Вы находитесь на странице: 1из 12

Langmuir 1999, 15, 533-544 533

A Molecular Model for Adsorption of Water on Activated


Carbon: Comparison of Simulation and Experiment
C. L. McCallum,† T. J. Bandosz,‡ S. C. McGrother,† E. A. Müller,§ and
K. E. Gubbins*,⊥
School of Chemical Engineering, Cornell University, Ithaca, New York 14853,
Chemistry Department, City College of New York, Convent Avenue at 138th Street,
New York, New York 10031, Departamento de Termodinámica y Fenómenos de Transferencia,
Universidad Simón Bolı́var, Caracas 1080-A, Venezuela, and Department of Chemical
Engineering, North Carolina State University, Raleigh, North Carolina 27695

Received May 19, 1998. In Final Form: September 28, 1998

Experimental and molecular simulation results are presented for the adsorption of water onto activated
carbons. The pore size distribution for the carbon studied was determined from nitrogen adsorption data
using density functional theory, and the density of acidic and basic surface sites was found using Boehm
and potentiometric titration. The total surface site density was 0.675 site/nm2. Water adsorption was
measured for relative pressures P/P0 down to 10-3. A new molecular model for the water/activated carbon
system is presented, which we term the effective single group model, and grand canonical Monte Carlo
simulations are reported for the range of pressures covered in the experiments. A comparison of these
simulations with the experiments show generally good agreement, although some discrepancies are noted
at very low pressures and also at high relative pressures. The differences at low pressure are attributed
to the simplification of using a single surface group species, while those at high pressure are believed to
arise from uncertainties in the pore size distribution. The simulation results throw new light on the
adsorption mechanism for water at low pressures. The influence of varying both the density of surface sites
and the size of the graphite microcrystals is studied using molecular simulation.

1. Introduction reduction, only source (a) above is important, and the


The adsorption behavior of water on activated carbons modeling is relatively straightforward. Such systems have
is qualitatively different from that of simple fluids, such been successfully modeled using both point charge and
as nitrogen, carbon dioxide, or hydrocarbons. This major off-center square well models of water in slit graphitic
difference arises from two sources: (a) for water the fluid- pores.1-4 Agreement with experiment is good.4 In these
fluid interaction is much more strongly attractive than cases there is almost no adsorption of water at low-to-
the interaction of water with the carbon surface, in contrast moderate relative pressures, but there is an onset of pore
to the situation for simpler fluids, where the reverse is filling that occurs suddenly at high relative pressures.
the case, and (b) the adsorption behavior for water is The adsorption behavior is of type V in the IUPAC
largely controlled by the formation of H-bonds with classification.5 Such type V behavior can be described even
oxygenated groups on the surface. Thus, the pore-filling with simple Lennard-Jones interactions,6 and requires
mechanism for water in activated carbons is distinctly only a fluid-fluid interaction that is strongly attractive
different from that for most fluids. For simple fluids pore compared to the fluid-solid interaction.
filling occurs via the formation of a fluid monolayer on the Activated carbons, in which a significant density of
surface, often followed by a second and possibly further oxygenated sites are present on the carbon surfaces, are
layers, prior to capillary condensation and pore filling. By considerably harder to model and to characterize. Mac-
contrast, water molecules first adsorb onto oxygenated roscopic phenomenological models have been proposed7,8
surface sites, and these adsorbed water molecules then which give a reasonable fit to the experimental data
act as nuclei for the formation of larger water clusters; outside the low-pressure region, but these give little insight
eventually these clusters connect, either along the surface into the underlying molecular mechanisms of adsorption
or across the pore, and pore filling usually occurs. When and are mainly useful for data interpolation. A limited
the density of oxygenated sites on the surface is ap- number of attempts to model water on activated carbons
preciable, the pore filling seems to occur by a continuous at the molecular level through molecular simulation have
filling process without capillary condensation. been reported. Segarra and Glandt9 modeled the activated
carbon as made up of graphite platelets having dipoles
In this paper we address the problem of modeling this
adsorption process at the molecular level. For graphitic (1) Antonchenko, V. Y.; Davidov, A. S.; Ilyin, V. V. Physics of Water;
carbons, in which all oxygenated groups have been Naukova Dumka: Kiev, 1991.
(2) Ulberg, D. E.; Gubbins, K. E. Mol. Simul. 1994, 13, 205; Mol.
removed from the surface through heat treatment or Phys. 1995, 84, 1139.
(3) Maddox, M.; Ulberg, D. E.; Gubbins, K. E. Fluid Phase Equilibr.
* To whom correspondence should be addressed. 1995, 104, 145.
† Cornell University. Current addresses: C. L. McCallum, Intel (4) Müller, E. A.; Rull, L. F.; Vega, L. F.; Gubbins, K. E. J. Phys.
Corporation, 5000 W. Chandler Blvd., Chandler, AZ 85226. S. C. Chem. 1996, 100, 1189.
(5) Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti,
McGrother, Department of Chemical Engineering, North Carolina R. A.; Rouquérol, J.; Siemineiewska, T. Pure Appl. Chem. 1985, 57, 603.
State University, Raleigh, NC 27695. (6) Balbuena, P. B.; Gubbins, K. E. Langmuir 1993, 9, 1801.
‡ City College of New York.
(7) Dubinin, M. M.; Serpinsky, V. V. Carbon 1981, 19, 402.
§ Simon Bolivar University.
(8) Talu, O.; Meunier, F. AIChE J. 1996, 42, 809.
⊥ North Carolina State University. (9) Segarra, E. I.; Glandt, E. D. Chem. Eng. Sci. 1994, 49, 2953.

10.1021/la9805950 CCC: $18.00 © 1999 American Chemical Society


Published on Web 12/15/1998
534 Langmuir, Vol. 15, No. 2, 1999 McCallum et al.

