluid-wall forces for such small micropores; at this pore
width the potential wells for the two walls overlap and
reinforce each other strongly. The pronounced peak at
rS*= 0.45corresponds to near completion of pore filling.
For iY* = 5 (Figure 121,a large micropore, the isotherms
are of type IV and show a fit-order capillary condensation
transition. In the plot of isosteric heat t h i s appears as a
nonhorizontal tie line, since the two coexisting adsorbed
phases have different values of qst. The values of qat*are
lower than for the H*= 2.5 pore, a result of the relatively
weaker fluid-wall forces for the larger pore. Results for
H* = 10,a mesopore, are shown in Figure 13. The form
of the isosteric heat curve is similar to that for H*= 5,but
the values are smaller, reflecting the larger pore and weaker
fluid-wall forcesfor the inner layere of adsorbedmolecules.
However, the bosteric heat remains larger than the bulk
heat of condensation, except for a point near rS*= 0.8
where QSt* = qmnd*. The isotherms are of type VI, and at

'F72I
I
I

P/Po
Figure 12. (a, top) Isoeteric heat and (b, bottom) adsorption
isotherms for H*= 5. Other conditionsare the same as in Figure
11. The dashed line is the tie line for capillary condensation.

0.2

0.4

11.

T+ values somewhat below 0.8 (e.g. T* = 0.7 in Figure 13b)


there is a fiit-order layering transition.
The behavior of the isosteric heat in these cases can be
given a relatively simple physical interpretation as follows.
the last
From (13)we have qat = HQ)- R(a)= Ute) - Ua);
step follows since the term P ( W - Va))is usually small
compared to the change in internal energy. Thus the
isosteric heat results mainly from the change in energy
from transferring a molecule from the bulk gas phase,
where it experiences only the fluid-fluid intermolecular
potential with its neighboring molecules, to the adsorbed
phase where the density may be much higher and the
adsorbed molecule now feels the solid-fluid forces as well
as any fluid-fluid forces. In the zero pressure limit qh will
measure the difference between the average fluid-fluid
interaction in the bulk gas and the solid-fluid interaction
in the pore. Since the latter is significantly stronger than
the fluid-fluid interaction for c d c f f = 0.4348,we observe
a relatively large qato* value of 12.0 for H* = 2.5, and
smaller values of 9.83 and 9.74 for H* = 5 and 10,
respectively, where the potential wells a t the two walls do
not significantly overlap. As the pressure, and hence the
adsorption, increases the interaction with other fluid
molecules in the bulk gas phase increases only slowly;
however,although the interaction of the adsorbed molecule
with the wall is little changed, ita interaction with other
fluid molecules increases rapidly, leading to a rise in qlt*.

Langmuir, Vol. 9,No. 7, 1993 1809

Adsorption Behavior of Simple Fluids

I-,*

PORE

BULK GAS

0.5

1.0

0
0

-0

"+mrF

--:

0.4

0.6

0
0

y
0 0
00000

0.8

1.5

2.5

2.0

r,'

0
0

25 I

0 0
O
0
0
O
0
0

I. 5

Figure 14. Physical picture of the bosteric heat of adsorption,


and its relation to the intermolecular interactions involved, for
the conditions of Figure 12 with H* = 5.

t
1

I
I

I
I

1.0

02

0.6

0.4

0.8

1.0

P/PO

Figure 16. (a, top) Ieosteric heat and (b, bottom) adsorption
isotherms for H* = 7.5. Other conditions ae in Figure 15.

