You are on page 1of 8

Carbon Vol 26, No. 3. pp. 261-274.

1988

Printed
mGreatBritain.

REVIEW

ARTICLE

ADSORPTION AND STRUCTURE


MICROPOROUS CARBONS *
School

of Materials

BRIAN MCENANEY
Science, University of Bath,
(Received

IN

Bath BA2 7AY, U.K

19 Oclober 1987)

Abstract-The
microporous
structure of carbons consists of a tangled network of defective carbon layer
planes in which micropores
are the spaces between the layer planes. Adsorption
of gases in micropores
is characterized
by (1) strong adsorption
at low pressure due to overlap of force fields from opposite
pore walls, (2) activated diffusion effects caused by constrictions
in the microporous
network.
and (3)
molecular
size and shape selectivity
(molecular
sieving). The surface fractal dimension
of activated
carbons decreases
from three to two with increasing
activation,
indicating that activation
smoothens
pore surfaces. Calculated adsorption potentials for slit-shaped model micropores show that adsorption
potentials
are enhanced
by a factor of up to 2 and enable critical dimensions
for diffusion of gases
through micropores
to be estimated. The Brunauer-Emmett-Teller
equation is unsuitable for analyzing
adsorption
with a significant microporous
contribution
but may be used to estimate the nonmicroporous
surface area, provided that the microporous
contribution
can be removed. The Dubinin-Radushkevich
and Dubinin-Astakhov
equations
have been more successful when applied to microporous
carbon5
because they reflect the influence of adsorbent
heterogeneity.
as they result from an approximation
to
the generalized
adsorption
isotherm (GAI). More exact solutions of the GA1 enable adsorption
energy
distribution
functions to be obtained.
The possibility of extracting
micropore
size distributions
from
adsorption
measurements
is briefly considered.
Key Words-Adsorption,

microporous

carbons,

activated

microstructure

micropores.
Definitions
of pore sizes follow recommendations
of the International
Union of Pure
& Applied Chemistry (IUPAC): micropores-width
less than 2 nm; mesopores-width
2 to 50 nm:
macropores-width
greater than 50 nm[3].
The major sources of industrial, microporous
active carbons are coals (lignites, bituminous
coals,
and anthracites),
peat, wood, and a wide range of
organic by-products
of industry
and agriculture.
About 90% of active carbons are produced in granular or powder form, with most of the remainder in
pelleted form[4]. In addition to these conventional
sources. an increasingly
wide range of specialized
industrial
microporous
adsorbents
has been produced (e.g., active carbon textiles[5,6],
and molecular sieve carbons[7,8]).
During carbonization
the
organic precursor
is thermally
degraded
to form
products that undergo either condensation
or volatilization reactions,
the competition
between these
processes determining
the carbon yield. The carbon
residue is formed by condensation
of polynuclear
aromatic compounds
and expulsion
of side chain
groups. However, industrial carbons retain a :,ignificant concentration
of heteroelements.
especially 0
and H, and many contain mineral matter. The adsorptive capacity of the carbonized materials is usually too low for practical applications,
so porosity in
the carbon is developed
by activation (i.e., by reaction of the carbon with oxidizing gases (e.g., Hz0
or COz) or by incorporation
of inorganic additives
(e.g., ZnCL) prior to carbonization.
For coal-based

1. INTRODUCTION

The literature
on adsorption
in porous carbons
ranges widely from theoretical
studies of adsorption
from the vapor and liquid phases to technical studies
of the many applications
of carbon adsorbents.
This
article is not a comprehensive
review of the subject,
nor is it a complete
survey of the many isotherm
equations
that have been proposed
to analyze adsorption in microporous
carbons. Instead, it is confined to consideration
of the effects of the microporous structure of carbons on adsorption
of gases.
Although
most nongraphitizing
(hard)
carbons
contain numerous
micropores,
not all are practical
adsorbents,
since the micropores
in them may not
be accessible to gases to a useful degree. However.
much valuable structural
information,
relevant to
microporous
carbon adsorbents,
has been obtained
from the study of materials
such as poly(acrylonitrile)-based
(PAN-based)
carbon
fibers[l]
and glasslike (vitreous)
carbons[2].
Many industrial porous carbons
(e.g., coal-based
active carbons) in addition to micropores
contain macropores
and mesopores
that may be important
in adsorption. However,
the role of these larger pores is not
considered
in this article, since the effectiveness
of
most active carbons used for gas adsorption depends
primarily on the nature and extent of the accessible

*Based on a plenary review lecture given


86, 30 June 1986, Baden-Baden,
FRG.

carbons,

at Carbon

267

268

B. MCENANEY

active carbons, Table 1 illustrates how the proportion of the different pore types varies with the nature
of the precursor[4]. By judicious choice of the precursors and careful control of carbonization and activation, it is possible to tailor active carbons to suit
particular applications.
2.

