You are on page 1of 8

JOURNAL OF CHEMICAL PHYSICS VOLUME 114, NUMBER 14 8 APRIL 2001

Adsorption of linear hydrocarbons in zeolites:


A density-functional investigation
Lubomir Benco,a) Thomas Demuth, and Jürgen Hafner
Institut für Materialphysik and Center for Computational Materials Science, Universität Wien,
Sensengasse 8, A-1090 Wien, Austria
François Hutschka
Total Raffinage Distribution, Centre Europèen de Recherche et Technique, B. P. 27,
F-76700 Harfleur, France
Herve Toulhoat
Institut Français du Pétrole, F-92852, Rueil-Malmaison Cedex, France
共Received 11 October 2000; accepted 23 January 2001兲
An extensive first-principles periodical study of adsorption properties of linear hydrocarbons in
zeolites is presented. The applicability of density-functional theory to weak interactions is inspected
within both local-density 共LDA兲 and generalized-gradient 共GGA兲 approaches for C1 to C6 linear
hydrocarbons. The LDA adsorption energies are due to the overbinding ⬃2.5 times larger than the
GGA values. A compact diagram is constructed showing the increase of the adsorption energy with
the length of the adsorbed molecule and with the concentration of acid sites in the zeolite support.
The flow of the electron density induced by the adsorption indicates that the adsorption on the acid
site is realized through the hydrogen bonding between the OH group and the CH3 group. The pattern
of the reconstructed bonding, however, is more complex than that of the simple hydrogen bond. The
regions of redistributed electron density within the adsorbed molecule are spread over the whole
CH3 group and the adjacent C–C bond. The off-centering of the reconstructed regions from atomic
positions is in good agreement with recent 13C measurements, showing only slight variation of
chemical shifts with the hydrocarbon length for both proton-free and the protonated forms of
zeolites. © 2001 American Institute of Physics. 关DOI: 10.1063/1.1355769兴

I. INTRODUCTION complex defect structures produced by the postsynthetic pro-


cessing of zeolites causes the interaction energy to increase
Zeolites are microporous materials widely used in tech- proportionally to the concentration of defects.
nological processes as multipurpose materials.1 They are Hydrocarbons 共HC兲 are raw materials for many techno-
very good sorbents with a high internal surface. Their
logically important products and their adsorption in zeolites
channel-like structures make the transport of molecules in-
is of continuous attention to both experimentalists and theo-
side or outside the crystals an easy process. Because of the
reticians. The adsorption isotherms of C5 –C10 normal al-
high stability and numerous imperfections such as Si/Al sub-
kanes in silicalite crystals are measured by Sun et al.3 The
stitutions compensated with acid protons or with extra-
adsorption of C3 –C6 normal alkanes in acidic zeolites is
framework cations, zeolites perform well as heterogeneous
studied by Eder et al.,4 and Denayer et al. have recently re-
catalysts.
ported on the adsorption of n-alkanes at high temperatures.2
In a series of intrazeolite processes through which reac-
tants are converted into products, the adsorption of mol- The experimental data indicate that adsorption energies of
ecules to the internal surface represents an important first the HC are much smaller than those of small polar molecules
step. The strength of the interaction depends on the compo- such as ammonia and water. The adsorption of polar mol-
sition of the zeolite being the weakest for the purely siliceous ecules is accomplished through relatively strong N–H¯O
compounds and proportionally increased with the increased and O–H¯O hydrogen bonds 共HB兲 with typical adsorption
concentration of the acid sites. For different structure types energies of ⬃200 kJ/mol 共NH3 5兲 and ⬃100 kJ/mol
different adsorption energies have been measured. In the (CH3OH, 6 H2O7兲. The strength of the adsorption forces act-
more compact structures 共ZSM-22, MOR兲 the contact be- ing on the HC is documented, e.g., by the adsorption energy
tween the framework and the adsorbed molecule is tighter of propane. For acidic ferrierite 共Si/Al⫽45兲 the measured
and leads to higher adsorption energies. On the contrary, value is 49 kJ/mol.4 For n-hexane Eder et al.4 deduced the
adsorption energies for the more open structures with smaller dependence of the heat of adsorption on the zeolite frame-
framework densities 共Y,X兲 are smaller.2 A wide range of work density. Supposing that a similar dependence is valid
also for other n-alkanes, the estimate for the adsorption of
a兲 propane in acidic gmelinite is ⬃40 kJ/mol and a much lower
Electronic mail: lubomir.benco@univie.ac.at. Permanent address: Institute
of Inorganic Chemistry, Slovak Academy of Sciences, Dubravska cesta 9, adsorption energy is expected for the purely siliceous struc-
SK-84236 Bratislava, Slovakia. ture.

0021-9606/2001/114(14)/6327/8/$18.00 6327 © 2001 American Institute of Physics

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
6328 J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Benco et al.

