Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

THERMODYNAMICS

Thermodynamics of Adsorption
in Porous Materials
A. L. Myers
Dept. of Chemical Engineering, University of Pennsylvania, Philadelphia, PA 19104

Thermodynamic equations are deeloped for adsorption of multicomponent gas mix-


tures in microporous adsorbents based on the principles of solution thermodynamics.
The conentional spreading pressure and surface area ariables, which describe 2-D
films, must be abandoned for adsorption in micropores, in which spreading pressure
cannot be measured experimentally or calculated from intermolecular forces. Adsorp-
tion is diided into two steps: (1) isothermal compression of the gas, (2) isothermal
immersion of clean adsorbent in the compressed gas. Thermodynamic functions (Gibbs
free energy, enthalpy, and entropy) from solution thermodynamics proide a complete
thermodynamic description of the system. Applications are described for characteriza-
tion of adsorbents, gas storage at high pressure, mixture adsorption, enthalpy balances,
molecular simulation, adsorption calorimetry, and shape selectiity in catalysis.

Introduction
Brief history of adsorption thermodynamics
The concept of surface excess in adsorption was intro- ing, because the evolution of a heat of adsorption requires a
duced by J. Willard Gibbs, but the interpretation and appli- change in loading. The existence of other heats of adsorption
cation of the Gibbsian version of thermodynamics of surfaces such as the differential heat of adsorption equal to q st y
was impeded by the abstruse nature of his writing Gibbs, RT ., the equilibrium heat of adsorption obtained by per-
1928.. The thermodynamics textbook by Lewis and Randall forming the differentiation in Eq. 1 at constant spreading
1923. exerted an enormous influence upon the development pressure, and others Valenzuela and Myers, 1989. adds to
of thermodynamics in the first half of the 20th century. How- the confusion. Agreement within the adsorption community
ever, their short section on thermodynamics of surfaces was on the thermodynamic definition and physical meaning of the
devoted entirely to the surface tension of liquid-vapor inter- energy of adsorption is long overdue.
faces. The revision of Lewis and Randall by Pitzer and Brewer Important contributions to the thermodynamics of physical
Lewis et al., 1961. devoted an entire chapter to surface ef- adsorption by D. H. Everett, T. L. Hill, and L. E. Drain were
fects, which concluded with a five-page section on adsorption summarized by Young and Crowell 1962.. The thermody-
of gases. In particular they defined the isosteric heat qst . namics of physical adsorption on solid adsorbents is based on
the concept of an inert adsorbent and the introduction of two
new variables: surface area A. and spreading pressure II..
ln P
q st s RT 2 A typical equation Eq. 15 of Chapter 3. from Young and
T / n
1.
Crowells book is

This definition of isosteric heat obtained by differentiating a


series of adsorption isotherms at constant loading n. is still dFs sy S s dT qVs dP q AdP q m s dn s 2.
used today. Unfortunately, Eq. 1 applies only to a pure, per-
fect gas and its connection with the enthalpy of the adsorbed
phase and its extension to the case of real gas mixtures has where Fs sGs q P A is a free energy and the subscripts refer
led to considerable confusion. The terminology isosteric to the surface phase. This 2-D, surface thermodynamics ap-
heat intended to mean lines of constant loading is mislead- proach is standard in adsorption Ruthven, 1984; Yang, 1987;

AIChE Journal January 2002 Vol. 48, No. 1 145


Valenzuela and Myers, 1989; Suzuki, 1990.. The problem with Table 1. Integral and Differential Thermodynamic Functions
the 2-D approach is that it requires a series of assumptions of for Single-Component Adsorption.U
unknown validity: inert adsorbent, pure perfect gas, negligi- Function Symbol Eq.
ble volume of adsorbed phase, and so on. More disturbing is a
Pn
the impossibility of calculating or measuring the spreading Surface potential F y RT H0 df
pressure inside a micropore. The external surface area of f
nonporous solids can be measured by microscopy, but the Free energy of immersion DG imm F q PV s
definition and measurement of the surface area of nanome- F
Enthalpy of immersion D H imm yT 2 q PV s
ter-sized micropores is fuzzy. T T P
The topic of adsorption has been omitted from chemical F
Entropy of immersion D S imm y
engineering textbooks on thermodynamics since the first one T P
by Dodge 1944.. Phase equilibrium is properly concentrated f
upon vapor-liquid systems because of their importance in dis- Integral free energy DG a a
n RT ln qF
f8
tillation. Less attention is devoted to equilibria in liquid-liquid a

and liquid-solid systems and adsorption thermodynamics was Integral enthalpy DHa H0n D h adna
ignored entirely until 1996. The reason for this neglect of D H a y DG a
physical adsorption is that the fundamental equations of so- Integral entropy DSa
T
lution thermodynamics developed for vapor-liquid equilibria f
do not apply to adsorbed phases, which require special treat- Diff. free energy chem. potential . D g a RT ln
f8
ment and the introduction of surface variables P and A. as ln f
described in recent editions of Smith et al. 1996, 2001.. Diff. enthalpy D ha y RT 2
T na
D ha f
Diff. entropy D sa y R ln
T f8
Adantages of solution thermodynamics
U
The integral enthalpy and heat of immersion are related by D H a s
Innovations in adsorption technology such as pressure- D H imm q n ah R y PV s and therefore D H a sD H imm for a perfect gas at
swing adsorption require chemical engineers to perform mass low pressure. For a perfect gas, f s P and except for very high pressure
and energy balances and calculate phase equilibria for ad- the Poynting term PV s is negligible. All functions are normally negative
in sign.
sorption systems. The objective here is to show how the fa-
miliar principles of phase equilibria and solution thermody-
namics established for vapor-liquid equilibria can be applied thalpy of solution thermodynamics is smooth and well-de-
directly and rigorously to physical adsorption from gaseous fined under all conditions.
mixtures, while avoiding the undefined variables of 2-D sur- 4. The conventional treatment of thin films is for adsorp-
face thermodynamics. tion of perfect gases on inert, planar solids at low pressures
Since engineers and chemists have been measuring gas ad- of a few bars. The solution thermodynamics approach applies
sorption in porous adsorbents for many years using the to adsorption of fluid mixtures in porous materials at high
metholodogy developed for thin films, the case needs to be pressure up to 1,000 bars.
made that the solution thermodynamics approach is superior 5. The solution thermodynamics approach leads naturally
not just in the nuances of interest to thermodynamic purists, to the immersional functions: free energy, enthalpy, and en-
but also in everyday practical usage. Some of the advantages tropy of immersion of the clean adsorbent in the bulk fluid.
of solution thermodynamics over the conventional 2-D ap- The obvious physical significance of these functions simplifies
proach are: their application to the solution of practical problems. The
1. Confusion over different heats of adsorption see concepts of spreading pressure, differential entropy, and iso-
above. is eliminated by the solution thermodynamics ap- steric heat developed in the conventional 2-D approach are
proach. In solution thermodynamics, the enthalpy of the ad- difficult to understand, even for experts.
sorbed phase is measured relative to a well-defined reference 6. Experimental measurements of adsorption in porous
state and there is no confusion about different types of en- materials using volumetric or gravimetric methods require
thalpy. careful attention to the placement of the Gibbs dividing sur-
2. The 2-D approach is based upon surface area, which face between the gas phase and the solid adsorbent. Conven-
for microporous materials cannot be measured experimen- tional 2-D treatments of adsorption do not apply to porous
tally or calculated theoretically. The suitability of approxi- materials.
mate procedures such as the BET Point B method Young 7. Since the spreading pressure inside a micropore con-
and Crowell, 1962. for estimating surface area has been ar- taining a few dozen molecules cannot be calculated from in-
gued for generations. A better characterization of an adsorb- termolecular forces, theory cannot be compared with experi-
ent would be the maximum adsorption of a supercritical gas ment using the conventional 2-D approach. Solution thermo-
such as argon at room temperature, or the enthalpy of im- dynamics provide a simple connection between theoretical
mersion of the adsorbent in a particular fluid. absolute variables total energy, total number of adsorbate
3. The isosteric heat defined in conventional 2-D treat- molecules. and experimental excess variables.
ments of adsorption has a singularity goes to infinity. at the The thermodynamic functions for adsorption developed in
point where the amount adsorbed reaches a maximum Salem the next section are summarized in Table 1, and a sample
et al., 1998; Siperstein et al., 2001., whereas the integral en- calculation for the Langmuir model is given in the Appendix.

