Download as pdf or txt
Download as pdf or txt
You are on page 1of 93

This article was downloaded by: [Yale University Library]

On: 15 August 2013, At: 01:03


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Catalysis Reviews: Science and


Engineering
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lctr20

Basic Zeolites: Characterization and Uses


in Adsorption and Catalysis
a
DENISE BARTHOMEUF
a
Laboratorie de Réactivité de Surface URA 1106 CNRS, Universié
Pierre et Marie Curie, 4 Place Jussieu, 75252, Paris, Cedex 05,
France
Published online: 16 Aug 2006.

To cite this article: DENISE BARTHOMEUF (1996) Basic Zeolites: Characterization and Uses
in Adsorption and Catalysis, Catalysis Reviews: Science and Engineering, 38:4, 521-612, DOI:
10.1080/01614949608006465

To link to this article: http://dx.doi.org/10.1080/01614949608006465

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Basic Zeolites:
Characterization and Uses in
Adsorption and Catalysis
Downloaded by [Yale University Library] at 01:03 15 August 2013

DENISE BARTHOMEUF*

Laboratoire de RCactivitC de Surface


URA 1106 CNRS
UniversitC Pierre et Marie Curie
4 Place Jussieu
75252 Paris Cedex 05. France

I. INTRODUCTION .......................................... 522

I1. GENERAL ZEOLITE FEATURES ............................ 523

111. BASIC SITES IN ZEOLITES ................................ 528


A . Bronsted Sites ......................................... 528
B . Structural Basicity (Intrinsic Framework Basicity) ........... 528
C. Clusters of Oxides and Hydroxides ....................... 530
D . Ionic and Metal Alkali Clusters .......................... 531
E . Species and Clusters Connected to Basic Centers . . . . . . . . . . . 533

IV. BASICITY CHARACTERIZATION ........................... 543


A . Theoretical Approach ................................... 543
B . Experimental Characterization ............................ 546

V. BASIC ZEOLITE-ADSORBED PHASE INTERACTION:


MOLECULAR APPROACH ................................. 560
A . Benzene .............................................. 560
B . Other Hydrocarbons .................................... 565

*Present address: 16 rue FranGois Gillet. 69003 Lyon. France .


521

Copyright 0 1996 by Marcel Dekker. Inc.


522 BARTHOMEUF

C. Other Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566


VI. BASICITY AND ADSORPTION ............................. 567
A. General Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
B. Separation of Mixtures .................................. 568
C. Competition of Adsorption .............................. 572
VII. BASICITY AND CATALYSIS ............................... 576
A. Alkylation Reactions ................................... 578
B. Dehydrocyclization of n-Alkanes ......................... 583
C. Hydrogenation-Dehydrogenation Reactions . . . . . . . . . . . . . . . . 584
D. Condensation Reactions ................................. 587
E. Miscellaneous Reactions ................................ 588
Downloaded by [Yale University Library] at 01:03 15 August 2013

VIII. .........................
ACIDIC VERSUS BASIC ZEOLITES 589
IX. CONCLUSIONS AND PERSPECTIVES ....................... 592
ACKNOWLEDGMENTS .................................... 595
REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 595
NOTE ADDED IN PROOF .................................. 610

I. INTRODUCTION

The presence of basic centers in some oxides has been recognized for a
long time as being important in catalysis [l-41.Usually both basic and acid
sites exist simultaneously. The two centers may work independently or in a
concerted way. For instance, in alcohol transformation, dehydration is favored
on acidic sites and dehydrogenation on basic centers [3,5].A large variety of
materials are cited as having basic character. They include single-metal oxides
(MgO, CaO, ZnO), supported alkali metals (Na/MgO, WK,CO,), mixed-
metal oxides (Mg0-A1203, ZnO-SiO,, MgO-TiO,), zeolites (X and Y sat-
urated with alkaline cations of low electronegativity), hydrotalcite-type an-
ionic clays, asbestoslike materials, carbon-supported basic catalysts, and basic
organic resins.
In the field of zeolites, the dependence of selectivity in catalysis in the
alkylation of toluene with methanol upon the acidic or nonacidic character
of the solid was first mentioned by Sidorenko et al. [6]. This was further
studied in detail, and the formation of ethylbenzene and styrene was linked
to basic sites in X and Y exchanged with K, Rb, or Cs cations. The production
of xylenes was related to the acidic character of Li- and Na-zeolites [7]. The
influence of cations in directing the prevailing acidic or basic character was
also shown in isopropanol dehydration (acidic centers) and in dehydrogena-
tion (basic sites) [5].
BASIC ZEOLITES 523

In adsorption processes, most of the zeolitic adsorbents are nonprotonic


and may present a basic character.
The importance of basic zeolites, both from a fundamental and an ap-
plied point of view, suggests that a better understanding of which parameters
determine the acidobasic character of zeolites is certainly requisite. The next
paragraphs describe, first, the general aspects of zeolite properties.

11. GENERAL ZEOLITE FEATURES

The Si-Al zeolites (aluminosilicates) consist of a network of TO, tet-


rahedra (T = Si or Al) linked by one oxygen. This determines a tridimensional
Downloaded by [Yale University Library] at 01:03 15 August 2013

framework with cages and channels, the size of which is comparable to that
of molecules (Figs. 1-4) [S-121. In modified zeolites the T atoms in TO,
tetrahedra may be B, Ga, Fe, Ti, V, or Ge, for instance [13]. In the alumi-
nophosphate family (AlPO-n), Al and P are predominant. The replacement
of P or/and Al generates a large variety of materials like SAPOs (Si-, Al-,
P-based solids), MeAPOs and MeAPSOs (Me = Co, Fe, Mg, Mn, Zn, . . .),
ElAPOs and EIAF'SOs (El: As, B, Be, Ga, Ge, Li, Ti, . . .) [14-171).
The replacement in a zeolite framework of an atom by another one with
a lower valency (for instance Si4' by A13+;P5+ by Si") creates a negative
charge on the framework which has to be neutralized by a proton or a metal
cation. The appearance of the positive and negative charge generates the

Sodalite
cage I

\
\

FIG. 1. Scheme of the faujasite structure with oxygen types (0)


and cation
locations (0)(from Ref. 8).
524 BARTHOMEUF
Downloaded by [Yale University Library] at 01:03 15 August 2013

FIG. 2. Structure of zecite beta: (a) 100 projection (polytype A); (b) porl on
of tetrahedral framework that defines the linear channel (from Refs. 9 and 10).

cancrinite cage

(a) (b)

FIG. 3. Structure of zeolite L (a) position of cations in hexagonal prism (A),


cancrinite cage (B) and large channel (C, D, E); (b) projection along c axis. Cations
in A, B, C , and E and oxygen atoms at position M are not accessible from large
channel (from Ref. 11).

acidobasic character of the zeolites. In fact, two main features characterize


these properties. A purely structural one resulting from the specific connection
of TO, tetrahedra and the second, physicochemical one, arising from the
chemical composition. It will be seen that both features may govern the
acid-basic character of zeolites. The first feature is linked to geometric fac-
tors (bond angles, bond lengths, spatial distribution of charges) and the second
BASIC ZEOLITES 525
Downloaded by [Yale University Library] at 01:03 15 August 2013

FIG. 4. Structure of mordenite (along c axis) with potential locations of cat-


ions. Oxygen atoms at position N are not accessible from the channel (from Ref.
12).

one to physicochemical properties (electronegativity, polarizing power of


ions, ionicity).
The framework is formed in the faujasite (FAU) structure, as an ex-
ample, of the building units sodalite and hexagonal prisms (Fig. 1). Figures
2, 3, and 4 describe, respectively, the structures of beta, L, and mordenite.
The vertices of the polyhedra are occupied by the Al and Si atoms. The
oxygen atoms belong to the tetrahedra TO, and are located at the apex of the
TOT angles, the angles of which vary from around 130" to 170". The bond
angles are interesting to consider since they affect the charge born by the
oxygen atoms [18], i.e., their basicity. The values depend on the crystal struc-
ture considered and for a given structure on the location of the oxygen atom.
For example, in FAU (Fig. 1) there are four different types of oxygen, giving
four TOT angles.
The number of oxygen types may be considerably higher than four as
seen from Table 1. This should give a large range of possible TOT angles
and consequently of charges on the corresponding oxygen.
It has to be noted that the TOT angle is also dependent on the T-0
bond length, which varies with the nature of T atoms. In faujasite, X and Y
differ only by the Al content. One may not expect similar values for TOT
angles whether Si or Al is present since the Al 0 bond (0.174 nm) is longer
than the Si-0 one (0.160 nm). In SAPO-37, which has the FAU structure,
the P-0 bonds (0.152 nm) will also give rise to different angles. The bond
angles and bond lengths are related to the sp hybridization of binding orbitals
[19] and to the bond ionicity [20]. For pure SiO, structure, a theoretical
calculation has shown that the framework ionicity increases in the order
FAU c LTL < MFI c OFF, MAZ c ERI < MOR [20].
In relation to the crystalline structure, several parameters may affect the
overall distribution of charges in the framework and consequently the basicity.
Downloaded by [Yale University Library] at 01:03 15 August 2013

TABLE 3
Characteristics of Usual Zeolites
Largest
0 atoms pore
Number of in largest aperture Usual
Zeolite" %/Ab A1/AI + Sib 0 types' apertured (4 name
FAU(X) 1.2 0.45 4 12 7.4 Faujasite
FAU(Y) 2.4 0.29 4 12 7.4 Faujasite
MAZ 2.6 0.28 6 12 7.4 Mazzite or
Omega
LTL 3 0.25 6 12 7.1 L
OFF 3.5 0.22 6 12 6.7 Offretite
MOR 5 0.17 10 12 6.5 X 7.0 Mordenite
BEA 14 0.07 15, 17 12 7.5 x 5.7 Beta
6.5 X 5.5
MFI 10 0.1 26 10 5.3 X 5.6 ZSMJ
5.1 X 5.5
MEL 10 0.1 15 10 5.3 x 5.4 ZSM-11
CHA 2 0.33 4 8 3.8 Chabazite
LTA 1 0.5 3 8 4.1 A
ERI 3 0.25 6 8 3.6 X 5.1 Erionite
"Nomenclature of International Zeolite Association [23].
bUsual values. May vary with synthesis conditions or further modifications.
'From Ref. 20, except BEA [lo].
dThe windows are usually referred to by this number (12R, 10R, 8R: R for ring).
BASIC ZEOLITES 527

In basic zeolites, by definition not protonic, the negative charges of the frame-
work are compensated by metal cations, usually monovalent. The potential
location of these cations is well defined and corresponds to the optimization
of crystal energy [21]. Figure 1shows the main cation sites known in faujasite
(I, 1’, TI, 11’, 111). Figures 3 and 4 describe the cation locations in L and
mordenite. The cations are mobile and may move from one site to the other
upon heat treatment or adsorption of certain molecules such as water, organ-
ics, etc. The location of the cations affects the TOT angles. In addition, their
location in the various possible sites and their displacement from one site to
another one in the presence of an adsorbed phase may influence the overall
distribution of charges in the framework. The negative charge born by the
oxygen atoms will then vary with the nature, valency, content, location, and
Downloaded by [Yale University Library] at 01:03 15 August 2013

mobility of the cations.


Another typical feature of zeolites is the existence in cages and channels
of fields and field gradients created by the charges distributed on the walls
of the cavities [19,22]. Modifying the distribution of these charges, by chang-
ing the acid-base properties through the Al content, the cation identity, or
the activation temperature, gives rise to new fields and field gradients which
possibly affect differently the molecules in cages, i.e., their reactivity.
The porous system of zeolites is defined both by the number of TOT
bonds in a ring (see Fig. 1 and Table 1) and by the size of the aperture (Table
1) [23]. Molecular sieving depends on the size of the ring and of the cages
relative to that of the molecules (adsorbate or reactant) which enter the zeo-
lite. It follows that the introduction of large cations in order to change the
acidobasic character may strongly modify the diffusivity of molecules in
pores, independently of any chemical effect.
The second main characteristic feature of zeolite concerns the chemical
composition. Changing this composition is the usual way to modify the prop-
erties of a given zeolite structure in order to generate specific selectivities in
adsorption or catalysis. The main application of zeolites employs the protonic
character in a large variety of hydrocarbon reactions (cracking, isomerization,
or alkylation). Other reactions involve basic or oxidation sites in zeolites or
bifunctional centers (metal plus protons). The present paper focuses on the
main chemical parameters which affect the basicity.
When the framework negative charge is compensated by cations with a
low electronegativity, the charge on the oxygen may be high enough to create
basic properties. It will be detailed further that increasing the Al content
strengthens the basic character [24], and that replacing Al and/or Si by other
atoms (Ga . . . and/or Ge, P . . .) having different electronegativities or sizes
may also modify the basicity [24-261. Other zeolite features are also depen-
dent on chemical composition. The covalency/ionicity of the bonding in zeo-
lites was already mentioned as dependent on the zeolite structure [20] and is
connected to the nature of the T atoms [27]. The field and field gradient in
cages vary with the Si/Al ratio in Si-Al zeolites and are also varied by
replacing Al or Si with other atoms [22].
528 BARTHOMEUF

All these general characteristics of zeolite properties indicate that the


intrinsic basic character of the framework oxygen is linked primarily to the
charge depending on both structure and chemical composition. It is accom-
panied by other effects (covalency/ionicity, spatial distribution of charges
modifying field and field gradient). All these parameters determine the reac-
tivity of basic sites in adsorption and catalysis. In addition, it will be shown
that additional basicity in zeolites may arise from different basic sites, which
are not framework oxygens (i.e., oxide clusters, etc).

111. BASIC SITES IN ZEOLITES


Downloaded by [Yale University Library] at 01:03 15 August 2013

Up to now the basic properties of zeolites have been reported mainly


for the Si-Al systems. The oxygen atom in the Si-0-Al species bears a
negative charge which may generate a basic character. The basicity may also
originate from other sites like basic hydroxyls, encaged oxide clusters, sup-
ported metals, or reducing centers. It may also be associated with the acidity
in acid-base pairs.

A. Bronsted Sites

The negatively charged lattice of Si-Al zeolites does not lead to the
existence of basic framework OH- groups. The OH- groups reported are
linked to extraframework species. For instance, small clusters of MgO or
CaO in faujasite cages were shown to generate basic hydroxyls [28,29]. They
give infrared bands vibrating at 3685 (MgY) and 3675 (Cay) cm-I, origi-
nating [30] from Reaction (1):
M2+(H,0) * M2+ OH- + H' (1)
Positively charged lattices, like those in hydrotalcite-clay systems, for
example, are not reported in the zeolite field [31]. One might expect that in
the large variety of approaches developed for synthesizing new zeolitic ma-
terials [15-17,32,33], one of them would give rise to a positively charged
framework with anionic exchange properties. This would generate Bronsted
basicity in OH groups.