smeared out around the edges of the platelets to represent


the surface functional groups. More recent work10 suggests
that this model of the surface sites is insufficiently
inhomogeneous to predict the correct trends in low-
pressure adsorption and in heats of adsorption. Maddox
et al.3 and Müller et al.4 have used models in which the
pores are approximated by slits with graphite walls, on
which are embedded surface sites. In the work of Maddox
et al. the sites were assumed to consist of -COOH groups
and were modeled using optimized potentials for liquid
simulations (OPLS),11 while water was modeled using a
TIP4P potential model;12 these potential models consist
of Lennard-Jones terms for the dispersion and overlap
interactions, plus Coulomb interactions between discrete
charges placed in the molecules to mimic the H-bonding.
Surface sites were placed in a regular array on the graphite
surface. The adsorption isotherms of water at 298 K in a
2 nm pore with various site densities were obtained. A
more detailed study was carried out by Müller et al.4 for
a wide range of surface site densities at 300 K. In their
work the H-bonding between water molecules and between
water and a surface site was modeled using off-center
square well potentials in place of charges. This approach
had two advantages. First, the square well potential is
short-ranged, thus eliminating the need for time-consum-
ing Ewald lattice sums to correct for long-range Coulomb
interactions; this greatly speeds up the simulations.
Second, the square well model allows for highly localized
attractive interactions, which may more realistically model Figure 1. Two possible mechanisms for water adsorption onto
H-bonding. Results were obtained for both regular arrays activated carbon walls, for an intermediate site density. Scheme
of surface sites, and for randomly placed sites. Cooperative A corresponds to sites regularly arranged, and therefore spaced
well apart. Scheme B shows a random array at the same site
adsorption effects were found to be very important, with density. Clearly cooperative bonding is more prevalent in the
water clusters forming readily in surface regions where latter scheme. The schematic adsorption isotherm shown for
two or more sites were placed at a separation suitable for Scheme A assumes that the water-surface site bonding energy
both water-site and water-water bonding. This model is greater than the water-water value, i.e., HB HB
sf > ff . If the
captures the effects of the strongly orientation-dependent, HB HB
reverse holds, i.e., ff > sf , the slope in region I will be less
short-range H-bonds, as well as the heterogeneous nature than that in region II for scheme A. Shaded circles are water
of the surface. However, the intermolecular potential molecules, with newly adsorbed molecules shaded darker; open
parameters used by Müller et al. were not optimized for circles are surface-active sites.
water/activated carbon systems, and no comparison with
experimental data was made. This was due in part to a work we have therefore tried to capture the essential
lack of appropriate data. Although a considerable amount features of hydrogen bonding of water with the surface
of data on water/carbon systems has been published, there sites and with other water molecules, the strongly
are little data in the critically important low-pressure heterogeneous nature of the surfaces, and the distribution
region for well-characterized materials. of pore sizes. Simplifications introduced include the use
In this paper we (a) report new experimental data for of slit-shaped pores, the assumption that the surface sites
water adsorption at low pressure on a carefully charac- can be represented as being of a single species with a
terized carbon, (b) describe a molecular model of water on single H-bonding energy, the use of a single graphite
activated carbons that we believe is more realistic than microcrystal size, and the use of simulations for a few
the ones3,4,9 used previously, and (c) carry out molecular pore widths, together with interpolation between these,
simulations for this model, and compare these model to represent the continuous pore size distribution of the
results with the experimental ones. It should be empha- real carbon.
sized that the complexity of the system is such that some Before describing the methods and results, it is useful
simplifications in the model are essential at this stage. to speculate on the possible adsorption mechanisms for
Complexity in the actual physical system includes poorly this system, particularly at low pressure. For simplicity,
characterized materials, a fairly wide range of pore sizes we assume that only one water molecule can adsorb at a
and shapes, and a lack of precise knowledge of the species given site. Two possible scenarios are illustrated in Figure
and location of the surface groups. Added to these problems 1, for the case of an intermediate site density. In the
are the difficulty of the molecular simulations in a system simplest scenario, A, it is assumed that the surface sites
in which complex arrays of H-bonds are formed; excep- (open circles) are arranged in a regular array on the
tionally long runs on large systems are needed to overcome surface. At the lowest pressures single water molecules
ergodic problems and ensure results that are representa- will adsorb on individual surface sites (stage I), with
tive of a macroscopic system. Such studies are at the limits bonding energy HB sf . As the bulk gas pressure is in-
of capability of current supercomputers. In the current creased, more surface sites are occupied, until at a pressure
corresponding to stage II all sites are occupied by a single
(10) Gordon, P. A.; Glandt, E. D. Langmuir 1997, 13, 4659. water molecule. The pressure region up to stage II is the
(11) Briggs, J. M.; Nguyen, T. B.; Jorgensen, W. L. J. Phys. Chem. Henry’s law region for this heterogeneous system, and
1991, 95, 3315.
(12) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D. J. Chem. the adsorption isotherm will be linear with a slope
Phys. 1983, 79, 926. characterized by the value of HB sf . As the pressure is
Adsorption of Water on Activated Carbon Langmuir, Vol. 15, No. 2, 1999 535