146 January 2002 Vol. 48, No. 1 AIChE Journal


The theoretical development section is followed by numerous The total amount of gas in the system n t . and its bulk den-
applications in adsorption technology. sity r g . are measured experimentally. The volume of the gas
phase V g . increases as the dividing surface in Figure 1b
Theory moves from left to right, while the amount adsorbed n a. de-
creases from a positive value to zero and finally becomes neg-
Defining adsorption in porous materials ative as the dividing surface enters the solid phase. Thus, the
An adsorption system contains two macroscopic phases: a amount adsorbed n a. depends upon the location x o . of the
gas phase and a solid phase. For typical adsorbents such as dividing surface.
zeolites or activated carbon, adsorption occurs inside the ad- The location of the dividing surface is fixed by selecting a
sorbent in micropores, mesopores, and macropores. Com- reference gas for which the amount adsorbed is defined to be
mercial adsorbents are manufactured not as single crystals, zero
but as small particles which are usually shaped into larger
particles using binders. Adsorption occurs mainly within the
n a s n t yV gr g s 0 for He . 4.
pores of the adsorbent, but may also occur on its external
surface and occasionally in the binder.
In order to divide the adsorbate molecules and their prop- It follows that
erties into two phases, it is necessary to distinguish adsorbed
molecules from gas-phase molecules. A methane molecule lo- nt
cated inside a micropore 4 A from its internal surface can V gs for He . 5.
reasonably be classified as adsorbed. A methane molecule lo-
rg
cated at a distance of 100 A from the surface of the adsor-
bent can reasonably be classified as belonging to the gas Equation 4 provides a method of measuring adsorption rela-
phase. Ambiguity arises for intermediate cases. Is a methane tie to helium by measuring the volume of the gas phase V g .
molecule located at a distance of 10 A from the surface of using helium as a reference component. V g , called the dead
the solid adsorbed? Fortunately, the problem of defining a space or void volume, is normally measured at the standard
boundary between the adsorbed and gas phases was solved conditions of room temperature and low sub-atmospheric .
by Gibbs 1928., whose solution was to propose the construc- pressure. The Gibbs procedure of defining adsorption rela-
tion of a mathematical dividing surface between the two tive to a reference component is independent of whether he-
phases. Gibbs was intentionally vague about the exact loca- lium actually adsorbs. Helium was selected as the reference
tion of this dividing surface, which is advantageous because it gas, because it is a small, inert molecule, but a smaller inert
would be difficult to define a dividing surface within the mi- molecule if one existed. would measure a slightly larger dead
cropores of an adsorbent. space than helium. The point is that all experiments and the-
Figure 1a shows the density profile close to the surface for oretical calculations should be based upon the same refer-
the case of single-component adsorption. The density of the ence state.
adsorbate r g . is constant in the gas phase and constant Referring again to Figure 1b, the gas volume V g . is re-
zero. within the solid. The amount adsorbed within the in- quired for volumetric experiments and the solid volume V s .
terface n a. obviously depends upon the location of its is required for gravimetric experiments. V g includes the pores
boundaries. Figure 1b illustrates the Gibbs model in which of the adsorbent, and V s is the skeletal volume of the ad-
the actual interface is replaced by a single dividing surface sorbent. The skeletal volume, which is the total volume of the
located somewhere within the interfacial region. The gas solid minus its pore volume, is measured by the buoyancy
phase with volume V g has a homogeneous density r g . up to force exerted in helium gas. The volume displaced by the solid
the dividing surface. The mass balance for adsorption is V s . is the slope on a plot of apparent weight vs. density of
helium. Thus, gravimetric and volumetric experiments are
n a s n t yV gr g 3. based upon the same assumption: that helium gas does not
adsorb at the standard conditions of room temperature and
low pressure.
Instead of defining adsorption by Eq. 3, why not measure
the total amount of gas contained in the micropores? Unfor-
tunately, an experimental method for measuring the total
amount of gas adsorbed in the pores does not exist. The stan-
dard volumetric technique measures the total amount of gas
in the system but cannot distinguish between molecules in
the pores and molecules in the gas phase. The gravimetric
method also fails to measure total adsorption in the pores
because of the necessity for a buoyancy correction to the ap-
parent weight.
Adsorption measured relative to helium by Eq. 3 is called
Gibbs excess adsorption. The excess adsorption is the total
amount of gas in the pores minus the amount that would be
present if the pores were filled with gas at the equilibrium
Figure 1. Profile of gas density at gas-solid interface. bulk density.

AIChE Journal January 2002 Vol. 48, No. 1 147


Thermodynamic properties of porous materials ficient of expansion of the solid, which is of order
According to the Gibbs model of adsorption depicted in 1rV . Vr T .P s10y5 to 10y3 Ky1. Changes in V s may be
Figure 1b, any extensive thermodynamic function may be significant when making gravimetric measurements over a
written wide range of temperature. Molecular simulations of the he-
lium pore volume of silicalite Siperstein et al., 2001. indicate
Z t s Z aq Z g q Z s 6. that the pore volumes at 300 and 400 K are 0.175 and 0.146
cm3rg, respectively. The decrease in helium pore volume is
The gas-phase portion is due to helium-solid intermolecular forces, not expansion or
contraction of the solid. Determinations of void volume
Z g sV gr g z g 7. should always be accompanied by the temperature of the
measurement. At high pressure in the range 100]1,000 bar,
where r g is the molar density of the bulk gas and z g is the the adsorption isotherm is highly sensitive to the void vol-
molar value of Z in the bulk phase at the same values of the ume.
intensive variables T, P, yi .. The molar function z g T, P,
yi . is determined from independent PVT measurements on
Fundamental equation for energy
the bulk gas. As explained in the previous section, V g is
determined from Eq. 5 using helium gas. Specifically, from Adsorption in microporous adsorbents falls naturally into
Eq. 7 the framework of solution thermodynamics, with the distinc-
tion that the solvent is a solid. The starting point is the fun-
n ig sV gr g yi damental differential equation for the energy of a micro-
U g sV gr g u g porous adsorbent containing C gaseous adsorbates Callen,
1985.
S g sV gr g s g
V g sV gr g g sV g 8. C
dUsTdSy PdV q mi dn i q m dm 10.
The solid-phase term Z s . in Eq. 6 is determined from inde- is 1
pendent measurements of the solid adsorbent in its pure
standard state. The intensive variables of the condensed phase are the tem-
From Eq. 6 perature T ., the pressure P ., the chemical potentials of the
adsorbates m i ., and the chemical potential of the solid ad-
n ai s n ti y n ig sorbent m .. The chemical potentials of the adsorbates are in
Jrmol and the chemical potential of the adsorbent is in Jrkg.
U a sU t yU g yU s
The extensive variables of the condensed phase are the inter-
S as S t y S g y S s nal energy U ., entropy S ., the amount number of mols. of
V a sV t yV g yV s s 0 9. each adsorbate n i ., and the mass of solid adsorbent m..
Equation 10 is written for the solid phase plus the adsorbed
a form which emphasizes the excess nature of the Gibbs phase, that is, the entire condensed phase.
model for which the adsorbed phase is a mathematical plane The particle-size distribution, composition, and structure
of zero volume V a s 0.. According to Eq. 9, the energy of of the individual adsorbent particles are assumed to be uni-
the adsorbed phase U a. includes the energy change of the form throughout the system. Thus, the external surface area
solid adsorbent relative to its standard state in addition to of the adsorbent and the extensive thermodynamic properties
the energy of the adsorbate molecules. A similar statement are directly proportional to the mass of adsorbent present.
applies to the entropy. Solution thermodynamics is unable to Stated mathematically
distinguish energy changes in the solid from energy changes
of the adsorbate molecules. U h S, h V , h n i , h m . sh U S, V , n i , m . 11.
Referring again to Figure 1b, the total volume is divided
into two parts: the gas phase V g . and the condensed phase
V s .. The condensed phase includes the adsorbed phase with The total energy U of the solid phase is a first-order, homo-
volume V a s 0 by definition. The Gibbs model ignores possi- geneous function of the variables S, V, n i , and m; that is to
ble changes in the volume of the condensed phase in re- say, doubling the values of S, V, n i , and m at constant values
sponse to increased pressure or because of swelling in the of the intensive variables T, P, m i , m . doubles the value of
case of polymeric adsorbents. Since the volumetric method is U. It follows from Eq. 11 and Eulers theorem for homoge-
based upon a constant value of V g and the gravimetric neous functions Callen, 1985. that
method is based upon a constant value of V s, deformation or
swelling of V s would be revealed by lack of agreement of C
volumetric and gravimetric isotherms. Experimental evidence UsTSy PV y mi n i q m m 12.
of such a discrepancy has not yet been found. is 1