B. Structural Basicity (Intrinsic Framework Basicity)

The framework oxygens bearing the negative charge of the lattice are
the structural basic sites. In most structures all the oxygen atoms are acces-
sible to adsorbates. For example, this is shown in Fig. 1 for the four oxygens
of faujasite. In less open structures, some oxygen atoms belong to cages
which are too small to be accessible. Examples of this case are atoms M in
LTL zeolite (Fig. 3) and N in mordenite (Fig. 4). It follows that only a part
of all the existing basic oxygens will interact with adsorbates or reactant in
BASIC ZEOLITES 529

these zeolites. Another characteristic of the oxygen sites is that they are fixed
between the two T atoms. In other words, unlike the protons in acid zeolites,
the oxygen sites are not mobile. This allows HC ions to move to reactants,
while in comparison the molecules have to approach the lattice oxygen in a
configuration that is favorable for the formation of a reaction intermediate.
One may expect a more demanding character in this case than when the
proton is involved in the formation of an activated complex.
The zeolite chemical composition and the structure type affect the oxy-
gen basicity. In Si-Al zeolites the most highly negatively charged oxygens
belong to the AlO, tetrahedra [34]. This selects which oxygen sites are basic
among all the existing oxygens of the lattice (i.e., Si-0-Si). The charge
on oxygen, expressing the basic strength, can be calculated theoretically. This
Downloaded by [Yale University Library] at 01:03 15 August 2013

is detailed in Section IV together with the influence of the chemical com-


position on this charge. Another factor influencing the oxygen charge is the
TOT bond angle. The electronic charge on oxygen (basic strength) varies as
the TOT angles are narrower and the TO distances longer [18,34]. It changes
in opposite direction to the case of the positive charge of the proton belonging
to the OH groups in protonic zeolites [35-371. In nonprotonic zeolites the
various TOT angles (up to 26 in ZSM-5) (Table 1) give rise to a large variety
of possible basic strengths. Depending on the Al location in the lattice, on
the nature of the cation (i.e., identity, content, valency, or location), and on
the accessibility of framework oxygens, the basicity of specific oxygen atoms
actually involved in adsorption or catalytic processes is not reliably obtained
from bond angle measurements. For instance, in Fig. 1 the TO,T angle de-
pends on whether the T atom is Al or Si, and whether the cationic sites SI,
SII, ,5111 are occupied or not by a given metal cation. No experimental tech-
nique is currently able to give the values of the corresponding TOT angles
for each of the 192 oxygens of a faujasite unit cell. X-ray diffraction (XRD)
or neutron diffraction, for instance, only give an average value. The case is
different from that of oxide surfaces-for instance, MgO, where different
basic sites are created at morphological defects such as edges, corners, steps,
and kinks [38,39]. In zeolites the basic sites are inherent to the crystalline
structure and to the chemical composition, even if no morphological defects
are present. During adsorption and catalysis, the reactant is able to choose
which oxygen to interact with, depending on the oxygen atom charge, access,
and neighboring configuration. Another important characteristic feature of
zeolites is that extraframework cations are mobile and, in the presence of an
adsorbed phase, may reach locations different from those of the starting ma-
terials [40-421. Due to lattice relaxation [43] this modifies the framework
angles and consequently the distribution of oxygen charges. This interactive
reactant-framework behavior is very important in the dynamic adsorption-
desorption process. Nevertheless, it is not much studied due to the difficulty
of experimentally characterizing specific TOT angles in the lattice.
Weakly basic oxygen atoms were reported to exist in N O 4tetrahedra in
the materials of the Alp0 family [44].
530 BARTHOMEUF

C. Clusters of Oxides or Hydroxides

Very small oxide particles can be encapsulated in zeolite cages


[28,29,45-501. Some oxide materials are known as basic catalysts (MgO,
CaO, ZnO) [ 2 ] . Dispersed in zeolite cages they may form small oxide clusters
with basic properties. Several modes of formation of these clusters are de-
scribed. A first approach is valid for polyvalent cations, giving rise to the
known hydrolysis reaction [30]:

Me"+ + H,O * Me'""''0H + H' (2)


This reaction is displaced to the right for the alkaline earth cations in the
Downloaded by [Yale University Library] at 01:03 15 August 2013

order Mg > Ca > Sr > Ba [51]. Upon heating, dehydroxylation occurs, form-
ing M,,,O oxides [52]. It was shown early that MgO and CaO clusters were
generated this way in MgY or CaY upon heating above 773-873 K [45].
Combined electron spin resonance (ESR), infrared, and XRD techniques
showed that these clusters are located inside the zeolite supercages. They
contain a number of MO species estimated to lie between 2 and 10. The
results demonstrate that Mg and Ca ions in Y zeolite may easily leave their
location at the exchange sites to form clusters of basic oxides [45].
A second approach to introduce basic clusters into zeolites is applicable
to a large number of cations. The zeolite may be impregnated, within a com-
patible pH range, with solutions of salts or hydroxides. For instance, the
addition of hydroxides of alkaline cations, mainly KOH or CsOH, was shown
to increase the chain selectivity in the alkylation of toluene with methanol,
typical of basic catalysis [53-571 or to facilitate the formation of anionic Pt
carbonyl clusters of the Longoni-Chini type [58] which are stable in basic
media. Under activation conditions, hydroxides, oxides, or carbonates are
formed [53,54,57,59]. Cs oxide is also proposed to be obtained by decom-
position of cesium acetate in NaY and NaX [46,47,49,60,61]. These basic
sites are very active in the dehydrogenation of isopropanol to acetone
[46,47,60], isomerization of 1-butene, and Knoevenagel condensation reac-
tion of benzaldehyde with ethylcyanoacetate [49,61,62]. Examples are given
where an excess of alkaline cations generates strong basicity [55,56,63], very
likely through formation of oxides.
Clusters of MgO and M 2 0 (M = Na, K, Rb, Cs) are prepared in Y and
X solids, respectively, by soaking the zeolites in solutions of magnesium
dimethoxide (alcoholic solution) or alkali metal acetates (aqueous solution)
[50]. Strong basic sites are obtained in the Mg case only if ensembles of Mg
and 0 form an MgO lattice, while isolated M 2 0 species produce strong ba-
sicity with the alkali metal oxides [50].
Two general trends are observed in the properties of these materials.
Firstly, exchanged zeolites are less basic than those which in addition contain
clusters of oxides. Secondly, carbonates are formed very easily from these
oxides even with atmospheric CO,.
BASIC ZEOLITES 531

D. Ionic and Metal Alkali Clusters

It has been known for a long time that strongly basic oxides can be
obtained by interaction of alkali metal with oxides like MgO or Al,O,
[4,64-671. Single metals [64,66], or binary mixtures of metals or salts [65,67]
are used. The metal is introduced as a vapor [64,67] or solid [67], or upon
NaN, decomposition [68].
Interaction of alkali metal vapors with zeolites generates colored prod-
ucts which may possess basic properties. It was first reported that
Nai' and Na:' paramagnetic centers were formed in alkali X and Y zeolites
[69]. Simultaneously, small neutral metal particles may be generated inside
and/or outside the zeolite framework [70-751. The interest in these ionic or
Downloaded by [Yale University Library] at 01:03 15 August 2013

neutral alkali metal clusters has grown rapidly in recent years since these
materials have application not only in catalysis but also as components of
advanced magnetic, optical, and electronic materials where they may act as
quantum dots [76,77]. The formation of Naz+, Na;+, Na;+, Ki+ was reported
in zeolites A, X, and Y in addition to that of Na, K, Rb, and Cs metal particles
in the same zeolites [69-75,78421. Zeolites A favors the appearance of
Rb;+ (n = 1or 2) [83] and Csi' or Cs:+ ionic clusters [84], while only Rb and
Cs metal particles are described in faujasite [71,79,80]. A Cs continuum
[(Cs,,,)"+ per unit cell of zeolite] is described [85]. K clusters exist in de-
hydrated L [86]. Rb ions interact with Ag clusters in A zeolite [87]. A Namb
alloy confined in NaY has been shown to exist [77]. A detailed catalytic study
showed that the active sites in basic catalysis are the basic framework oxygen
close to neutral Na: clusters entrapped in the large zeolite cages. These neutral
clusters are apparently responsible for a sharp ESR signal at g = 2.003 (Fig.
5) [72-751. The ionic Na clusters [Na:-'"] showing a 13 lines hyperfine
ESR signal at g = 2.003 (Fig. 5) are inactive in basic catalysis [72]. All these
results tend to show that in order to increase the basicity of the materials, the
formation of ionic clusters should be avoided, and that of neutral alkali metal
particles favored [72-751.
Different methods are described to generate these clusters. Interaction
of the alkaline ion-exchanged zeolites with metal vapors is commonly used
[69-71,77,83,84,88]. They are also obtained from organoalkali metals in sol-
vents [89,90], from Na or K stirred in the solid state with NaY [90b], or by
using sodium solutions in liquid ammonia [91]. Following the early work of
Fejes et al. [92], whose aim was to poison the acid sites of zeolites, the
decomposition of sodium azide (between 523 and 673 K) successfully led
[55,56,72-75,784301 to the generation of ionic and neutral Na clusters (either
separately or together) in zeolites [72-75,781. Other alkaline metal azides are
also described [79,80]. The azides may be mixed with the dry zeolite or
impregnated from a methanolic solution. The thermal decomposition of the
azides is strongly dependent on experimental conditions [93] so that the type
of cluster obtained (ionic or neutral) in the zeolites varies depending on the
532 BARTHOMEUF

-
Downloaded by [Yale University Library] at 01:03 15 August 2013

200G
DPPH 1
FIG. 5 . First derivative of ESR spectra of samples of NaY reacted with so-
dium vapor: (a) Na:' ionic cluster (from Refs. 69 and 70); (b) Na" cluster (from
Refs. 70 and 72). The marker corresponds to a DPPH signal at g = 2.0036.

method of adding the azide, the heating rate, the presence of 02,etc. [72-
751. Concerning more specifically the production of neutral clusters (i.e., of
basic sites), it is shown that the ease of their formation depends both on the
zeolite used (Al content, cation identity) and on the way the metal is added.
For instance, decomposition of NaN, in zeolites Y, X, A, and L gives rise to
Na metal clusters [72-75,781. Exposure of Y to alkali metal vapor gives metal
clusters only with Rb-exchanged Y when contacted with Rb vapor, and not
for Na or K vapor in presence of the corresponding NaY and KY zeolites
[70,71]. In the case of X zeolites treated with various metal azides, decom-
posed in situ, Rb and Cs clusters are easily formed. The Rb or Cs metal
atoms may arise either from the reduction of zeolite-exchanged ions regard-
less of the choice of metal vapor or from the metal vapor contacted with LiX,
NaX, or KX (except Cs/KX) [79,80]. It is proposed that the zeolite framework
modifies the gas-phase electron-transfer energetics. Several parameters are
involved in this modification: electron-transfer distance, electron transfer via
the zeolite framework, ion size effect on electrostatic interactions with the
oxygen framework, large cation blocking of a-cage entrances, and ion/atom
migration after electron transfer [79].
The results suggest a strong influence of the zeolite cationic form on
the reducibility of alkali metal atoms. In many cases, Rb and Cs metals are
easily formed. This is in line with the suggestion of the formation of K, Rb,
or Cs metal vapor upon heating the K-, Rb-, or Cs-exchanged zeolites above
600 K [59]. This would not occur for Li or Na forms. This behavior may be
explained by the lowest electronegativity of Rb and Cs atoms (Table 2).
BASIC ZEOLITES 533

TABLE 2
Sanderson Electronegativities, S
Atom
0 H Si Al Li Na K Rb cs
s" 5.21 3.55 2.84 2.22 0.74 0.70 0.42 0.36 0.28
S' 3.654 2.592 2.138 1.714 0.670 0.560 0.445 0.312 0.220
"From Ref. 181.
'From Ref. 182.

E. Species and Clusters Connected to Basic Centers


Downloaded by [Yale University Library] at 01:03 15 August 2013

1. Acid-Base Pairs
Acid-base pairs are well known in oxides [l-41. The existence of acid
and basic sites was reported in zeolites [5,6]. A rank of acid-base pairs is
given in Table 3 for various alkali metal forms of X and Y [24]. The cation
acts as a Lewis acid [51,94] and the close framework oxygen as a base. For
a given Al content the acid character prevails for cations with a high elec-
tronegativity. The basic properties increase in parallel with the Al level. It is
seen from Table 3 that at the boundary some zeolites have an amphoteric
character (KY, NaX). They behave as acids or bases depending on the shape,
size, or chemical properties of the molecules or reactants in the cages. The
involvement of acid-base pairs in ZSM-5 is described in the catalysis of
propene aromatization [95], and in transformations of methylbutynol and
monoethanolamine and the condensation of acetone [96].

2. Electron Donor-Acceptor Sites


Electron donor-acceptor sites are known to exist in aluminosilicates
(amorphous or crystalline). Their dependence on chemical properties of the
zeolite is often correlated with the acid and basic behavior. The Lewis acidity
is usually connected to electron acceptor sites [97-1001, and basicity to elec-
tron donor centers [101,102]. For instance, in zeolites, perylene forms Pet
radical ions and tetracyanoethylene is ionized t o the TCNE- radical anions.
Both may be studied by ESR (Table 4).

3. Metal Carbonyls
A very large amount of work has been devoted to the study of the
formation and properties of transition metal complexes in zeolites [103- 1091.
The zeolite acts as a solvent, an anion, and a ligand [103-1051. Some authors
point out the role of protonic acidity in addition to the specific properties of
the complexes in catalytic reactions [104,105]. It appears that the basicity of
zeolites is not much considered despite the fact that many complexes are
studied in nonprotonic, eventually basic zeolites.
534 BARTHOMEUF

TABLE 3
Acidobasicity Scale of Y and X Zeolites Exchanged with Alkaline Cations”
LiY NaY NaX KY
Acid strength’
Basic strength
KY RbY NaX KX RbX Csk
“From Ref. 24.
bPyridine titration.
‘Pyrrole infrared study.
Downloaded by [Yale University Library] at 01:03 15 August 2013

TABLE 4
Number of TCNE- Radical Anions Formed on Zeolites LTL Treated at 773 K“
and Average Charge on Oxygen
Zeolite
HL LiL KL CSL
Spinsjg x lo-’’ 1.25 2.60 2.85 6.09
-ti: 0.229 0.316 0.350 0.356
“From Refs. 153 and 154.
bCalculated average charge on oxygen using Sanderson electronegativities of Ref. 181.

Among all the complexes, metal carbonyl are of particular importance


as potential catalysts for hydrogenation, isomerization, hydroformylation, and
carbonylation reactions [ 1041. It is usually proposed that ion pairing occurs,
in which the zeolite cation interacts with the metal carbonyl through CO
[110,111]. The cation acts as a Lewis acid site, and the oxygen of the bridging
carbonyl as a Lewis base [110]. The zeolite basicity was considered only
recently in studies on the carbonyls of Pt [112-1151, Mo [116-1201, 0 s
[121,122], Rh [123,124], or Ir [58,125]. Only research involving basic sites
in zeolites is discussed in what follows.
Looking first at the increase of basicity with Al content, some changes
in properties are reported. For Pd carbonyls in NaY and CaA (5A type), it
was suggested that the higher wavenumbers of terminal CO in CaA possibly
arise from the lower nuclearity of the clusters, the presence of Ca instead of
Na, or from the lower Si/Al ratio (higher Al content) [126]. The thermal
stability of Fe(CO), is higher in NaX than in Nay, reflecting a stronger in-
teraction of the carbonyl with the Al-rich zeolite [127]. The thermal stability
of Mo(CO), increases and that of Mo(CO), decreases as the Al content in-
creases in different Na faujasites [120,128]. Considering these results, it has
to be noted that a higher Al content means a larger amount of cations, and
the effect of the zeolite could arise from this higher density of charges. Nev-
BASIC ZEOLITES 535

ertheless, examples are known with Fe [129,130], Pt [112], and Mo


[117,119,120] carbonyls showing an influence of the zeolite basicity for the
same density of charges, i.e., same Al content but different cation identities.
A n influence of the alkaline cations (Li' > Na' >> K' > Rb', Cs') in
the photoinduced catalytic activity of Fe(CO), in Y zeolites is explained by
the stabilization of the coordinatively unsaturated species (formed upon ir-
radiation) by coordination with the lattice oxygen [130]. The higher thermal
stability of the Fe-CO bond in CsY compared to NaY is proposed to originate
from the higher electron donor properties of Cs-FAU [129]. With Y zeolites,
only the alkaline cation form gives rise to anionic Pt carbonyl clusters. Nei-
ther HY nor NH,Y show the bright color typical of these clusters or the
infrared bands specific to Longoni-Chini Pt carbonyls and known to be
Downloaded by [Yale University Library] at 01:03 15 August 2013

formed in basic solutions [112]. The formation mechanism of these anionic


Pt carbonyl clusters, [Pt (CO),(~,CO),]f-, in zeolites strongly depends on the
experimental conditions and on the basicity of the zeolite [112,114,115,131].
The CO infrared wavenumber depends on the charge on Pt and is related to
the mean polarizability of the Pt cluster [132].
In the case of Mo carbonyls, Y and X zeolites exchanged with the series
of alkaline cations shows that the ease of partial decomposition of Mo(CO),
to the subcarbonyl species is promoted by the basic nature of the framework
oxygen (evaluated by 01,binding energy), whereas the thermal stability of
Mo(CO), is enhanced by the basic sites [117,119,120]. In addition, correla-
tions are found between the O,, binding energy and both the CO wavenumber
[119] and the M0(3d,,~) binding energy [119,120]. Figure 6 shows the de-
crease in the symmetric u(C0) band of Mo(CO), parallel to that of the 01,

of d -
binding energy (i.e., increase in oxygen basicity). This suggests that the extent
IT* electron donation from the molybdenum to CO increases with
zeolite basicity. This is in line with a strong donation of lone pair electrons
of the negatively charged zeolite oxygen to Mo in the Mo(CO), carbonyl.
The results are confirmed in Fig. 7, showing for this carbonyl a decrease in

O(k) BE (eV)

FIG. 6. Changes in the wavenumber of symmetric CO band in Mo(CO), as


a function of 01,binding energy in faujasites X and Y exchanged with alkaline
cations (from Ref. 120).
536 BARTHOMEUF

FIG. 7. Changes in binding energies of Mo (3d5,J in Mo(CO), (H) and


Mo(CO), (0)as a function of the 01,binding energies of faujasites X and Y ex-
Downloaded by [Yale University Library] at 01:03 15 August 2013

changed with alkaline cations (from Ref. 120).