THE STRUCTURE

OF MICROPOROUS

CARBONS

Microporous
carbons have a very disordered
structure as revealed by high resolution electron microscopy and various model structures have been
proposed based on studies of polymer carbons[2,9],
carbon fibers[l],
carbonized
coals and kerogens[lO,ll], and active carbons[l2,13]. Although the
models differ in detail, the essential feature of all of
them is a twisted network of defective carbon layer
planes, cross-linked by aliphatic bridging groups.
The width of layer planes varies, but typically is
about 5 nm[12]. Simple functional groups (e.g.,
C-OH,
C===O)and heteroelements
are incorporated into the network and are bound to the periphery of the carbon layer planes. Functional groups
can have an important influence on adsorption[l4],
but detailed consideration of their effects is outside
the scope of this review. In microporous carbons the
layer planes occur singly or in small stacks of two,
three, or four with variable interlayer spacings in the
range 0.34 to 0.8 nm[ll]; Fryer[l2] found an average
layer plane separation of 0.7 nm for a steam-activated anthracite carbon, and separations in the range
0.6 to 0.8 nm for Carbosieve. There is considerable
microporosity in the form of an interconnected network of slit-shaped pores formed by the spaces between the carbon layer planes and the gaps between
the stacks. Thus the widths of pores formed by interlayer spacings (typically from about 0.34-0.8 nm)
are significantly less than 2 nm, the arbitrarily defined[3] upper limit for micropore widths. Constrictions in the microporous network are particular
features of the structure that control access to much
of the pore space. Constrictions may also occur due
to the presence of functional groups attached to the
edges of layer planes[l5,16] and by carbon deposits
formed by thermal cracking of volatiles.
3. ADSORPTION

IN MICROPOROUS

CARBONS

The complex and disorganized structure of microporous carbons, in which the dimensions of the pore
space are commensurate with the dimensions of adsorbate molecules, makes interpretation of adsorpTable 1. Pore volumes (ml/g) for coal-based active carbons[4]
Coal Precursor
Anthracite
Bituminous
Lignite
Blended

Micropores

Mesopores

Macropores

0.51
0.43
0.22
0.42

0.07
0.17
0.58
0.11

0.11
0.26
0.32
0.33

tion in these materials very difficult. First, gases are


strongly adsorbed at low pressures in micropores,
because there is enhancement of adsorption potential due to overlap of the force fields of opposite
pore walls (see below). Second, constrictions in the
microporous network cause activated diffusion effects at low adsorption temperatures when the adsorptive has insufficient kinetic energy to penetrate
fully the micropore space; this poses problems when
using N, at 77 K as the adsorbate[l7]. Third, microporous carbons can exhibit molecular sieve action
(i.e., the selective adsorption of small molecules in
narrow micropores[7]). Carbons also exhibit molecular-shape selectivity by preferential adsorption of
flat molecules, as expected from the slit shape of
the micropores[l8,19].
Molecular sieve action can
be exploited to effect separation of gas mixtures
by microporous carbons (e.g., oxygen and nitrogen separation from air using the pressure swing
method[7,20,21]).
4. FRACTAL

ANALYSIS OF MICROPOROUS

CARBONS

Fractal geometry has been applied to many


branches of science in recent years, including the
properties of porous solids[22]. The surface of a solid
may be characterized by a surface fractal dimension,
Ds, which is a measure of surface irregularity or
roughness. For a smooth surface the value of Ds
approaches 2, the dimension of a Euclidean surface.
For a very irregular surface, the value of Ds approaches 3, the dimension of a Euclidean volume;
thus 2 5 Ds s 3. The value of Ds for a fractal object
may be estimated by measuring its surface area with
different yardsticks. In the case of adsorbent surfaces, the yardsticks are usually adsorptive molecules of different size, the value of Ds being found
from the increase in surface area of the object with
decreasing molecular size. Using this approach, Avnir et al. [23] have shown that for a graphitized carbon
black the value of Ds is close to 2, as expected for
a smooth, nonporous surface. For a series of activated charcoals, the value of D, decreased from
about 3 to about 2 with increasing activation. They
concluded that, in addition to increasing pore widths
and adsorptive capacity, the process of activation
also smoothens pore surfaces. The surface fractal
dimension of a porous solid can also be estimated
by measuring its surface area as a function of particle
size. Using this method, Fairbridge et af.[24] obtained a value of Ds = 2.48 for a Syncrude coke
from N2 and CO, surface areas. The concept of fractal dimension provides an interesting new perspective on the properties of porous solids, which is being
actively developed at present.
5.