increased further, water molecules will adsorb by forming


H-bonds to preadsorbed water molecules (stage III); the
bonding energy will be HB ff , producing a second linear
region of the adsorption isotherm, with a slope somewhat
different from that of the initial Henry’s law region. A
further pressure increase will lead to the formation of
larger water clusters. As these clusters grow, the op-
portunities for cooperative adsorption effects will increase;
that is, a water molecule adsorbing onto these preexisting
clusters will often be able to position itself so that it forms
two or more H-bonds with neighboring adsorbate mol-
ecules. In this stage (IV) the adsorption energy will be
much greater, leading to a steep rise in the adsorption
isotherm. This stage is followed by a rapid increase in
water adsorption with pressure, with bridging between Figure 2. Pore size distribution of the Norit activated carbon
adsorbed clusters on the same pore wall, or on opposite N2, as determined by nitrogen adsorption at 77 K using nonlocal
walls. Soon after this the pore fills with water molecules density functional theory. The vertical dashed lines divide the
(stage V). PSD into three regions, which form the basis for choosing
discrete pore sizes for study by simulation.
In scenario B the surface sites are placed randomly on
the edges of the graphite microcrystals. In this case some using the Dubinin-Radushkevich (DR) method.16 These were
sites will be relatively isolated from others; for such sites found to be SBET ) 860 m2 g-1 and Vmic ) 0.426 cm3 g-1,
the adsorption mechanism will be similar to that for respectively. The pore size distribution (PSD) was determined
scenario A. However, some sites will be close together and from the adsorption isotherm using the density functional theory
method (DFT),17,18 which has been shown to be more reliable for
can participate in a cooperative adsorption mechanism at small pores than semiempirical methods.19 The PSD determined
certain pressures. At the lowest possible pressures water in this way is shown in Figure 2 and displays sharp peaks at
molecules will bond singly to individual sites, as in scenario approximately 0.65 and 1.35 nm. There is a long tail to the
A. However, once a few such water molecules are adsorbed distribution, which extends to 8.0 nm in Figure 2; pore widths
(stage I), there will be opportunities for ensuing water greater than 8.0 nm account for less than 2.6% of the total pore
molecules to engage in cooperative bonding (i.e., to bond volume.
simultaneously to a surface site and to a preadsorbed water Surface Characterization. The densities of acidic and basic
molecule on a neighboring site). Such adsorption produces surface sites were determined using Boehm20 and potentiomet-
ric21,22 titration. The methods and full details of these determi-
an energy release of (HB HB
sf + ff ) and will result in a steep nations have been described previously15 and are not repeated
increase in the slope of the isotherm (stage II). Eventually, here. The total site density was found to be 0.675 sites/nm2, with
locations where such cooperative bonding can occur will the groups being mainly basic in nature; the oxidation method
be saturated, and further adsorption will be by water- used resulted in only small changes in the acidity of the initial
water bonding as in scenario A (stage III). Subsequent Norit sample. The species present on the surface of the N2 carbon
stages in pore filling would be the same as those in scenario are characterized by pKa values of 3.3, 4.8, 7.4, and about 10.
A. Sorption of Water. The water sorption isotherm was obtained
More complex scenarios can be imagined and are likely at 298 K using a Micromeritics ASAP 2010 sorptometer equipped
with a vapor sorption kit. Temperature was controlled by a Fisher
to be more typical of real physical systems. Thus, in Isotemp thermostatted system. Before the measurement the
practice several species of oxygenated groups are usually sample was heated at 393 K and outgassed to a constant vacuum
present, and these may have different values of HB sf . In of 10-6 Torr. HPLC grade water was used as an adsorbate. Prior
addition, some of these groups can H-bond to more than to the experiment dissolved gases were removed by consecutively
one water molecule. The initial pressure at which coop- melting and freezing the water several times, followed by
erative bonding occurs, as well as the range of pressures outgassing. Each point of the isotherm was recorded after
where this phenomenon is observed, will depend on the sufficient time to ensure that equilibrium had been reached. The
surface density of sites. For higher site densities coopera- initial dose amount was 0.1 cm3. It took several days to measure
the sorption at the lowest pressures (P/P0 e 0.02). The desorption
tive bonding is expected to commence at lower pressures. process in this pressure range was not determined because of
2. Experimental Method and Characterization the extremely slow equilibration. Thus, it took about 55 h to
Materials. As an adsorbent Norit activated carbon was chosen desorb water from a relative pressure of 0.02 to 0.01.
(Sorbonorit 2-A5998); this material is referred to as N2 in this 3. Model
(and earlier) papers. The precursor for this material was a peat Pore Size Distribution. The activated carbon is
moss. It was oxidized with 30% hydrogen peroxide at 323 K for modeled as being made up of noninterconnected slit pores
2 h. The details of the procedure applied were described
elsewhere.13-15 We also considered the use of nitric acid as an (16) Gregg, S. J.; Sing, K. S. W. Adsorption, Surface Area and Porosity,
alternative oxidizing agent. However, this strong oxidizer was 2nd edition; Academic Press: New York, 1982.
found to lead to a dramatic reduction in the total pore volume, (17) Olivier, J. P.; Conklin, W. B. Determination of pore size distri-
and so this approach was not pursued. butions from density functional theoretical models of adsorption and
Sorption of Nitrogen. The nitrogen adsorption isotherm was condensation with porous solids. Presented at the 7th International
Conference on Surface and Colloid Science, Compiegne, France, 1991.
measured at 77 K using a Micromeritics ASAP 2010 sorptometer. (18) Lastoskie, C. M.; Gubbins, K. E.; Quirke, N. J. Phys. Chem.
Prior to the adsorption measurements the sample was heated at 1993, 97, 4786.
393 K and outgassed at this temperature to a constant vacuum (19) Lastoskie, C. M.; Quirke, N.; Gubbins, K. E. In Equilibria and
of 10-6 Torr. The isotherm was used to calculate the specific Dynamics of Gas Adsorption on Heterogeneous Solid Surfaces; Rudzinski,
surface area using the BET method and the micropore volume W., Steele, W. A., Zgrablich, G., Eds.; Elsevier: Amsterdam, 1996; p
745.
(20) Boehm, H. P. In Advances in Catalysis; Academic Press: New
(13) Bandosz, T. J.; Jagiello, J.; Schwarz, J. A. Anal. Chem. 1992, 64, York, 1966; Vol. 16.
891. (21) Bandosz, T. J.; Jagiello, J.; Contescu, C.; Schwarz, J. A. Carbon
(14) Jagiello, J.; Bandosz, T. J.; Schwarz, J. A. Carbon 1992, 30, 63. 1993, 31, 1193.
(15) Bandosz, T. J.; Jagiello, J.; Schwarz, J. A.; Krzyzanowski, A. (22) Jagiello, J.; Bandosz, T. J.; Schwarz, J. A. Carbon 1994, 32,
Langmuir 1996, 12, 6480. 1026.
536 Langmuir, Vol. 15, No. 2, 1999 McCallum et al.

having a distribution of pore widths given by the experi- pore (and thus varies with pore width), H0 ) limPf0 dP/
mentally determined PSD shown in Figure 2. The PSD dN is the Henry’s law constant,
from experiment gives information on the effective, or
“available” pore width w. This quantity is assumed to be -1 + (1 + 4Kξ)1/2
related to H, the distance separating the planes through Ψ) (5)
the centers of the first layer of C atoms on opposing walls, 2K
by
and
w ) H - σc (1)
NmN
where σc is the effective diameter of a carbon atom. The ξ) (6)
Nm - N
amount of fluid adsorbed at a given pressure may be
written where N is the amount adsorbed at pressure P.
To improve this interpolation scheme, a fourth pore
N) ∫0 f(w) N(w) dw