Theoretical ab initio works investigating a contact be- tions of chemical processes. Subsequent comparative studies
tween the zeolitic framework and the HC include molecules on technological zeolite are performed with mordenite. We
of various complexity, starting with methane8 and ending have compared structural properties as well as energetics of
with long linear hydrocarbon chains9 or unsaturated protonation of the two structures19,20 complemented by cal-
branched molecules like isobutene.10 The HC to zeolite in- culations of the stretching frequencies of hydroxyl groups. A
teraction has been, however, inspected mostly by means of dynamical study of H2O adsorbed in Na-free gmelinite has
the molecular cluster approach, which uses a fragment of the demonstrated the existence of the spontaneous proton trans-
zeolite containing the active site as a representative model of fer 共PT兲 between O sites.21 A similar scenario of the PT is
the zeolite. The fragment is cut out from the parent structure found for the Na zeolite.22
with the hydrogen atoms saturating terminal bonds. Two The present report describes the adsorption of linear hy-
main disadvantages of this approach are 共i兲 a relaxation of drocarbons in zeolites. The periodical approach is used to
the cluster resulting in the lost of resemblance to the envi- preserve the surface structure of the zeolitic framework and
ronment of the active site, and 共ii兲 omitting of long-range to include the long-range electrostatic effects. Two basic
electrostatic effects. Both relaxation producing the more DFT approaches, LDA and GGA, are used to calculate the
tight contact between the adsorbent and the adsorbate and adsorption energies. Weak interactions between the hydro-
neglection of electrostatic repulsion lead to adsorption ener- carbon molecules and the zeolite framework represent a
gies too high compared with experimental data. The former probe for the applicability of commonly used functionals to
drawback is eliminated by usage of a constrained cluster. An adsorption phenomena in zeolites.
influence of the long-range electrostatic interactions is taken
into account via embedding techniques11,12 based on the II. STRUCTURE AND MODEL OF ADSORPTION
combination of the quantum mechanics 共‘‘inner’’ region of The simulation of the adsorption of hydrocarbons in a
interest兲 and the molecular mechanics 共less important outer zeolite is modeled on the gmelinite structure. Gmelinite is a
region兲. Application of both corrections end with adsorption rather rare natural zeolite with the chemical composition
energies diminished to values compared reasonably with ex- Na8共AlO2兲8共SiO2 ) 16 . The primary building blocks 共SiO4 and
perimental data.10 AlO4 tetrahedra兲 are stacked into hexagonal prisms which
The density-functional theory 共DFT兲13 and the general- are secondary building units isomorphous with those of the
ized gradient approximation 共GGA兲 for the exchange technologically important faujasites.23 Parallel linkage of
functional14 have been used in numerous applications in hexagonal prisms leads to a hexagonal structure ( P6 3 mmc).
which DFT compares well with experiment and with the The set of irreducible atomic positions contains only one
most accurate ab initio calculations for properties such as tetrahedral site 共Si/Al兲 and four oxygen sites.24 The largest
structure and bond energy and are now in routine use for a aperture is a ⬃7 Å channel circumscribed by a ring of 12
number of fundamental properties of chemical and physical SiO4 tetrahedra 共12-membered ring: 12MR兲, which runs par-
systems. Despite the successes, problems remain. The GGA allel to the c axis. From the main channel perpendicular to
considerably improves bond distances and bond energies for the c axis the smaller eight-membered rings 共8MR兲 lead to
strong interactions of both covalent and ionic character com- the gmelinite cages which is the only type of cage in the
pared to the uncorrected local density approach 共LDA兲. Ap- structure. Figure 1 presents the structure, the orientation of
plicability of GGA to weak intermolecular forces has been 12MR and 8MR, the situation of the gmelinite cage, and the
tested on rare-gas diatomic molecules.15 Though GGA func- positions of irreducible O atoms. The x-ray structure refine-
tionals significantly improve the LDA values of bond ment performed on natural gmelinite with an Si/Al ratio of
lengths, binding energies, and vibrational frequencies, the approximately 2 and containing the corresponding number of
agreement is still not quantitative. Interaction energy com- counterions and approximately 24 water molecules per cell
prises a high percentage of dispersion energy 共more than yields cell dimensions of a⫽13.756 Å and c⫽10.048 Å. 24
90% in He2 兲. Because dispersion energy scales with r ⫺6 , it The simulations of the adsorption of n-hydrocarbons are
is sensitive to interatomic distances. Undervaluation of bond performed with the fixed volume and shape of the experi-
lengths which often occurs in GGA, e.g., for Ar and Kr,15 mental unit cell of the hydrated zeolite.24 The experimental
then leads to too small GGA interaction energies. Investiga- volume was shown to be only slightly larger than that ob-
tion of alkali metal adsorption at the surface of MgO16 have tained by the optimization within the GGA approach19 共cf.
shown that application of GGA functionals decreases inter- similar results for mordenite20 and chabazite25兲. Because the
action energies by a factor of ⬃1.6 共PBE9617兲 and by ⬃2.7 adsorption in a zeolite always leads to slight expansion of the
共BLYP18兲. crystal lattice, the cell volume determined on hydrated
We have performed an extensive study of properties of samples reasonably compares with the optimized unit cell of
zeolite structures aiming at increasing knowledge of in- the zeolite. The hydrocarbons are placed in the 12MR. The
trazeolite chemistry. Within the periodical approach the to- most probable position of the adsorbed molecules, trans-
pology of the inner surface of zeolites is preserved and long- ported into the structure by the stream of gaseous or liquid
range electrostatic effects are taken into account. Both a hydrocarbons flowing along the main channels, is parallel to
model structure and a zeolite of technological importance are the main channel. The position of the hydrocarbon molecule
under study. Gmelinite as a model structure represents, due is therefore chosen parallel to the c axis in which any energy
to its relatively high symmetry and the small number of at- loss due to the deformation of the the linear chain of the
oms per cell, a framework convenient for numerical simula- molecule is avoided. The acid proton is located at the O4