0 0

fac mer

FIG. 8. Proposed configurations for Mo(CO), entrapped in faujasites ex-


changed with alkaline cations (OF- : basic framework oxygen) (from Ref. 120).

M0(3d~,~) binding energy as the zeolite basicity rises. This suggests (Fig. 8)
a direct coordination of the zeolite oxygen atoms to the molybdenum [120].
This may explain the stabilization of Mo(CO), carbonyls as the basicity of
the zeolite increases [117,119]. In zeolites these carbonyls show a selectivity
different from that observed for Mo/HY systems in the hydrogenation of 1,3-
butadiene to cis-2 butene [117,119]. Mo(CO), carbonyls have no significant
donor-acceptor interactions with the zeolite oxygen and probably no strong
interaction with the zeolite cations.
Stabilization of 0 s carbonyls in NaY zeolite with increased basicity is
strongly evidenced [121,122]. The anion clusters resemble those in basic so-
lutions or on basic MgO surface and have different properties from those
formed in the protonic HY zeolite (better thermal stability and selectivity in
catalytic reaction) [ 1221. Similarly, Rh carbonyls are highly stabilized in NaY
with enhanced basicity (i.e., after treatment with NaN,). They give similar
products (methanol and ethylene glycol) in the hydrogenation of CO, as ob-
served for the Rh carbonyl clusters in basic solutions, but different from those
of Rh/GA1203 [123,124]. The anionic carbonyl clusters [Ir6(CO)15]2-are
stabilized in NaX. Their chemistry is similar to that in basic solutions and on
the basic surface of MgO. Infrared and x-ray absorption fine structure
BASIC ZEOLITES 537

(EXAFS) studies suggest an interaction through ion pairing with the Na'
cation of the zeolite [58]. The anionic Pt carbonyls [Pt3(C0)& CO),]f- are
stable in NaY or CsNaX up to 473 K [112].
In conclusion, transition metal carbonyls similar to those observed in
basic solutions may be obtained in basic zeolites. Their high stability is de-
scribed as arising from their interaction with basic sites. In catalysis they give
selectivities different from those obtained with carbonyls in protonic zeolites.
Better knowledge of the generation of basicity in zeolites (see Sections I11
and IV) will give general rules for the formation of new transition metal
carbonyl clusters combined with shape selectivity and diffusion constraints
in zeolites. This is a very promising way towards new catalysts. In addition,
better attention could be paid to the cationic, neutral, or anionic character of
Downloaded by [Yale University Library] at 01:03 15 August 2013

the metal carbonyls in relation to the acid-base properties of the zeolite. This
might well be an important parameter for the generation and stability of the
metal carbonyls.

4. Supported Metals
The case of supported alkali metals has already been described (Sect.
1II.D). Transition metals are considered in the following.
A very large amount of work has been devoted to the preparation and
the study of properties of highly dispersed metals in zeolites. Reviews were
published in the field [111,133,134]. The present paper is limited to connec-
tions between metal particle characteristics and the basicity of zeolites. Most
of the following results are confined to Pt metal particles for the metal, and
to L, X, or Y zeolites.
It was shown early that, in addition to intrinsic electronic effects due to
the small size of the particles, a specific influence of the zeolitic environments
on the metal clusters can be seen for any zeolite type [133]. A main idea was
that, in the case of Pt/HY, the interaction of the Pt particle with acid sites
resulted in an electron deficiency of the metal compared to Pt/SiO,
[133,135,136]. A little later, Pt highly dispersed on alkaline L zeolites was
shown to exhibit very interesting activity in dehydrocyclization of n-paraffins.
The reaction was shown to be monofunctional on the platinum 11371. It was
proposed that Pt particles acquired specific properties in these basic zeolites,
due either to a typical crystalline configuration and/or to an electron excess
[113,138,139]. This last hypothesis was suggested from shifts in CO infrared
bands to lower wavenumbers. It was suggested to arise from an electron
transfer from highly negatively charged framework oxygen to the metal par-
ticles [113,138,139]. A similar explanation was also given for the lowering
of the hydrogenation activity of Ru supported on Y zeolites in the order
NaY > KY > CsY. The more electron donor zeolite CsY is the less hydro-
genating catalyst [140]. Further studies of Pt L zeolites exchanged with var-
ious cations and characterized using infrared spectroscopy of adsorbed CO
[ 1411, competitive hydrogenation of toluene and benzene [141], x-ray ab-
538 BARTHOMEUF

sorption, x-ray absorption near-edge structure (XANES), and extended x-ray


absorption fine structure [142], tend to suggest that an electron charge transfer
occurs from the zeolite to the Pt. The infrared approach of CO adsorption
applied to Pt/Y and Pt/X zeolites [143-1451 exchanged with the alkaline
cations shows a regular shift of vco to lower wavenumbers (Fig. 9) as the
oxygen basicity increases. In addition to bridged CO, two bands of linear CO
are seen around 2000-2060 cm-' (Fig. 10). A detailed scanning electron
microscopy (STEM) study of particle size and clusters location shows a nar-
row distribution of sizes around 12 A and the absence of Pt outside the zeolite
crystals [143,144]. The two bands of linear CO cannot then be ascribed to
platinum particles different in size or location. The experiments are conducted
in conditions which lead to a high CO dispersion on the metal. This is
Downloaded by [Yale University Library] at 01:03 15 August 2013

checked by the constant value of vco as a function of CO loading, demon-


strating that the CO molecules are isolated [143,144]. It has to be kept in
mind that if care is not taken to have such isolated CO molecules adsorbed
on the metal, then CO-CO coupling may increase vco [146-1481. The ex-
istence of the two linear bands of CO cannot then reflect different CO-CO
interactions, Fig. 10. The high frequency band (HF) corresponds to the usual

N
mol. h-1
2050
200

2000
r(

'fj
W

8
1950 100

1900

PiHY PtHNaY I'INaY PtNaX PtCsX

FIG. 9. Changes in the CO wavenumber [HF (A) and LF (B) bands of linear
CO] and catalytic activity at 298 K in benzene hydrogenation per surface Pt atom
(C) as a function of framework oxygen charge (from Refs. 143 and 145).
BASIC ZEOLITES 539

2063 2049

1778

I , , , I I
J ' S I I

200 ZOO0 1900 1800 2200 2000 1900 1800 cm

FIG. 10. Occurrence of one band of linear CO in HY and of two bands in


Downloaded by [Yale University Library] at 01:03 15 August 2013

PtHNaY (B), Pt NaY (C), and Pt NaX (D) in addition to bridging hydroxyls at 1763
or 1778 cm-'. Note the different scales for A, B and C, D (from Ref. 144).

linear CO band adsorbed on platinum. The shift to lower wavenumbers as


the zeolite basicity increases may reflect an interaction of the metal with
negatively charged atoms (framework oxygen) 1144,1451. This is in line with
the opposite shift to high wavenumbers for Pt clusters interacting with acid
sites [147]. It might, alternatively, result from a polarization of the metal
particles by the electrostatic field created in the cages by the cations and the
charges on the walls of the cavities [143]. No direct electron transfer from
the positively charged cation to the metal particle could be easily envisaged
as sometimes postulated [149]. The value of this HF band for PtNaY (2049
cm-') [144] is very close to the case where there is no CO-CO coupling and
no metal support interaction (around 2052 cm-I for Pt/Al,O, and isolated CO
molecules [147]). It follows that for zeolites giving u,, lower than around
2050 cm-', for isolated CO molecules one may expect particles with electron
excess or polarized in the zeolite field. The low frequency band (LF) of linear
CO is a new band observed mainly in basic zeolites [113,144,145,149-1521.
It appears to depend more strongly than the HF band on the sample type and
experimental conditions. It was assigned to CO adsorbed on Pt strongly in-
teracting with framework oxygen [144], or to CO interacting with cations
[151,152] or with framework oxygen [149]. Nevertheless, this band is not
observed In the absence of Pt. This does not make clear its assignment to the
interaction of CO with framework zeolite atoms. Similar HF and LF bands
are observed for CO adsorbed on PtL basic zeolites [113]. The regular shift
of the two bands to low wavenumbers as the cations are exchanged from Li
to Cs in the alkaline series 11131 may be related to the increase in the electron
donor character of this zeolite [153,154]. Table 4 shows an increase in this
property measured by the ESR signal of tetracyanoethylene anion radicals
(TCNE-) formed upon reaction of the TCNE with the L zeolites [153-1551.
In I r U , [156], as the K/AI ratio increases, the uco wavenumber is shifted to
low values and the intensity of the vco band decreases. The 300 cm-' uco
540 BARTHOMEUF

shift is explained by an increase in the electron density of Ir clusters due to


their interaction with zeolites sites.
For both linear HF and LF bands it was shown that water decreases
[ 1441 or increases [ 1521 the wavenumber. The decrease in vco is in line with
the known influence of electron donor molecules adsorbed on Pt in the pres-
ence of isolated CO [133,144,146,147]. In the case of the vco increase, the
two bands are assigned to CO adsorbed, not on Pt but on the cations [152].
In contrast to Refs. 149 and 152, the direct connection between Pt and CO
is evidenced by the very sensitive diffuse reflectance spectroscopy technique
which detects the Pt-C stretching vibration between CO and PtNaX [157a]
comparable to the one observed in Pt/Mg (A1)O [157b]. The observed wave-
number, 515-530 cm-' is found to be about 60 cm-' higher for the small Pt
Downloaded by [Yale University Library] at 01:03 15 August 2013

clusters in NaX (1-2 nm) compared to the case of larger particles (2-3 nm)
or of bulk metal (460-470 cm-I). This stronger bond is attributed to a neg-
ative charge on the Pt particles in NaX [157a]. The change in electronic state
of Pt in basic zeolites was confirmed by an x-ray photoelectron spectroscopy
(XPS) study of Pt/L exchanged with Li, Na, or Rb cations [158]. The presence
of a small amount of Pt'+ (electron deficient) was seen but the larger part of
the metal was present as Pts- (electron rich). It was also observed that the
fraction of partial negative charge (ti-) increases in parallel with the basicity
of the support [158]. EXAFS applied to PtKL showed, as in Ref. 142, a
strong metal-support interaction modifying the electronic properties of the
metal [159]. This decreases the H, chemisorption capacity. In the case of
PtKL [150] and PdNaY [111,160], metal adducts, Pt,Hm or (Pd,Hm)"+, are
proposed to be formed. The protons result from the reduction of Pt2' or Pd".
In fully protonic zeolites the number of protons is high. This induces an
electron deficiency in noble metal particles [161]. In the case of basic PtKL,
the high-frequency infrared bands of adsorbed CO decrease from about 2066
to around 2000 cm-l. This value of 2000 cm-', which is lower than 2050
cm-', suggests an electron excess on Pt or Pd particles as explained above.
It follows that a large part of the Pt particles would interact with basic sites
and generate the CO bands at the lower frequencies. The hypothesis that few
electron-deficient particles exist in addition to the electron-rich particles
would be in line with the presence (as already noted) of a small amount of
Pts+ and a large amount of Pts- in Pt alkaline L zeolites as seen by XPS
[158] and with the hypothesis of particle polarization in the cages [143]. For
Pd NaY the weak electron-deficient character of Pd particles is seen by XPS.
It is attributed to the interaction of Pd with the protons arising from the
palladium reduction [160]. One should remember that NaY is at the border
of basic and acidic zeolites (Figs. 9 and 11; Table 3). The interaction of Pt
or Pd metals with basic framework sites should then be very weak and should
strongly depend on the relative position of the metal clusters in the pores and
of the framework oxygen which are the more basic. The influence of various
parameters is seen in the decomposition of Pd(NH,),CI, complexes and the
reducibility of Pd" ions in zeolites PdNaY, PdNaX, and PdCsX. A correlation
BASIC ZEOLITES 541

c
w
;Om4
0
C
0
&
L.
0.3
01
5
0.2
Downloaded by [Yale University Library] at 01:03 15 August 2013

FIG. 11. Changes in the calculated charges on proton (A) and oxygen (B-
G ) as a function of Al content. Zeolites exchanged with proton (A, B), Li (C), Na
(D), K (E), Rb (F), Cs (G) (from Ref. 162).

is seen with the density of charges (X > Y), the strength of basicity, and the
steric hindrance due to the cations [163].
The metal-support interaction may be changed not only due to the more
or less basic character of framework oxygen but also due to the direct mod-
ification of the metal by an adsorbed phase or another metal. The possible
change in the electron-deficient character of Pt adsorbed on acidic zeolites
caused by adsorption of electron acceptor or donor molecules is well known
[141,1641.
Electron acceptors like sulfur increase the electron deficiency while elec-
tron donors like NH, or benzene decrease it [164]. It was shown in the case
of basic faujasite [143,144] that adsorption of NH, and H,O (donor mole-
cules) shifts the HF and LF bands of linear CO to lower wavenumbers, while
that of 0, (electron acceptor) increases the wavenumbers [144]. This may be
easily explained by an increased electron excess on Pt upon adsorption of
NH, or H,O on the platinum particles and by a decrease for 0,.
Several examples described the influence of a second metal in Pthi-
metallic clusters in basic zeolites. A clear study by XANES of a series of
Pt-Ni bimetallic KL catalysts shows the electron transfer from Ni to Pt in
the bimetallic clusters as determined in the presence of H2 [165]. In the case
of PtFe/KL zeolite, Mossbauer results show an interaction between Fe and
Pt [166]. Pd-Fe bimetallic particles are formed inside NaX zeolite cages.
This promotes the methanol synthesis [167]. Pt-Re clusters are supported on
NaY made basic by treatment with CsOH. It is shown that Re significantly
inhibits Pt cluster agglomeration during the treatment in H, [168]. Similarly,
Pt-Ni clusters in Pt/KL are significantly less subject to sulfur-induced particle
growth than their monometallic (Pt) counterparts [169].
The results indicate that one may expect to monitor the electronic state
542 BARTHOMEUF

of Pt by selective modifications by an adsorbate or by formation of a bime-


tallic phase. They also suggest that reactant or product may modify the elec-
tronic state of Pt in the course of the catalytic reaction itself, a point which
is usually not considered.
Many papers deal with the case of Pt supported on alkaline zeolite L.
An interesting modification of the Pt/L system is obtained by treatment of
KL with CF,Cl, which gives electron-rich particles [176]. Another important
aspect of PtL catalysts is its poisoning by adsorbed sulfur. It is well known
that Pt highly dispersed in acidic zeolites is sulfur resistant [164,170]. This
is explained by the presence of a very weak bond between electronegative S
atoms formed upon decomposition of sulfured compounds and the electron-
deficient Pt particles [164]. There is a small electron transfer from Pt to S.
Downloaded by [Yale University Library] at 01:03 15 August 2013

One may expect that, by contrast, a strong bond would exist between S and

sulfur sensitive [171-1741. This is related to a strong Pt-


electron-rich Pt clusters. In fact, Pt highly dispersed on basic zeolites is very
S electron transfer
[ 1621. The well-known Pt agglomeration upon sulfur poisoning [ 172,1731
may then be explained by a consecutive weakening of the Pt framework bond.
It is also proposed that sulfur may compete with Pt as an electron acceptor
for the electron transfer from the zeolite to Pt [175]. More recent studies
show that Pt would remain metallic when sulfur is chemisorbed [172-1741,
or that the sulfur resistance is increased in Pt-Ni clusters as compared to Pt
particles (less metal particle growth) [169].
No general relationship is described between metal particle size and the
basicity of zeolites. In basic zeolites a very high dispersion of Pt is observed
[171,172,174,177,178]. This suggests an anchoring of small Pt particles re-
sulting from their interaction with basic sites.
Easy autoreductions are described for Pt2+ ions in Kbeta [179] and for
Ruf cations in NaY [180].
As a general conclusion, this section on basic sites may be summarized
as follows:
Framework oxygens are the potential structural (or intrinsic) basic
sites. Their strength depends on the chemical composition and TOT
angles.
The overall zeolite basicity may be increased by addition of species
trapped in the cages. These species bring their own intrinsic basicity
(oxides, hydroxides) or interact with framework oxygen increasing
their charge (metal alkali clusters).
The basic sites (oxygen atoms) are associated with Lewis acid sites
(cations), forming acid-base pairs.
A correlation is usually seen between basicity (two-electron transfer)
and reducing activity (one-electron transfer).
Trapped metal carbonyls acquire properties of such species stabilized
in basic media.
Transition metals clusters in the cages (Pt, Ru, Ir . . .) behave differ-
BASIC ZEOLITES 543

ently from such clusters in acidic zeolites, suggesting an excess of


electrons in these particles trapped in basic zeolites.