CALCULATIONS

OF ADSORPTION

POTENTIALS

There have been several theoretical studies of adsorption potentials in model micropores[25-271. Adsorption potentials were calculated by Everett and

Adsorption and structure in microporous carbons

269

Table 2. Calculated critical micropore dimensions for diffusion of gases


Adsorptive
Slit width (nm)

co
0.541

co2
0.542

Pow1[27] for cylindrical pores and aiso for slit-shaped


pores formed between semiinfinite slabs (9:3 potentials) and between single layer planes (10:4 potentials). The 10:4 potential function appears to be the
more appropriate model, considering the structure
of microporous carbons, in which dit-shaped pores
are separated by one to four layer planes[ll].
Potentials have a single minimum for narrow pores and
two minima in wide pores. Figure 1 shows the minimum 10:4 potential in the model pore, Q *, relative
to that on the plane surface, Q, plotted against the
pore half-width, d, relative to the collision radius of
the adsorptive molecule, r,. For 10:4 potentials the
two minima occur when d/r, _Z 1.140. Q*/Q is a
maximum of 2.00 at d/r, = 1.00, and Q* JQ decreases to about 1.1 at d/r, = 1.50. At d/r, less than
1.00 (i.e., between a and b, Fig. l), the enhancement
of adsorption potential decreases rapidly due to the
short range repulsive terms in the 10:4 potential
function. Recent measurements~2gl of heats of adsorption at low surface coverage of a range of hydrocarbons on microporous carbons and on carbon
blacks confirm that enhancement of adsorption energy in micropores is by a factor of about 2.
Pores with enhanced adsorption potential have
been termed ultramicro~res
and wider pores, up to
2 nm wide, have been termed supermicropores[29].
The definition of the size range of ultramicropores
is not precise and depends on the size of the adsorptive molecule, but, for the purposes of illustration, may be taken as the range corresponding to ac
in Fig. 1. For Ar (r. = 0.340 nm) and Xe (r. = 0.375
nm), ultramicroporous slit widths, W, = 2d, are 0.58
to 1.02 nm and 0.64 to 1.13 nm, respectively. W, is
the internuclear distance between C atoms in opposite walls of the model pore. If slit widths are

2.0

1.a

02
0.544

022

corrected for the finite radius of the C atoms to give


W,, in the notation of Everett and Powl[27]. then
for Ar and Xet W, = 0.24 to 0.68 nm and 0.30 to
0.79 nm respectively. A further approximate correction can be made for the compressibility of the
adsorptive in the force field of the micropore[27], to
give the available ultramicroporous pore width, W,;
for Ar and Xe, W, = 0.34 to 0.78 nm and 0.40 to
0.89 nm, respectively; these values of pore width are
within the range of interlayer spacings measured by
high resolution electron microscopy[ 11,12],
Calculated adsorption potentials in model micropores have been used to estimate critical dimensions
for diffusion of gases. For the Everett-Pow1 10:4
potentials, this dimension corresponds to dir,, at
point a in Fig. 1. For pore dimensions less than this
value, diffusion of gases becomes activated. The
value of this dimension is critical for the ability of
molecular sieve carbons to separate gases of different moiecutar size. The results of recent calculations[30,31] for diffusion of several monatomic and
polyatomic molecules through slit-shaped model micropores are in Table 2. The value of the critical
pore dimension for Ar is in very good agreement
with that obtained from the Everett-Pow1 10:4 potentials and the values for Nz and OZ are consistent
with experiments on the separation of these gases
from air[20]. These results also appear to resolve an
anomaly resulting from estimations of critical molecular dimensions using gas phase data. These yield
a smaller dimension for NZ than for CO?; this is contrary to experiment, since CO, is known to diffuse
more rapidly through carbon molecular sieves than
NJ171.
As useful as model pore calculations are in elucidating the nature of adsorptive-adsorbent
interactions in micropores, they are highly idealized when
applied to microporous carbons. For example, the
models ignore factors such as (1) the highly defective
nature of the carbon layer planes, (2) edge effects
resulting from their finite size, and (3) the role
of heteroelements
and polar surface groups (e.g.,
C%O) in the adsorption process. The calculations
also relate to the interaction of an isolated adsorptive molecule with the pore wall and so cooperative
effects during micropore filling are not considered.
6.

THE BRUNAUER-EMMETT-TELLER
(BET) EQUATION

Fig. 1. Adsorption potentials Q * in model slit-shaped micropores, relative to the adsorption potential Q on a free
surface, as a function of pore half-width, d, relative to the
collision radius of the adsorptive, r,; calculated using 10:4
potential functionsf271.