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Adsorption of linear hydrocarbons in zeolites 6329

FIG. 1. Hexagonal structure of gmelinite. Top view showing large 12MR FIG. 2. Fragment of the structure showing the location of the adsorbed
channels 共a兲. Side view of the framework structure 共b兲 displaying the 8MR hydrocarbon molecule, the Al atoms, and the acid protons, respectively. Top
and the gmelinite cages. Short arrows indicate the four inequivalent O sites view 共a兲 and the side view 共b兲.
surrounding the tetrahedral site.

bon, and radii of r s ⫽r p ⫽1.1. a.u. are used for hydrogen.


site, the OH group pointing to the center of the main Brillouin-zone sampling is restricted to the ⌫ point. The total
channel.19 To facilitate a maximum contact of the adsorbed energy is converged to 10⫺5 eV; the convergence is im-
molecule with two acid sites 共AS兲, a second proton is placed proved using a modest smearing of the eigenvalues. The op-
at a relatively short distance from the first one located on timization of atomic positions within the unit cell is per-
another O4 atom. Two Al atoms corresponding to two AS formed combining both the quasi-Newton and the conjugate-
are placed in second-neighbor positions, obeying the Löwen- gradient algorithms. In relaxed positions the residual
stein rule for the distribution of the Al-sites. The fragment of analytical Hellman–Feynman force acting on atoms is not
the structure, the position of the adsorbed molecule, the lo- larger than 0.1 eV Å⫺1.
cation of the acid protons and of the Al atoms, respectively,
are displayed in Fig. 2.
IV. RESULTS AND DISCUSSION

III. COMPUTATIONAL DETAILS A. Adsorption energies


The adsorption of HC in a purely siliceous zeolite is
Periodic first-principles calculations are performed
realized via the contact between the zeolite O sites and the
within the density-functional theory 共DFT兲 using the Vienna
hydrogen atoms of the HC. In a protonated structure the
ab initio simulation package VASP.26,27 The simulations use
strongest interaction occurs between the Brønsted O–H
both the local-density 共LDA兲 and gradient-corrected 共GGA兲
group of the zeolite and one carbon atom of the HC. Figure
density functionals in standard versions proposed by Ceper-
3 schematically compares these two types of bonding to the
ley and Alder and parametrized by Perdew and Zunger28 and
ordinary O–H¯O hydrogen bond and orders the bonding
by Perdew and Wang.29 The valence-electron wave functions
contacts according to a decreasing bond strength. All of them
are expanded in terms of plane waves,30,31 and the electron–
belong to the category of the hydrogen bonds.36 Compared to
ion interaction is described by Blöchl’s projector augmented
the typical O–H¯O interaction 关Fig. 3共a兲兴 the other two
wave 共PAW兲 technique32,33 applied to ultrasoft
contacts, O–H¯C 关Fig. 3共b兲兴 and C–H¯O 关Fig. 3共c兲兴, are
pseudopotentials.34,35 The calculations are performed with a
much weaker due to much lower electronegativity of carbon
plane wave cutoff energy of E pw⫽400 eV. Cutoff radii of
compared with oxygen. Adsorption energies are calculated
r s ⫽r p ⫽r d ⫽1.9 a.u. for the pseudo-wave functions are used
using the expression
for aluminum and silicon, respectively. Cutoff radii of r s
⫽1.3 a.u. and r p ⫽r d ⫽1.5 a.u. are used for oxygen and car- E ads⫽E 共 zeo⫹HC兲 ⫺E 共 zeo兲 ⫺E 共 HC兲 , 共1兲

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
6330 J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Benco et al.

FIG. 3. Schematic illustration of different hydrogen bonds 共HBs兲 in zeolites.