I
I

I
I

P/PO

Figure 15. (a, top) bosteric heat and (b,bottom) adsorption


~ 0.9462, and Tz = 0.8, for various
isotherms for H* = 5, u d / u =
values of the ratio e$-.
Dashed linea show capillary condensation.
This rise in qat* becomes most pronounced as a layer of
adsorbed molecules nears completion (see the regions
leading up to the peake at rn*values of approximately
0.45-0.5 in Figures l l a , 12a, and 13a). This is because as
the layer nears completion, a molecule that is added to

this layer enters a cavity in the layer where it is surrounded


by other fluid molecules and so b in a deep fluid-fluid
intermolecular potential well due to its neighbors. Once
this adsorbed layer is full, any further adsorbed molecule
must go into a second adsorbed layer, where it experiences
considerably weaker solid-fluid forces, as well as weaker
fluid-fluid forces, since there are few fluid molecules in
the second layer. This leads to a sharp decrease in qSt*
following completion of the first layer, resulting in the
pronounced peaks in qat* seen near rn*= 0.5 in the figures.
For the smallest pore, H* = 2.5, the pore is full at this
point. For H* = 5 there is a further rise corresponding to
increasing density in the second layer prior to capillary
condensation. Further increase in pressure beyond capillary condensation leads to a further sharp rise in qb*,
since the added molecule must enter a cavity in which the
tightly packed surrounding fluid molecules provide a deep
potential well. For H*= 10the resulta are similar to those
at H* = 5, except that a second adsorbed layer is formed
* 1,leading to a second, smaller peak in qnt* before
at rn*
capillary condensation. The smaller values of the boateric
heat for the larger pores, particularly at pressures significantly greater than zero, are due to the fact that the
molecules experience weaker solid-fluid interactions on
average as the pore gets larger. This physical picture of
the various features of the isosteric heat is illustrated
schematically in Figure 14.
The effect of varying the strength of the solid-fluid forces
is shown in Figures 15and 16. For H*= 5 (Figure 15) the
adsorption isotherm behavior changes from type V at e&/
eyt = 0.1 to a type intermediate between IV and V at e d e i n
= 0.2, to type IV at enf/eff = 0.3. The qualitative behavior