Another consideration is the variation of V g and V s with


temperature. Although both volumes are functions of tem- Adsorption thermodynamics deals with systems such as a
perature, their temperature dependence is weak. V s in- packed column or sample cell containing a fixed mass of ad-
creases with temperature according to the temperature coef- sorbent. The solid phase is open with respect to the adsor-

148 January 2002 Vol. 48, No. 1 AIChE Journal


bates, but closed with respect to the adsorbent. The re- Substracting Eq. 19 from Eq. 18, and using Eq. 15
versible addition of solid adsorbent to the system correspond-
ing to the m dm term in Eq. 10 is not physically realizable, so C
dms 0 and the mass of adsorbent is a constant. The exten- dU a sTdS a q mi dnai 20.
sive variables U,S,V,n i . are converted to extensive variables iq1
per unit mass so that Eq. 12 may be written in the asymmet-
ric, but more useful, form Equations 16 and 20 are the integral and differential equa-
tions governing the adsorbed phase. Changes in the proper-
C
ties of the adsorbent from its standard-state values U S, S s .
UsTSy PV q mi n i q m 13.
are included implicitly in U a and S a. In surface thermody-
is 1
namics, the adsorbent is assumed to be inert so that its chem-
In Eq. 13 and, henceforth, all extensive variables U,S,V,n i , ical potential is unperturbed by isothermal adsorption m s
and so on. are written per unit mass of adsorbent and called m s .. Assuming the adsorbent to be inert is acceptable for ad-
mass extensie ariables. For the adsorbent in its pure stan- sorption on the external surface of solid particles, but, for
dard state at the equilibrium pressure and temperature microporous adsorbents, the concept of an inert adsorbent
must be abandoned. The surface potential F . in Eq. 16 van-
ishes only in the trivial case of no adsorption.
U s sTS s y PV s q m s 14.

The pure solid is assumed to be incompressible so that V s s Enthalpy and free energy
V sU , U s sU sU , and S s s S sU , where the asterisk refers to the The Legendre transformations for the auxiliary functions
clean adsorbent in acuo. However, the pressure affects the are Callen, 1985.
enthalpy and free energy so H s s H sU q PV s, G s sG sU q
PV s, and m s s m sU q PV s. The free energy G s . of the pure
H sUq PV
adsorbent is equal to its chemical potential m s .; the units of
both are Jrkg. The PV s term is a Poynting correction which F sUyTS
accounts for hydrostatic pressure de Azevedo et al., 1999..
Gs F q PV s H yTS 21.
Given the equilibrium values of T and P, any mass exten-
sive function of the adsorbed phase is obtained by subtract-
ing the function for the pure solid from the corresponding where H is enthalpy, F is Helmholtz free energy, and G is
function for the condensed phase Gibbs free energy. From Eq 16.

U a sUyU s U a sTS a q m i n ai q F
S a s Sy S s
H a sTS a q m i n ai q F
V a sV yV s s 0
F a s m i n ai q F
n ai s n i 15.
G a s m i n ai q F 22.
Subtracting Eq. 14 from Eq. 13, and using Eq. 15

C In Eq. 22 and, henceforth, summation over the C adsorbates


a
U sTS q a
m i n ai q F 16. is implicit. The enthalpy is equal to the internal energy, be-
is1 cause the volume of the adsorbed phase is zero. The
Helmholtz and Gibbs free energies are equal for the same
where reason. The free energy of the adsorbed phase consists of
two parts: the free energy of n ai mols of each species in the
F s m y ms . 17. equilibrium gaseous state plus an additional term, the surface
potential F ., which is zero if no adsorption takes place.
is called the surface potential Sircar and Myers, 1973a. or This is the point at which surface thermodynamics diverges
grand potential. If no adsorption occurs, then m s ms and the from solution thermodynamics. Surface thermodynamics,
surface potential is zero. Writing Eq. 10 for the condensed which is the standard approach adopted in monographs on
system containing a constant mass m of adsorbent adsorption Ruthven, 1984; Suzuki, 1990; Yang, 1987., as-
sumes an inert adsorbent with chemical potential which is
C independent of loading. It is difficult at first to accept the
dUsTdSq mi dn i 18. fact that the chemical potential of a porous adsorbent varies
is 1
with the isothermal loading of the adsorbate molecules. In
molecular simulation, it is generally assumed that the adsorb-
The PdV term vanishes because V sV s is assumed constant. ent generates an invariant potential field. However, the exis-
For the same mass of adsorbent in its standard state tence of a potential field is not inconsistent with changes in
free energy, which contain both energetic and entropic con-
dU s sTdS s 19. tributions. From the perspective of solution thermodynamics,

AIChE Journal January 2002 Vol. 48, No. 1 149


the intimate contact of the adsorbate molecules with the tion isotherm
atoms of the porous material alters the chemical potential of
the solid. Consider the immersion of clean adsorbent in a dF syn ai d m i constant T . 28.
compressed gas held at constant pressure and temperature.
Since the variables pressure, temperature, and chemical po-
tential of the compressed gas are fixed during the immersion Replacing chemical potential by fugacity
process, the only intensive variable capable of change is the
chemical potential of the adsorbent. dF sy RT n ai d ln f i constant T . 29.
Surface thermodynamics is valid when the surface area can
be measured experimentally. For example, the external sur- Since surface potential is a state function, the integration for
face area of nonporous carbon black can be measured by mi- F is independent of the path. The thermodynamic consist-
croscopy. In this case, the model of an inert adsorbent with ency of experimental data may be tested by comparing values
adsorption occurring in a 2-D film on its external surface is of F obtained for different paths Myers and Sircar, 1972..
realistic. However, the concepts of surface area and spread- Integrating for pure-component adsorption from the unad-
ing pressure lose their physical meaning in porous adsorbents sorbed state at zero pressure where m s m s s m sU and F s 0
such as zeolites.
From Eq. 22
P na
F sy RT HP s 0 df constant T . 30.
H a sU a f
23.

and For a perfect gas

G a s F a sU a yTS a 24. P na a d ln P
F sy RT H0 P
dP sy RT H0n d ln n a
dn a constant T .
From Eqs. 20, 23, and 24, it follows that 31.

dU a sTdS a q m i dn ai
In the solution thermodynamics approach, the chemical po-
a
dH sTdS a
q m i dn ai tential of the nonvolatile adsorbent is determined indirectly
by integrating the chemical potential of the gas. An analo-
dF a sy S adT q m i dn ai gous integration is used to calculate the chemical potential of
a strong electrolyte in aqueous solution from the change in
dG a sy S adT q m i dn ai 25. chemical potential of the water vapor de Azevedo et al.,
1999..
The surface potential F . term in Eq. 22 for the integral The concept of the adsorbent as a solvent suggests that its
functions does not appear explicitly in the differential equa- chemical potential at the limit of zero loading should be given
tions for the adsorbed phase. by Raoults law. At the limit of zero pressure, the slope of the
Equations 22 and 25 are the basic integral and differential adsorption isotherm is given by Henrys law: n a s KP. Since
equations for the adsorbed phase. Any extensive property of the adsorbed gas obeys Henrys law, the solvent the solid
the condensed phase is the sum of the property in the ad- adsorbent . follows Raoults law over the same dilute range
sorbed phase H a, S a, and so on. plus the corresponding
property of the solid adsorbent in the absence of adsorption RT RT
H s, S s, and so on. as described by Eqs. 14 and 19. F s m y mss ln x s s ln 1y x a .
M M
RT
Surface potential fy x a fy n aRT 32.
M
The surface potential, which is the chemical potential of
the solid adsorbent relative to its pure standard state, is ob-
where x s is the mol fraction of solid adsorbent, x a is the mol
tained from Eq. 22
fraction of adsorbate, and n a is mols of gas per unit mass of
adsorbent. The unknown. molecular weight of the adsorbent
F sG a y m i n ai 26. M . cancels in the result. At the limit of zero pressure, the
surface potential decreases in direct proportion to the amount
Differentiating and substituting Eq. 25 for dG a gives adsorbed. The limiting relation F sy n aRT may be com-
pared to the corresponding equation for a 2-D perfect gas
from surface thermodynamics: P As n aRT. Thermodynam-
dF sy S adT yn ai d m i 27. ics is indifferent to the name of the property the P A prod-
uct or the chemical potential of the adsorbent ., but the physics
This equation is analogous to the Gibbs-Duhem equation for of adsorption in micropores is best described within the
a liquid mixture with the yVdP term replaced by dF. At framework of solution thermodynamics because spreading
constant temperature, Eq. 27 reduces to the Gibbs adsorp- pressure in a micropore is undefined.