Ordinary O–H¯O HB for adsorbed water 共a兲, the O–H¯C contact be-
tween the acid site and a hydrocarbon 共b兲, and the C–H¯O contact between
the purely siliceous zeolite and the hydrocarbon molecule.

where E(zeo⫹HC) is the energy of the relaxed zeolite struc-


ture containing the HC molecule adsorbed in the main chan-
nel, E(zeo) is the energy of the adsorbate-free zeolite struc-
ture, and E(HC) is the energy of the HC molecule optimized
in vacuo. Geometries of both free HC molecules and ad-
sorbed complexes are given below. No volume (p•⌬V) and
entropy (T•⌬S) terms are considered in the calculation of
adsorption energies.
Figure 4 displays the adsorption energies as a function of
the HC length for the purely siliceous structure 共a兲, for one
acid site/cell 共b兲, and for two acid sites/cell located in the
framework at the short distance of ⬃4.8 Å 共c兲. For a series of
C1 –C6 molecules adsorbed in the purely siliceous structure,
a quasilinear dependence is observed for both GGA and
LDA. The values calculated using the GGA result from a
tedious structure optimization often ending in numerous lo-
cal minima of the potential energy surface. Figure 4共a兲
shows a slight increase between C1 and C2 . The adsorption
energies for longer HC then oscillate around the value of ⫺4
kJ/mol and do not increase for longer HC. Within the LDA
the corresponding adsorption energies are much higher than
those obtained with GGA, e.g., for ethane the GGA and
LDA values are ⫺4.7 and ⫺11.2 kJ/mol, respectively. A
modest increase of the LDA adsorption energies between C1
and C2 is followed by a quasilinear increase for higher HC
with an increment of ⬃⫺4.3 kJ/mol per C atom. The quasi-
linear behavior of LDA energies agrees reasonably with the
experimental data obtained for adsorption of n-alkanes in
acid zeolites2,4 as well as with the adsorption data for
silicalite.3 The linearity is due to the homogeneous increase FIG. 4. Adsorption energies of linear alkanes. Adsorption of C1 to C6 in the
purely siliceous structure 共a兲, C3 to C6 to one acid site 共b兲, and C3 to C6 to
of the lateral interactions between adsorbed molecules and
two acid sites 共c兲.
zeolite structure with every CH2 group of the HC. Much
lower values of the GGA adsorption energies calculated for
weak interactions are in agreement with the former results.
The applications to rare-gas diatomic molecules15 and alkali nonlinear increase between C4 and C5 . The reason for the
metal adsorption at the surface of MgO16 have documented step-like dependence is another contact of the adsorbed mol-
that when GGA functionals overvalue bond lengths the long- ecule to the second acid site. While C3 and C4 molecules
range van der Waals forces otherwise making substantial make only one contact, the C5 and C6 molecules are long
contributions to the adsorption energy vanish due to too large enough to make contacts to both acid sites. The gain in ad-
interatomic distances. sorption energy through the formation of this contact is
Energies calculated for one acid site/cell 关Fig. 4共b兲兴 are ⬃⫺25 kJ/mol 共LDA兲 and ⬃⫺9 kJ/mol 共GGA兲, i.e., almost
quasilinear for GGA and LDA, both shifted to higher adsorp- the same as for the first bond to the acid site. This increase of
tion energies by an almost constant increment of about 28 the adsorption energies is expected for protonated structures
kJ/mol 共LDA兲 and 8 kJ/mol 共GGA兲 compared with corre- with high concentration of acid sites 共Si/Al⫽1兲 provided that
sponding values for the purely siliceous structures. This the geometry of the adsorbed molecules and the distribution
proves that the strength of the O–H¯C interaction 关Fig. of the acid sites allow for a simultaneous formation of two
3共b兲兴 is larger than that of the C–H¯O bond 关Fig. 3共c兲兴. unstrained bonds. No such data, however, are available to be
Figure 4共c兲 shows that on structures with two acid sites compared with our calculated adsorption energies.
the adsorption energies are further increased, displaying a Figure 5 comprehends the adsorption data as a function

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Adsorption of linear hydrocarbons in zeolites 6331

TABLE I. Geometry parameters of a propane molecule in vacuo 共distances


in Å, angles in deg, and deviation 共␦兲 in percent兲. Notation according
to Fig. 6.