1810 Langmuir, Vol. 9, No. 7, 1993

Balbuem and Gubbine

of the isosteric heat is similar to that shown for H* = 5


in Figure 12, but the values of qst are smaller than shown
16t
i
a
there because of the weaker fluid-wall forces. We again
Itn
find a peak at an adsorption near to rs*= 0.4 corresponding
to filling of the first adsorbed layer for e,f/eff values of 0.2
and 0.3, followed by capillary condensation. This first
peak is weaker and broader than in Figure 12 and has
disappeared by the time the e d e f f ratio has fallen to 0.1;
.Eh 8
for this latter value, no monolayer is formed prior to
capillary condensation (see Figure 15b). The values of
5
0
qsto* are 2.03,4.4, and 6.83 for e d e f t values of 0.1,0.2, and
a
4
0.3, respectively. For the weakest value of the fluid-wall
forces qst lies below the bulk heat of condensation for all
coverages prior to capillary condensation, a reflection of
I
~
~
I
I
I
the fact that the fluid-wall interactions are weaker than
0 0.2 0 0.2 0 0.2 0 0.2 0.4 0.6 0.0 1.0
the fluid-fluid ones. Results for a mesopore, H*= 7.5, are
P/ P"
shown in Figure 16, and are generally similar to those of
Figure 15, except that the values of the isosteric heat are
Figure 17. Experimental adsorption botherme for argon on
even lower due to the larger pore width. The isotherms
thermal carbon black at -195 "C. The dashed line ia for the
untreated, heterogeneous carbon and the solid lines are for
are again of types V, intermediate between IV and V, and
carbons that have been treated in vacuo at the temperatures (in
V for esf/cff ratios of 0.1, 0.2, and 0.3, respectively. The
"C) shown. Reprinted with permission from ref 27.
physical interpretation of the results shown in Figures 15
and 16 is the same as that given above for Figures 11-13,
the principal difference being the smaller values of the
loooo
1500'
3.5
isosteric heat due to the weaker fluid-wall forces.
I
The results shown in Figures 11-13,15, and 16 are all
-UNTREATED
for the intermediate temperature, T* = 0.8. For other
temperatures the results are qualitatively similar, the
structure shown by qatbeing stronger at lower temperatures
and less pronounced at higher ones. For supercritical
temperatures, e.g. T* = 1.4, there are no phase transitions
and the maxima and minima in qst are less apparent.
Qualitative comparisons of the results presented here
for isosteric heats of adsorption with experimental data
on real materials are complicated by the fact that many
real materials are heterogeneous, because of a distribution
of pore sizes and/or because of chemical heterogeneity of
the surface. These effects of heterogeneity are most
pronounced at low pressures, since at suchpressures strong
adsorption will occur in the smalleat micropores or on the
most strongly adsorbing surface sites. The effect of such
heterogeneity will usually be to obscure the structure seen
in the curves of qst at low adsorption in Figures 11-13,15,
and 16; in particular, the region leading up to monolayer
I
formation will be modified, with the sharp rise during
0
I
2
0
1
2
3
monolayer formation observed for single homogeneous
v/v,
pores being masked. In the experimentsthis is most easily
Figure 18. Experimental isoeteric heats of adsorption for argon
seen for systems with type I1 and VI isotherms. An
on untreated and treated thermal carbon black at -196 O C . Here
example is shown in Figures 17 and 18, where adsorption
u ia the total gas volume adsorbed and um is the adsorbed volume
isotherms (Figure 17) and isosteric heats (Figure 18) are
correspondingtothe complex monolayer,both at STP. Reprinted
shown for argon on untreated and graphitized c a r b ~ n s . ~ - ~with permission from ref 28.
The untreated carbon is highly heterogeneous, the isofor the bulk fluid (cf.Figure lla), due to the deep potential
therm is type I1 and the isosteric heat falls monotonically
well
from the fluid-wall forces; (b) for large micropores
for coverages up to the monolayer. However, after heat
and
mesopores, following the sharp maximum in q,t
treatment to high temperaturesmost of the heterogeneity
corresponding to monolayer formation there is a rapid
is removed, the isotherm becomes of type VI, and the
decrease to values approaching the bulk heat of condenisosteric heat rises to a maximum corresponding to
sation (cf. Figures 12a and 13a); and (c) for weak fluidformation of the monolayer before falling sharply on
wall
forces (cf. Figures 15 and 16) qat is small and lies
completionof the monolayer; these features are similar to
qmnd at the lower coverages, rising with coverage
below
those found here for monodisperse slit pores (cf. Figures
until
it
is above qwnd. An example of the latter experi11-13). Other qualitative features of the experimental
mentally
observed behavior for water on nonporous
results30 for qst are similar to those found here for
polymer surfaces is shown in Figure 19.
monodisperse slit pores. In particular: (a) for micropores,
the qstvalues are large, well above the heat of condensation
5. Fluid Phase Equilibria in Slit Pores
In this section we discus the effectsof the pore variables
(27) Polley, M. H.; Schaeffer W. D.;Smith, W. R. J. Phys. Chem. 1958,
Y

57, 469.
(28) Beebe, R. A.; Young, D. M. J. Phys. Chem. 1954,58,96.
(29)Avgul, N. N.;Kiselev, A. V. In Chemistry and Physics of Carbon;
Walker, P. L., Ed.; Dekker: New York, 1970; Vol. 6, p 1.

(30)Gregg,S.J.;Sing, K.S . W.Adsorption, SurfaceArea andPomity,


2nd ed.; Academic Prese: London, 1982. In particular BBB Figurea 2.11
to 2.16, 2.24, 2.26, 4.6, 4.7, 6.2, and 5.4.

Adsorption Behavior of Simple Fluids

Langmuir, Vol. 9, No. 7,1993 1811

,
n /n,

01

02

03

04

I /H*

Figure 21. Capillary condensation pressures versus 1/H* for


various fluid-wall strengths,together with the Kelvin equation
~ 0.9462.
result, for P = 0.6 and u t / u =

cb)
Figure 19. Experimental isostericheats of adsorption for water
on (a) poly(methy1 methacrylate)and (b) bis(A-polycarbonate)
(Lexan)at 20 "C. The adsorption isothermsare of type In.Here
n/n,,,is the total moles adsorbed relativeto the molar adsorption
for the full monolayer. Reprinted with permission from ref 30.
Copyright 1982 Academic Press.