(2) width, H ) 0.79 nm, was studied for the entire pressure
range 0 < P < 1. The four simulated adsorption isotherms,
where f(w) is the PSD and N(w) is the amount adsorbed H ) 0.79, 0.99, 1.69, and 4.5 nm, were fitted to eq 4 to
in the pores of width w at the given pressure. yield values of the parameters H0, K, and Nm. The resulting
In principle, it is possible to calculate N by carrying out smooth fits showed good agreement with the simulation
simulations to determine N(w) for a large number of fixed data in the steeply rising region of the isotherms; the main
pore widths H covering the range of widths shown by the discrepancy between eq 4 and the simulations occurred
experimentally determined PSD. The total adsorption can in the low-pressure region, where the equation gave
then be determined from eq 2, using the experimentally adsorbed amounts that were too low. The parameters H0,
determined PSD, f(w). At low pressures a few pore widths K, and Nm varied smoothly with w and were fitted to simple
suffice for such a scheme, since the isotherm changes relations. Using these relationships, together with eq 4,
slowly with pressure. For higher pressures a much larger enabled us to construct adsorption isotherms for any pore
number of pore widths is needed since pore filling occurs, width; these, together with the experimental PSD, made
and the adsorption changes rapidly. Since the simulations it possible to carry out the integration of eq 2.
are lengthy, even on the fastest supercomputers, it is not Surface Groups. Groups such as hydroxyl, carbonyl,
feasible at present to carry out such calculations for a carboxylic, phenolic, carbonyl, lactonic, chromene, and
large number of pore widths, and some interpolation pyrone have been identified on active carbon surfaces.
scheme must be devised that enables the overall adsorp- The relative and absolute amounts of these groups depend
tion to be calculated using results for a small number of on the carbon precursor material and on the activation
pore widths. We have employed two such schemes, one conditions employed in the creation of the porous sample.
suitable for low pressures and the other for higher X-ray diffraction studies show that these oxygen-contain-
pressures. ing surface groups are located around the edges of the
In the low-pressure region (0 e P/P0 e 0.02) we carbon microcrystallites.24 The experimental character-
approximate the integral of eq 2 by a finite sum over ization of surface chemistry can be achieved using a variety
discrete, but representative, pore widths: of chemical and physical techniques. Frequently, different
methods lead to quite disparate results, presumably
N ) c1N(w1) + c2N(w2) + ... + cnN(wn) (3) because of the complexities of the carbon structure.
The carbon used here has a total estimated surface group
where wi are the selected pore widths and ci are weighting density of 0.675 sites/nm2, with about 3 times as many
factors indicative of the total pore volume represented by basic groups (e.g., pyrone) as acidic surface sites (e.g.,
the individual pore widths. On the basis of the PSD shown carboxylic); both can bind water, since water can act as
in Figure 2, three values were chosen for the pore widths either a Lewis acid or base. To simplify our model, we
to be studied: H ) 0.99, 1.69, and 4.5 nm. These widths assume that the mixture of surface groups can be modeled
represent regions 1, 2, and 3 in Figure 2, respectively. by an effective surface in which all oxygenated groups
The weighting factors, ci, for these three pore sizes are have the same binding energy, HB sf , with water molecules;
taken to be proportional to the area under these three moreover, we assume that each surface group can only
regions of the PSD: their values are c1 ) 0.518, c2 ) 0.371, bond with one water molecule. We refer to this as the
and c3 ) 0.111, respectively. “effective single group” model. In reality, there will exist
As mentioned above, this weighted-averaging procedure a mixture of groups of different species on the surface,
breaks down in the pore-filling region because pores of with sites which have somewhat different binding energies
different width fill at different pressures. Weighted and some with more than one binding site. We postulate
averaging using a few pore widths necessarily leads to a that our model can represent this surface reasonably well
stepped isotherm. For these higher pressures, therefore, with a value of HB
sf that is a suitably weighted mean of all
we use the semiempirical equation of Talu and Meunier23 these site-water interactions.
to interpolate among the adsorption isotherms for the pore In our model, surface groups are represented as OH
widths studied. This three-parameter equation was groups and consist of a Lennard-Jones (LJ) sphere to
developed specifically for associating systems. It is represent an oxygen atom, and a single square well site
placed on this that represents the H atom and can interact
H0Ψ with similar sites on water molecules. Surface groups are
P) exp(Ψ/Nm) (4)
1 + KΨ placed randomly on a square lattice that is superimposed
on the carbon surfaces, with the restriction than no two
where K is a parameter related to the equilibrium constant groups can be closer than 21/6σ, where σ is the LJ diameter
for dimer formation, Nm is the saturation capacity of the
(24) Bansal, R. C.; Donnet, J. In Carbon Black; Donnet, J., Bansal,
(23) Talu, O.; Meunier, F. AIChE J. 1996, 42, 809. R. C., Wang, M., Eds.; Marcel Dekker: New York, 1993.
Adsorption of Water on Activated Carbon Langmuir, Vol. 15, No. 2, 1999 537

Table 1. Maximum Site Density n for Different


Microcrystal Edge-Lengths L, for a Random Placement
of Sites
L/nm n/nm-2 L/nm n/nm-2
1.0 9.66 2.5 4.34
1.5 6.89 3.0 3.67
2.0 5.33 10.0 1.15

Table 2. GEMC and NPTMC Simulation Results and


Experimental Data for Coexisting Liquid Water and
Water Vapor at 298 K (The Accuracy Is the Simulation
Value as a Percentage of the Experimental Value)
property experiment simulation accuracy
Fvapor(mmol/m3) 1.2798 1.267 ( 0.004 99.1 ( 0.3%
Figure 3. Model of the water molecule used in the simulation. Fliquid (mmol/m3) 55389 54400 ( 600 98 ( 1%
The oxygen is represented by a Lennard-Jones sphere of P0 (atm) 0.03125 0.0307 ( 0.0006 98 ( 2%
diameter σLJ; the association sites (hydrogen atoms and lone-
pair electrons) are tetrahedrally arranged square well sites of a range of values of HBff . These simulations yielded the
diameter σHB HB
ff and well depth ff . vapor pressure and coexisting densities of the gas and
for the O atom. The lattice lines define the edges of the liquid phases. The value of HB ff was then chosen to give
graphite microcrystals, and the length L of a microcrystal the best agreement with experimental data for these
edge is retained as a variable. Values of L between 1 and properties. This procedure gave a value of HBff /k ) 3800
10 nm have been reported, on the basis of diffraction K; agreement with experiment was within 2% for each of
studies; however, most studies have shown values of L in the coexistence properties (see Table 2).
the range 1-3 nm. Experimental site densities range from The interaction of a water molecule with the walls is of
0 to 2.65 sites/nm2. We have chosen L to be 2.0 nm for the form
most of the calculations in our study (the effect of varying
site density is considered in section 5). The arrangement usf(r1ω1) ) uCf(z) + uLJ(r) + uHB
sf (r1ω1) (9)
of groups on the surface is kept fixed as H is varied, to
eliminate differences due to site location. where uCf is the interaction energy of the water with the
Intermolecular Potentials. Water is modeled as a carbon atoms, uLJ is the LJ interaction between the LJ
LJ sphere, with four tetrahedrally arranged square well center in the water molecule and the LJ centers that
sites4,25 (Figure 3). Two sites represent hydrogen atoms, represent the O atoms in the OH groups attached to the
and two mimic the lone pairs of electrons. Bonding only walls, and uHBsf is the H-bond interaction between the
occurs between unlike sites. The sites are located at a water molecule and a site on the wall; r1 and ω1 denote
distance of 0.42σ from the center of the LJ sphere; this the location and orientation of the water molecule, z is the
placement is consistent with the observation of a peak in distance of the center of the water molecule from the
the O-H pair correlation function for water at 0.19 nm, nearest point on the wall (taken to be the plane through
as found from neutron diffraction experiments.26 Thus, the centers of the carbon atoms forming the first layer),
the fluid-fluid pair intermolecular potential energy and r is the distance between the centers of the LJ sites
between two water molecules can be expressed as in the water molecule and the OH group on the wall. The
interaction between water and the carbon atoms is

[( ) ( ) ]
σff σff
12 6
modeled using the 10-4-3 potential of Steele:27
uff(rω1ω2) ) 4ff - + uHB
ff (rω1ω2) (7)