IV. BASICITY CHARACTERIZATION

The oxygen charge which in zeolites characterizes the basicity may be


calculated as in any compound [181,182]. It is important to have information
on both the strength and the number of basic sites in a basic material
[24,155,162,183]. As mentioned in the preceding section, all the framework
oxygens are potential basic sites. Only the oxygens which acquire the highest
negative charge have a true basic character. This property may arise from
changes in the chemical composition (A1 content, cation identity and number)
Downloaded by [Yale University Library] at 01:03 15 August 2013

and/or in bond angles and bond lengths in the structure. The methods used
for the characterization do not always give information on both the strength
and amount of basicity. It is difficult at the present time to locate exactly, on
an experimental basis, the basic oxygens in the framework.

A. Theoretical Approach

A large amount of research has been devoted to theoretical calculations


on the acidity of zeolites [18,34,35,37,184-1901 using ab initio or semiem-
pirical approaches. The objective is not to evaluate the oxygen charge per se.
Nevertheless, this estimation may be obtained as an indirect result in these
works [18,34,184,186,188,189] and general trends are seen. The charge on
oxygen changes as the SiOAl angle becomes larger and TO distances become
smaller [181.
Ab initio calculations performed on small clusters show that the absolute
value of the charge on oxygen (i.e., the basic strength) increases for the
cationic forms in the order H c Li c Na [186]. Incorporation of Ga or Ge in
clusters with Si atoms increases the oxygen basicity [188].
Mortier [191] was the first to introduce in the field of zeolites the use
of the Sanderson principle of equalization of electronegativities in a com-
pound [181]. The intermediate electronegativity Sintof a given material re-
flects the mean electronegativity reached by all the atoms as a result of elec-
tron transfer during the formation of the compound.
For a compound P,Q,R,, the intermediate electronegativity Sintis given
by
,)r 1/@ +q +r )
S i n t = (S$ SY, S (3)
where S, denotes the electronegativity of atom j . The numerical values of S,
are given in Table 2 for the two scales proposed by Sanderson [181,182]. It
has to be noted that for zeolites where T atoms are not Si or Al-i.e., for
substituents like Ga, B, Fe, or Ge-the values of the Sanderson electronega-
tivities of these elements given in Refs. 181 and 182 might have to be cor-
rected. For instance, it was shown experimentally [192] that in zeolites the
544 BARTHOMEUF

electronegativity of Ga is 1.59, and not 2.42 as given in Ref. 182. The value
is lower than that of Al (1.74, Ref. 182). No precise information on “ex-
perimental’ ’ values of electronegativity in zeolites is available for the other
possible framework atoms.
Even assuming that the right atom electronegativity values are used, one
has to keep in mind that the approach takes into account only the chemical
composition of the zeolite. No effect of the structure is considered [193].
Nevertheless it is a very powerful method. The intermediate electronegativity
gives access to the mean charge on an atom j in the framework by
Downloaded by [Yale University Library] at 01:03 15 August 2013

with S, = 5.21 for oxygen [181], this gives


Sint - 5.21
6, =
4.75
Figure 12 relates this charge to the ionization potential of the cations
[191]. It shows that in the alkaline series the average basicity follows the
order of cationic zeolite form, i.e., Li < Na < K < Rb < Cs. The calculation
applied to zeolites with increasing A l content is shown in Fig. 11 [162]. It
shows that in protonic zeolites the known increase in acid strength as the Al
content decreases is accompanied by a decrease in basic strength (Fig. 11,
curves A and B). For all the zeolites investigated, the basicity increases in
parallel with the Al content. The dashed horizontal lines in Fig. 11 limit the
domain where it was shown experimentally that the oxygen charge is high
enough to create basic properties as detected, for instance, in catalysis

W
c3
&
4
z
u
2
-0.40 -
Fa Li

8 -0.30 - Ba (‘1
*X
0

Zn
-0.20 I I I
I I I I I I I
2 4 6 8 10 12 14 16 18
IONIZATION POTENTIAL

FIG. 12. Changes in the calculated oxygen charge of fully exchanged Y ze-
olites with the ionization potential [first (I) and second (11) ionization potentials] of
the cations (from Ref. 191).
BASIC ZEOLITES 545

[162,194]. It may be noted that we may not expect a significant average


basicity of oxygen atoms in Si-rich zeolites. In addition, for these materials,
the differences between the various cationic forms should be very small in
this composition range (very close curves in Fig. 11 at low Al content).
Nevertheless, it will be shown that in this Al range some zeolites present a
high experimental basicity. This is discussed in sections IV.B.2.c and VI.B,
and is related to an effect of the zeolite structure (bond angles). The order
obtained in Figs. 11 and 12 for different exchanged forms is the same as that
obtained from ab initio calculations [186].
It was shown experimentally that the basic oxygens are adjacent to the
cations [195,196]. As a result, a S,,,calculation is proposed for the moieties
including only 6 T atoms and 12 oxygens close to the cation, i.e., for the
Downloaded by [Yale University Library] at 01:03 15 August 2013

formula Sis+41,01zM,. This is assumed to lead to the local oxygen charge


[ 195,1971.
A more sophisticated approach, the EEM calculation (electronegativity
equalization method), includes the influence of the crystalline structure
[20,193,198]. It gives access to the framework ionicity [20,193,198] and to
the charges on specific atoms, for instance, on the four different 0 atoms of
the faujasite structure [198]. It was shown that the charges on the faujasite
oxygen change differently from oxygen to oxygen when the compensating
cations are Na, K, Rb, or Cs. This charge increases for some oxygens (02,
0,) while it decreases for 0, and 0, as the zeolite is exchanged from Na
to Cs [198]. The EEM approach suggests [20] that the framework ionicity
decreases in the following order for the various structures (pure SiOz
composition):
MOR > LTA > ERI > CHA > OFF,
MAZ > MFI > LTL, MEL > FAU
It is not related to the framework electronegativity calculated with the EEM
method:
MOR < MEL < MFI < MAZ < ERI < OFF < LTL < LTA < FAU
Independently of these theoretical approaches using ab initio, semiem-
pirical or electronegativity calculations, it was shown that the topological
distribution of Al and Si atoms in the structure affects the location and charges
in the framework [199,200]. Isolated Al (i.e., with no other Al as second
neighbors) gives rises to strong acidity of the proton in the Si-OH--A1
species. This involves a weak basicity of the oxygen. Calculations showed
that a maximum is expected in the number of such strong protons as the Al
atomic fraction reaches values below around 0.15 (Si/Al = 6). The Al content
at this maximum depends on the topology of each structure, i.e., on the con-
nectivity of the atoms. The experimental acidity measurements verify quite
well these features [200-2071. One may expect that for the alkaline cation
forms of zeolites, a minimum in the concentration of weak basic sites should
546 BARTHOMEUF

be found in the same Al level range. As pointed out in the comments on Fig.
11, this is an Al domain where the basicity of zeolites theoretically should
be small.

B. Experimental Characterization

Two experimental approaches give access to the basic strength. First,


direct methods of measurements can be carried out. In the second approach
a study of adsorbed probe molecules is used.

1. Direct Methods
Downloaded by [Yale University Library] at 01:03 15 August 2013

a. TOT Bond Angles (XRO). A correlation between TOT bond angles


and oxygen charges was suggested some time ago [MI. It allows estimation
of the rank of the acidity of the OH groups and correlative basicity of oxygen
from the angles deduced from XRD studies. The measured angles represent
only an average for any oxygen type considered. This method does not dis-
tinguish Si-0-Al from Si-0-Si species for the same framework oxy-
gen site (for instance 0, or 0, . . . in faujasite). It follows that up until now
this approach has not been applied very successfully.
b. NMR Chemical Shifts. The nuclear magnetic resonance (NMR)
chemical shift (6) of an atom may characterize the charge and the bond angle
associated with that atom in a crystalline structure [208]. "0 NMR is able to
distinguish the various species Si-0-Si (6, = 44 to 52 pprn), Si-0-Al
(6, = 31 to 40 ppm) and Si-0-Ga (6, = 28-29 ppm) in zeolites A, Y, X,
and sodalite [209,210]. However, no relationship is found with the Al (or Ga)
content or with the structure of the zeolite. The differences in chemical shifts
(6) as the T atom is changed from Si to Al and Ga might reflect a higher
negative charge of the oxygen when one moves from Si to Ga. This is in
line with an increased basicity as Si is replaced by Al in Si-Al zeolites [24]
(Fig. 11) or as Al is replaced by Ga in Si-Ga materials [211]. This also
agrees with the lower experimental electronegativity of Ga compared to that
of Al [192].
The search for correlations between 29Sichemical shifts (13,~) and the
characteristics of the zeolite structures showed that aSiis related to Si-0-T
bond angles and Si-0 bond lengths. For a given structure and Si(4Al) species
a linear correlation shows that is more negative as the Si-0-Al angle is
broader [208,212,213]. A semiempirical quantum-chemical calculation al-
lows one to propose that the Si chemical shift decreases approximately lin-
early with increasing effective (orbital) electronegativity, EN, of the oxygens
in the four Si-0 bonds around a Si atom. Up until the time of writing no
study has been devoted to the search of possible correlations between ex-
perimental oxygen basicity and tisi.
c. Binding Energies (XPS). The first application of XPS to the study
of zeolites demonstrated that in NaA, Nay, NaX, and NaMOR the Si/Al ratio
BASIC ZEOLITES 547

at the crystal surface was roughly twice that of the bulk [214]. This indicates
that in case of such heterogenous materials, one has to be careful in making
correlations between XI’S results and bulk properties.
The use of XPS to evaluate the binding energies of 01,was reported in
Ref. 215 for NaA, Nay, NaX; and in Ref. 216 for the same zeolites in addition
to ZSM-5 and CaX. A correlation was made between the 01,binding energy
and the basic strength of framework oxygen [217,218]. As one may expect,
the 01,binding energy decreases almost linearly (rise in oxygen basicity)
with the increase of the Al/Si atomic ratio of zeolites Y, X, and A. A study
of zeolites in the Li-, Na-, K-, and Cs-exchanged forms confirms the enhanced
oxygen basicity as the electronegativity of the cation decreases from Li to Cs
(Table 5) [217]. The decrease (Fig. 13) in Ols binding energy as Cs progres-
Downloaded by [Yale University Library] at 01:03 15 August 2013

sively replaces Na in NaY [116] is in agreement with theoretical calculations


(Figs. 11 and 12; Table 4) or other experimental measurements [5,24]. Ex-
tension of these results to other zeolite structures or cations (NaZSM-5,
Na-mordenite CaX, Ca A) [27] shows a linear relationship between the 01,

TABLE 5
01,Binding Energies (XPS) for X and Y Zeolites”

BE (ev)” BE (ev>
Y H-76“ 532.5 X H-78 532.2
Li-45 532.2 Li-57 531.6
Na- 100 531.8 Na-100 531.1
K-51 531.9 K-68 531.3
CS-8.4 531.9 (3-50 530.9
34 531.8
58 531.5
“From Ref. 217.
’O,, binding energy references to C,, = 285 eV.
‘Degree of ion exchange. The difference with 100 corresponds to Na
ions.

i532M
6531 0 50

Cs exchanged (%)
100

FIG. 13. Change in 01,binding energy for CsNaY as a function of Cs ex-


changed (%) (from Ref. 116).
548 BARTHOMEUF

binding energy and the average oxygen charge calculated with the Sanderson
electronegativities, as in Fig. 11 [27]. Similar trends and very comparable
numerical values were obtained more recently for X, Y, mordenite, L, and
ZSM-5 [196,219]. ZSM-5 gives the highest 01,binding energy and the value
is almost the same for the different alkaline cations. This is in line with
theoretical calculations (Fig. 11) which predict a very similar oxygen basicity
for the various alkali-cation forms of zeolites having a low Al content.
It may be important to avoid the presence of coke in XPS studies since
a splitting of 01,peak may result from the charge donation from the coke to
the framework oxygens [220]. It is also observed that using a high level of
accelerating potentials (8 or 10 V) involves the reduction of Na' ions to the
metal [221].
Downloaded by [Yale University Library] at 01:03 15 August 2013

A study of silicoaluminophosphate molecular sieves showed that the


presence of phosphorous generally lowers the 01,binding energy, i.e., in-
creases the negative charge on the framework oxygen [222].