The equation
most widely used to analyse adsorption isotherms to obtain surface areas, the BET
equation[e.g.,
321, is subject to severe limitations
when applied to microporous carbons. Values of surface areas, up to about 4000 m2/g for some highly
activated carbons[33] are unrealistically high, since

B. MCENANEY

270
800

800

8
9

480

-8 -

200

I0

0.5

I.0

I.5

alpha

layer

plane,

40

60

80

InaPolP

surface area for an extended graphite


counting
both sides, is about 2800

m2/g. Adsorption in micropores does not take place


by successive buildup of molecular layers, as supposed by the BET theory. Instead, the enhanced
adsorption potential in micropores induces an adsorption process described as primary or micropore
filling[29]. Because of its widespread use for other
adsorbents, the BET surface area will continue to
be used for microporous carbons, but its notional
character should be recognized.
The BET equation can be applied with more confidence to adsorption of gases on the nonmicroporous surface of carbons (i.e., on mesopores,
macropores, and the external surface), provided that
the microporous contribution to adsorption can be
effectively removed. Three types of method have
been proposed to isolate microporous adsorption:
(1) comparative methods, (2) preadsorption techniques, and (3) isotherm subtraction. In comparative
methods, adsorption on a microporous adsorbent is
compared to adsorption on a nonporous reference
adsorbent of well-defined surface area. The most
widely used comparative methods are the t plot[34]
and the a plot[35]; these methods have been discussed in detail by Gregg and Sing[32]. The statistical
thickness of the adsorbed layer, t, is defined as t =
VPIS, where VP is the volume of adsorptive adsorbed
at pressure P and S is the total BET surface area of
the reference adsorbent. Sing[35] proposed an alternative parameter a = V/ (V at Pl PO = 0.4). Both

t and a are essentially normalizing parameters for


producing a reduced isotherm. The choice of a suitable reference material has provoked debate. Lecloux and Pirard[36] proposed that the choice should
be dictated by the value of the BET C constant, but
Gregg and Sing[32] argue persuasively that the reference material should be selected on the basis of
chemical similarity to the test material. Graphitized
carbon blacks were originally used [34,37] as reference materials for carbons, but, it has been recently
suggested that nongraphitic carbon substrates (a
heat-treated carbon[38] and sooty silica[39]) are
superior reference materials for microporous carbons.
For microporous carbons, t plots and a plots are
of similar form. Figure 2 shows a plots for adsorption
of Ar at 77 K on four coal-based activated carbons[40]. The steep rises at low a values are attributed to micropore filling and the linear regions at
high a values to adsorption on the nonmicroporous
surface. The micropore volume, V,, is estimated by
extrapolation of the linear region to a = 0 and the
nonmicroporous surface area, S, can be estimated
from the slope of the linear portion of the a plot;
the ratio S/S provides a useful index of the relative
contributions of microporous and nonmicroporous
adsorption. Values of these parameters are in Table
3, which shows that these carbons are dominantly
microporous.
In preadsorption methods[38-411, micropores are
filled at room temperature with a strongly adsorbed
vapor, usually n-nonane, which is retained in micropores on subsequent outgassing. The nonmicroporous surface area is then measured by adsorption of

Table 3. Parameters obtained from u-plots (Fig. 2) for adsorption of Ar at 77 K on coal-based activated
carbons[40]
Carbon
ABl
AB2
AB3
AB4

Fig. 3. Adsorption of CO, at 195 K on an activated cellulose triacetate carbon heat-treated to 1600C plotted in
DR coordinates to illustrate the principle of isotherm subtraction[Q].

Fig. 2. Alpha plots for adsorption of Ar at 77 K on four


..
coat-based active carbons, ABl-AB4,
plotted using a
graphitized carbon black, Vulcan-3G, as reference adsorbent[40].

the calculated

20

S(m2/g)

S(m*/g)

V&nl/g)

SIS

595
773
877
1006

47
80
61
74

0.24
0.31
0.37
0.41

0.08
0.10
0.07
0.07

Adsorption and structure in microporous carbons


N, at 77 K, while the micropore volume can be obtained by comparison of N, adsorption before and
after adsorption of n-nonane.
Isotherm subtraction[43] or decomposition[44] is a simple method for
estimating S from a single isotherm. The principle
of the method is illustrated in Fig. 3 in which the
isotherm is plotted in Dubinin-Radushkevich
coordinates (see Section 7). The low pressure part of
the isotherm, ab, which is dominated by micropore
filling, is extrapolated to high pressures, bd, using
the DR equation, and subtracted from the total high
pressure isotherm, bc, to give a residual, nonmicroporous isotherm, which can be analysed by the BET
equation to give S. Martin-Martinez and co-workers[45] have carried out an extensive comparison of
the n-nonane preadsorption and isotherm subtraction methods. They conclude that for carbons with
narrow micropores the two methods give very similar
results; when the distribution of micropore sizes is
wider, the preadsorption technique fails because nnonane is not retained in some wide micropores on
outgassing. For superactivated carbons with a very
wide range of pore sizes, neither technique was able
to separate microporous and nonmicroporous
adsorption.