[( ) ( )
r r
2 σCf 10 σCf 4
where r is the vector joining the centers of the two water uCf(z) ) 2πFCCfσCf2∆ - -
5 z z

( )]
molecules, ωi denotes the orientation of water molecule
i, ff and σff are the usual LJ well depth and molecular σCf4
diameter parameters, and the H-bond term is given by (10)
3∆(z + 0.61∆)3
uHB HB HB
ff (rω1ω2) ) -ff if rRβ < σff where the density of the carbon is FC ) 114 nm-3, the
) 0 otherwise (8) separation of the graphite planes is ∆ ) 0.335 nm, and
Cf and σCf are the LJ parameters for the interaction
where rRβ is the distance between the center of a square between carbon atoms and water. The carbon-carbon
well (SW) site R on molecule 1 and a SW site β on molecule potential parameters are taken as27 Cf/k ) 28 K and σCf
2. ) 0.340 nm. Fluid-solid cross parameters are calculated
The LJ parameters, and also the HB size parameter, using the usual Lorentz-Berthelot rules:
were taken from previous studies:4 σff ) 0.306 nm; ff/k
) 90.0 K; σHB Cf ) (CCff)1/2
ff ) 0.0612 nm. The calculations are sensitive
to the value of the strength of the H-bond, HB ff , and we
therefore carefully fit this parameter to best reproduce σCf ) (σCC + σff)/2 (11)
the gas-liquid coexistence properties of bulk water. A
series of Gibbs ensemble and isothermal-isobaric simu- The H-bond term in eq 9 is the same square well function
lations were undertaken at 298 K for this water model for given by eq 8, but with subscripts ff replaced by sf. The
active sites on the carbon surface are represented by LJ
(25) Müller, E. A.; Gubbins, K. E. Ind. Eng. Chem. Res. 1995, 34,
3662. (27) Steele, W. A. The Interaction of Gases with Solid Surfaces;
(26) Soper, A. K.; Phillips, M. G. Chem. Phys. 1986, 107, 47. Pergamon Press: Oxford, 1974.
538 Langmuir, Vol. 15, No. 2, 1999 McCallum et al.

maximum displacement and rotation are very small


compared to values encountered for nonbonding systems.
In these simulations the number of adsorbed molecules
fluctuates during the simulation. Calculation of the
average of this number for a range of chemical potentials
enables the adsorption isotherm to be constructed.
The walls of the slit pores lie in the x-y plane. Normal
periodic boundary conditions, together with the minimum
image convention, are applied in these two directions. For
low pressures, P/P0 e 0.02, the length of the simulation
cell in the two directions parallel to the walls was
maintained at 10 nm for each of the pore widths studied,
to maintain a sufficient number of adsorbed molecules.
For higher pressures where more water molecules were
present, the minimum cell length in the x and y directions
was 4 nm. A potential cutoff of 5σ was used. For the system
studied HBff /kT ) 12.8. While this value is not high
Figure 4. Representation of the activated carbon pore used enough to require biased sampling methods,4 long runs
in the MC simulations. The structureless walls are decorated are needed to ensure ergodicity. Associating water mol-
with active sites that are taken to be OH groups; these are ecules tend to remain in energetically favorable configu-
represented as Lennard-Jones oxygen atoms (large open circles) rations for many MC steps, and the system thus requires
and square well hydrogen atoms (small filled circles). The pore many steps to reach a true equilibrium state. In our runs
width H is the distance between the plane through the centers 500 million MC steps were used for equilibration, followed
of the carbon atoms in the first layer on opposite walls of the
pore. The lower figures (A and B) show typical arrangements
by a further 500 million steps for property averaging.
of sites on the surface for n ) 0.675 sites/nm2 for L ) 2.0 (A) Shorter runs than this were not adequate to sample
and 1.0 (B) nm. desorption events. The average number of water molecules
in the simulation cell varied from a few molecules at the
spheres with an attached single square well bonding site lowest pressures to a few hundred or thousands of
(Figure 4); these represent hydroxyl groups. The LJ sphere molecules when the pores were full; filled pores contained
represents the O atom of the OH group and its center is about 320 molecules for a pore width of 0.79 nm, 460 at
placed at a distance of 0.1364 nm above the plane through 0.99 nm, 830 at 1.69 nm, and 2100 at H ) 4.5 nm.
the centers of the first-layer carbon atoms. The LJ Calculations were carried out on the Cornell Theory Center
parameters for this LJ sphere are the same as those for IBM SP2.
the LJ site in the water molecule, i.e.,  ) ff ) 90.0 K and In determining the adsorption isotherm, we commenced
σ ) σff ) 0.306 nm. The center of the square well site is with the cell empty; a value of the fugacity corresponding
placed at a distance of 0.42σ from the center of the LJ to a low pressure was chosen and the average adsorption
sphere and is placed as far from the wall as possible (i.e., determined from the simulation. The final configuration
on the normal to the wall which passes through the center generated at each stage was used as the starting point for
of the LJ sphere). The wall-site square well interaction simulations at higher fugacities. The pressure of the bulk
has the same range as those on the water molecules (i.e., gas corresponding to a given chemical potential was
σHB
sf ) 0.0612 nm) and mimics a hydrogen atom, so that
determined from the ideal gas equation of state. Gas-
it is only able to form bonds with the lone-pair electron phase densities corresponding to the range of chemical
sites on the water molecules. Determination of the potentials studied were determined by carrying out
appropriate strength of the wall-fluid hydrogen bond is simulations of the bulk gas. These were found to agree
discussed below. with those calculated from the ideal gas equation within
the estimated errors of the simulations.
4. Molecular Simulation Method
5. Results
Calculations were carried out using the grand canonical
Low-Pressure Region (P/P0 e 0.02). At sufficiently
Monte Carlo (GCMC) method.28 This method is convenient
low pressures the adsorbate molecules interact solely with
for adsorption studies, since the chemical potential µ,
the pore walls, and fluid-fluid interactions can be
temperature T, and volume V are specified and kept fixed
neglected. The simulations show that this region lies at
in the simulation. Since µ and T are the same in the bulk
relative pressures P/P0 below about 0.003, and this seems
and adsorbed phases at equilibrium, the thermodynamic
to be confirmed by the experimental results. Experimental
state of the bulk phase is known in such simulations. Three
measurements were carried out at relative pressures down
types of molecular moves are attempted: molecular
to 0.002, corresponding to 6.3 × 10-5 bar (0.048 Torr),
creation, molecular destruction, and the usual Monte Carlo
which is approaching the limit of sensitivity of the
translation/rotation moves. Each of these three moves is
sorptometer. At relative pressures above 0.003 the ex-
attempted with equal frequency. The type of move is
perimental adsorption isotherm (Figure 5) shows a sharp
chosen randomly to maintain microscopic reversibility.
increase in slope. This steeper region persists to relative
The probability of successful creations or destructions is
pressures of about 0.01, after which the slope decreases
strongly dependent on the density of the system. The
again. The quantity plotted in Figure 5 is the excess
maximum allowable rotation and displacement of a
adsorption, Γexcess ) (N - Nbulk)/A, where N is the number
molecule are adjusted so that the combined move has an
of molecules adsorbed, Nbulk is the number of molecules
acceptance probability of about 40%. This value should
in a volume of bulk gas equal to that of the pore at the
ensure the most efficient probing of the phase space
same temperature and chemical potential, and A is the
distribution. It is noteworthy that the values of the
surface area.
(28) Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids; The simulation results are sensitive to the value chosen
Clarendon Press: Oxford, 1987. for HB
sf /k. We first made a rough estimate of this param-
Adsorption of Water on Activated Carbon Langmuir, Vol. 15, No. 2, 1999 539