'

# . I

O'o*l
0.06

p* 0.04

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

T'

Figure 20. Capillary condensation pressures for a pore of H*


= 6,u d u =
~ 0.9462, together with the vapor pressure curve for
the bulk fluid.

and state conditions on capillary condensation and layering


transitions in the pores. The capillary condensation
pressures as a function of temperature for a pore of width
H* = 5 are shown in Figure 20 and are compared with the
condensation pressures for the bulk fluid. The curves
shown for ed/w values of 0.1, 0.2, and 0.4348 end at a
capillary critical point; however, for esf/eff = 0.01 we were
unable to solve the density functional theory equations
for T* values above about 1.0. For the stronger wall forces
shown in Figure 20 there is capillary condensation as
expected, but the weaker wall forces (for e,f/cff values of
0.01, and for 0.1 at temperatures below about T* = 0.75)
capillary evaporation (type VII) is observed. Such behavior is qualitatively predicted by the Kelvin equation,
even though this equation is not otherwise reliable for
such small pores. Thus, on combining the Kelvin and
ClausiusXlapeyron equations, the differencebetween the
pore condensation temperature, Tpc,and the bulk fluid
condensation temperature, Tbc, at the same pressure, is
given byS1
(31) Evans,R.; Marini Bettolo Marconi, U. J. Chem. Phys. 1987,86,
7138.

(T,"- T<)/T< = 27k COS e/Hqcon,j

(19)
where yk is the liquid-gas surface tension and 8 is the
contact angle made by the liquid-gas interface with the
solid surface. For weak walls, where partial wetting occurs
and 8 lies between r / 2 and T , cos 8 is negative and the pore
condensationtemperature will be lower than the bulk value
(capillary evaporation, type VII). For more attractive
walls, where 8 lies between zero (complete wetting) and
d 2 , the pore condensation temperature will be greater
than the bulk value (capillary condensation).
In Figure 21 the effects of pore width on the capillary
condensation pressure is shown for various wall strengths
at T* = 0.6, together with the results predicted by the
Kelvin equation. As expected, the Kelvin equation is in
poor agreementwith the density functional theory results
for pores of width less than about H*= 20, except for e d / e ~
values close to 0.15. The discrepancies are especiallylarge
for the smallest pores and the stronger wall strengths.
Similar results are found for T* = 0.8, where the Kelvin
equation is in rough agreementwith the density functional
theory results for esf/cff values close to 0.1 but shows large
discrepancies for other conditions. Use of the modified
Kelvin equation results in only a slight improvement to
the r e ~ u l t s .For
~ most cases of practical interest ed/eff is
0.2 or greater, and the Kelvin equation will predict capillary
condensation pressures that are too large for a given pore
width. For low temperatures and weak substrates Tarazona et ai." have discussed the accuracy of the Kelvin
equation using the density functional theory.
The PC-P relations for layering transitions are shown
in Figure 22 for a pore of width H* = 10 and for various
fluid-wall strengths. The qualitative trends seen for the
layering transitions are similar to those found experimentally; for example,Bassignana and Larhef2obtained
similar results for ethylene on two different surfaces, PbI2
and graphite, the former being a weaker wall than the
second. Critical temperatures corresponding to the first
and second layering transitions are shown for H* = 10 and
various values of the fluid-wall strength in Figure 23. The
first layering transition is observed for values of ed/cff of
0.19(*O.l) and greater, while the second layering transition
only occurs for values of this ratio greater than 0.38f 0.2.
For the stronger fluid-wall strengths, the critical temperatures for the two layering transitions are similar, as
expected. For pores larger than H* = 10 further layering
transitions are found at the lower temperatures and for
sufficiently strong wall strengths.
(32) Baesignana, I. C.; Larher,Y.Surf. Sci. 1984,147, 48.