150 January 2002 Vol. 48, No. 1 AIChE Journal


Degrees of freedom and independent ariables Since equality of chemical potentials m ig s m ia corresponds to
The number of degrees of freedom from the Gibbs phase equilibrium dG t s 0., the use of the notation m i for the
rule is chemical potential in either phase is justified.

F sC q2y P 33. Determination of oid olume


The measurement of Gibbs excess variables according to
where C is the number of chemical components present and Eq. 9 requires prior determination of the void volume dead
P is the number of phases. Since one component is the ad- space. of the apparatus using Eq. 5
sorbent
nt
F sC q3y P sC q3y2 sC q1 34 . V gs for He . 42.
rg

if C is the number of adsorbates present. For example, for The standard procedure is to measure the void volume using
binary adsorption, there are three degrees of freedom. The pure helium gas at low pressure and room temperature T8.
natural independent variables for the integral functions are by assuming that helium does not adsorb. n t is the total num-
T, P, y 1; the dependent variables are then n ai , D H a, D S a, F, ber of mols of helium admitted to the sample cell per unit
and so on. The natural independent variables for the differ- mass of adsorbent; V g is the specific void volume of the ap-
ential functions are T, n1a, n 2a ; the dependent variables are paratus measured in cm3 per unit mass of adsorbent; r g is
then P, y 1, and so on. the helium density. At low pressure, use of the perfect-gas
law in Eq. 42 gives
Equality of chemical potentials at equilibrium
Up to this point it has been assumed without proof that n t RT8
the chemical potential in the adsorbed phase m a. is equal to V gs for He . 43.
P
the chemical potential in the gas phase m g .. Here, the as-
sumed equality of chemical potentials is verified by minimiz- The experimental determination of dead space is based on
ing the Gibbs free energy of the total system at constant tem- Eq. 43. The theoretical value is calculated from statistical
perature and pressure. From Eq. 6 thermodynamics, which at the limit of low pressure gives for
the excess adsorption
G t sG a qG g qG s 35.
BP
nas 44.
Since these are extensive functions per unit mass of adsorb- RT8
ent, G s is the chemical potential of the adsorbent in its stan-
dard state where B is called the adsorption second virial coefficient
Steele, 1974.
dG s s d m s sy S sdT qV sdP 36.
1 yE r .rk T 8

Differentiating Eq. 35
Bs
m
Hw e y1 x dr 45.

E is the gas-solid potential energy of a single molecule, and


dG t s dG a q dG g q dG s 37. m is the mass of a representative sample of solid adsorbent.
The integration is performed over the gas phase including
From Eq. 25 the pore volume. In order for theory to mimic experiment,
the second virial coefficient for helium must be zero so that
dG a sy S adT q m iadn ai 38.
1 yE r .rk T 8
V gs He dr for He. 46.
For the bulk gas phase Sandler, 1998. m

dG g sy S g dT qV g dP q m ig dn ig 39. E is the gas-solid potential energy of a helium atom. The


integration in Eq. 46 is performed over the entire condensed
phase as well as the gas phase, since the exponential vanishes
Minimizing G t at constant T and P under the constraint
within the solid where where f `.

n ti s n ai q n ig s constant 40. Integral and differential properties of the adsorbed phase


gives The integral functions free energy, enthalpy, and entropy,
which arise naturally from solution thermodynamics, are
missing from the traditional surface thermodynamics ap-
dG t s m iadn ai q m ig dn ig s m ig y m ia . dn ig s 0 41. proach based upon spreading pressure, 2-D films, isosteric

AIChE Journal January 2002 Vol. 48, No. 1 151


heat, and other differential quantities. A special feature of At low pressure, the Poynting correction PV s . for the en-
the integral functions is that their natural independent vari- thalpy and free energy may be ignored. Thus, disregarding
ables temperature, pressure, and gas-phase composition. are the PV s . nuisance term, the surface potential is equal to the
controllable experimentally. The integral functions are free energy of immersion. The free energy of immersion is
needed for engineering calculations and are useful for char- negative because adsorption is spontaneous. The enthalpy of
acterization of adsorbents, as shown below. The integral immersion D H imm . may be measured directly by calorimetry
functions for the adsorbed phase are defined relative to the see below. or indirectly by differentiating the surface poten-
perfect-gas reference state at the same temperature tial according to Eq. 50. Since the heat of immersion is
exothermic, D H imm is negative in sign. Since the adsorption
DG a sG a yn ai m i8s D H a yTD S a process is associated with a loss of entropy, the entropy of
immersion D S imm . is also a negative quantity. The free en-
D H a s H a yn ai h i8 ergy and enthalpy of immersion in a pure liquid can be used
D S a s S a yn ai si8 to predict adsorption from liquid mixtures Sircar and Myers,
47.
1973b..
The surface potential F . and associated immersional
The quantities mTi , hTi , and sTi refer to the molar values in the functions DG imm ,D H imm ,D S imm . are more closely related to
perfect-gas reference state. The integral free energy and en- the adsorption process than the integral functions DG a,
thalpy DG a, D H a. are measured in joules per kilogram of D H a, D S a. in Eq. 48, which contain an additional term for
adsorbent. Substituting for G a in Eq. 47 from Eq. 22 isothermal compression of the bulk gas

DG a sn ai m ig q mTi . q F Pyi
DG comp sn ai m ig y mTi . s RT n ai ln qn ai g iR
PT
F
D H a sn ai h ig y hTi . yT 2 D H comp sn ai h ig y hTi . sn ai h iR
T T P, yi
Pyi
F D S comp sn ai sig y sTi . sy Rn ai ln qn ai siR 51 .
DS a
sn ai sig y sTi .y 48. PT
T P, yi
The compression terms for the bulk gas are obtained from
The overline notation for partial molar variables in the bulk partial pressures Pyi . and from residual functions g iR,h iR, siR .
gas phase h ig , sig . is omitted for the chemical potential m ig . Smith et al., 2001. which vanish for a perfect gas.
since its partial molar character is understood. The expres- The integral and immersional functions derived from solu-
sion for D S a is obtained by combining Eq. 47 with the partial tion thermodynamics are related to the molar and differen-
differential of Eq. 27 with respect to temperature; then D H a tial functions generated by 2-D surface thermodynamics. The
s DG a qTD S a. integral functions in Eq. 48 are converted to molar variables
The equations for DG a, D H a, and D S a contain two parts: by dividing each function by the total amount adsorbed n at s
1. changes for isothermal compression of the gaseous ad- i n ai
sorbates from their perfect-gas reference states to the equi- DG a
librium pressure; 2. changes for isothermal, isobaric adsorp- D g as s D h a yTD s a
tion F and its derivatives.. Consider isothermal immersion n at
of clean, evacuated adsorbent into the compressed gas; the DHa
free energy of immersion is D has
n at