7. THE

DUBININ-RADUSHKEVICH

(DR)

EQUATION

Because adsorption in microporous carbons occurs by primary or micropore filling, Dubinin and
his group originally modeled microporous adsorption using the Polanyi potential theory. Dubinin
found empirically that the characteristic curves for
adsorption on many microporous carbons could be
linearized using the DR equation[46]:

271

P/P,,, V,, is the micropore volume, A = RT ln(P,l


P) is the adsorption potential, where R is the gas
constant and T the absolute temperature, and E is
an energy constant. Dubinin showed that the energy
constant E could be factorized into a characteristic
energy, E,, which relates to the adsorbent, and an
affinity coefficient, B, which is a constant for a given
adsorptive. Essentially, B is a shifting factor that
allows the characteristic curves (linearized by the
DR equation) for different adsorptives on the same
adsorbent to be superimposed.
For some carbons the DR equation is linear over
many orders of magnitude of pressure. For others,
however, deviations from the DR equation are
found[47]. In such cases the Dubinin-Astakhov
(DA) equation has been proposed[46], in which the
exponent two in the DR equation is replaced by a
third adjustable constant n.
Figure 4 shows the effect of n on the shape of the
DA isotherm when plotted in DR co-ordinates. For
n < 2, the isotherm is convex to the abscissa; such
deviations from the DR equation are found for wellactivated carbons with a wide range of micropore
sizes[48]. For IZ > 2, the isotherm is concave to the
abscissa and such deviations from the DR equation
are found for unactivated, or slightly activated carbons with a narrow range of micropore sizes. Caution is needed in interpreting deviations from the
DR equation because some are due to activated diffusion. These deviations are characterized
by a
downward deflection in the DR plot at low PIP,,,
Fig. 5.

8. THE

GENERALIZED
ISOTHERM

V = V,, exp - [A/E]*,

(1)

where V is the amount adsorbed at relative pressure

20

40

60

80

ln2P,/

ADSORPTION
(GAI)

Because of its empirical origins and the numerous


types of deviation from it, the DR equation has not
met with general acceptance. However. a justifica-

-1.4
-_I

IO
In2 P,I P

15

Fig. 4. The influence of the parameter


n in the DA equation on the shape of the isotherm plotted in DR co-ordinates; E = 10 kJ/mol,
T = 77 K.

Fig. 5. Adsorption
of Ar at 77 K on activated
carbons
from: Saran, n ; almond shells, 0; olive stones, A. The
downward
deflections
at low Pi PO(high In* PO/P) are due
. ..^_. P.,. .^7
to activated dittuslon(4U,4X].

B. MCENANEY

272

tion for the DR and DA equations can be found by


considering the generalized adsorption isotherm
(GAI). The structures of microporous carbons are
highly disordered and their surfaces are energetically
heterogeneous so that adsorption on them may be
modeled by the GAI[49], in which it is assumed that
adsorption on sites having energy, q, can be represented by a local isotherm G(P, q); the overall
isotherm, 0(P), is given by
G(P)

= lu@YP, q)f(q)

&,

P E 4

(2)

where f is the site energy distribution function defined on a domain fl and 4 is the domain of P; this
integral equation is the GAI. Solving the GA1 for
f, given 0 and G, is ill-posed, that is, small perturbations in 8 (e.g., experimental errors) can result
in widely different solutions for f. The general approach to solving the GA1 is to restrict the class of
possible solutions by imposing constraints on f that
correspond to assumptions about or prior knowledge
of the system. The solution is then considered to be
the f in this restricted class that gives the best fit to
the experimentally determined total isotherm (e.g.,
by least squares). One approach is to constrain f to
be a smooth, analytic function which, with a suitable
equation for G, allows the GA1 to be integrated to
give an analytic function for 8. A recent example
of this approach was presented by Sircar[SO], who
chose the Langmuir equation for the local isotherm
and a gamma-type function for f. A solution to the
GA1 can also be obtained by a numerical method
called regularization[5 1,521. A third method used to
solve the GA1 is the condensation approximation
(CA)[53] in which the local isotherm is approximated by a step function. Applying the CA to the
Langmuir local isotherm results in the relationship
A = q - go, where go is a constant, minimum ad-