Exp. 共Ref. 37兲 LDA ␦ GGA ␦


C–C 1.526共2兲 1.509 ⫺1.11 1.528 ⫹0.13
C2–H 1.096共2兲 1.107 ⫹1.00 1.102 ⫹0.55
C1–Hs 1.089共9兲 1.102 ⫹1.19 1.099 ⫹0.92
C1–Ha 1.094 1.105 ⫹1.01 1.101 ⫹0.64
C1–H 1.091共10兲a 1.104a ⫹1.19 1.100a ⫹0.82
C–C–C 112.4共2兲 111.9 ⫺0.44 112.7 ⫹0.27
H–C2–H 106.1共2兲 105.5 ⫺0.57 106.0 ⫺0.09
Ha–C1–Ha 107.3 107.0 ⫺0.28 107.4 ⫹0.09
Ha–C1–Hs 108.1 107.7 ⫺0.37 107.6 ⫺0.46
H–C1–H 107.7共10兲a 107.5a ⫺0.19 107.5a ⫺0.19
C2–C1–Hs 111.8共10兲 112.5 ⫹0.63 111.9 ⫹0.09
FIG. 5. Adsorption energies as a function of the hydrocarbon length and the C2–C1–Ha 110.6 110.8 ⫹0.18 111.1 ⫹0.45
concentration of acid sites. a
Average value.

of the acid site concentration for C3 –C6 molecules. It dem-


onstrates that at all concentrations the GGA adsorption ener- siliceous structure does not induce any remarkable deforma-
gies are approximately 2.5 times lower than those by LDA. tion of the adsorbed molecule. The distances and angles of
More pronounced LDA-calculated adsorption energies dem- the Ha atoms which are in contact with the zeolite are similar
onstrate that the strength of the interaction between the HC to those of the free molecule 共cf. Tables II and I兲. The
and the zeolite framework is approximately doubled for the O1¯Ha distances to the framework O sites, however, are
acid zeolite compared with the purely siliceous structure. considerably shorter within LDA. The adsorption to the acid
With the increase of the acid site concentration the gain in site gives rise to changes which are observable in both LDA
energy is slowed down, probably due to steric constraints. In and GGA. These changes, however, are more pronounced
structures with a concentration of up to one acid site per within LDA. The OH-to-C1 contact, which represents the
adsorbed molecule the adsorption energy increases almost strongest interaction, keeps the adsorbed molecule at a dis-
linearly with increased length of the HC. For higher acid site tance shorter than that in the siliceous structure 共cf. averaged
concentrations, however, a pronounced nonlinearity appears, LDA values of 2.593 and 2.623 Å, respectively兲. The inter-
distinguishing between molecules according to the number action leads to a widening of the Ha–C1–Ha angle and to
of contacts with framework acid sites 共C3 ,C4 —one contact; the elongation of the C1–Ha bonds. On the side of the zeo-
C5 ,C6 —two contacts兲. lite the O1–H bond is elongated, giving evidence for the
formation of the O–H¯C hydrogen bond.
B. Structure of adsorption complexes The adsorption of the hydrocarbon induces a deforma-
tion of the zeolite structure as well. Table II reports the angle
As a response to the interaction with the zeolite support
O1⬘ –O4–O1 characterizing the curvature of the inner sur-
structural changes are induced in an extent which depends on
face of the zeolite 共cf. Fig. 6兲. The comparison of this angle
the interaction strength. This section reports details of the
with the values in parentheses, which correspond to the
geometries of the adsorbate sorbent complex at the example
adsorbent-free zeolite structures, shows that both the sili-
of the propane molecule.
ceous and the acid zeolite are deformed in the same manner.
1. Molecule in vacuo The uniform decrease of this angle indicates that in response
Geometry parameters of the propane molecule optimized to the local adsorption, a global deformation of the frame-
in vacuo by means of PAW-projected pseudopotentials work occurs producing an ellipsoidal shape of the main
within both LDA and GGA are compiled in Table I together channel. Different O1⬘ –O4–O1 angles observed for the acid
with experimental data.37 These values indicate that for in-
teratomic distances the GGA performs slightly better than
the LDA. All the LDA distances deviate by about 1% from
the experimental value, the C–C bond length being underes-
timated and the C–H bonds overestimated by this amount.
Within the GGA all distances are slightly overestimated and
compare reasonably to experimental data. The angles calcu-
lated within both LDA and GGA are of acceptable quality.

2. Adsorbed molecule
The geometrical arrangement of the contact site between
the zeolite support and the hydrocarbon molecule is dis-
FIG. 6. Geometry of the contact site between zeolite and the hydrocarbon
played in Fig. 6 and the geometry parameters are collected in molecule. Purely siliceous zeolite 共a兲, bonding at the acid site 共b兲. For the
Table II. The weak interaction between the molecule and the numerical values of bond-distances and angles, see Table II.

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
6332 J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Benco et al.

TABLE II. Geometry parameters of the adsorbate/sorbent complex for pro-


pane adsorbed in gmelinite 共distances in Å and angles in deg兲. Notation
according to Fig. 6.