1812 Langmuir, Vol. 9, No. 7, 1993


I

Balbuena and Gubbim


1.2 I

T * = 0.5

P/ Po
0.8

ESf/ff

I .o

0.6

- 1st

= 0.I

0.9

0.2

LT

2nd LT
-25 1

08

1.0

1.2

1.4

1.6

1.8

2.0

2.2

/?.=

I/T*

Figure 22. Pressuretemperature relations for f i s t and second


layering transitions in a pore of width H* = 10 and ut/un =
0.9462. Curves end at a critical point on the left; values of the
critical temperature are shown for each curve.

0.06 -

0.04
P/ Po

0.4348

06

0.02
H a = IO

kTc 'Sf f

0-

0.8

Figure 25. Phase coexistence curves for capillary condensation


in pores of H* = 5 and udun = 0.9462, shown as pressure va
average pore density. Tie lines, together with their temperature,
are included.

1st LT

0 .i

0.E
0.5

0.4

0.3

0.6

Esf /Ef f

Figure 23. Reduced critical temperaturesfor the first and second


layering transitions for a pore ef width H* = 10 and u,r/utt =
0.9462. The f i t layering transition only occura for eden values
above about 0.2, and the second layering transition only occurs
for values of this ratio above about 0.4.
I

OJ

1.0

0.1

0.2

0.3

0.4 0.5

0.6

0.7

0.8

0.9

1.0

P;

T*

Figure 26. Phase coexistence envelopes for capillary condensation (cc) and f i t (1st LT) and second (2nd LT) layering
transitions for potential parameters correspondingto the methane/carbon system. Results for pore of widths H* 2.5,5, and
10 are shown.

0.7

0.4
0

0.2

0.4

0.6

0.8

I .o

P"

Figure 24. Phase coexistence curves for capillary condensation


in porea of H*= 5 and uduff= 0.9462, for two fluid-wall strengths.
The bulk fluid vapor-liquid curve is shown for comparison.
Dashed curves show estimates of the critical region (see text).

phase diagramsfor the confiied fluids are shown


in Figures 24-26, where the densities of the coexisting
phases are shown for various wall strengths and pore sizes.
The averagepore density,pp*' = pplug, is defmed to exclude
the dead volume near the walls (roughly z < 0.7utt)33*34

Since the mean field approximation used in the density


functionaltheory leads to large errors in the criticalregions,
we extrapolated the results near to,but below, the critical
region using the scaling law, Ap' = AIT - T,P,where Ap'
is the density difference between the two phases and /3
was set equal to 0.125. These approximatecriticalregions
are shown dashed. Critical points were estimated with
the aid of the law of rectilinear diameters.
In Figure 24 we show the phase envelopes for capillary
condensation for two different wall strengths for a pore
(33) Specoviua, J.; Findenegg, G. H.Ber. Burwen-Gee. Phye. Chem.
1978,82,174; 1980,84,690.
(34)Tan, Z.;Gubbine, K. E.J. Phys. Chem. 1990,94,6061.

Adsorption Behavior of Simple Fluids

Langmuir, Vol. 9, No. 7, 1993 1813

0.5
0

0.I

0.2

I /H+

0.3

0.4

0.5

Figure 27. The effect of pore width on the capillary critical


temperature, T,*= kTJeff, for parameters Correspondingto the
methanelcarbon system.