experimental system. The simulations for HB sf /k ) 4800


and 4900 K show similar results, but with less adsorption,
and with the increase in slope delayed to higher pressures.
Snapshots of typical molecular configurations in the
pore reveal that all adsorbed water molecules are bonded
onto a surface site for the pressure range shown in Figure
5; however, at the higher pressures shown such adsorbed
water molecules may also be H-bonded to other water
molecules, where two surface sites are spaced a suitable
distance apart. For pressures below P/P0 ≈ 0.003 water
molecules are adsorbed onto single sites, with no water-
water bonding (Figure 6a). Since all water-surface site
bonds have the same energy in our effective single group
model, this very low pressure region corresponds to
Henry’s law regime, and the isotherm is linear. The
experimental data in this region is sparse and it is not
Figure 5. Adsorption of water at 298 K at low relative pressure possible to detect with confidence a linear region. We
(P/P0 e 0.02). Open circles are the experimental data (the solid
line is a guide to the eye), and solid points are simulation believe that even in this low-pressure region the experi-
results: squares, HB HB mental isotherm is likely to show significant nonlinearity,
sf /k ) 4800 K; circles, sf /k ) 4900 K;
HB
triangles, sf /k ) 5000 K. Data points are the weighted
since different chemical groups on the carbon surface will
average of simulation results for the three individual pore interact with water with different bonding energies. At
widths. Error bars denote one standard deviation of the pressures somewhat above P/P0 ≈ 0.003 the snapshots
simulated values. show that cooperative adsorption effects are important.
This corresponds to the region of the isotherm with
eter. Typical H-bond strengths obtained by fits of simula- increased slope. At these pressures newly adsorbed water
tions to bulk thermodynamic properties are HB sf /k ) 3643
molecules seek surface sites on which they can simulta-
K for H-bond donor group/water and 2673 K for H-bond neously bond to both the surface site and a preadsorbed
acceptor group/water, respectively. These values were water molecule, thus releasing an amount of energy HB sf
obtained for acetic acid-water complexes.11,29 These values + HB
ff . An example of this behavior is shown in Figure 6b.
are less than the value of 3800 K for water-water Thus, the adsorption mechanism in our model is similar
interactions. However, the HB sf /k value for our model is to that shown in stages I and II of scheme B of Figure 1.
expected to be larger than this, since it only allows for one At still higher pressures, beyond those shown in Figure
group-water H-bond for each group, whereas the real 5, snapshots of the simulations indicate that cooperative
carbon surface will have groups that can form multiple bonding of the type shown in Figure 6b is complete, and
H-bonds with water molecules. Carboxyl groups, for adsorption involves H-bonding of additional water mol-
example, can H-bond with two water molecules, one ecules to either isolated surface sites or to preadsorbed
through the H-bond donor OH site and the other with the water molecules through water-water bonds.
H-bond acceptor site dO. Similarly, we expect two H-bond The most noticeable difference between the simulation
acceptor sites for lactonic groups, one H-bond donor site results and the experimental ones is the persistence of
for phenolic groups, and so on. In our “effective single the steep region of the isotherm to too high of a pressure
group” model, the appropriate value of HB sf /k will be a in the case of the model. The effective single group model
weighted average of the total bonding energies for the apparently provides too many possibilities for cooperative
various surface groups for the real carbon. Using the bonding. A possible reason for this defect may be that on
experimental values of group densities for the various the real carbon surface many groups have more than one
species present, we estimate the value of the average bonding site, which is not accounted for in our model. In
H-bond strength to be HB sf /k ) 4949 K. This estimate is addition, steric hindrance is likely to limit cooperative
in agreement with our initial simulations, which showed bonding to a greater extent in real carbons than in our
that a value in the range HB sf /k ) 4800-5000 K gives the model.
best fit to the low-pressure data. Higher Pressures (0 e P/P0 e 1.0). On the basis of
We have carried out simulations in the pressure range the results at low and moderate pressures, we adopted a
0 e P/P0 e 0.02 for three pore widths, H ) 0.99, 1.69, and value of the surface bonding strength of HB sf /k ) 4800 K,
4.50 nm, and for three values of bonding strength, HB sf /k since this seemed to give the best overall agreement with
) 4800, 4900, and 5000 K. The site density was fixed at experiment for relative pressures above 0.01. Simulations
0.675 sites/nm2, and the length of the graphite microc- were conducted for pressures over the range 0 e P/P0 e
rystals was 2 nm in all systems studied. For each of the 1 for pores of width H ) 0.79, 0.99, 1.69, and 4.50 nm; site
three bonding strengths we calculated the isotherm for density and microcrystal length were kept fixed at 0.675
the PSD of Figure 2, using eq 3. These results are shown sites/nm2 and 2.0 nm, respectively. The number and
in Figure 5, along with the experimental data. None of arrangement of the sites were the same for the different
the three simulated isotherms reproduce the data over pore widths.
the full pressure range. With HB sf /k ) 5000 K the lowest The simulation results for the four pore widths are
pressure part of the isotherm is correctly reproduced, and shown in Figure 7. Since the pore capacity expressed as
the sharp increase in slope is obtained at the correct Γexcess varies widely with pore width it is more convenient
pressure (P/P0 ≈ 0.004). However, the higher slope to display the results in Figure 7 as the excess density in
persists to too high of a pressure, the slope being reduced the pore, defined as
only for pressures above about P/P0 ≈ 0.03-0.04, whereas
this reduction in slope occurs at P/P0 ≈ 0.01 in the
Γexcess
Fexcess ) (12)
(29) Gao, J. J. Phys. Chem. 1992, 96, 6432. w
540 Langmuir, Vol. 15, No. 2, 1999 McCallum et al.