sorption energy.
distribution

f(4) =

If

f is assumed to be a Weibull

n(q - So)-
E

exp -

[VI,

(3)

then integration of the GA1 gives the DA equation,


or the DR equation if n = 2[54]. Thus, go and the
DA parameters n and E represent, in an approximate way, the influence of adsorbent heterogeneity
on the adsorption process. The ability of the DA
and DR equations to reflect adsorption heterogeneity may explain their success in describing adsorption on a wide variety of porous and nonporous
solids[53].
Examples of the Weibull functions calculated from
eqn (3) are in Fig. 6, where distributions of the dimensionless parameter (q - q,,)lE are plotted for
different values of n. For constant q,, and E, the
mode increases as n increases from 1.5 to 3. This is
as expected in the context of microporous carbons,
since a high value of n is found for carbons with
narrow micropores. It is also consistent with a recent
example[55] of more exact solutions of the GA1 obtained by application of the method of regularization
to adsorption of Ar at 77 K on two activated carbons. The energy distributions from regularization,
Fig. 7, show that the PVDC-based carbon, which
has narrow micropores, has a higher modal value
of q than that for anthracite-based
AB carbon,
which has a wider range of pore sizes, extending to
mesopores.
9. ESTIMATIONS

OF MICROPORE

SIZES

A useful objective for adsorption studies is the


estimation of micropore sizes. In the case of carbons
exhibiting molecular sieve action, the most obvious,
if laborious, method to obtain a micropore size dis-

0.6
f(q)
0.4

O-5

l-0

(q- q&E
Fig. 6. The influence of the DA parameter n on the shape
of the Weibull distribution function f(q) of eqn (3).

Fig. 7. Distribution functions obtained by the method of


regularization for adsorption of Ar at 77 K on a PVDCbased carbon and an antracite-based (AB) active carbon[52].

Adsorption and structure in microporous carbons


tribution is to measure adsorption isotherms of gases
of different molecular size; a recent example of this
method was presented by Carrott and Sing[56]. Alternatively, a single parameter estimate of micropore
sizes can be obtained from the characteristic energy,
E,, of the DR and DA equations. From a comparison
of En with the average Guinier gyration radius,
Rg, obtained from small angle X-ray scattering
(SAXS)[57], Dubinin and Stoeckli[58] have proposed a direct inverse relationship
E,Rg = 14.8 2 0.6 (nm . kJ/mol)

(4)

A similar relationship was proposed between E and


the width of micropores accessible to molecular
probes, W,,,,
W,,,E, = K (nm . kJ/mol),

(5)

where Kl E. is a weak function of Eo. McEnaney[59]


has shown recently that the SAXS and molecular
probe data may be correlated equally well with E.
using a two-constant, semilogarithmic equation:
W, = 4.691 exp( - O.O666E,) (nm).

(6)

Considering the limited nature of the experimental


data, these estimates of micropore size must be regarded as approximate, as is emphasized by their
authors.
The distribution function f, eqn (3) contains information that relates to the disordered structure of
microporous carbons, but extracting useful information from it is not easy. Contributions to f from
microporous carbons include (1) structural heterogeneity arising from the enhancement of adsorption
potentials in micropores and (2) surface heterogeneity arising from the presence of imperfections and
heteroatoms on the carbon surface. If the component of f due to structural heterogeneity could be
isolated, then a micropore size distribution can be
obtained by relating adsorption energy to micropore
size. In recent examples of this approach[60,61], eqn
(4) was used to relate pore size to E,, and the condensation approximation was used to obtain f. The
development of such ideas is a subject of active study
at present and may lead to better descriptions of
relationships between adsorption and structure in
microporous carbons than exist at present.
10.

SUMMARY

The microporous structure of carbons consists of


a tangled network of defective carbon layer planes.
Micropores are the spaces between the layer planes
with dimensions estimated by high resolution electron microscopy to be about 0.34 to 0.8 nm. Adsorption
of gases in microporous
carbons
is
characterized by (1) strong adsorption at low pressure due to enhanced adsorption potentials caused
by overlap of the force fields from opposite pore