of H* = 5. The mean densities in the pore for both


coexisting phases increase with wall strength. In the case
of the strongest wall, c,f/ctf = 0.4348,the density of the
gas phase in the pore is particularly high because a
layering transition occurs prior to capillary condensation.
For this case, the density of the liquid phase in the pore
is higher, on average, than the bulk density, implying a
more ordered state than in the bulk phase. For relatively
weak walls, such as esf/cff = 0.15,the densities of the two
phases in the pore lie below the bulk values. The effect
of fluid-wall strength on the capillary condensation is
shown in terms of pressure vs mean pore density in Figure
25. Tie lines connecting coexisting phases are shown, along
with their temperature. The changes in density of the
two coexisting phases with wall strength for a given
temperature can be seen by comparingtie-lines at a given
temperature. The temperature dependence of the capillary condensation pressure is strongly affected by the
fluid-wall strength and changes sign for e,f/cffvaluesbelow
about 0.14. This inversion effect is well-known for weak
walls and can be qualitatively explained in terms of the
Kelvin e q ~ a t i o n . ~In~ Figure
. ~ ~ 26 we show the phase
coexistence envelopes for parameters corresponding to
methane in graphitic carbon pores, for both layering
transitions and capillary condensation, and for severalpore
sizes. Finally, in Figure 27 we show the efect of pore size
on the capillary critical temperature with parameters for
methane in graphitic carbon pores.
6. Conclusions and Recommendations

We have presented a detailed analysis of the classes of


adsorption, heat of adsorption, and phase transition
behavior to be expected for simple fluids in slit pores,
based on nonlocal density functional theory. The theory
is known to be quite accurate from testa against molecular
simulations, even for very small slit pores. Ita principal
limitation is that it does not readily predict the formation
of solid adsorbed phases. Freezing transitions are known
to occur in adsorbed layers at sufficientlylow temperatures,
from both simulation^^^ and e ~ p e r i m e n t . Any
~ ~ such
transitions are unlikely to have a very significant effect on
our conclusions for the adsorption isotherms, since the
density changes involved are small. The effect of freezing
(35) Sullivan, D. J. Chem. Phys. 1981, 74, 2604.
(36) Rhykerd, C.;Tan, Z.;Pozha~,L. A.; Gubbins, K. E. J . Chem. SOC.,
Faraday Trans. 1991,87,2011. Jiang, S.; Rhykerd, C.; Gubbins, K. E.
Mol. Phys., in press.
(37) Kim, H.K.;Zhang,Q. M.; Chan, M. H. Phys. Reu. B 1986,34,
4699.

transitions on the heat of adsorption curves will be more


pronounced, leading to sharp peaks or discontinuities in
qst. When making comparison of theory with experiment,
it should be noted that the predicted critical temperature
for the bulk fluid is usually somewhat different from the
experimental value, since this is a mean field theory; some
authors have preferred to make such comparisons at the
same TIT,, rather than at the same T.17J8 Despite these
limitations of the theoretical approach, we believe it to be
much more accurate and reliable than other approaches
currently used to routinely interpret absorption behavior.
Other limitations of the model considered here, not
connectad with the use of densityfunctionaltheory, include
(a) neglect of networking and pore blocking effects, (b)
neglect of chemical heterogeneity, (c) restriction to slit
pores and neglect of pore geometry heterogeneity effects,
(d) restriction to fluids of simple spherical molecules.
Effects of nonuniform pore sizes can be readily accounted
for if the pore size distribution is known, using eq 1; the
use of this theory to determine pore size distributions from
experimental adsorption isotherm data has recently been
reported.* We are in the process of investigatingthe effects
of changing to a cylindrical pore geometry98 and of the
presence of networking. Despite these limitations the
results shown here should provide a useful framework
against which qualitative comparisons with experiment
can be made, particularly for materials in which the pores
are slitlike.
The main conclusions of our calculations are as follows
(we use the terms strong and weak walls to signify
large and small values for the ratios csfleff, respectively,
where it should be noted that it is the ratio of the solidfluid and fluid-fluid intermolecular energies thaC is
important):
1. Type I adsorption isotherm behavior is characteristic
of micropores, in which H*is less than about 4.5. In welldefined slit pores phase transitions from a gaslike phase
to one with a liquidlike mean density occur even for very
small pores, provided the temperature is below the pore
critical temperature. The phase transition behavior is
complex, and for certain wall strengths the pores may fill
c~ntinuously;~
however, for most situations a phase
transition can be expected at temperatures below the pore
critical temperature. For supercritical temperatures we
refer to this class as Is,; the amount adsorbed is plotted
against either P or PIP,,where P, is the critical pressure.
2. Types I1 and IV are observed for relatively strong
walls, I1 occurring for nonporous solids and IV for
mesoporous solids at subcritical temperatures. For nonporous surfaces class I1 is observed at intermediate and
higher temperatures and may show finite adsorption (type
IIf)at the bulk saturation pressure for temperatures below
complete wetting. Class IIf is also found for porous
materials at supercritical temperatures.
3. Types I11 and V are observed for relatively weak
walls, I11 being found for nonporous solids and V for
mesoporous solids. Again I11 can show finite adsorption
(IIIf) at the bulk saturation pressure for nonporous
materials below the wetting temperature and for porous
materials at supercritical temperatures.
4. Types VI and VIf are found for nonporous materials
with reasonably strong walls at low temperatures; it will
also occur for mesoporous materials (H*above about 4)
at low temperatures (e.g. T* of about 0.5-0.6) and for larger
mesopores and macropores at intermediate temperatures
(e.g. T* of about 0.8). At the lowest temperatures a variety
(38) Balbuena, P. B.; Gubbins, K. E. COPS I11 Conference, Marseille,
May 1993, in press.