Purely siliceous zeolite Acid zeolite

Hydrocarbon GGA LDA GGA LDA

C–C 1.527 1.508 1.529 1.509


C2–H 1.103 1.107 1.101 1.105
C1–Hs 1.099 1.103 1.098 1.102
C1–Ha 1.101 1.106 1.107 1.115
C–C–C 112.4 112.4 113.3 112.3
H–C2–H 105.9 105.8 106.4 106.1
Ha–C1–Ha 107.2 106.9 111.7 113.1
共107.4兲a 共106.7兲a

Contact site
O4共zeo兲¯C1 3.514 3.150 3.085 2.941
O4–H 0.996 1.014
H¯C 2.090 1.927
O1¯Ha 2.908 2.604 2.781 2.669
O1⬘¯Ha 2.994 2.642 2.672 2.517
O1¯Ha 2.951b 2.623b 2.727b 2.593b

zeolite
O1⬘ –O4–O1 158.6 156.3 163.4 162.1
共159.4兲c 共158.4兲c 共164.2兲c 共163.9兲c
a
Values in parentheses refer to the contact-free methyl group.
b
Average value.
c
Sorbent-free zeolite structures.

zeolite compared to the siliceous structure are due to the


local deformation raised by the Al/Si substitution.

C. Bonding to zeolite framework


The formation of a chemical bond between the zeolite
support and the adsorbed molecule is accompanied by a re-
distribution of the electron density. The change of bonding is
displayed by the differential charge density calculated as the FIG. 7. Difference electron densities of the propane molecule adsorbed in
difference between the electron density of the adsorbed com- the purely siliceous zeolite. Spheres show atomic positions of the adsorbed
plex and that of the adsorbent-free support and of the free molecule and the zeolite fragment adjacent to the contact site. Light gray
regions indicate a gain and the dark regions a depletion of the electron
nonadsorbed molecule, all systems preserving the relaxed density. LDA 共a兲 and GGA 共b兲 approaches 共isosurfaces of ⫹10 e/Å3 and
atomic positions of the adsorbed complex. ⫺7.5 e/Å3兲.

1. Purely siliceous zeolite


Figure 7 displays the difference electron densities of the increase of the electron density between the two Ha atoms
propane molecule adsorbed in the purely siliceous structure. close to the C1 atom and a polarization of the framework O4
Only minor changes of the electron density are induced atom on which the charge density is withdrawn from the side
within both LDA 关Fig. 7共a兲兴 and GGA 关Fig. 7共b兲兴 as shown of the zeolite and accumulated towards the CH3 group of the
by the low-value isosurfaces of ⫹10 and ⫺7.5 e/Å3 共cf. iso- adsorbed molecule. The other end of the molecule makes
surfaces of the acid zeolite displayed below兲. A different another contact between the second C1 atom and the frame-
pattern of stabilizing interactions is obtained within LDA work O3 atom of 3.343 Å. For this weaker contact Fig. 7共a兲
and GGA, none of them indicating a hydrogen bonding of shows a charge redistribution of an extent comparable to that
the type O共zeo兲¯H–C. The GGA leads to a very small within the shortest contact between O4 and C1. The electron
charge redistribution not localized on atoms. The LDA func- density is again accumulated on the framework O3 atom, but
tional leads to a more pronounced charge flow in correspon- depleted from the C–H bond.
dence with higher LDA adsorption energies 共Figs. 4 and 5兲
and shorter O4共zeo兲¯C1 distances 共Table II兲. The bonding
is of clear directional character localized between the O at- 2. Acidic zeolite
oms of the zeolite and the CH3 group of the hydrocarbon The charge flow induced by the interaction of the pro-
molecule 关Fig. 7共a兲兴. The changes are of comparable extent pane molecule with the zeolite acid site is displayed in Fig.
on both the adsorbed molecule and on the zeolite framework. 8. A distinct redistribution of the charge density is obtained
The shortest contact between C1 and O4 induces a slight for both LDA 关Fig. 8共a兲兴 and GGA 关Fig. 8共b兲兴 functionals

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Adsorption of linear hydrocarbons in zeolites 6333

nounced feature of the hydrogen-bonded hydrocarbon mol-


ecule is the large region of increased electron density be-
tween the two Ha atoms elongated towards the zeolite acid
site. An increase of similar extent and shape is found for a
benzene molecule adsorbed in acid mordenite, as well.39 This
accumulation of the electron density weakens the C–H
bonds and leads to an increased Ha–C1–Ha angle 共cf. Table
II兲. The off-centered position of the two depleted regions
indicates that the transferred electron density comes from
nonbonding regions of the Ha atom 共Fig. 8, left兲 and of the
Hs atom. The third depleted region is localized directly on
the C–H bond. The depletion of the electron density within
the CH3 group polarizes the C–C bond whose electron den-
sity is shifted towards the CH3 group.
Figures 7 and 8 show that regions of substantial differ-
ence electron densities within hydrocarbons adsorbed in both
purely siliceous and acid zeolites are typically not localized
on atoms. Recently, adsorption properties of linear alkanes in
ZK-5 zeolites were investigated by means of 13C NMR
measurements.40 The NMR chemical shifts, which are very
sensitive to the change of bonding on the measured atomic
center, show only slight variation for hydrocarbons of differ-
ent length in both the proton-free and protonated form of the
zeolite. This agrees well with our results displaying the
change of bonding within the terminal CH3 group which is
localized close to the H atoms or within the C–C and C–H
bonds but leaves the electron density around the carbon at-
oms unchanged.