213

walls, (2) activated diffusion effects caused by constrictions in the microporous network, and (3) molecular size and shape selectivity, with preferential
adsorption of flat molecules. These properties have
enabled molecular sieve carbons to be developed for
the separation of gases. Fractal analysis of adsorption in microporous carbons suggests that the surface
fractal dimension D, decreases from about 3 to about
2 with increasing activation, indicating that activation smoothens pore surfaces.
Adsorption potentials may be calculated for slitshaped model micropores formed between singlelayer planes. These calculations show that adsorption potentials in micropores are enhanced by a
factor of up to 2 compared to the free surface and
lead to critical dimensions for diffusion of gases
through micropores. The nature of the adsorption
process in micropores renders the Brunauer-Emmett-Teller (BET) equation unsuitable for analyzing adsorption isotherms with a significant microporous contribution.
The BET equation can be
used to estimate the nonmicroporous surface area,
provided that the microporous contribution to adsorption can be removed. Various techniques have
been developed to do this such as t plots, LYplots,
n-nonane preadsorption, and isotherm subtraction.
The Dubinin-Radushkevich
(DR) and DubininAstakhov (DA) equations have been more successful when applied to adsorption on microporous carbons, because they reflect the influence of adsorbent
heterogeneity on the adsorption process. The DR
and DA equations result from an approximation to
the generalized adsorption isotherm (GAI); more
exact solutions of the GA1 enable site-energy distribution functions for microporous carbons to be
obtained. Micropore size distributions for molecular
sieve carbons can be obtained from measurements
of adsorption of molecules of different size. Single
parameter estimates of pore sizes can be obtained
from the DR constant E,,. In principle, micropore
size distribution functions can also be obtained from
the energy distribution function f of the GAI.
Acknowledgments-Thanks
go to T. J. Mays for valuable
discussions, to S. Ah for the use of unpublished data. and
to J. Wang for help in the production of the figures.
REFERENCES

1. D. J. Johnson, J. Phys. D. Appl. Phys. 20,286 (1987).


2. G. M. Jenkins, K. Kawamura, and L. L. Ban, Proc.
Roy. Sot. A327, 501 (1972).
3. K. S. W. Sing, D. H. Everett, R. A. W. Haul, L.
Moscou, R. A. Pierotti, J. Rouquerol, and T. Siemienewska, Pure and Appl. Chem. 57, 603 (1985).
4. J. Wilson, Fuel 60, 823 (1981).
5. E. G. Doying, US Patent 3 256 206 (1965); E. M.
Peters, US Patent 3 235 323 (1966); A. Bailey and
F.A.P. Maggs, British Patent 1 310 101 (1971).
6. R. N. Macnair and G. Arons, In Carbon Adsorption
Handbook (Edited by P. M. Cheremisinoff and F. Ellerbush) Ann Arbor, Michigan, pp. 819-860 (1978).
7. H. Juntgen, Carbon 15, 273 (1977).
8. J. W. Neely, Carbon 19, 27 (1981).

274

B. MCENANEY

9. L. L. Ban, D. Crawford, and H. Marsh, J. Appl. Cryst.


8, 415 (1975).
10. G. R. Millward and D. A. Jefferson. In Chemistry and
Physics of Carbon (Edited by P. L. Walker, Jr. and P.
A. Thrower), Vol. 14, pp. l-82. Marcel Dekker, New
York (1978).
11. A. Oberlin, M. Villey and A. Combaz, Carbon 18,
347 (1980).
12. J. R. Fryer, Carbon 19, 431 (1981).
13. H. Marsh, D. Crawford, T. M. OGrady, and A. Wennerberg, Carbon 20, 419 (1982).
14. B. R. Puti, In Chemistry and Physics of Carbon (Edited
bv P. L. Walker. Jr.) Vol. 6. DD. 191-282. Marcel Dekker, New York; (1970).

15. H. P. Boehm and M. Voll, Carbon 7,227 (1970); idem.,
Carbon, 8, 481 (1971).

16. H. P. Boehm, A. Vass, and R. Kollmar, Extended


Abstracts, 18th American Carbon Conference, Worcester Polytechnic Institute, Worcester, MA, USA pp.
86-87 (1987).
17. P. L. Walker, Jr. and M. Shelef, Carbon 5, 7
(1967).

18. J. R. Dacey and M. J. B. Evans, Carbon 9,579 (1971).


19. S. S. Barton, M. J. B. Evans and B. H. Harrison, J.
Colloid Interface Sci. 49, 462 (1974).

20. H. Juntgen, K. Knoblauch, and K. Harder, Fuel 60,


817 (1981).
21. J. Koresch and A. Soffer. J. Chem. Sot. Faradav
Trans. Z 76, 2472 (1980).

22. D. Avnir, D. Farin, and P. Pfeifer, Nature 308, 261


(1984).