2. Use of Probe Molecules


The choice of a good probe for the measurement of basicity is a major
problem. It should react specifically with the basic sites under consideration
(framework oxygen, basic hydroxyls, oxide clusters; see Sect. 111) while at
the same time it should not decompose or polymerize.
A variety of probes are known for the characterization of basic sites on
oxides (0,- and OH-) [2,223-2261. Deuterated chloroform [227] and pyrrole
[25,228] interact with basic sites through H bonding. CO can interact with
basic 0,- sites forming carbonites COi- ions [225]. CO, [225,229] forms a
variety of carbonates including mono- and bicarbonates, polydentates, or hy-
drogenocarbonates. Six different modes of CO, adsorption have been pro-
posed (interaction with the cation or the oxygen of the oxides [230], Fig. 14).
SO, may give sulfites (SO:-) or hydrogenosulfites [225,231,232]. Acetonitrile
gives rise to the amide or to the (CH,CN)- anion species [225]. Various
alcohols, isopropanol, methylbutinol [226,233,234], hexafluoroisopropanol
[225,235], or a mixture of cyclohexanone-cyclopentanol [236] have also
been described as basic site probes. Organic acids like benzoic acid
[7,55,237], acetic acid [237], and trichloroacetic acid [238] have also been
used. Furthermore, boric acid trimethyl ester detects different types of Lewis
basic sites with various strengths [239], while phenol [240] decomposes very
easily, which precludes its practical use.
Some of these molecules have been used for the characterization of the
basicity in zeolites. Some others such as benzene appear to be specific to
zeolites of the faujasite type [ 143,1531. Several techniques, mainly thermal
or spectroscopic, have been applied to this kind of study.
a. Colored Indicators Method. A variety of colored indicators have
been proposed to evaluate the basic strength of solids and to titrate the amount
of sites [2,7,55,223,224,238]. Table 6 reports the changes in color and the
BASIC ZEOLITES 549

H
I

f
c
Downloaded by [Yale University Library] at 01:03 15 August 2013

(4 (b) (c)

0
II
0/"\ 0

(
1 I
-M-0-M-)

0
II
i
0
..
2... *y0
C

I
0
I
0 I
\c&o
I
( -M-0-M-) ( - 0-M-0-) ( -M-0-M-)

(d) (e) (0

FIG. 14. Interaction of CO, with an oxide (from Ref. 230). (a) Adsorption
on a hydroxyl group with formation of a superficial hydrogenocarbonyl ion. @)
Adsorption on the metal cation and dissociation of the resulting bonding. (c) Ad-
sorption on the metal ion and the neighboring oxygen ion and formation of a biden-
tate carbonate. (d) Adsorption on the oxygen vacancy and formation of a superficial
carbonyl group. (e), ( f ) Adsorption on the metal ions with participation of oxygen
in excess and formation of a monodentate carbonate.

respective pK, (of these indicators). Quantitative measurements by titration


with benzoic acid give very small amount of sites in alkali-cation faujasites
(<0.3 mEq/g) [7,55,237]. Surprisingly, the number increases upon exchange
of alkaline cations by protons [237].In fact, the results probably demonstrate
that steric effects are important when metal cations are present in the cavities
so that the titrating agent and the large indicator molecules do not easily enter
550 BARTHOMEUF

TABLE 6
Indicators Used for the Measurement of Basic Properties of Solids"
~~ ~

Indicator Acid form Basic form PK6


Bromothymol blue Yellow Green 7.2
0. cresol phtalein Colorless Red-violet 8.2-9.8
Thymol-phtalein Colorless Blue 9.3-10.5
2,4,6-Trinitroaniline Yellow Reddish orange 12.2
2,4-Dinitroaniline Yellow Violet 15.0
4-Chloro-2-nitroaniline Yellow Orange 17.2
4-Nitroaniline Yellow Orange 18.4
4-Chloroaniline Colorless Pink 26.5
Diphenylmethane Colorless Yellowish orange 35.0
Downloaded by [Yale University Library] at 01:03 15 August 2013

Cumene Colorless Pink 37.0


"From Refs. 2, 7, 55, 223, 224.
"pPK, of indicator, BH (= pK,,,,)

into the pores. Only qualitative information may be obtained on the number
of sites. More precise information on basic strength is instead obtained with
thymolphtalein [7]. KX and RbX give the color of the basic form while LiX
does not. With the series of indicators listed in Table 6, the following order
of zeolite basicity is obtained (55):
Zeolite: NaX c CsNaX(CsC1) c Cs beta (CsOH), CsNaX (CsOH)
pK: 7.2 c 8.2 < 18.4
The pK, gives the upper basic strength detected. The oxygen basicity is
lower when Cs ions are introduced with CsCl rather than with CsOH. The
hydroxides trapped in the cages are then strongly basic. Despite its low Al
content, Cs beta is as strongly basic as CsNaX. This is in line with results
of benzene adsorption [241] and catalytic results [211]. A comparison [237]
of NaX and NaY confirms that NaX has stronger basic sites. These results
are in agreement with theoretical calculations, except for beta, whose specific
case is discussed in Sections IV.B.2.c and V1.B.
The basic strength of solids may be expressed with values of H-, where
H- = pKsH, i.e., the pK of indicator BH as listed in Table 6. Another method
determines the basicity of samples (H, scale) by using the Hammett indica-
tors, which are different from those of Table 6 and are usually employed for
acidity characterization. Both acidic and basic properties can be determined
in this way on a common Ho scale. A H,,, value represents the equal strong-
est & value of both acid and basic sites. A unique Ho,max is found for every
solid [224,242]. No such application is known at the present time in the field
of zeolites.
b. Thermal Methods. Thermogravimetry, calorimetry, and thermopro-
grammed desorption can give very valuable information on the interaction of
acid molecules with basic sites.
BASIC ZEOLITES 551

The adsorption of benzene on zeolites was studied calorimetrically [243]


or from isotherms determined by chromatography [244]. The enthalpy and
entropy of adsorption depend on the benzene loading and on the cations in
the zeolite [244]. It is shown that benzene interacts usually both with cations
and framework oxygen, and that the thermal effect is an average of these
interactions.
Several studies have used CO, as the acidic probe. The general features
observed in oxides are accepted as such in zeolite characterization. The
method may give access to the strength and to the number of basic sites
present. The basic strength is evaluated from the desorption temperature of
CO,. For instance, the increasing order of basicity of alkaline earth oxides
determined in this way was MgO < CaO < SrO c BaO, with MgO desorbing
Downloaded by [Yale University Library] at 01:03 15 August 2013

CO, at the lowest temperature [224,245,246]. This order is in agreement with


that expected from the electronegativity scale. In the field of zeolites, the
strength of the CO, interaction is usually related to the heat of adsorption
[247-2491. It was shown that the energy of adsorption decreases for zeolites
exchanged with Li, then Na, then K. This is quite opposite to the order
expected from theoretical calculations of basicity (Fig. 11 and Sect. 1V.A) or
from the catalytic properties of these zeolites (Sect. VII). An analogous order
was also obtained with L zeolites, the Na form giving the higher heat of
adsorption [250]. Similarly, since the desorption temperature of CO, is higher
for NaX than for CsNaX, it would suggest that the former zeolite is the more
basic [49,251]. This also completely disagrees with the known order of ba-
sicity derived by theoretical or experimental methods (Sections 1V.A and VII;
Fig. 11).These surprising results suggest that in zeolites the process of CO,
adsorption is more complex than that occurring on simple oxides. The si-
multaneous participation of several types of adsorption sites, as suggested in
oxides [230] (i.e., various cation locations and various oxygens Fig. 14)
would be very important. This may explain why CO, measures not only the
basicity of the zeolites but rather an overall property. The polarizing power
of the cation plays a very important role in this interaction [249]. In fact,
infrared studies showed the formation of ill-defined carbonatelike species
[28,29]. The adsorption of CO, appears to be much more connected to an
interaction with the cations rather than with basic oxygens. It follows that
CO, cannot be considered as a specific probe to study the framework basicity
of zeolites. At the same time, the order of the CO, desorption temperature in
a series of variously exchanged CsNaY samples does follow the theory [236].
Some studies are related to CO, adsorption on zeolites containing oxide
clusters [49,50,252-2541. It is very clear that the presence of these oxide
clusters enhances the amount of CO, adsorbed. In addition, this amount of
adsorbed CO, increases linearly in CsNaX with the simultaneous rise in added
cesium oxide phase [49,254]. The clusters trapped in the faujasite supercage
very likely form Cs carbonate-like species upon CO, adsorption [49,50]. At
present no information is available on the exact nature of this carbonate. It
is not obvious if the rise in CO, adsorption reflects a preferred interaction
552 BARTHOMEUF

with the cation, the oxygen, or both ions in the oxide cluster since all these
parameters increase in parallel. More experiments are needed before a clear
picture of the system will be available and before it can be concluded that
only the basicity of the oxide clusters is measured in this way.
c. Spectroscopic Methods. (i) Infrared. Infrared spectroscopy is the
most commonly used technique for the study of adsorbed probes in zeolites.
Only recently, studies were specifically performed with the aim of character-
izing zeolite basicity [24,25,195,197,228,237,255]. The IR spectroscopic
probes used in these studies include:
CO,. Using CO, as a probe to characterize the cation locations in
cages, it was shown early that in addition to physisorbed COz, some CO,
forms carbonates upon adsorption at temperatures around 673-773 K in Y
Downloaded by [Yale University Library] at 01:03 15 August 2013

zeolites saturated with divalent cations [28,29,256-2581 or in NaA [259,260].


Bands appear in the 1440-1500 cmpl range. Various carbonatelike species
are formed (uni- or bidentate, bicarbonate, etc.) depending on the experi-
mental conditions, such as the zeolite pretreatment, temperature of the CO,
interaction, presence or absence of water, etc. [28,29,257,258]. Framework
oxygens and water are mentioned as being responsible in the carbonate for-
mation [28,29,257,258]. CO, may also react with Bronsted basic sites formed
in Reaction (1).This results in the generation of carbonated species and acidic
OH groups in CaY and MgY zeolites [28].
The complexity of the stability of the various infrared bands [28,29]
confirms that CO, is not a good specific probe for the study of zeolite basicity.
PyrroZe. Pyrrole (C,H,NH) is an amphoteric molecule and interacts
with framework oxygens [25,228]. The NH infrared vibration decreases from
3430 cm-' in the pure liquid down to around 3200 cm-I upon the NH . . . 0
interaction with the framework oxygens [25,228]. This molecule is reported
to be dissociatively chemisorbed on some oxides [225]. On basic zeolites, if
care is taken (i) to use freshly distilled pyrrole and (ii) to record the spectra
rapidly after the adsorption, no extra bands of modified pyrrole are seen
[24,183,195,197,255]. Polymerization is usually not observed under these
conditions [24,25,183,255] while it is observed with basic oxides other than
zeolites [228]. A weak polymerization may occur with LiY or NaY when
heated [219].
The basic strength may be obtained from the value of the shift of the
NH vibration upon interaction with zeolite basic sites [25] by analogy with
that observed in solution [261,262]. The first study on zeolites [25] showed
that the order of basicity was NaY c NaX c NaGeX for these three zeolites,
in agreement with calculated basicities (Sect. 1V.A). The replacement of Si
by Ge in GeX increases the basic strength. Not knowing the electronegativity
of Ge in a zeolite structure (see Sect. IV.A), it is not possible to evaluate
theoretically its basicity. Nevertheless, the pyrrole measurements appear val-
uable since they agree with other experimental results like the order of acidity
of the protonic forms of these zeolites [263], which is reverse to that of
basicity [25]. The sodium-exchanged Ge-beta is also a basic zeolite as seen
BASIC ZEOLITES 553

in catalytic experiment [211]. Further research applied to various cationic


zeolites with different structures showed a good relationship between the shift
AuNH and the calculated oxygen charge [24] or intermediate electronegativity
[183] for X, Y, L, and ZSM-5-type zeolites exchanged with alkaline cations
(Fig. 15). ZSM-5, which has a low calculated basicity, does not shift the NH
band of pyrrole. By contrast mordenite and beta show a AuNH different from
zero while their Al content (and correspondingly the calculated SinJwould
suggest no basicity. Compared to Y and L, on the one hand, and ZSM-5, on
the other, one would not expect a shift of the vNH vibration. Other experi-
ments like benzene adsorption [241], CS aromatics separation (see Sect.
VI.B), or basic catalysis [211] show the existence of a significant basicity in
Na and Cs-beta, confirming the pyrrole results. The interaction of benzene
Downloaded by [Yale University Library] at 01:03 15 August 2013

with oxygen of the 12R window in mordenite [153] is also in line with a
basicity of framework oxygen. The pyrrole experiments of Fig. 15 clearly
show that not only electronegativity but other important parameters must be
considered, when comparing the basicity in different structures. The influence
of bond angles [162,211], framework ionicity [162,193], and Al pairing, at
least for mordenite [162], has been invoked. Mordenite has a high intrinsic
framework ionicity [20], which may explain the higher negative charge on
its oxygen. The alkali-cations forms have also a highly asymmetric structure
with cations at nonsymmetric locations [264]. This may create a small dipole
moment which can polarize pyrrole molecules [24] and modify their reactiv-
ity. For high-silica zeolites, in addition to their weak basic strength, they also
show a weak NH band intensity, reflecting a low density of sites [195].
For some zeolites (e.g., X, Y, mordenite, ZSM-5) containing two dif-
ferent cations, two vNH bands are observed. The simultaneous observation
of two N1, binding energies in XPS suggests that the basic sites are adjacent

- BEEF1
-0.4 -0.3 90
-AV I 1

y LTL
I I
MOR
I

u
X 1 -

cm-1

200

100
554 BARTHOMEUF

to the cations, and that each cation type generates a specific basic strength
on the close framework oxygen in a given zeolite structure [195,255]. The
amphoteric character of pyrrole may be used to characterize the Lewis acidity
of the alkaline cation simultaneously with the basic strength of the neigh-
boring oxygens [195,255]. The approach is extended to X zeolites with al-
kaline earth cations [197].
Benzene. Benzene was successfully used for the characterization of
faujasites basicity and of some other zeolites with 12R windows. The mea-
surement is based on the detailed study of the infrared bands of CH out of
plane (u, + q7and uI0 + ul,) and C-C (ulO)vibrations. It was shown that
upon interaction of benzene with cations, acting as Lewis sites, the CH bands
are shifted by around 20 to 40 cm-' while the C-C band did not move
Downloaded by [Yale University Library] at 01:03 15 August 2013

[183,241,265-2691. The interaction of the CH group of the aromatic ring


with the framework oxygen of the 12R window shifts the CH band by around
50 to 100 cm-' and the C-C band by 2 to 7 cm-'. Figure 16 depicts the

1815

1985 1844
I

2017 1877

-
1 I I I I I I I
2200 2000 1900 1800 cm-l
u
ACID SITES
BASIC SITES

FIG. 16. Infrared spectra in the CH out-of-plane range of benzene liquid (A)
and adsorbed in NaY at various loadings in molecules per supercage. G: 1.1, F: 2.1,
E: 2.9, D: 3.4, C: 3.7, B: 4.1 (from Refs. 143 and 265).
BASIC ZEOLITES 555

shape of the spectra as a function of the benzene loading for Nay. It shows
the bands at 1844 and 1985 cm-' (shifts of around 25-30 cm-') assigned to
the interaction with the cations and the bands at 1877 and 2017 cm-' (shifts
of around 60 cm-') assigned to the interaction with oxygen. The decompo-
sition of the spectra taken after completely filling the zeolite with benzene
gives access to the amount of benzene adsorbed on the cations and oxygen,
respectively. This also represents the number of zeolite sites interacting with
benzene assuming one molecule of aromatic per site as is usual in any titra-
tion. The method is then able to give information (i) on the content of basic
sites from the amount of benzene adsorbed on oxygen and (ii) on their
strength from the shift of the C-C band reflecting a distortion of benzene
[183,241,265-2691. As to the amount of sites present, Table 7 reports values
Downloaded by [Yale University Library] at 01:03 15 August 2013

for faujasite-type zeolites with various Al contents. The solids are ranked in
order of increasing calculated basicity (or charge on oxygen) from the top to
the bottom. There is a rather good correlation between these values and the
increasing amounts of benzene interacting with the oxygens of the 12R win-
dow for this large variety of Al contents and cation nature. Figure 17 gives
another example where the oxygen basicity is increased by exchanging the
Na cations of NaX with Cs ions [266]. The amount of benzene adsorbed on
oxygens increases with the oxygen charge. At high Cs contents steric effects
might disturb the location in the 12R window. The results of Table 7 and
Fig. 17 are in line with the observed basicity trends as determined by theo-
retical approaches (Fig. l l ) , XPS measurement of 01,binding energy (Fig.
13) [116,1961, and pyrrole measurements [24,162,183,255].
The strength of the zeolite basicity may be evaluated from the shifts
of the C-C vibration after saturation of the zeolites with benzene [266].

TABLE 7
Amounts of Benzene Adsorbed per Supercage on Cations and Oxygens of 12R
Window for Various Faujasite-Type Zeolites (Saturation of the Solids)"
~~

Alfu.c. Cation Window -62 Ref.