DG imm s G a yn ai m ig . q G s yG sU . 49. DSa


D s as 52.
n at
DG imm is the change in the free energy of the condensed
phase caused by adsorption, measured relative to compressed The molar integral functions D g a and D h a have units of
gas and evacuated adsorbent. The first term on the righthand Jrmol.
side of Eq. 49 is the free energy of adsorbed gas relative to The differential functions for component i in a multicom-
bulk compressed gas; the second term is the free energy of ponent mixture are obtained from the integral functions. Dif-
the adsorbent in its standard state T, P . relative to the evac- ferentiation of the functions DG a, D H a, and D S a in Eq. 47
uated state T,in acuo.. Combining Eqs. 47]49 and observ- gives
ing that G s yG sU . s PV s
DG a fi
D g ia s s m ia y mTi s RT ln
n ai T , n aj fT
DG imm s F q PV s i

F D H a ln f i
D H imm syT 2 q PV s D h ai s s h ai y hTi sy RT 2
T T P, yi
n ai T , n aj T n ai , n aj

F D S a fi
DS imm
sy 50. D sia s s sia y sTi sy RT ln 53.
T P, yi
n ai T , n aj T fT
i n ai , n aj

152 January 2002 Vol. 48, No. 1 AIChE Journal


The relations s G ar n ai .T, n ai s m ia and S ar n ai .T , n aj s integrating the differential enthalpies; for example, from Eq.
y m iar T . n ai , n aj necessary for the derivation of Eq. 53 were 53 for a binary mixture
obtained from the total differential for G a in Eq. 25. The
independent variables for the differential functions are tem-
d D H a . s D h1adn1a q D h 2a dn 2a const. T . 57 .
perature and mol numbers n ai .. f i is fugacity and fT i is the
standard-state fugacity 1 bar.. The overline notation in these
functions such as D h ai . is used to distinguish differential Since D H a is a state function, the integral is independent of
functions from molar integral functions D h a.; the differen- the path and the differential enthalpies are related by
tial character of the chemical potential is implicit. The differ- Maxwell-type equations
ential functions D g ia and D h ai have units of Jrmol and, like
the integral and molar integral functions, are based on the D h1a D h 2a
perfect-gas reference state. For a pure bulk fluid, the partial s 58.
n 2a T , n 1a
n1a T , n 2a
molar and molar functions are equal. The molar and differ-
ential functions for the adsorbed phase are unequal, even for
a pure adsorbate. The differential mixture enthalpies whose absolute values are
The relationships in Eq. 53 are similar to those for partial called isosteric heats in the literature of adsorption. are func-
molar quantities in a bulk fluid. However, for a bulk fluid, tions of loading and can be measured experimentally or pre-
the partial molar derivatives are at constant pressure; for an dicted from single-gas adsorption data Siperstein and Myers,
adsorbed fluid, the partial derivatives are at constant loading. 2001.. For pure-component adsorption, the integral is
A sample calculation of the integral, immersional, molar
integral, and differential functions summarized in Table 1 is a

provided in Appendix A. The immersional functions of Eq. D H as H0n D h dn a a


const. T . 59 .
50 are more closely related to the adsorption process than
the other functions, which contain additional terms for
isothermal compression of the gas. The differential enthalpy D h a is a function of loading n a.,
but its variation with temperature is weak and the assump-
tion of its constancy over a moderate range of temperature is
Differential enthalpy and isosteric heat a useful approximation. For pure-component adsorption of a
Confusion about the meaning of isosteric heat is perfect gas, Eq. 56 gives for constant D h a
widespread in the adsorption literature. For example, Eqs.
29]46 in the revised version of Lewis et al.s 1961. textbook P D ha 1 1
gives the following relation between the differential entropy ln T s y constant n a . 60.
and the isosteric heat q st . P R T TT

q st f which provides the function P T . given a reference point


D s as y R ln T 54. PT TT . at the same loading n a..
T f

Equation 54 has a sign error: the first term should be


Heat capacity
y q strT . cf. Table 1.. Errors like this one could be avoided A useful and frequently used approximation is that the dif-
by replacing the ill-defined isosteric heat with the differential ferential D h a. and integral D H a. enthalpies of the ad-
enthalpy of adsorption. Differential enthalpy is not a heat of sorbed phase are independent of temperature, at least over
adsorption, the value of which would depend upon the path. some modest interval of temperature. With this approxima-
Differential enthalpy is a state function which can be mea- tion, the heat capacity at constant loading from Eq. 70 sim-
sured either by calorimetry see below. or by differentiating a plifies to
series of adsorption isotherms at constant loading
Ha
ln f i C pa s s n ai cTp . i 61.
T n ai
D h ai sy RT 2 55. i
T n ai , n aj

so that the heat capacity of the adsorbate is equal to its heat


For a pure, perfect gas capacity in the perfect-gas state. If in addition the gas phase
is ideal, the heat capacity of the system follows from Eq. 6
ln P
D h a sy RT 2 56. C pt sC pg qC pa qC ps s n ig cTp . i q n ai cTp . i
T na
i i

Except for the minus sign, this equation is identical to Eq. 1. qC psU s nti cTp . i qC psU 62.
The isosteric heat is a positive quantity by definition, but the i
differential enthalpy of adsorption is negative exothermic..
The integral enthalpy can be calculated by differentiating The heat capacity of the adsorbent in its pure standard state
the surface potential F T, P, yi . according to Eq. 48, or by C ps . is equal to its heat capacity in acuo C psU .. Thus, the

AIChE Journal January 2002 Vol. 48, No. 1 153


heat capacity of the entire system condensed phase and gas Adsorption of gas mixtures
phase. may be estimated from the heat capacity of the evacu- Equations 48 and 50 apply to multicomponent adsorption
ated solid adsorbent and the ideal-gas heat capacities of the of an imperfect gas, but the integration for the surface poten-
adsorbates Sircar, 1991.. tial term according to Eq. 29 requires experimental, isother-
mal mixture data for the loading as a function of fugacity.
Equation of state Similarly, the integration for the integral enthalpy by Eq. 57
The intensive variables pressure, temperature, and gas- requires differential enthalpies for adsorbed mixtures. Exper-
phase composition determine the gas-phase properties den- imental mixture isotherms and enthalpies are seldom avail-
sity, molar enthalpy, chemical potential, and so on., the able so reliable methods of predicting mixture data from sin-
solid-phase properties in acuo, and the adsorbed-phase gle-gas adsorption isotherms are essential Siperstein and
properties from Eq. 9. The equation of state for the adsorbed Myers, 2001..
phase is the surface potential F T, P, yi .. From Eq. 31, the
adsorption isotherm for a pure, perfect gas is Adsorption at high pressure
At high pressure, according to Eq. 3, the total amount of
a
P F gas contained in the micropores n t . tends toward a limit
n sy 63.
RT P T called the saturation capacity, while the density in the gas
phase r g . increases without limit. Eventually, when the
The integral enthalpy from Eq. 48 for a pure, perfect gas is density in the gas phase increases with pressure at the same
t p.
rate as the absolute density in the pores nrV , the excess
a
F adsorption n . passes through a maximum and then begins
D H a syT 2 64. to decline with pressure. At very high pressure, when the ab-
T T P solute density in the pores is equal to the bulk-gas density,
the excess adsorption is zero. Although adsorption loses its
The surface potential of an adsorbent in a pure liquid deter- potential for storage and separation applications under these
mines the selectivity for adsorption from liquid mixtures conditions, high-pressure adsorption has been studied exten-
Sircar and Myers, 1973b.: the higher the value of < F < in a
sively Benard and Chahine, 2001..
pure component, the greater its preferential adsorption from Once the adsorption isotherm has passed through a maxi-
liquid mixtures. The surface potential is also required for cal- mum, the amount adsorbed becomes an invalid independent
culations of mixed-gas adsorption Siperstein and Myers, variable because it is not single-valued. The differential en-
2001.. thalpy isosteric heat. and differential entropy have no mean-
Example. The equation of state at the limit of zero pres- ing under these conditions. However, the integral functions
sure is given by Eq. 32: F sy n aRT. Using Eq. 63 DG a, D H a, D S a., for which the independent variable is the
pressure, are well defined even when the excess amount ad-
na na sorbed passes through a maximum and declines to negative
lim s 65.
P0 P P T
values.

which is true if and only if n a s K T . P. Thus, the equation Enthalpy and entropy balances
of state predicts Henrys law for the adsorption isotherm. The solution thermodynamics approach to adsorption gives
From Eqs. 59 and 64 the thermodynamic properties of the entire system as the sum
over the gas, adsorbed, and solid phases. The most important
ln n a D ha property is the enthalpy for energy balances; calculations of
lim s . 66. lost work and efficiency are based upon entropy balances.
P0 T P RT 2
The enthalpy and entropy functions for the entire system are
given by Eq. 6
Equation 66 for calculating the zero-pressure limit of the dif-
ferential enthalpy from the temperature coefficient of ad-
sorption should be more accurate than the usual procedure H t s H g q H aq H s
of extrapolating Eq. 56 or differentiating Henry constants
S t s S g q S aq S s 68.
Valenzuela and Myers, 1989.

d ln K Gas Phase. The thermodynamic properties for the gas


lim D h a s RT 2 . 67. phase are conveniently calculated from perfect-gas heat ca-
P0 dT
pacities and residual functions Smith et al., 2001.