1814 Langmuir, Vol. 9, No. 7, 1993


of other transitions can ale0 occur, such as freezing and
wetting transitions; these have not been studied in this
work.
5. Type W occurs for very weak walls in either
micropores or mesopores at subcritical temperatures and
corresponds to negligible amounts of adsorption for
pressures below the capillary evaporation pressure.
6. Isosteric heats of adsorption show sharp peaks
corresponding to layer completion (especiallypronounced
for the monolayer) and discontinuities for phase transitions. In general they show more pronounced signatures
of the physical phenomena involved than do the adsorption
isotherms.
The work presented here also suggests several recommendations for experimental studies and presentation of
adsorption data:
(a)For microporousmaterials the measurementsshould
be extended to pressures as low as PlP" = 10-8 in order
to capture information about the smallest pores. For
strong walls even lower pressures may be needed.' This
is well below the sensitivity of present day commercial
sorptometers, which seem to be limited to PlP" down to
about 106 with reasonable precision. While it should be
possible to achieve such low pressures, very careful
outgassing procedures would be needed.
(b) Adsorption isotherms for micropores should be
plotted using an expanded linear scale for PlP", so that
the region of pore filling can be clearly seen. For wellordered materials phase transitions can then be clearly

Balbuena and Gubbina

seen as discontinuities in the amount adsorbed. Some


authors have preferred to plot adsorption against log(Pl
PO); this makes it possible to show the isotherm over large
ranges of PlP" but has the disadvantage of changing the
shape of the isotherm, particularly at the low adsorption
end, making it more difficult to distinguish between classee
IV and V, for example.
(c) Type VII, though uncommon, should be added to
the classification scheme.
(d) The pore sizes dividing micropores, mesopores, and
macropores should be specified in reduced units, i.e. as
H* and not H. We recommend H* = 4.5 as being the
dividing line between micropores and mesopores, rather
than H = 20 A as used in the IUPAC classification.
(e) Wherever possible, it is highly desirable to combine
measurements of the isosteric heat of adsorption with
measurements of adsorption isotherms. The heats show
strong signatures of the various physical phenomena
occurring, in some cases (e.g. monolayer completion,
freezing transitions, etc.) much stronger than the adsorption isotherms. Such a combinationof the two measurements provides a clearer picture of the phenomena and
better eatimates of pore size distribution, as well as a
stronger test of any theoretical model.

Aoknowledgment. Helpful discussions with K. S.W.


Sing are gratefully acknowledged. We thank the National
Science Foundation (GrantNo. CTS-9122460)for support
of this work.

Das könnte Ihnen auch gefallen