NaYD,' 7 0.45 -+ 0.1 0 0.232 143
NaYD,, 27 2 5 0.2 0 0.275 265
RbYD,, 27 1.75 2 0.2 0 0.303 265
NaY 56 3.5 +_ 0.4 0.6 * 0.2 0.350 265
KY 56 3.1 -+ 0.4 0.5 2 0.2 0.382 143
RbY 56 2.8 & 0.4 0.6 5 0.2 0.380 265
NaX 86 3.65 -+ 0.3 0.45 2 0.3 0.410 266
RbX 86 2.2 2 0.2 1.3 2 0.2 0.450 143
cs5,x 86 1.8 -+ 0.3 1.2 5 0.15 0.463 266
"From Refs. 143, 265.
*Average oxygen charge calculated from electronegativities of Ref. 181 (see Sect. IVA).
'Dealuminated Y.
556 BARTHOMEUF
Downloaded by [Yale University Library] at 01:03 15 August 2013

0 t
,,,
,
1
,, 2 3 4 5 CS+lS.C.

FIG. 17. Changes in the amount of benzene adsorbed in molecules per su-
percage in 12R windows (A), cations (B), or total (C) as a function of the average
number of Cs cations per supercage. Numbers in brackets represent the calculated
negative oxygen framework charge (from Ref. 266).

Table 8 reports an example for X zeolites exchanged with alkaline cations.


The bands are shifted to higher wavenumbers as the calculated charge on the
oxygen increases. The shift is rather small but is meaningful. A similar shift
is observed in the series of CsX described in Fig. 17 as the Cs content rises
[266]. The shifted band is observed (separately from the 1479 cm-' band of
benzene adsorbed on cations) only for the more basic zeolites and when the
CH bands show benzene in the 12R window. Another approach to basic
strength evaluation consists in following the disappearance of benzene bands
(CH or C-C) as a function of evacuation temperature for various zeolites
[143,269,270]. Figure 18 shows that CsX retains benzene in the 12R windows
at a considerably higher temperature than NaX, in accordance with its higher
basic strength [1431.
Various Probes. The infrared spectra of SO, [231] or acetic acid [237]
adsorbed on NaY and NaX have been reported. Similarly with the case of
CO, [28], SO, interaction with basic hydroxyls may generate acidic OH
groups [231]. This formation of acidity also occurs for H,S [271,272] upon
decomposition to H' and HS-. The interaction of HCN with X and Y zeolites
exchanged with the alkaline cations correlates with the basicity calculated or
evaluated from pyrrole adsorption [273]. The adsorption of H, or CO at low
temperature studied by infrared [274] indicates that H, acts as a probe for
surface acid-base pairs while CO interacts only with the Lewis acidic sites
BASIC ZEOLITES 557

TABLE 8
Wavenumbers of the C-C Vibration (q9)of Benzene
Adsorbed on X Zeolites“ and Average Charge on Oxygen
v19h
(cm-l)
Zeolite Cation Oxygen -6;

LiX 1479 NOd 0.406


NaX 1479 NOd 0.410
Kx 1479 1481 0.452
RbX 1479 1485 0.450
CS5,X 1479 1486 0.463
Downloaded by [Yale University Library] at 01:03 15 August 2013

“From Refs. 143, 266.


b1479 cm-’ for liquid benzene.
‘Average charge on oxygen calculated from Sanderson electro-
negativities of Ref. 181.
dNot observed. Probably in the 1479 cm-’ band.

FIG. 18. Decrease in amount adsorbed upon evacuation, for 1 h at each tem-
perature, for benzene interacting with the framework oxygen (12R window) in NaX
(A) and CsX (B) (from Ref. 143).

(cations). Nevertheless, the infrared wavenumber of both adsorbed molecules


(H, and CO) is modified by the Si/Al ratio, i.e., by the basic strength of the
oxygen atoms [274]. A correlation is also reported between the CO infrared
wavenumber and the calculated oxygen charge of X and Y zeolites exchanged
with polyvalent cations [191]. It results very likely from the indirect influence
of oxygen atoms as in Ref. 274. All these reported secondary reactions and
indirect interactions suggest that these probes should be considered with care
in the characterization of zeolite basicity.
(ii) XPS. The adsorption of pyrrole was followed by X P S on X and
Y zeolites exchanged with the series of alkaline ions [196,219,275]. XPS
peaks of polymerized pyrrole were observed only when the LiY and NaY
samples containing adsorbed pyrrole were heated at 338 or 523 K [219]. This
result confirms that pyrrole may be adequately used as a probe provided there
is a short contact time at room temperature [24,25,183,195,255]. The N,,
558 BARTHOMEUF

binding energy of pyrrole is lower for X than for Y and decreases in the
series of alkaline cationic forms from Li to Cs. A good correlation, seen in
Fig. 19 [219], is obtained with the shift of the NH infrared vibration of pyrrole
similar to that described in Fig. 15. In addition the deconvolution of XPS
peaks shows the presence of two peaks whenever the zeolite contains two
kinds of alkali cations. They were assigned to pyrrole adsorbed on the oxygen
atoms close to each specific alkali cations [196,219]. Combined with infrared
studies the results show that pyrrole is a sensitive probe of basicity, and as
indicated by XPS experiences, a charge transfer occurs from the framework
oxygen to the pyrrole molecule. This shifts the N1, level to lower energies
and weakens the N-H bonding [196,219].
(iii) ESR. The oxidizing and reducing properties of protonic zeolites
Downloaded by [Yale University Library] at 01:03 15 August 2013

have been studied for a long time [97-99,101,102]. However, not much in-
formation is available on nonprotonic or basic zeolites. A recent study was
carried out on LTL zeolite exchanged with various alkaline cations. Tetra-
cyanoethylene (TCNE) was used to measure the reducing site concentration
based on the number of TCNE- radical anions formed [153,154]. The values
reported in Table 4 are in line with the results (- lo1’ spins/g) usually mea-
sured for oxidizing and reducing centers in zeolites [97-1021. The number
of TCNE- anions formed increases in the order H < Li -= K < Cs, parallel
to the calculated charge on the oxygen. This is in agreement with the general
trend observed between the basic and reducing properties [101,102].
(iv) Other Techniques. Raman spectroscopy was used mainly to char-
acterize Al distribution or cations in faujasites [276]. One may hope that this
approach and others (such as UV-vis) could bring some information related
to zeolite basicity.

398 399 400 401 402


NIS Binding Energy (eV)

FIG. 19. Correlation between wavenumber of NH stretching vibration and


N1,binding energy of pyrrole adsorbed on X (0)
or Y (0)zeolites exchanged with
alkaline cations (from Ref. 219).
BASIC ZEOLITES 559

4. Catalytic Approach. Several reactions give selectivities which may


characterize basic catalysis in zeolites [277,278] (see Sect. VII). Examples
are the alkylation of toluene with methanol [ 6 ] ,the dehydrogenation of iso-
propanol [5], the transformation of methylbutynol [234], the cyclization of
diketones [63], the condensation of benzaldehyde with esters [279,280], and
the hydrogen transfer from cyclopentanol to cyclohexanone [236]. These
methods show, from their reaction products, the very likely presence of basic
sites. Up to now only partial evaluation of the number and strength of these
sites has been carried out from the catalytic experiments.
In conclusion, the characterization of basicity in zeolites is carried out
using various approaches. The method may concern the structural basicity or
that of a mixed system, i.e., zeolite plus occluded species.
Downloaded by [Yale University Library] at 01:03 15 August 2013

Theoretical calculations (ab initio, semiempirical, equalization of


electronegativities) can give access to a specific charge of each
framework oxygen type or to an average charge. This is an estimate
of the basic strength.
The calculations using the principle of equalization of electronega-
tivities indicate a way to increase the structural (or intrinsic) basicity
by decreasing the framework electronegativity. This may be achieved
by introduction into the zeolite of atoms of low electronegativity like
Al, and exchangeable cations (increase in the oxygen basicity in the
series of cationic forms Li < Na < K < Rb < Cs), substitution of Si
by Ge, of Al by Ga, etc.
Experimental characterization uses either direct methods (mainly
XPS) or indirect ones with the help of probe molecules. The inter-
action of the probe with the basic site may be observed directly (as
for colored indicators) or characterized by specific techniques (such
as calorimetry infrared, XPS, ESR) or catalytic tests.
The 01,binding energy (XPS) of framework oxygen shows an in-
crease in basicity in line with the one predicted from the electroneg-
ativity approach (high Al content, cations of low electronegativity).
Using probes, this is confirmed, for example, by the colored indicator
method, infrared study of pyrrole or benzene adsorption, and XPS
determination of N1, binding energy of adsorbed pyrrole. Neverthe-
less, beta and mordenite are more basic than predicted.
Few probes are reliable for the characterization of zeolite basicity.
Among them pyrrole and benzene have already given valuable
information.
The use of CO, as a probe is not recommended (at least in calorim-
etry or infrared) due to the large variety of possible interactions with
oxygen atoms in oxides. It might give some useful information when
oxide clusters are present. However, the results should be interpreted
with care.
560 BARTHOMEUF

The basic strength of the zeolite systems containing occluded species


is increased in the presence of oxide or hydroxyde clusters.
Catalytic reactions are a very informative tool for the classification
of basic zeolites.
Reducing centers connected to basicity are studied by ESR.

V. BASIC ZEOLITE-ADSORBED PHASE INTERACTION:


MOLECULAR APPROACH

Basic zeolites are used in processes (such as adsorption and catalysis)


where adsorbed molecules enter the cages. It is then of interest to study the
way in which these molecules are retained in the zeolites in order to better
Downloaded by [Yale University Library] at 01:03 15 August 2013

understand which sites interact with the adsorbates and thereby activate the
molecules for catalysis, for instance.
The molecular approach to the study of zeolite/adsorbed phase interac-
tion has developed rapidly in the last few years due to the progress in physical
techniques and to the growing use of interactive molecular graphics. Applied
to molecules adsorbed in zeolite cages or channels, it gives an increasing
amount of information on aspects such as the location of molecules, their
mobility, extent of clustering, etc.
The results correlated mainly to chemical properties of nonprotonic mo-
lecular sieves are considered here.
It was proposed early that cations in zeolites are adsorption sites for
molecules with TT electrons [94,244,281-2871 giving rise to a “specific in-
teraction” [282]. Many molecules of this type have been studied in particular
for benzene, which has been most extensively studied. Since this aromatic
interacts with both cations and basic oxygens, its location in the zeolite
cages-i.e., its further reactivity-depends on a balance for adsorption on
each type of site.

A. Benzene

A large part of catalytic or adsorption processes involve aromatics. Ben-


zene is chosen as the simplest aromatic molecule.
Different zeolite types have been considered in the interaction with ben-
zene including mordenite [288], LTL [289,290], beta [241,291], EMT [292],
and FAU [198,266,286,293-3101. Studies were carried out using various
techniques like infrared [ 143,162,183,241,265-267,286,291,292,296,302~,
UV-visible [294,296], laser Raman [293], neutron scattering [288,299],
neutron diffraction [289,290,297,298,306], ’H and I3C NMR [283,285,295,
303,304,3071, ’H NMR [300,308,309,311,312], fluorescence yield near edge
spectroscopy (FYNES) [313], and theoretical calculations [198,2871. Molec-
ular dynamics studies [301,305,309,310] compare models to experimental
results.
While cations have been accepted for a long time as adsorption sites
BASIC ZEOLITES 561

[244,282,287], Lechert et al. (295) were the first to propose the location of
benzene in the 12R window of NaX from NMR studies. This was shown to
occur also in NaY by neutron diffraction and neutron scattering [297-2991
and by infrared spectroscopy in various Y and X exchanged with the series
of alkaline cations [143,183,265-2671 in Na- and Cs-beta [241] and Na-EMT
[292]. Powder neutron diffraction [289] or infrared [153,270,314,315] did not
detect benzene in the 12R aperture of KL. Only Na cations are also mentioned
as adsorption sites in mordenite studied by quasi-elastic neutron scattering
[288].
The 12R window might not be a systematic site for benzene adsorption,
or the various techniques may not have the same sensitivity for the different
systems benzeneheolite [314].
Downloaded by [Yale University Library] at 01:03 15 August 2013

The two main modes of benzene adsorption, cation and 12R aperture,
are schematized in Fig. 20. The interaction of the ‘TT electron cloud with the
cation is well documented 1244,287,289,293,294,297,2981. It involves the
Lewis character of the ions [94], which is parallel to the order of decreasing
electronegativity (Table 2). The adsorption of benzene in the 12R window
may result from an interaction of the CH group of the aromatic ring with six
basic oxygen atoms of the aperture. This location has been deduced from
NMR [295], neutron diffraction [297,298], inelastic neutron spectra [299],
infrared studies [ 143,183,241,265-267,2921, and theoretical calculations
[198,3011.
The benzene-zeolite interaction was seen by UV-visible [294] or in-
frared [143,183,265-267,286,3021 spectroscopies. Figure 21 shows that the
wavelength of the ‘TT-T*transition of benzene (high loading) increases with
the cation radius [294]. A correlation was noted at a loading of one molecule
of benzene per cage between the UV diffuse reflectance and the laser Raman
wavelengths [293]. The infrared shift of the CH out-of-plane band decreased

(4 (b)
FIG. 20. Schematic representation of benzene interacting with zeolite sites.
(a) cation in a hexagonal face interacting with 7r-electrons of benzene; (b) CH of
benzene in interaction with the oxygen atoms of a 12R window. The T atoms (Si or
Al) are at the apex.
562 BARTHOMEUF

Na K Rb Cs
Downloaded by [Yale University Library] at 01:03 15 August 2013

Ionic Radius (A)

FIG. 21. Changes in the wavelength of the T - T * transition of benzene with


the cation radius in X or Y zeolites (adsorption of four molecules per cage) (from
Ref. 294).

from Li to Cs in the alkaline ions series [143,183,265,266,302].The UV [294]


and infrared [183,265,266] results were also found to depend on the Al con-
tent, with higher wavelengths shifts being observed in X compared to Y.
The location of benzene in the 12R window of faujasite-type zeolite has
been clearly described using various techniques [ 143,183,198,265-267,295,
297-2991. The changes in the interactions with the chemical properties of
the zeolite (such as Al content or cation nature) were studied mainly by
infrared [143,183,265-2671 and to a lesser extent by NMR [304] spectros-
copies. The quantitative study of the CH out-of-plane infrared bands in the
range 1800-2100 cm-' gives very precise information on the amount of
benzene adsorbed at equilibrium on each site at each loading. The influence
of benzene loading is mainly considered here. Figure 16 reports, for example,
the changes in the absorbance for different loadings of benzene in Nay. At
low loadings benzene is adsorbed first on Na cations. Adsorption in the 12R
window starts when the benzene content is increased. This is in reasonable
agreement with results obtained from other techniques [298,303,304]. The
study of a large number of zeolites [143,183,265-2671 shows two main fea-
tures. Firstly, at low coverage the preference for adsorption on basic sites
(12R window) increases with the average calculated basicity as the Al content
rises and when cations with low electronegativity like Cs are exchanged in
the material [265,266]. A sample like Cs,,Na,,X does not retain benzene on
cations below a loading of one molecule per supercage [266]. The aromatic
is adsorbed only in the 12R window. Secondly, the quantitative estimate at
high coverage indicates also that the adsorption of benzene on oxygen atoms
of the 12R aperture is the highest for the more basic zeolites (Table 7, Fig.
BASIC ZEOLITES 563

13). Results obtained with a purely siliceous faujasite-i.e., with no cations,


no basicity-show that benzene is highly mobile. There are no specific sites
for adsorption as seen by powder neutron diffraction [306], 'H solid-state
NMR [308,309], or molecular dynamics [309]. A calculation using the elec-
tronegativity equalization method (EEM) suggests that in a USY (no Al and
no cations), benzene could be located in the 12R window 11981. It also con-
cludes that the field gradient in the aperture is an important parameter for
directing the adsorption of benzene. Figure 22(a) shows that the mean field
gradient on the benzene C-H bonds for cationic adsorption sites decreases
from Na to Cs in the alkaline series (as the Lewis acidity decreases). For
adsorption in the 12R window [Fig. 22(b)] the field gradient increases from
Na to Cs (increase in the calculated oxygen basicity) [198]. The results pre-
Downloaded by [Yale University Library] at 01:03 15 August 2013

sented strongly suggest that, not only in adsorption processes but also in
catalysis [22] the location of benzene (or other aromatic molecule) on one