Applications T
H g s n ig hTi q HT 8 cT R
p i dT q h i
.
Characterization of adsorbents i
Table 1 summarizes the integral and differential thermody- dT P
T
namic equations needed for characterization of adsorbents S g s n ig sTi qHT 8 c T
p i . y R ln q siR 69 .
and for engineering calculations. i T P8

154 January 2002 Vol. 48, No. 1 AIChE Journal


h8 and s8 are molar enthalpy and molar entropy, respectively, so
at the reference temperature T8. and standard pressure P8..
cTp is the ideal-gas heat capacity. h R and s R are the residual dn g sy dn a 75.
enthalpy and entropy, respectively, at temperature T ; the
residual functions are zero for a perfect gas. As usual, the Combining Eqs. 23, 47, 53, 73 and 75 for the case of single-
mass extensive variables H g , S g, n g . refer to a unit mass of component adsorption
solid adsorbent.
Adsorbed Phase. Similarly, the thermodynamic properties
dQ dU g
for the adsorbed phase are calculated from ideal-gas heat ca- D has q y h8 76.
a
pacities and the integral functions of Eq. 48 dn dn g

T In general, the derivative dU grdn g . depends upon the equa-


H a s n ai h i8q HT 8 c p8 . i dT qD H a tion of state of the bulk gas, but for a perfect gas U g s n g u8
i
and
T dT
S a s n ai si8q HT 8 c 8. p i
qDSa 70. dU g
i T s u8s h8y RT 77.
dn g
Solid Adsorbent. The thermodynamic properties of the
solid adsorbent in its standard state at the equilibrium tem- For a perfect gas, Eq. 76 simplifies to
perature and pressure are
dQ
D has y RT 78.
T U dn a
H s s H sU q HT 8 C ps . dT q PV s
Since adsorption is exothermic, the differential heat dQrdn a
T U dT is negative.
S s s S sU q C ps
HT 8 . 71.
T The measurement of the differential enthalpy of adsorp-
tion from mass and energy balances applied to a calorimeter
where the asterisk refers to the properties of the adsorbent is straightforward compared to derivations for the measure-
per unit mass in acuo. Equation 71 is not based on the as- ment of the ill-defined isosteric heat Young and Crowell,
sumption that the solid adsorbent is inert; isothermal changes 1962; Siperstein et al., 1999..
in the enthalpy or entropy of the solid adsorbent induced by
adsorption are included in the D H a and D S a functions for Molecular simulation of adsorption
the adsorbed phase.
The grand canonical partition function McQuarrie, 1976.
used in grand canonical Monte Carlo GCMC. simulations is
Calorimetry
The differential enthalpy of adsorption defined in Eq. 53 Js eyE j N ,V .rk T e Nmrk T 79.
can be measured by calorimetry. An isothermal batch j, N
calorimeter consists of a dosing cell and a sample cell con-
nected through a valve. When the valve is opened, an incre- The independent variables are volume V . temperature T .,
ment of gas expands from the dosing cell into the sample cell and chemical potential of the adsorbate molecules m .. The
and a portion of the increment adsorbs. The total energy of grand potential G sUyTSy m N is related to the partition
the calorimeter is function by

U t sU g qU a qU s qU cell 72. G sy kT ln J V , T , m . 80 .

U cell T . includes the walls of the sample and dosing cells and For bulk fluids, G sy PV. For 2-D surface thermodynamics,
the valve. U s T . is the energy of the adsorbent in its standard G syP A, where P is spreading pressure and A is surface
state. According to the first law for a closed system area. Within the framework of solution thermodynamics, the
grand potential of an adsorbed phase is equal to its surface
dU t s dU g q dU a s dQ 73. potential, G s F.
The volume is a simulation box containing a representive
sample of the microporous adsorbent. The dependent vari-
Equation 73 and the following equations are for constant
ables are the specific potential energy f . in Jrkg and the
temperature. The differentials dU s and dU cell vanish under
specific total absolute. amount of adsorbate in the pores n t .
isothermal conditions and no work is done on the composite
in molrkg. These total variables must be converted to excess
system. The mass balance is
functions, especially the excess amount adsorbed n a. and the
integral enthalpy D H a., for comparison with experiment. For
n t s n g q n a s const. 74. single-gas adsorption, from Eqs. 8 and 9

AIChE Journal January 2002 Vol. 48, No. 1 155


n a s n t yV gr g 81. wise arbitrary, loading for the standard state in the adsorbed
phase
and
D g a sy RT ln K qconst. 88 .
U a sU t yV gr g u g yU s 82.
Comparison of two adsorbates in identical standard states
V g is calculated from the adsorption second virial coefficient
gives
of helium, Eq. 46. The potential energy f . is the total en-
ergy of gas-gas and gas-solid interactions relative to the per- K1
fect-gas reference state and the solid adsorbent in acuo D g 1a y D g 2a sy RT ln 89.
K2
f sU t y n t u8yU s 83.
The comparison in Eq. 89 is independent of the standard
A combination of Eqs. 81]83 with Eqs. 23 and 47 gives state chosen for the adsorbed state. This equation was used
recently Schenk et al., 2001. to compare the Gibbs free en-
D H a s f q PV g y n t RT yV gr g h R 84. ergies of formation of various alkane molecules in zeolites
using Monte Carlo calculations. For example, the free energy
of formation of 3,3,5-trimethylheptane relative to n-decane
where h R is the residual enthalpy in the bulk gas phase Smith
at 415 K is 33 kJrmol in MFI, but near zero in FAU or in the
et al., 2001.. For a perfect gas, h R s 0.
gas phase. Thus, MFI strongly favors the formation of n-de-
Equations 81 and 84 are key equations for converting simu-
cane relative to 3,3,5-trimethylpentane, because the shape of
lation variables to experimental variables. The correction term
n-decane is commensurate with the pore shape Schenk et
containing the pore volume V g . is sometimes negligible at
al., 2001..
low temperature and low pressure, but dominates at very high
pressure and causes the excess amount adsorbed n a. to pass
through a maximum. The density factor r g . in the correc- Conclusions
tion term of Eq. 81 suggests that the difference between ab- Adsorption in microporous adsorbents can be treated as a
solute and excess functions is zero at low pressure since particular case of solution thermodynamics in which the sol-
lim P 0 r g s PrRT. Actually, the Henrys constants and the vent is a solid adsorbent. Spreading pressure and surface area
zero-pressure differential enthalpies differ for absolute and variables is unnecessary.
excess adsorption Talu and Myers, 2001. and the magnitude An obvious difference between adsorbed and bulk fluids is
of the difference increases with temperature. the relative importance of the interfacial region. In vapor-
The absolute differential energy w fr n t xT may be calcu- liquid equilibrium, two macroscopic phases are separated by
lated directly from fluctuations in f and n t Nicholson and a gas-liquid interfacial region which is too small to influence
Parsonage, 1982., but there is no simple relationship between the thermodynamic properties of either phase. In adsorption
this simulation variable and the differential enthalpy D h a of equilibrium, two macroscopic phases gas and solid adsorb-
experiment. The most direct conversion is the definition D h a ent. are separated by a gas-solid interfacial region, which may
s D H ar n a.T using excess variables from Eqs. 81 and 84. contain more molecules than the bulk gas phase.
A subtle difference between adsorbed and bulk fluids is
Shape selectiity in catalysis observed in the differential variables. For a pure bulk fluid,
the partial molar variables are identical to the molar quanti-
The Henry constant K . defined by
ties, for example, the partial molar enthalpy Hr n.T, P is
equal to the molar enthalpy Hrn.. For adsorbed fluids, the
na
K s lim 85 . differentials are not partial molar variables because the pres-
P0 P sure is not held constant for the differentiation. Conse-
quently, for a pure adsorbed fluid see Figure 4., the differ-
can be related to free energy by combining Eqs. 48 and 51 ential entropy D s . is not equal to the molar entropy D Srn..

