GubbinsPore2.pdf

download GubbinsPore2.pdf

of 14

Transcript of GubbinsPore2.pdf

  • 7/25/2019 GubbinsPore2.pdf

    1/14

    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 reportedfor 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; he range of variables correspondingto each of the

    six isotherm types is determined, and the underlying factors leadingto each of the types are elucidated.

    In additionto the six ypes of the 1985classification,a seventh ype is identified, corresponding to capillary

    evaporation. A similarstudy and classification s reported for the heata of adsorption and 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 experi-

    mental 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 elative pressures approaching unity the curve may reach

    a limiting value or rise if large pores are present. Types

    I1

    and I11are 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 adsorption sotherms

    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

    such materials he 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 me-

    chanical theories, particularly density functional theory,

    + Present address: Department of Chemical Engineering,Uni-

    versity 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. 940,62, 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.

    0743-7463/93/2409-1001 04.00/0

    I

    I n

    1

    MICROPORES

    SUBSTRATE

    WEAK Y

    LAYER ING

    Relative

    pressure,

    P/ Po

    Figure

    1. The six types

    of

    adsorption isotherm accordingto he

    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 neglects pore blockingandnetworking

    and assumes the heterogeneity isdue onlytoadistribution

    of pore sizes and chemically heterogeneous sites on the

    surface, one can approach the problem by writing the

    adsorption

    rs

    n the form

    where H s the pore

    width,

    ed is the attractive energy

    between an adsorbed molecule and a chemically hetero-

    geneous 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 th e real material, as

    (4)

    Laetoekie, C.;

    Gubbine,

    K.

    E.; Quirke,

    N.

    angmuir, in

    prese.

    0

    1993

    American Chemical Society

  • 7/25/2019 GubbinsPore2.pdf

    2/14

    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

    functionP(H,ed)or even the simpler pore size distribution

    P(H)

    xcept for some rather easily characterized materials.

    Inview of this rather unsatisfactory situation, 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 eval-

    uated. Accordingly, in

    h i s

    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),nd 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, espectively. Since

    heterogeneity is absent in our model, the relation between

    adsorptiontype,phase transitions, etc. and the underlying

    molecular and pore properties can be clearly seen. We

    firstdetermine the range of parameter space ( k

    Tleff,

    H/

    ft,

    e,f/eft, and uduft)correspondingto each of the adsorption

    types of the W A C classification; in doing this we

    also

    introduce a new type, VII,which correspondstocapillary

    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

    same parameters. 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-component luid at fixed

    chemical potential p and temperatureT,the totalvolume

    of the system being

    V.

    The fluid inside the pores feels the

    presence of the solid surfacesasan 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 Lennard-

    Jones U)otential, given by

    u )

    = uU&)

    - ~ ~ , ~ ( r , )

    f r < r ,

    = O if r > rc (2)

    where r , = 2 . 5 ~s the cutoff radius and uu s the full LJ

    potential,

    (3)

    The advantage of using the cut and shifted potential is

    thatcomparisonsof he theoretical resultawithmolecular

    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 fullU otential. When theoretical calculations

    are made for the adsorption sotherm with the two potential

    modelsof eqs 2 and 3 using the same potential parameters,

    the isotherm for the full U model is shifted t o lower

    pressures than for the cut and shifted

    U

    nd 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

    uU,ff

    = 4e[(uff/r)12- (u , /~ )~ I

    Balbuena and Cubbins

    2.0 1

    r:

    1.5

    .

    '.O

    0

    10-3

    10-2

    10-1

    IO0

    P/ Po

    FYgum 2.

    Adsorption isotherms for

    fluids

    n a slitpore of width

    H*

    6 at

    Tz

    = 0.8, d e n =

    0.3, d a n = 0.9462. Resulta forboth

    the full LJ (rc*= -) and the cut and shifted LJ

    potential (re*

    = 2.6) are shown. Vertical lines

    are

    capillary condensation; the

    approximate exten t of thermodynamic hyekree isisale0 shown.

    model is similar

    to

    the shift in the condensation ransition

    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 fullU model isused

    We neglect the lateral solid structure of 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 isthen performed over the planes of molecules in the

    solid, the separation between planes being A. Thisyields

    the 10-4-3 potential,

    tp&)lkT

    =

    A[ -(-)losf - ( -

    02

    3A(0.61A+ z ) ~

    2

    where A = 2.rrps(cdkT)(usr)2(A)nd isthe solid density.

    The cross-parameters are calculated according

    to

    the

    Lorentz-Berthelot rules, esf = ( ~ ~ e f t ) l / ~ ;d

    =

    uw + q ) / 2 .

    The external potential involves several nputs, two of which

    are characteristic of the surface itself: the solid density

    and the separation between layers. Inall 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

    isthe potential exertad by one wall, the external potential

    for the slit geometry is

    (5)

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

    according to

    V* m

    =

    9,(d

    +

    4#

    -

    2 )

    where Fa*

    =

    I',uf?, p*

    =

    puf?, H*

    =

    Hiaft ,and z*

    =

    daft .

    HereFB s the numberofmolecules adsorbed per 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. 19 73,36 , 317; The Interaction of Gosee

    with Solid Surfaces; Pergamon: Oxford,

    1974.

  • 7/25/2019 GubbinsPore2.pdf

    3/14

    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 calculations we have fixed the value

    of uBf/uR

    = 0.9462,

    which corresponds

    to

    theU 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,

    he 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 modi-

    fication

    to

    changing the solid-fluid interaction parameter

    tsf. Changing the bSf/bff ratio has additional effects, since

    it is raised to various powers, asshown by eq

    4.

    This ratio

    gives also the relative range of the potentials, which has

    been shown7

    o

    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 Func tiona l Theory . Several theories

    have been used for inhomogeneous fluids, particularly

    integral equation and density functional heories. We have

    adopted the latter approach, since i t 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 functional of 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 same value 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 - Jd rp (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 . ~ J ~he perturbation term

    involves the attractive potential ua+,t(r

    rl).

    Langmuir, Vol. 9, No. 7,

    1993

    1803

    thereby neglecting correlations due to attractive forces,

    so that

    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

    r

    re

    where rm s the value of the

    U

    otential a t the minimum.

    The attractive term is treated in mean field approximation,

    -eff

    -

    uLJVC)

    uatt= uLJ(r ) uLJ(rc)

    rm

    < r