W
\

2
4
0.05

0.00 I I I 1
0.1 0.15
ionic radius (nm)

0.06
(b) f csx

r( 0.04
4
.
a,

>
0.02 - cv
NaX
NaY 6
f.... l RbY

0.00
0.1 0.15
ionic radius (nm)
FIG. 22. Calculated mean field gradients on the CH bonds of benzene as a
function of the ionic radius of the alkaline and alkaline earth cations. (a) cation site;
(b) 12R window site (from Ref. 198).
564 BARTHOMEUF

site or the other depends to a large part on the chemical properties of the
zeolite and on its loading, i.e., on the aromatic partial pressure. The aromatic
will be in competition for location at a given site with other reactants or
adsorbates even during adsorption of mixtures or catalytic processes.
One should also recall that molecules are mobile inside the cavities
[290,300,301,308,309,312].The picture presented here on benzene location
implies that benzene moves around its axis and may jump from site to site.
At increasing loadings the mobility of benzene decreases [290,298,304] and
it moves to the center of the cavities. This is explained by the formation of
clusters, and some molecules of this species interact with the cations while
others interact with the 12R window (Fig. 23) [298,299]. This easy agglom-
eration of benzene might introduce some mistakes in experiments if care is
Downloaded by [Yale University Library] at 01:03 15 August 2013

not taken to check whether equilibrium is reached in all the cavities and if
the sample is not heterogeneous with respect to the adsorbate, i.e., with some
cavities empty and some cavities containing clusters of benzene [303,316].
From NMR [303,316] and infrared [265,266], general trends are observed.
For rather large amounts of zeolite (1 g in NMR), it is necessary to heat the
benzene/zeolite system at 523 K for 10 h to render the adsorbate distribution
uniform [316]. Nevertheless, at loading of one molecule per supercage a
clustering of benzene molecules is already observed [303]. For infrared
studies, smaller amounts of zeolite are required (ca. 15-20 mg of zeolite). It
was carefully checked that 1-h equilibration at room temperature is conven-
ient [265,266] and gives the same results as equilibration times between 48
h and 1 week at 300 K or 393 K.

. Window

Cation

FIG. 23. Clustering of benzene molecules in NaY upon interaction with the
cations (0)and 12R windows (from Ref. 299).
BASIC ZEOLITES 565

B. Other Hydrocarbons

The adsorption of a variety of hydrocarbons in nonprotonic zeolites has


been studied on a molecular scale. The mobility of adsorbed molecules (such
as hexamethylbenzene or naphtalene) was studied by 'H NMR [290,300] and
other techniques including I3C NMR [317].
Only a few studies have dealt with the location of the adsorbates and
their interaction with the acid or basic sites of the zeolite. In that sense, it
was mentioned early that the adsorption of hydrocarbons in NaY slightly
distorts the framework [40]. Proton multiple quantum NMR gives information
on the distribution of molecules (isolated molecules or clusters) in NaY
[317,318]. Powder neutron diffraction locates p-xylene and m-xylene in the
Downloaded by [Yale University Library] at 01:03 15 August 2013

BaX zeolite on the Ba cations in SII sites [319]. At high loading a part of
the p-xylene molecules do not interact through the 11-electron with the Ba
cations but are located in the neighborhood of the 12R window where they
may be slightly tilted. m-Xylene is always located near the Ba cations [319].
Infrared studies of toluene, xylenes, and ethylbenzene adsorbed in NaY and
KY indicate that these molecules interact only with the Na or K cations [143].
Similarly, ethylbenzene interacts mainly with the cations in Na-beta and Cs-
beta [241]. It follows that one may predict that in adsorption of mixtures
some competitive adsorption will exist not only between molecules adsorbed
on the same type of sites (cations) but also between the aromatics (C, and
C,) and molecules adsorbed on both cations and window sites. For instance,
one may expect that ethylbenzene may displace benzene from cation sites
moving the C , to the 12R window. This is described in detail in Sect. V1.C.
One may also expect that the location and interaction of molecules in
the cages and channels of zeolites strongly affect their activation for catalytic
reactions. Acidic and basic sites are known as catalytic centers but very little
is known at a molecular level on the zeolite-reactant interaction. One ex-
ample is provided in the case of the alkylation of toluene with methanol,
where the possible interaction of each reactant with the zeolite is proposed
(Fig. 24) [320a].
In the absence of acidic or basic adsorption sites, no specific zeolite-
adsorbate interaction is noted. This is the case of benzene and o-xylene ad-
sorption in AlP0,-5 studied by infrared [268]. This situation is similar to that
of benzene in siliceous faujasite [306,308,309]. The aromatic fills the pores
where it is highly mobile.
A general comment on the adsorption of hydrocarbons on zeolites is
related to the studies of heats of adsorption. These heats include the inter-
actions with all possible sites (including cations and oxygens) and give an
overall view at any loading. Since the early results of Dzhigit et al. [321] a
large number of works have related the initial heat of adsorption of hydro-
carbons to the charge density of cations. This was, for instance, verified for
methane adsorbed on Y or X exchanged with divalent cations (except Mg)
[322]. One may speculate as to whether this relation is as simple as it looks.
566 BARTHOMEUF

-I--
6-7h
Downloaded by [Yale University Library] at 01:03 15 August 2013

FIG. 24. Interaction of formaldehyde and toluene with acid and basic sites
in zeolites: (a) from calculations [320]; (b) scheme (a) modified to represent oxygen
basic sites and cationic Lewis acid sites (CS+,for instance). The oxygen and cations
may belong to the zeolite or to added clusters of oxides.

The adsorption may well be connected to an overall property of the zeolite,


such as electronegativity, as it was shown to occur for the activation energy
of desorption [323]. This would take into account the possible adsorption on
acidic and basic sites.

C. Other Molecules

Compared to the high level of information available at a molecular scale


on the state of hydrocarbons (mainly aromatics) adsorbed on nonprotonic
zeolites, very little is known for other molecules. A large number of works
have been devoted to the study of pyridine adsorption on protonic zeolites.
With alkaline forms, pyridine was proposed to interact with zeolitic cations
as suggested a long time ago from infrared studies [94], the strength of the
bonding increasing with the Lewis acidity of the cations. However, powder
neutron diffraction at 4 K does not locate pyridine at low coverage near the
cation in NaY but in the 12R window [324]. In fact, the pyridine adsorbed
on the cations might not be detected by this technique. NMR spectra and 13C
chemical shifts reveal that the nature of pyridine closely resembles that of
BASIC ZEOLITES 567

the neat liquid in NaY or siliceous faujasite [308]. In gallozeolite L, powder


neutron diffraction shows that the nitrogen atom is coordinated to the K
cation, but the molecule also lies close to the channel wall in order to optimize
its interaction with the framework [325]. The situation is still unclear and the
results might depend on the time scale of each physical method used to
characterize the absorbed phase.
Deuterated aniline adsorbed in a YbNaY has two locations. The first site
is the Na cation in the S,, position where aniline interacts through its IT-
electrons. The second is in the 12R window, where the interactions with the
framework oxygen cause a slight deformation of the framework geometry.
The ND, group lies above the ring plane formed by the 0, oxygen [326] (see
Fig. 1). Pyrrole, an amphoteric molecule, was proposed from infrared studies
Downloaded by [Yale University Library] at 01:03 15 August 2013

to interact through its NH group with basic oxygens in nonprotonic zeolites


[24,25]. Additional XPS studies of the N1, changes in binding energy with
cation nature in X and Y nicely confirm this first suggestion [196,219].
In conclusion, the growing field of research on the zeolite-molecule
interaction at the molecular scale is providing extremely valuable information.
It is becoming possible, for example, to understand and “see” the mode of
interaction, the zeolite sites which are important, and the location of the
adsorbate with regards to the site. Combined experimental and theoretical
approaches are important. They bring a detailed and powerful understanding
of the adsorption phenomenon at the molecular level which can be extended
to a more global view.

VI. BASICITY AND ADSORPTION

A. General Aspects

Nonprotonic zeolites are commonly used in adsorption processes in or-


der to avoid any catalytic transformation of the adsorbates on acid sites. Many
zeolites (such as A, X, Y, mordenite, and beta) in various cationic forms are
described as adsorbents for practical applications or fundamental studies
[327,328]. In the separation of mixtures (liquids or gases), two main param-
eters govern the separation process. At first, kinetic limitations are important
in favoring the diffusion of some reactants. This is related to the sieving,
molecular packing, and shape selectivity effects. A second feature is con-
nected to the chemical properties of the reactants. Subtle differences in the
properties of the various adsorbed components of a mixture may generate
different interactions with the adsorption sites in the zeolites. The acidobasic
character of the adsorbent may be important at this stage. Only this second
aspect is considered here and some examples are given. The basicity of the
zeolite is easily expressed through the calculated Sintzeolite intermediate elec-
tronegativity parameter. It was shown that for various compounds (such as
568 BARTHOMEUF

hydrocarbons, alcohols, water), the activation energy of desorption can be


correlated to Sin,[323].
The study of benzene adsorption in faujasites showed that the aromatic
may be adsorbed on cations in S,, sites and/or in the 12R window
[143,183,241,265-267,287,295,297-299,306]. It was shown that the location
depends on a balance between the acidity of the cation and the basicity of
the oxygen, and also on the loading [183,241,265-2691.

B. Separation of Mixtures

In a mixture of components the selectivity of adsorption is expressed by


Downloaded by [Yale University Library] at 01:03 15 August 2013

a separation factor a between two components A and B [268,329-3351. The


aAiB value corresponds to the ratio of the amounts (or concentrations) in the
zeolite and in the liquid mixture, where aNB= (Azeol/Aliq)/(Bzeol/Bliq). Values
of a higher than 1 indicate preferential adsorption of A over B in the zeolite.
In the adsorption of naphtalenes, the separation coefficient a between
methyl-2 and methyl-1 naphtalenes ( c x ~ M N / o ~ ~ M N ) increases as Sin,decreases
upon alkaline and alkaline earth ion exchange. Two different curves are ob-
tained for X and Y [336]. The results may be interpreted as indicating a
preference for adsorption of isomer 2 as the basicity of the zeolite rises.
Another example concerns the separation of C, aromatics, which is of
major importance for industry. The three 0-,m-, and p-xylenes and ethylben-
zene have very close boiling points. Selective adsorption on zeolites is very
attractive in that it avoids difficult distillation or crystallization separations.
A very large number of patents are related to the influence of the zeolite
structure, the cation, and the S i / A ratio on the selectivity. In the faujasite
series, KY preferentially adsorbs p-xylene [330,337], NaY m-xylene [334],
and RbX and CsX ethylbenzene [333]. Ethylbenzene is rejected from CaY
or CaX 13321. Several parameters must be invoked to explain the results such
as the packing of aromatics [329,338], the basicity of hydrocarbons, and the
ionic potential of exchanged cations [338]. Significant trends were obtained
between separation efficiency and acid-base properties [278,337]. In the X
and Y zeolites it was proposed [278] that the order of preference for specific
isomers may be related to a balance between the zeolite acidobasicity ex-
pressed as Sin$ and the relative ?r basicity of the aromatics [162,278]. This
last factor increases from ethylbenzene (1.5) to p-xylene (1.6), o-xylene (1.8),
and finally to m-xylene (2.0) (Table 9) [339]. This correlation is seen, for
instance, in Table 9, where for a Cs-beta zeolite (i.e., for given acid-base
adsorbent properties) ethylbenzene is progressively preferred over the other
aromatics as their basicity rises [335]. The zeolite which has basic character
[241] rejects the more basic aromatics. Other kinds of examples are given in
Table 10, which reports changes in selectivities of various faujasites for a
given feed (equimolar mixture of C, aromatics). Fourteen zeolites of the fau-
jasite type with different Al and cation contents are listed in Table 10. The
BASIC ZEOLITES 569

TABLE 9
Changes on Cs-Beta in Separation Factor of Ethylbenzene and Various Aromatics
(a = eb/aromatic)" as a Function of the IT Basicity of Aromatics
Relative'
Aromatic a = eb/aromatic" IT basicity
C6 Benzene 0.6 1.o
G Toluene 1.9 1.5
C8 p-Xylene 2.1 1.6
o-Xylene 6.1 1.8
m-Xylene 7.6 2.0
c,, Prehnitene 40 2.6
Isodurene 50 2.7
Downloaded by [Yale University Library] at 01:03 15 August 2013

"eb, ethylbenzene.
bEquimolar feed of Cs aromatics mixed with another aromatic in the ratio 1:2. Sepa-
ration measurement at room temperature (from Ref. 335).
'From Ref. 339. Value for ethylbenzene: 1.5

zeolites are ranked in the increasing order of Sint(decreasing basicity). The


more basic zeolites (samples 1-9) are all selective for ethylbenzene whatever
their Al level (X or Y zeolites), the nature of the alkaline or alkaline earth
cation (Mg, sample 8), or the presence or absence of Ga in the framework
(samples 3 and 7). The remaining zeolites (samples 10-12), with lower ba-
sicity, are no longer selective for ethylbenzene and prefer p-xylene, which is
less basic than m- and o-xylene. A sample with a different Al content is
included. Ba zeolites are known to be p-xylene selective [331], as seen, for
example, with sample 12. The Ba cation is more electronegative than Mg,
which may explain the different selectivities of samples 8 and 12. NaY and
LiY prefer rn-xylene [334]. They are the least basic zeolites and are even
rather acidic, Fig. 11. This selectivity for samples 13 and 14 for the most
basic aromatic m-xylene, is in line with the genera1 trends reported here.
Figure 25 describes the same trends for the zeolite beta, the more basic Cs
and Rb forms being the most selective for ethylbenzene [335]. The gallozeo-
lite beta gives similar selectivities as the Al form [335b], as was observed
for faujasite (Table 10, samples 3, 7).
All these results are summarized in Scheme 1 showing the opposite
trends for selectivity upon a change in both the acidobasicity of zeolites and
the basicity of the preferred adsorbed aromatic. The more acidic zeolites (high
Slnt)prefer the more basic aromatics, and reciprocally a high zeolite basicity
(low Sint)gives a high selectivity for the less basic hydrocarbons. The results
of Table 10 show that the selectivities do not depend on the size or number
of the cations (determined by the Al content). It should not depend strongly
on steric effects. The importance of the zeolite acid-basic properties is then
critical. Another parameter which appears to be effective is seen from Fig.
25. Beta zeolite has a calculated S,,,values of around 4.0-4.2 (mean oxygen
Downloaded by [Yale University Library] at 01:03 15 August 2013

TABLE 10
Faujasite Zeolites: Intermediate Electronegativity Sin,,Oxygen Charge, and Selectivity in C, Aromatics
Separation (Feed: Equimolar C, Aromatics)”
Oxygen
Zeolite Szeol charge eblpx eblmx eb/ox Selectivity Ref.
1 3.012 -0.463 1.4 1.5 1.5 333c
2 3.027 -0.461 3.3 4.1 1.8 333c
3 3.028 -0.461 3.1 4.2 2.2 333c
4 3.070 -0.451 2.3 3.6 1.9 333c
5 3.081 -0.448 2.3 2.5 1.6 eb 333c
6 3.081 -0.448 2.3 1.7 3.2 333c
7 3.082 -0.448 2.4 3.3 1.6 333c
8 3.185 -0.427 1.9 1.6 2.4 333c
9 3.251 -0.413 1.2 1.2 2.6 333c
pxleb pxlmx pxlox
10 K54NaZ 3.382 -0.385 2.0 7.8 7.0 PX 334b
11 &8(-410z)48(Si0z)144 3.499 -0.360 3.7 6.2 8.8 334b
12 Ba3,Rb9NalnX 3.510 -0.358 1.3 2.4 1.9 331
mxleb mxlpx mx/ox
13 NaY 3.538 -0.352 6.2 3.5 1.9 mx 334b
14 LiY 3.550 -0.349 5.0 2.5 1.2 334a
oxleb oxtpx oxlmx
15 3.3 3.2 3.3 ox 268
“eb, ethylbenzene; px, para-xylene; mx, meta-xylene; ox, ortho-xylene.
bThe S,,,and oxygen charge value of AlP04-5 cannot be compared to those of Si-A1 zeolites due to the presence of
3
M
a new atom (P) in the formula. s
BASIC ZEOLITES 571

CsRb K Na H
Downloaded by [Yale University Library] at 01:03 15 August 2013

4.0 4.1 4.2 5 It

FIG. 25. Changes in the selectivity a in C, aromatic separation at room tem-


perature as a function of zeolite beta intermediate electronegativity Sint(equimolar
feed of the four C,). (A) eblmx, (B) eblox, (C) eblpx (from Ref. 335).