na
DG a s n aRT ln q F. 86. Acknowledgment
KP8
Financial support by National Science Foundation Grant NSF
CTS-0080915 is gratefully acknowledged.
Using Eq. 32
Notation
DG a na
D g as
na
s RT ln
KP8 /
y1 . 87 . Asspecific surface area, m2rkg
Asconstant in Eq. A2, Jr mol ? K.
Bsadsorption second virial coefficient, cm3rg
Bsconstant in Eq. A2, Jrmol
Since the limit at infinite dilution n a 0. is D g a sy`, a cTp sheat capacity of perfect gas, Jr mol ? K.
standard state must be chosen. For example, for Ar on sili- C sconst. in Eq. A2, Pay1
calite at 32.68C, the Henry constant K s 0.00173 molr kg C p sspecific heat capacity, Jrkg ? K.
kPa. Dunne et al., 1996.. For P8s100 kPa and n a s 0.001 C ps .U sheat capacity of solid adsorbent in acuo, Jrkg ? K.
molrkg, D g a sy15.64 kJrmol. Selecting a small, but other- Esenergy, J

156 January 2002 Vol. 48, No. 1 AIChE Journal


f sfugacity, Pa Literature Cited
F sspecific Helmholtz free energy, Jrkg
Gsspecific Gibbs free energy, Jrkg

Benard, P., and R. Chahine, Determination of the Adsorption
Isotherms of Hydrogen on Activated Carbons Above the Critical
DG a sintegral free energy of adsorbed phase, Jrkg Temperature of the Adsorbate over Wide Temperature and Pres-
G sgrand potential UyTSy m N ., Jrkg sure Ranges, 17, Langmuir, ASAP article 2001..
g smolar Gibbs free energy, Jrmol Callen, H. B., Thermodynamics and an Introduction to Thermostatis-
g iR sresidual Gibbs free energy of ith gaseous component, Jrmol tics, 2nd ed., Wiley, New York pp. 35,59,137 1985..
D g a smolar integral free energy of adsorbed phase, Jrmol de Azevedo, E. G., J. M. Prausnitz, and R. N. Lichtenthaler, Molecu-
D g a sdifferential free energy in adsorbed phase, Jrmol lar Thermodynamics of Fluid-Phase Equilibria, 3rd ed., Prentice-
hsmolar enthalpy, Jrmol Hall, Upper Saddle River, NJ, pp. 16,511 1999..
h8smolar enthalpy of perfect gas, Jrmol Dodge, B. F., Chemical Engineering Thermodynamics, McGraw-Hill,
h g spartial molar enthalpy in gas phase, Jrmol New York 1944..
h iR sresidual enthalpy of ith component in gas phase, Jrmol Dunne, J. A., R. Mariwala, M. Rao, S. Sircar, R. J. Gorte, and A. L.
D h a smolar integral enthalpy of adsorbed phase, Jrmol Myers, Calorimetric Heats of Adsorption and Adsorption
D h a sdifferential enthalpy in adsorbed phase, Jrmol Isotherms: I. O 2 , N2 , Ar, CO 2 , CH 4 , C 2 H 6 , and SF6 on Silicalite,
H sspecific enthalpy, Jrkg Langmuir, 12, 5888 1996..
D H a sintegral enthalpy of adsorbed phase, Jrkg Gibbs, J. W., Collected Works of J.W. Gibbs, Longmans and Green,
k sBoltzmann constant, 1.3806=10y2 3 JrK New York 1928..
K sHenry constant, molr kg ? Pa. Lewis, G. N., and M. Randall, Thermodynamics and the Free Energy of
msmass of adsorbent, kg Chemical Substances, McGraw-Hill, New York, p. 247 1923..
M smolecular weight, kgrmol Lewis, G. N., M. Randall, K. S. Pitzer, and L. Brewer, Thermodynam-
Nsnumber of molecules ics, Second ed., Chap. 29, McGraw-Hill, New York 1961..
nsspecific amount adsorbed, molrkg McQuarrie, D. A., Statistical Mechanics, Harper & Row, New York,
P spressure, Pa p. 58 1976..
P8sstandard pressure, 1 bar Myers, A. L., and F. Siperstein, Characterization of Adsorbents by
Energy Profile of Adsorbed Molecules, Colloids and Surfaces A,
q st sisosteric heat, Jrmol
187188, 73 2001..
Qsspecific heat absorbed by system, Jrkg
Myers, A. L., and S. Sircar, A Thermodynamic Consistency Test for
Rsgas constant, 8.3145 Jr mol ? K.
Adsorption of Liquids and Vapors on Solids, J. Phys. Chem., 76,
ssmolar entropy, Jr mol ? K. 3412 1972..
s8smolar entropy of perfect gas, Jrmol ? K. Nicholson, D., and N. G. Parsonage, Computer Simulation and the
s g spartial molar entropy in gas phase, Jrmol ? K. Statistical Mechanics of Adsorption, Academic Press, London, p. 94
siR sresidual entropy of ith component in gas phase, Jrmol ? K. 1982..
D s a smolar integral entropy of adsorbed phase, Jrmol ? K. Ruthven, D. M., Principles of Adsorption and Adsorption Processes,
D s a sdifferential entropy in adsorbed phase, Jrmol ? K. Wiley, New York 1984..
Ssspecific entropy, Jrkg ? K. Salem, M. M. K., P. Braeuer, M. v. Szombathely, H. Heuchel, P.
D S a sintegral entropy of adsorbed phase, Jr kg ? K. Harting, and K. Quitzsch, Thermodynamics of High-Pressure Ad-
T stemperature, K sorption of Argon, Nitrogen, and Methane on Microporous Adsor-
T8sroom temperature; reference temperature, K bents, 14, 3376 1998..
usmolar internal energy, Jrmol Sandler, S. I., Chemical and Engineering Thermodynamics, Third ed.,
Usspecific internal energy, Jrkg Wiley, New York 1998..
smolar volume, m3rmol Schenk, M., B. Smit, T. J. H. Vlugt, and T. L. M. Maesen, Shape
V sspecific volume, m3rkg Selectivity in Alkane Hydroconversion, Angew. Chem. Int. Ed., 40,
V g sspecific void volume of system, m3rkg 736 2001..
V s sspecific volume of solid adsorbent, m3rkg Siperstein, F., R. J. Gorte, and A. L. Myers, A New Calorimeter for
x smol fraction in condensed phase Simultaneous Measurements of Loading and Heats of Adsorption
x sdistance perpendicular to Gibbs dividing surface, m from Gaseous Mixtures, Langmuir, 15, 1570 1999..
yi smol fraction of ith adsorbate in gas phase Siperstein, F., O. Talu, and A. L. Myers, Absolute Adsorption Ver-
m schemical potential of adsorbent, Jrkg sus Excess Adsorption, Fund. of Adsorption, 7, Nagasaki, Japan
2001..
m s schemical potential of pure adsorbent in std. state, Jrkg
m i schemical potential of ith adsorbate, Jrmol Siperstein, F. R., and A. L. Myers, Mixed-Gas Adsorption, AIChE
J., 47, 1141 2001..
Jsgrand canonical partition function
Sircar, S., Isosteric Heats of Multicomponent Gas Adsorption on
P sspreading pressure, Nrm
Heterogeneous Adsorbents, Langmuir, 7, 3065 1991..
r smolar density, molrm3 Sircar, S., and A. L. Myers, Surface Potential Theory of Multilayer
f sspecific gas-gas q gas-solid potential energy, Jrkg Adsorption from Gas Mixtures, Chem. Eng. Sci., 28, 489 1973a..
F ssurface potential, Jrkg Sircar, S., and A. L. Myers, Prediction of Adsorption at Liquid-Solid
Interface from Adsorption Isotherms of Pure Unsaturated Vapors,
AIChE J., 19, 159 1973b..
Subscript Smith, J. M., H. C. Van Ness, and M. M. Abbott, Introduction to
isrefers to ith adsorbate component Chemical Engineering Thermodynamics, Fifth ed., McGraw-Hill,
New York, pp. 536]553 1996..
Smith, J. M., H. C. Van Ness, and M. M. Abbott, Introduction to
Superscripts Chemical Engineering Thermodynamics, Sixth ed., pp. 131, 207, 585
2001..
asrefers to adsorbed phase Steele, W. A., The Interaction of Gases with Solid Surfaces, Pergamon
compsrefers to isothermal compression of gas Press, New York, p. 107 1974..
g srefers to gas phase Suzuki, M., Adsorption Engineering, Kodansha, Tokyo; Elsevier, New
imm srefers to isothermal immersion of pure adsorbent in com- York 1990..
pressed fluid Talu, O., and A. L. Myers, Molecular Simulation of Adsorption:
8srefers to value in perfect-gas standard state Gibbs Dividing Surface and Comparison with Experiment, AIChE
ssrefers to pure solid J., 47, 1160 2001..
sU srefers to pure solid in acuo Valenzuela, D. P., and A. L. Myers, Adsorption Equilibrium Data
t srefers to total value for system Handbook, Prentice-Hall, Englewood Cliffs, NJ, p. 2 1989..