V. CONCLUSIONS

A first-principles DFT study of the adsorption properties


of linear hydrocarbons in zeolites is presented. The weak
adsorbate–sorbent interaction energies are calculated within
both LDA and GGA approaches. The LDA adsorption ener-
FIG. 8. Difference electron densities of the propane molecule adsorbed to gies are ⬃2.5 times larger compared to the GGA values,
the acid site. LDA 共a兲 and GGA 共b兲 approaches 共isosurfaces of ⫹40 e/Å3 which is in agreement with other applications of the DFT to
and ⫺30 e/Å3兲. weak interactions. There are no direct experimental data for
the adsorption of the hydrocarbons in gmelinite. Adsorption
experiments performed on different structures of zeolites
which is slightly more pronounced for LDA. Both function- 共mordenite, ZSM-5, ferrierite兲 show that adsorption energies
als provide the same pattern of bonding restricted to the OH are similar to the LDA interaction energies. Experiments,
group on one side and to the terminal CH3 group on the other however, are performed on real crystals containing attractive
side. A rather complex charge redistribution occurs in the centers for the adsorption like framework defects and nano-
hydroxyl group, showing features typical of hydrogen- particles of extra framework aluminum oxides. Our calcu-
bonded OH groups.38 The area of the acid H atom is depleted lated GGA interaction energies indicate that much lower ad-
of electron density which is transferred towards both the sorption energies are expected for ideal crystal structures of
O–H bond and the CH3 group. A rather symmetrical redis- zeolites.
tribution on the O atom indicates a nonisotropic polarization In the relaxed structures of adsorbed molecules obtained
directed along the OH group, leaving the two covalent O–Si by GGA, the contact between the molecule and the support is
and O–Al bonds unaffected by the hydrogen bonding. On looser compared with corresponding structures calculated us-
the side of the hydrocarbon changes occur only in that part of ing LDA. Proportionally to the decreased strength of the in-
the molecule which is involved in the contact with the zeo- teraction, the adsorbed molecules are in the GGA much less
lite. In contrast to the typical O–H¯O hydrogen bond where deformed relative to geometries of the noninteracting mol-
the contact is realized through a single O atom, the hydro- ecules. The GGA adsorption energies are much smaller than
carbon molecule makes contact through the entire CH3 the corresponding LDA energies due to longer interatomic
group. Figure 8 shows that the increase of the electron den- distances and underestimation of long-range van der Waals
sity here is not centered on a single atom, but reflects a forces in the GGA. For the adsorption on the acid sites the
polarization of the whole methyl group. The most pro- LDA/GGA ratio is ⬃2.6 and much larger differences are

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
6334 J. Chem. Phys., Vol. 114, No. 14, 8 April 2001 Benco et al.