23. D. Avnir, D. Farin, and P. Pfeifer, J. Chem. Phys. 79,


3566 (1983).
24. C. Fairbridge, S. H. Ng, and A. D. Palmer, Fuel 65,
1759 (1986).
25. N. S. Gurfein, D. P. Dobychin, and L. S. Koplienko,
Russ. J. Phys. Chem. 44, 411 (1970).
26. H. F. Stoeckli, Helv. Chim. Acta 57, 2195 (1974).
27. D. H. Everett and J. C. Powl, J. Chem. Sot., Faraday
Trans. I 619, (1976).

28. P. J. M. Carrott and K. S. W. Sing, Chem. Ind. 360


(1986).
29. M. M. Dubinin, J. Colloid Interface Sci. 46,351(1974).
30. M. B. Rao, R. G. Jenkins, and W. A. Steele, Langmuir
1, 137 (1985).
M. B. Rao and R. G. Jenkins, Carbon 25,445 (1987).
;:: S. J. Gregg and K. S. W. Sing, In Adsorption, Surface
Area and Porosity, 2nd Edition, Academic Press, London (1982).
33. H. Marsh and T. G. Lamond, Carbon 1, 293 (1963).
34. J. H. de Boer, B. G. Linsen, Th. Van der Plas, and
G. J. Zondervan, J. Catal., 4, 649 (1965).
35. K. S. W. Sing, Chem. Ind. 1528 (1968).
36. A. Lecloux and J. P. Pirard, J. Colloid Interface Sci.
70, 265 (1979).

37. M. M. Dubinin, Carbon 23, 373 (1985).


J. M. Martin-Martinez,
C.
38. F. Rodriguez-Reinoso,
Prado-Burguete, and B. McEnaney, J. Phys. Chem.
91, 515 (1987).
39. P. J. M. Carrott, K. S. W. Sing, and J. H. Raistrick,
Colloids and Surfaces 21, 9 (1986).

40. S. A.-R. M. Ali, A study of microporous and nonmicroporous adsorption in activated carbons, MSc.,
Thesis, University of Bath, UK (1984).
41. S. J. Gregg and J. F. Langford, Trans. Faraday Sot.
65, 1394 (1969).

42. A. Linares-Solano, J. D. Lopez-Gonzales, J. M. Martin-Martinez, and F. Rodriguez-Reinoso,


Ads. Sci.
Tech. 1, 123 (1984).
43. S. Ali and B. McEnanev. J. Colloid Interface Sci. 107.
355 (1985).
44. H. F. Stoeckli and F. Krahenbuehl, Carbon 22, 297
2

(1984).

F. Rodriguez-Reinoso,
M.
45. J. M. Martin-Martinez,
Molina-Sabio, and B. McEnaney, Carbon 24, 255
(1986).

46. M. M. Dubinin, In Progress in Surface and Membrane


Science. Vol. 9 (Edited bv D. A. Cadenhead) Academic Press, New York, pp. l-70 (1975).

47. H. Marsh and B. Rand, J. Colloid Interface Sci. 33,


101 (1970).
48. K. J. Masters and B. McEnaney, J. Colloid Interface
Sci. 95, 340 (1983).

49. M. Jaroniec,

A. Patrykiejew

and M. Borowko, In
Vol. 14
(Edited by D. A. Cadenhead and J. F. Danielli), Academic Press, New York, pp. l-68 (1981).
50. S. Sircar, J. Colloid Znterfnce Sci. 101, 452 (1984);
idem.. J. Chem. Sot.. Faradav Trans. I. 80. 1101
(1984).
51. J. A. Britten, B. J. Travis, and L. F. Brown, A.1.Ch.E.
Progress in Membrane

and Surface Science,

Symposium Series No. 230, 79, 7 (1983).

52. T. J. Mays, B. McEnaney, and P. D. Causton, Langmuir 3, 695 (1987).

53. G. F. Cerofolini, Surface Sci. 24, 2391 (1971); idem.,


Thin Solid Films 23, 129 (1974).

54. H. F. Stoeckli, Carbon 19, 325 (1981).


55. T. J. Mays and B. McEnaney, In Characterisation of
Porous Solids. Elsevier, Amsterdam, in press.
56. P. J. M. Carrott. R. A. Roberts. and K. S. W. Sina.
In Characterisatton of Porous Solids, Elsevier, Amsterdam, in press.
57. M. M. Dubinin and G. M. Plavnik, Carbon 2, 261
(1968).

58. M. M. Dubinin and H. F. Stoeckli, J. Colloid Interface


Sci. 75, 34 (1980).
59. B. McEnanev. Carbon 25. 69 (1987).
60. M. Jaroniec and R. Made;, Carbon 25, 579 (1987).
61. M. Jaroniec, R. Madey, J. Choma, B. McEnaney, and
T. J. Mays, submitted to Carbon.

You might also like