AIChE Journal January 2002 Vol. 48, No. 1 157


Yang, R. T., Gas Separation by Adsorption Processes, Butterworths, the molar integral functions are obtained by dividing by n a
London 1987..
Young, D. M., and A. D. Crowell, Physical Adsorption of Gases, Chap.
3, Butterworths, London, p. 187 1962.. DG a P m
D g as a
s RT ln y RT ln 1qCP .
n P8 na
Appendix DHa
The enthalpy, free energy, and entropy functions are calcu- D has sB
na
lated for the Langmuir model of adsorption of a pure, per-
fect gas. Although the Langmuir model seldom fits experi- TD S a m P
TD s a s s Bq RT ln 1qCP . y RT ln A9.
mental data quantitatively, it predicts qualitatively the behav- n a
n a
P8
ior of the thermodynamic functions in microporous adsorb-
ents. The Langmuir adsorption isotherm is Equation A9 shows that the constant B in Eq. A2 is equal to
the molar integral enthalpy, which, in general, varies with
mCP loading, but is a constant for the Langmuir model. In prepa-
nas A1.
1qCP ration for calculating the differential functions, the pressure
is eliminated from Eq. A5 using A1
n a is specific loading in molrkg, m is the saturation capacity
in molrkg, P is the pressure, and C is a function of tempera- m 1 na
ture DG a sy mRT ln a
q n aRT ln ?
my n CP8 my n a
1 A10.
Cs e ArR eyBrRT A2.
P8
Using Eq. 53
where A and B are constants: it is shown below that A is the
molar integral entropy at saturation and B is the constant. P
differential enthalpy of adsorption. P8 is the standard-state D g a s RT ln
P8
pressure. Assuming a perfect gas, the surface potential is ob-
tained by substituting Eq. A1 in Eq. 31 D has B
P
F sy mRT ln 1qCP . A3. TD s a s By RT ln A11.
P8

From Eqs. 48 and 51 Taking the limit as P ` and na m in Eq. A9 gives for
the molar integral entropy
P
DG a s n aRT ln qF A4.
P8 B
lim D s a s q R ln CP8 . s A A12.
P ` T
Using Eq. A3
which shows that the constant A in Eq. A2 is the molar inte-
P
DG a s n aRT ln y mRT ln 1qCP . A5. gral entropy at saturation. Experimental data for adsorption
P8 of gases near their critical temperatures in zeolites Myers
and Siperstein, 2001. gives values in the range y10 R to
From Eqs. 48 and 51 for a perfect gas h R s 0. y12 R for the molar integral entropy at saturation. This is
comparable to a typical value of y10 R for the entropy of
F condensation of liquids at their normal boiling points Smith
D H a syT 2 A6. et al., 2001..
T T P
Finally, the immersional functions are calculated from Eq.
50 using A3
Using Eqs. A2 and A3
DG imm sy mRT ln 1qCP .
D H a s n aB A7.
mBCP
D H imm s
From Eqs. 47, A5, and A7 1qCP
mBCP
P TD S imm s q mRT ln 1qCP . A13.
TD S a s D H a y DG a s n aBq mRT ln 1qCP . y n aRT ln 1qCP
P8
A8. The nuisance term PV s in Eq. 50 is negligible at low pres-
sure. All three functions are negative in sign. The equation
Having calculated the integral functions DG a, D H a, and D S a, for the enthalpy of immersion has the same functional form

158 January 2002 Vol. 48, No. 1 AIChE Journal


Figure A1. Langmuir adsorption isotherm at 298.15 K. Figure A3. Differential ( D g a, D h a, D s a ) and molar
Differential enthalpy sy28 kJrmol; capacity s 5 molrkg; functions ( D g a, D h a, D s a ) at 298.15 K rela-
molar integral entropy at saturation sy10.5 R. tive to perfect-gas reference state.
Constants same as Figure A1.

as the adsorption isotherm. The expression for the entropy of


immersion contains two terms of opposite sign TD S imm s
Numerical example
D H imm y DG imm . so that TD S imm has a minimum. For ad-
sorption of subcritical fluids, condensation of the vapor at its Let constants Asy10.5R, Bsy28 kJrmol, and ms 5
vapor pressure may occur first, before the minimum is molrkg; let variables P8s1 bar and T s 298.15 K. From Eq.
reached. A2, C s 2.214 bary1 .

Figure A2. Integral functions ( D G a, D H a, D S a ) at Figure A4. Immersional functions ( D G imm , D H imm ,
298.15 K relative to perfect-gas reference D S imm ) at 298.15 K relative to compressed
state. gas and clean adsorbent.
Constants same as Figure A1. Constants same as Figure A1.

AIChE Journal January 2002 Vol. 48, No. 1 159


The Langmuir adsorption isotherm Eq. A1. is plotted on The immersional functions from Eq. A13 are plotted on
Figure A1. The integral thermodynamics functions for the Figure A4. For subcritical fluids, these functions terminate at
adsorbed gas DG a, D H a, D S a. from Eqs. A5, A7 and A8 are the vapor pressure. For supercritical fluids, there is no limit
plotted on Figure A2. Values for all three functions are rela- to the pressure, but the Langmuir model does not account
tive to the perfect-gas reference state at 298 K and 1 bar. for excess variables according to Eq. 3 and therefore fails at
The inset in Figure A2 shows that the DG a and TD S a func- high pressure.
tions intersect at low pressure 0.01 bar., because D S a has This example shows that the behavior of the integral func-
zero slope at the origin. tions DG a, D H a, D S a. plotted on Figure A2 and their dif-
The differential and molar integral functions from Eqs. A9 ferentials plotted on Figure A3 is complex, even for the
and A11 are plotted on Figure A3. The differential Gibbs Langmuir model. However, the immersional functions
free energy D g a. is equal to the chemical potential of the DG imm , D H imm , D S imm . on Figure A4 have simple shapes,
gas m a.. Notice that the differential and molar functions are finite non-zero slopes at the origin, and no singularities. The
unequal. The Langmuir model has the unusual property that free energy of immersion is the equation of state for the ad-
the molar and differential enthalpies are equal. For a real sorbed phase according to Eq. 63. The immersional functions
system, the differential enthalpy is neither constant nor equal are clearly defined physically as the change accompanying the
to the molar enthalpy. Furthermore, for a real system, the isothermal, isobaric contact of clean adsorbent with com-
differential enthalpy D h a. is undefined goes to infinity. at pressed fluid.
high loading approaching the saturation capacity. Manuscript receied Mar. 23, 2001, and reision receied July 9, 2001.

160 January 2002 Vol. 48, No. 1 AIChE Journal

You might also like