Figure 6. Snapshots of typical configurations for one of the activated carbon surfaces for a pore of width H ) 1.69 nm for HB
sf /k
) 4800 K: at low pressures, top figure, P/P0 ) 0.003; lower figure, P/P0 ) 0.017. The gray background is the carbon surface; the
black spheres embedded on this surface are the O atoms of the OH surface groups, the attached small white sphere being the square
well bonding site (or H atom). Water molecules are represented as a large white sphere (O atom) with four small attached spheres
arranged tetrahedrally, representing the two H atoms (white spheres) and two lone-pair electrons (gray spheres).

Figure 7. Simulated adsorption isotherms at 298 K for water into activated carbon pores of various widths: H ) 0.79, 0.99, 1.69,
and 4.50 nm. Points denote simulated values, and lines are the fits to the simulated data using the expression of Talu and Meunier
(eq 4).

where w is the effective pore width given by eq 1. As classification,5 and the pore filling pressure and pore
expected, the isotherms are of type V in the usual IUPAC capacity depend strongly on pore width. The narrowest
Adsorption of Water on Activated Carbon Langmuir, Vol. 15, No. 2, 1999 541

Figure 8. Snapshots of typical configurations for a pore of width H ) 1.69 nm for higher pressures than those shown in Figure
6: upper figure, P/P0 ) 0.20 (showing one of the carbon surfaces); lower figure, P/P0 ) 0.575 (showing both carbon walls). Key
as in Figure 6.
pores, H ) 0.79 and 0.99 nm, display much greater To calculate the adsorption isotherm for the experi-
adsorption at low pressures than the larger pores, and mentally determined PSD of Figure 2, the semiempirical
pore filling occurs at a low relative pressure of P/P0 ≈ 0.2. equation of Talu and Meunier23 was used to interpolate
In these smaller pores water molecules are able to form between the simulated isotherms for the four pore widths
“bridges” between opposing walls at lower pressures than shown in Figure 7, as described in section 3. The constants
for larger pores; the establishment of such a link between in this equation were fitted for each of the pore widths,
the pore walls is quickly followed by pore filling, with the and were found to vary smoothly with H. These fitted
bridging water molecules acting as the nuclei for further
curves are included in Figure 7. By interpolating between
adsorption.
these fitted values of the constants K, Nm, and H0 it was
Snapshots of typical configurations are shown for the
pore of H ) 1.69 nm in Figure 8 for an intermediate possible to estimate adsorption isotherms for a wide range
pressure of P/P0 ) 0.20 and a pressure near to pore filling, of pore widths, and to then carry out the integration of eq
P/P0 ) 0.575. At the lower pressure of 0.20 we see water 2 using the experimentally determined PSD. A comparison
molecules bonding to preadsorbed water molecules, in of the isotherm determined in this way from the simulation
addition to water molecules adsorbed to surface sites and results with the experimental one is shown in Figure 10.
the cooperative bonding seen in Figure 6b. This corre- The overall agreement is quite good. The pore filling
sponds to stage III of scheme B of Figure 1. At a reduced pressure is predicted well, although the simulated curve
pressure of 0.575 the water clusters have grown and is somewhat steeper than the experimental one in this
bridged across the pore walls (stage IV of scheme B in region. The predicted adsorption is too low at low
Figure 1). Although large water clusters of liquidlike pressures, and the predicted adsorption for the filled pore
density are present, there is still considerable empty space is also somewhat too low. The underestimation of the
in the pore at this stage. At this stage additional adsorption adsorption at low pressure, and the predicted steepness
of water molecules occurs very easily, with new molecules in the pore-filling region, are primarily a result of the use
forming several water-water bonds on adsorption with of the Talu-Meunier equation, as seen from the com-
molecules already previously adsorbed. Thus, the pore
parison of this equation with the simulation results shown
fills rapidly, and at a pressure slightly above 0.575 the
pore is full. in Figure 7. Errors in the experimentally determined PSD
For smaller pores the principal difference is that the could also contribute to discrepancies in the pore-filling
formation of bridges of water molecules between opposing region. At the highest pressures, near P/P0 ≈ 1, the
walls can occur more easily, and so appears at lower simulated adsorption is approximately 6% too low. The
pressures. This is seen in the series of snapshots for a Talu-Meuinier equation describes this region well. Since
pore of width 0.99 nm in Figure 9, where bridging between the model used for water predicts the bulk density
walls occurs at a pressure of P/P0 ) 0.05, and pore filling accurately, the discrepancy in this region most likely
has occurred at a reduced pressure of about 0.25. results from uncertainty in the PSD for large pores; a
542 Langmuir, Vol. 15, No. 2, 1999 McCallum et al.

Figure 9. Snapshots of typical configurations for a pore of width H ) 0.99 nm: top figure, P/P0 ) 0.010; middle, P/P0 ) 0.050;
bottom, P/P0 ) 0.30. Both carbon walls are shown.

Figure 11. Influence of microcrystal length, L, on low-pressure


adsorption, for a pore width H ) 1.69 nm, site density ) 0.675
Figure 10. Adsorption of water from experiment and simula- sites/nm2, T ) 298 K, and HB
sf /k ) 5000 K. Circles show results
tion at 298 K. The solid line and open circles represent the for L ) 1.0 nm and squares are for L ) 2.0 nm.
experimental data; the dashed line and filled circles show the
simulated isotherm, obtained from eq 2 using simulated
isotherms for discrete pore widths together with the experi- is a substantial decrease (by up to 45%) in the adsorption
mental PSD. as the crystal length is reduced from L ) 2.0 to 1.0 nm.
The study of typical molecular configurations shows that
small increase in the fraction of large pores would lead to cooperative bonding of the type shown in Figure 6 is much
a significant increase in the adsorption capacity here. less for a microcrystal length of 1.0 nm. For example, if
Influence of Microcrystal Size on Adsorption. For we consider a point on each isotherm with the same
a constant site density, reducing the length of the amount of adsorption, N ≈ 30, the percentage of site-
microcrystals leads to an increase in the average spacing bonded molecules participating in cooperative bonding is
between active sites (see Figure 4). With sites further about 60% for L ) 2.0 nm compared to 33% for L ) 1.0
apart, we expect less cooperative bonding and less bridging nm. This observation is part of the reason for the greater
between adjacent sites. To investigate the importance of amount of adsorption at a given pressure in the case of
such effects, we carried out simulations for two micro- the larger microcrystal. Thus, the microcrystal length has
crystal lengths, L, of 1.0 and 2.0 nm at 298 K. In these a relatively large impact on the adsorption at low
calculations the site density was fixed at 0.675 sites/nm2, pressures, since it strongly affects the relative distances
the pore width was H ) 1.69 nm, and the fluid-wall between sites.
bonding strength was chosen to be HB sf /k ) 5000 K. The Effect of Site Density on Adsorption. To study the
simulation cell length was fixed at 10.0 nm, and calcula- effect of site density on adsorption, we carried out a series
tions were made for the pressure range P/P0 ) 0 to 0.10. of simulations in which the number of active sites on the
The results are shown in Figure 11. As expected, there pore walls was varied. For a pore having a width H ) 1.69
Adsorption of Water on Activated Carbon Langmuir, Vol. 15, No. 2, 1999 543