10
obtained for the adsorption of longer hydrocarbons in the P. E. Sinclair, A. de Vries, P. Sherwood, C. R. A. Catlow, and R. A. van
purely siliceous structures. Santen, J. Chem. Soc., Faraday Trans. 94, 3401 共1998兲.
11
Bonding effects visualized through difference electron P. Sherwood, A. H. de Vries, S. J. Collins, S. P. Greatbanks, N. A. Burton,
M. A. Vincent, and I. H. Hillier, Faraday Discuss. 106, 79 共1997兲.
densities show that within the GGA almost no charge redis- 12
I. H. Hillier, J. Mol. Struct.: THEOCHEM 463, 45 共1999兲.
tribution occurs in both the adsorbed molecule and in the 13
P. Hohenberg and W. Kohn, Phys. Rev. 136, 864 共1964兲; W. Kohn and L.
siliceous framework—in contrast to the LDA. There is no J. Sham, ibid. 149, 1133 共1965兲.
14
hydrogen bonding of the type C–H¯O. Within the LDA the J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson,
interaction of the framework with the adsorbed molecule is D. J. Singh, and C. Fiolhais, Phys. Rev. B 46, 6671 共1992兲; A. D. Becke,
J. Chem. Phys. 96, 2155 共1992兲; 97, 9173 共1992兲.
not realized through individual hydrogen atoms. Due to the 15
D. C. Patton and M. R. Pederson, Int. J. Quantum Chem. 69, 619 共1998兲.
contact with framework oxygen atoms, the entire terminal 16
J. A. Snyder, J. E. Jaffe, M. Gutowski, Z. Lin, and A. Hess, J. Chem.
CH3 group of the hydrocarbon molecule is polarized and Phys. 112, 3014 共2000兲.
17
participates in the formation of the adsorbate/sorbent bond. J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865
The adsorption on the acid zeolite induces extensive charge 共1996兲.
redistribution. The bonding is realized here through the acid
18
A. D. Becke, Phys. Rev. A 38, 3098 共1988兲; J. Chem. Phys. 96, 2155
共1992兲; C. Lee, W. Yang, and R. G. Parr, Phys. Rev. B 37, 785 共1988兲.
hydroxyl group. Again, the bonding leads to a redistribution 19
L. Benco, T. Demuth, J. Hafner, and F. Hutschka, J. Chem. Phys. 111,
of the electron density within the whole CH3 group and even 7537 共1999兲.
to a polarization of the adjacent C–C bond. The regions of 20
T. Demuth, L. Benco, J. Hafner, and H. Toulhoat, J. Phys. Chem. B 104,
maximum charge flow are not centered at atomic positions 4593 共2000兲.
21
and therefore not detectable by NMR measurements. L. Benco, T. Demuth, J. Hafner, and F. Hutschka, Chem. Phys. Lett. 324,
373 共2000兲.
22
L. Benco, T. Demuth, J. Hafner, and F. Hutschka, Chem. Phys. Lett. 330,
ACKNOWLEDGMENTS 457 共2000兲.
The work has been performed within the Groupement de
23
D. W. Breck, Zeolite Molecular Sieves 共Wiley, New York, 1974兲, p. 45.
24
E. Galli, E. Passaglia, and P. F. Zanazzi, N. Jb. Miner. Mh. 1982, 1145
Recherche Européen ‘‘Dynamique Moléculaire Quantique 共1982兲.
Appliquée à la Catalyse,’’ founded by the Council National 25
Y. Jeanvoine, J. G. Ángyán, G. Kresse, and J. Hafner, J. Phys. Chem. B
de la Recherche Scientifique 共France兲, the Institut Français 102, 5573 共1998兲.
du Pétrole 共IFP兲, TOTAL Recherche et Development, and 26
G. Kresse and J. Furthmüller, J. Comput. Mater. Sci. 6, 15 共1996兲.
the Universität Wien. Computing facilities at IDRIS 共France兲
27
G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 共1996兲.
28
J. P. Perdew and A. Zunger, Phys. Rev. B 23, 5048 共1981兲.
are kindly acknowledged. 29
J. P. Perdew and Y. Wang, Phys. Rev. B 45, 13244 共1992兲.
30
M. C. Payne, M. P. Teter, D. C. Allan, T. A. Arias, and J. D. Joannopo-
1
O. Haag, in Zeolites and Related Microporous Materials: State of the Art ulos, Rev. Mod. Phys. 64, 1045 共1992兲.
1994, edited by J. Weitkamp, H. G. Karge, H. Pfeifer, and W. Hölderich 31
D. J. Singh, Planewaves, Pseudopotentials and the LAPW Method 共Klu-
共Elsevier, Amsterdam, 1994兲, p. 1375. wer Academic, Norwell, MA, 1994兲.
2
J. F. Denayer, G. V. Baron, J. A. Martens, and P. A. Jacobs, J. Phys. 32
P. E. Blöchl, Phys. Rev. B 50, 17953 共1994兲.
Chem. B 102, 3077 共1998兲. 33
G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 共1999兲.
3
M. S. Sun, O. Talu, and D. B. Shah, J. Phys. Chem. B 100, 17276 共1996兲. 34
D. Vanderbilt, Phys. Rev. B 41, 7892 共1990兲.
4
F. Eder and J. A. Lercher, J. Phys. Chem. B 101, 1273 共1997兲. 35
G. Kresse and J. Hafner, J. Phys.: Condens. Matter 6, 8245 共1994兲.
5
D. J. Parillo and R. J. Gorte, J. Phys. Chem. 97, 8786 共1993兲. 36
6 J. E. Huheey, Inorganic Chemistry: Principles of Structure and Reactivity
S. P. Greatbanks, I. H. Hillier, and N. A. Burton, J. Chem. Phys. 105,
3770 共1996兲. 共Harper & Row, New York, 1978兲, p. 241.
7
J. Sauer, P. Ugliengo, E. Garrone, and V. R. Saunders, Chem. Rev. 94,
37
L. R. J. Lide, J. Chem. Phys. 33, 1514 共1960兲.
2095 共1994兲.
38
L. Benco, Eur. J. Mineral. 9, 811 共1997兲.
39
8
L. A. M. M. Barbosa, G. M. Zhidomirov, and R. A. van Santen, Phys. T. Demuth, L. Benco, J. Hafner, H. Toulhoat, and F. Hutschka, J. Chem.
Chem. Chem. Phys. 2, 3909 共2000兲. Phys. 114, 1 共2001兲.
9 40
J. M. Martı́nez-Magadán, A. Cuán, and M. Castro, Int. J. Quantum Chem. W. J. M. van Well, J. Jänchen, J. W. de Haan, and R. A. van Santen, J.
75, 725 共1999兲. Phys. Chem. B 103, 1841 共1999兲.

Downloaded 29 Feb 2008 to 130.192.146.150. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

You might also like