Figure 12. Effect of active site density on water adsorption


at 298 K in a pore of width 1.69 nm, with HB sf /k ) 5000 K and
a microcrystal length of 1.0 nm. Three site densities are shown:
0.3 (circles), 0.675 (squares), and 1.2 sites/nm2 (diamonds). For
the filled pore (not shown) the adsorption was about 3.3 × 10-2
mmol/m2 for all three cases.

nm and HB sf /k ) 5000 K, three site densities were


investigated: 0.30, 0.675, and 1.20 sites/nm2. The mi-
crocrystal length used for this study was 1.0 nm, and the
simulation cell length was held fixed at 3.0 nm. Adsorption
isotherms at 298 K were generated for the full range of
relative pressures, P/P0 ) 0 to 1.
Adsorption isotherms for the three site densities are
shown in Figure 12 for pressures up to the pore-filling
region. Three conclusions can be drawn. First, the pore-
filling pressure depends strongly on the site density,
occurring at a lower pressure for the higher site densities.
Second, the systems with higher site densities (0.675 and
1.20 sites/nm2) exhibit a gradual transition to pore filling,
while the system with a low site density shows a sharp
transition, which appears to be capillary condensation.
Finally, an increase in site density leads to an increase
in adsorption in the region prior to pore filling. This agrees
with experimental observations that adsorption is en-
hanced by an increase in the degree of oxidation of active
Figure 13. Snapshots of typical configurations just prior to
carbons (up to some limiting level, after which further
pore filling for a pore of width H ) 1.69 nm and HB
sf /k ) 5000
oxidation begins to destroy the porous structure and leads K. Results are shown for three site densities: n ) 0.30 (top
to a reduction in adsorption capacity). Each of these figure) at P/P0 ) 0.81, 0.675 (middle) at P/P0 ) 0.63, and 1.20
observations can be explained by the adsorption mech- (bottom) at P/P0 ) 0.45.
anism, as described below.
An increase in adsorption with site density occurs bond to surface sites, and there is almost no cooperative
naturally because there are more sites available. In bonding. Pore filling appears to occur by capillary con-
addition, at the higher site densities of 0.675 and 1.20 densation, the transition being from a state in which only
sites/nm2, cooperative bonding will come into play and a few water molecules are adsorbed onto active sites (top
lead to significant increases in adsorption. Since the snapshot in Figure 13) to one in which the pore is filled
possibilities for cooperative bonding are similar for these at a liquidlike density. At the higher site densities,
two higher site densities, the adsorption curves are very however, adsorption of single water molecules onto active
similar for relative pressures up to 0.10. For higher sites at low pressure is followed by the buildup of clusters
pressures adsorption of water onto surface sites (without (lower two snapshots in Figure 13), formation of bridges
cooperative bonding) continues up to pressures just below between pore walls, and finally pore filling. This mech-
pore filling, since the wall sites have a higher bonding anism of pore filling is clearly different from that at the
energy than the water-water H-bond. Since the pore with low site density, and the pore filling is a continuous process.
the site density of 1.20 sites/nm2 has more of these sites The results suggest that there may be a limiting site
available, adsorption is higher in this relative pressure density above which capillary condensation no longer
region (0.10-0.43). As a result of this increased adsorption, occurs, somewhere between n ) 0.3 and 0.675 sites/nm2
bridging of adsorbed water molecules between the two for this system.
pore walls occurs more readily for the highest site density.
Such bridging provides many opportunities for cooperative 6. Conclusions
water-water bonding (newly adsorbed water molecules
bonding to more than one preadsorbed molecule), and leads While the experimental isotherm is of class V as
to lower pore-filling pressures. expected, the low-pressure measurements exhibit con-
Snapshots of typical configurations for the three site siderable sensitivity to the surface chemistry, which
densities at pressures up to pore filling show a qualitative manifests itself in changes in slope at particular pressures.
difference in behavior at the lowest site density, 0.3 sites/ Although the model shows a very short Henry’s Law region
nm2 versus the two higher densities, 0.675 and 1.20 sites/ (for relative pressures P/P0 e 0.003), it seems unlikely
nm2 (Figure 13). At 0.3 sites/nm2 single water molecules that a significant Henry’s law region exists for the
544 Langmuir, Vol. 15, No. 2, 1999 McCallum et al.

experimental system, since a variety of surface groups Future experimental work on these systems should focus
are present having different adsorption energies. on carbons which have a narrow PSD and are well-
The model calculations show good overall agreement characterized. In particular, the use of neutron or X-ray
with the experimental data, and we believe they are diffraction to determine the average length (or distribution
significantly more realistic than previous attempts to of lengths) of the microcrystals would be valuable, as would
model this complex system at the molecular level. the use of spectroscopic techniques such as FTIR to
However, the effective single group model as applied here
determine the chemical nature and density of surface
still has significant defects. In particular it fails to predict
groups. Further work on the molecular model should
the very low pressure data accurately. While it describes
the increase in slope (seen experimentally at P/P0 ≈ 0.003) incorporate the possibility of groups having multiple
and the later decrease in slope, it predicts too large of a bonding sites, as well as sites of different bonding energies,
pressure range for this cooperative bonding region. We to better describe the low-pressure adsorption region.
believe this may be due to a neglect of groups that have Finally, these models should be tested against a variety
multiple bonding sites, and/or to a neglect of groups having of activated carbons, having differing surface chemistry
different bonding energies. A further possible cause may and site density.
be that the microcrystal length used (L ) 2.0 nm) is
inappropriate. Another unsatisfactory feature is the
method used here to describe the pore size distribution at Acknowledgment. This work was supported by the
the higher pressures. Because the semiempirical equation Department of Energy (Grant No. DE-FG02-88ER13974).
used to interpolate between results for different pore The simulations reported here were performed on the IBM-
widths does not describe the low-pressure region well, SP2 at the Cornell Theory Center, with the support of the
the resulting calculated isotherm for the experimental National Science foundation through a Metacenter grant
PSD is inaccurate in this region and also has too steep of (MCA93S011). S.C.M. also thanks the NSF for a CISE
a slope in the pore-filling region. This is not a defect of the Postdoctoral Fellowship. International cooperation was
model itself, but reflects the difficulty of simulating a made possible by a NSF/CONICIT U.S.-Venezuela Co-
sufficient number of pore widths to cover the PSD for this operative Research grant (No. INT-9602960).
system, which is broad. It could be diminished by the use
of a carbon with a narrower PSD. LA9805950

Вам также может понравиться