Zeolite S,,, < >


Acidity Basicity
Aromatics, mx ox PX eb"
IT basicity <
2.0 1.8 1.6 1.5
Selectivity for mx PX eb
adsorption on
zeolites Acidic zeolites Mildly basic zeolite Basic zeolites

"eb, ethylbenzene; px, ox mx: para-, ortho-, meta-xylene.

SCHEME 1. Dependence on acidobasicity of selectivity in the separation of


C, aromatics.

charge around 0.20), in a range where Fig. 11 suggests no basic character.


The experimental selectivities reported in Fig. 25 are characteristic of basic
zeolites. In addition, studies of pyrrole adsorption confirm the basic properties
of Cs-beta and Na-beta [241]. One may then consider that either defects site
572 BARTHOMEUF

in beta have basic properties and/or that the average value of oxygen charge
does not always apply to the actual atoms “seeing” the adsorbates. Due to
the correlation between bond angles and oxygen charge [MI, it may well be
that some TOT angles, characteristic of the beta crystalline structure, generate
a negative charge which becomes more important than the average calculated
one. These oxygens would be the ones involved in the adsorption of aro-
matics. The comparison of one family of zeolites to another one must take
this structured parameter into consideration.
The molecular approach described in Sect. V may be of great help in
understanding which zeolite atoms (oxygen, cations) are interacting with the
adsorbates. In this view, Table 10 reports for comparison the selectivity of
AlP0,-5 in the C, aromatics separation [268]. This zeolite material, which is
Downloaded by [Yale University Library] at 01:03 15 August 2013

neutral with no cations and no basic oxygens, preferentially adsorbs o-xylene.


This arises from the absence of chemical interaction. The physical adsorption
directs the selectivity towards the isomer with the highest cohesive energy.

C. Competition of Adsorption

The separation of mixtures just described results from a competition of


adsorption of the different components on the same or on various adsorption
sites. The practical use of adsorbents in industry always involves competitive
adsorption, not only between the adsorbates but also with the desorbent
[331-3351. A very large number of patents describe the choice of the best
desorbent which removes the adsorbed phase from the zeolite and improves
the separation coefficients a. The success of a separation process depends,
then, as much on the zeolite as on this desorbent. The desorbent may be an
organic (such as a hydrocarbon or alcohol) or an inorganic (such as water or
ammonia) molecule. As an example, Table 11 gives the influence of an ad-
ditional component on the a coefficients for the separation of the four C ,

TABLE 11
Effect in KY of an Additional Component on the Separation Coefficients a for an
Equimolar Feed of C, Aromatics“
Selectivities‘
Feed pxleb pxlmx pxlox
4C8 1.7 6.7 7.5
4C8 + benzeneb 0.5 1.5 1.9
4C, + methanol 2.2 5.5 5.1
4C, + pyrrole 2.5 7.9 6.5
4C8 + pyridine 2.4 7.7 9.1
“From Ref. 334b.
bThe selectivities with regard to ethylbenzene are eb/px = 2, eb/mx = 3, eb/ox = 3.8
‘eb, ethylbenzene; px, para-xylene; mx, meta-xylene; ox, ortho-xylene.
BASIC ZEOLITES 573

aromatics [334b]. Benzene gives strong modifications with even an inversion


of the ratio p-xylenelethylbenzene, the KY zeolite becoming ethylbenzene
selective. This reveals a strong competition between benzene and the C, aro-
matics. Benzene is the least basic, smallest, and more symmetric of the aro-
matics considered. This favors its preferential and stronger adsorption in the
12R window [1431. Considering the other aromatics, for instance, ethylben-
zene, it was shown by infrared spectroscopy that this hydrocarbon is adsorbed
mainly on cations in NaY and KY, and not in the 12R window [340]. The
simultaneous presence of benzene and ethylbenzene in the KY supercages
results in the adsorption of ethylbenzene on the cations only, while benzene
is shared between both cations and oxygens of the 12R window as described
in Sects. IV.B.2.c and V A full redistribution of aromatics occurs in the cage.
Downloaded by [Yale University Library] at 01:03 15 August 2013

A desorbent like methanol, pyrrole, or pyridine, which does not interact very
specifically with the oxygen of the 12R window, induces considerably smaller
changes in the ranking of selectivities.
Fundamental studies have distinguished the adsorption of benzene on
cations and on the oxygen of the 12R window [143,183,241,265,266,268,
269,295,2991. It was shown by infrared spectroscopy of adsorbed benzene
that the acidic sites of NaY (cations) may be poisoned by NH, [267]. The
competitive adsorption of both NH, and benzene (Fig. 26) leads to a dis-
placement of the aromatic from the acidic cationic sites (bands 1848 and
1986 cm-') to the 12R window bands (1880 and 2017 cm-l) [267]. A similar
effect was also observed for a dealuminated NaYD [143]. In contrast to this,
addition of HCl gas to CsNaX (a basic material) displaces the benzene in-

I1479

2200 2000 1900 1800 1700 1500 1400


v (cm-')

FIG. 26. CH out-of-plane infrared bands of benzene adsorbed on NaY for


pure benzene (1 molecule/unit cell) (A) and after addition of NH, so that NH,:C,H,
in the infrared cell equals 3.8 (B), 4.8 (C), 6.7 (D), and 36 (E) (from Ref. 267).
574 BARTHOMEUF

teracting with the basic oxygen of the 12R aperture (bands 1922 and 2053
cm-') to cationic sites (bands 1850 and 1995 cm-') (Fig. 27) [267]. Adsorp-
tion of HC1 on NaY gives a comparable effect [143]. It may be noted that
the proton of HCl interacts with the basic framework oxygens. This gives
OH groups detected by infrared [143]. With regards to water, it is known that
in separation processes its presence in small amounts may be necessary. In-
frared studies show that water interacts with both framework oxygens and
cations, resulting in a redistribution of benzene on the two adsorption sites
of NaY [143]. The adsorption of water on cations appears to be stronger than
that of ammonia [143]. These results confirm the importance of the acid-
basic character of the zeolite in determining the adsorption properties, and its
consequences in the interactions with competitive adsorbates and the redis-
Downloaded by [Yale University Library] at 01:03 15 August 2013

tribution of adsorbates on zeolite sites.


Figure 28 extends the results of Fig. 26 and gives a quantitative view
of this competitive adsorption for NaY [143]. At a loading of 1 molecule of
benzene per supercage-i.e., 8 per unit cell-around 20 molecules of NH,
start to displace the benzene from its location on Na cations in S,, sites and
move it to the 12R window. However, no benzene is desorbed, the total
amount of aromatic in the zeolite being the sum of the aromatic adsorbed on
the cation and on the oxygen of the large aperture (curve c, Fig. 28). In
contrast to Fig. 28, at higher aromatic loadings-for instance, 3.7 molecules
per supercage, which is close to the saturation limit of the cage (4.5
molecules)-the introduction of 50 molecules of NH, per unit cell leads to
the same displacement from the cation to the 12R window and in addition to
the departure of some benzene molecules into the gas phase [143]. Before

I1479
1850
1995 I
2053 I 1922 I

J
2200
I
2000
I
1900 1800
v (cm-')
-
1500 1400

FIG. 27. CH out-of-plane infrared bands of benzene adsorbed on Cs-NaX


for pure benzene (1.1 molecule/unit cell) (A) and after addition of HCI so that HCL:
C6H, in the infrared cell equals 1.2 (B) and 3.5 (C). From Ref. 267.
BASIC ZEOLITES 575

25 50
Downloaded by [Yale University Library] at 01:03 15 August 2013

0
adsorbed NH3 (mo1Ju.c.)

FIG. 28. Changes for NaY in the amount of benzene adsorbed per supercage
on cations (A) or oxygen (B) as a function of NH, introduced per unit cell. Total
amount adsorbed on cations and oxygen (C). The dashed line represents the benzene
introduced initially (1 mol/supercage) (from Ref. 143).

and after adsorption of NH, the population of benzene in its two adsorption
sites is, respectively, 3.3 and 2.0 molecules for the S,, Na cations and 0.4 and
0.7 for the 12R window. There is an increase of 75% of benzene retained in
the window while the total amount of benzene adsorbed decreases from 3.3 +
0.4 = 3.7 to 2.0 + 0.7 = 2.7 molecules per cage [143]. These results indicate
that the presence of NH, completely modifies the distribution of benzene in
the two adsorption sites.
Slightly different results are obtained for benzene adsorption on Na-
EMT. This zeolite adsorbs benzene mainly on cations [292] at any loading.
In Na-EMT (Fig. 29), containing 4.5 molecules of benzene per supercage
(i.e., close to the cage filling), addition of NH3 displaces the aromatic from
the cations to the 12R window, creating in this way new adsorption sites. No
benzene is desorbed from the zeolite even at high benzene loadings [153]. It
is suggested that the adsorption of NH, on the cation may change the distri-
bution of charges on the framework, enhancing the basicity of the oxygen
atoms [153]. With KL zeolite saturated with benzene, all the aromatic mol-
ecules are adsorbed on cations [315]. Upon adsorption of NH, they move to
the gas phase but not to the 12R window despite the fact that the calculated
(Fig. 11) and experimental [24] basicity of oxygen in KL is not very different
from that of Y or EMT zeolites [153]. Other parameters are very likely im-
portant in this case, for instance, the field gradient as already proposed to
explain adsorption of benzene in the 12R window [198], the presence of
specific TOT angles in the L channel with a small charge on oxygen, or steric
constraints which prevent the formation of benzene cluster like those de-
scribed in Fig. 23. All these results obtained with faujasites and EMT and L
576 BARTHOMEUF

5.0
c!
=eE 4.0
v

3.0
4t
2 2.0
2
s
5 1.0
2
0.0
0 20 40 60 80
adsorbed NH3 (mo1Ju.c)
Downloaded by [Yale University Library] at 01:03 15 August 2013

FIG. 29. Changes for Na-EMT in the amount of benzene adsorbed per su-
percage on cations (A) or oxygen (B) as a function of NH, introduced per unit cell.
Total amount adsorbed on cations and oxygen (C). The dashed line represents the
benzene introduced initially (4.5 mol/supercage) (from Ref. 153).

zeolites suggest that a molecular recognition effect governs the sitting of


benzene in the 12R aperture in zeolites. Both chemical properties and steric
constraints are important.
In conclusion, all the above examples of adsorbate displacements from
site to site or outside the zeolite may explain the influence of desorbents
(such as NH,, water, hydrocarbons, or aromatics) in the industrial separation
of mixtures. The competition of adsorption is then important to consider at
the molecular level in order to understand why some desorbents may greatly
change the selectivities of separations. This is an emerging field in funda-
mental research, in both experimental or theoretical studies. Several param-
eters have to be considered, including:
Basicity of framework oxygen
Lewis acidity of the cations
Steric configuration of the zeolite sites
Acidobasicity of the adsorbates
Shape of the adsorbates in the adsorbed state on specific sites
Competition of adsorption.
All these parameters govern the molecular recognition effect.

VII. BASICITY AND CATALYSIS

The success of zeolites in catalysis results from the combination of var-


ious properties which are very important for industrial applications. The main
use of the catalysts involves the protonic forms. Nevertheless, in the nonpro-
BASIC ZEOLITES 577

tonic materials some general features of zeolite properties also exist which
are effective and important for catalysis. Among these common characteris-
tics, the thermal stability of the crystalline structure is comparable for the
two chemical forms, acidic or not. It is very often higher in the metal cation
exchanged zeolites. The crystalline structure itself, which determines both
geometric (porosity) and energetic (electrostatic field and field gradient) prop-
erties, has also to be considered for catalysis in the field of basic zeolites. As
in protonic solids, the pore size (Table 1) and the shape of the channels or
cages determine molecular sieving and shape-selective catalysis. Neverthe-
less, the generation of basicity usually requires the exchange of protons by
large alkaline cations (Table 3). This reduces the pore volume and the win-
dows aperture [143]. Up until now this aspect has not been considered sys-
Downloaded by [Yale University Library] at 01:03 15 August 2013

tematically in order to direct the selectivity in basic catalysis with zeolites.


The addition of oxides or alkali metal clusters to increase the basicity of the
zeolite may decrease even more the available space. In this line, a first hy-
pothesis suggested that in Nay, (Cs,O), clusters may block access to the cages
[62a]. It was shown more recently [62b] that this effect might not be the
major one. A strong zeolite-oxide cluster interaction may strongly modify
the catalytic properties of the materials. Electrostatic fields and electrostatic
field gradients [22,341] in protonic zeolites, which polarize or ionize mole-
cules [22,341], also exist in nonprotonic forms [327]. As a consequence, one
may expect that reactants and products in cages are disturbed by the presence
of charges located on the walls of the cages and channels, and that this effect
strongly depends on the nature, content, location, and valency of the cations.
This should modify the selectivities in catalysis, as was experimentally evi-
denced for protonic zeolites (Si-Al forms [22,342] or SAPOs [343]). Finally,
and similar to the case of protonic catalysts, basic zeolites may replace dan-
gerous, corrosive, or toxic catalysts in some cases. They are potential envi-
ronmentally friendly catalysts.
With regards to catalyst deactivation, one may not expect coking since
it is due to carbenium ion reactions. An aging typical of anionic catalysis
may result, for instance, from polymerization reactions [344, 3451. Alterna-
tively, poisoning of basic sites by C02, which is very often present, may
considerably decrease the catalytic activity.
Venuto and Landis first reported a large variety of reactions catalyzed
by nonprotonic zeolites [346]. Despite the fact that the basic character of
these zeolites was not pointed out, its importance may be deduced from the
interesting results described in this review. Various papers report the catalytic
performances of zeolites with examples on nonprotonic materials [224b,347-
3571. More specific are reviews devoted to reactions catalyzed in basic zeo-
lites [134,277,278,358-3601. It is outside the scope of this paper to consider
all the information available. Only the more recent results which take into
account the basic sites of the zeolites are discussed. Kinetic and mechanistic
results are not within the scope of the present paper, since they are described
578 BARTHOMEUF

in other reviews. This part is focused much more on the zeolite properties
which generate specific catalysis rather than on the reactions themselves.
The examples given below are classified with regards to the type of
reaction involved. For some of them the addition of oxide or alkaline metal
clusters is needed to significantly improve the performances. This is exem-
plified, for instance, by the alkylation of toluene with methanol or formal-
dehyde [6,7,47,49,53-56,194,361-3641, the dehydrogenation of isopropanol
to acetone [46,360], or the reactions involving some transition metal carbon-
yls [58,122,123]. In some cases the presence of such an extraphase is not
mentioned per se but it exists anyway in the materials [63,365]. For some
other reactions the intrinsic framework basicity is strong enough to direct the
activity and selectivity, like in alkylation of aniline with methanol or di-
Downloaded by [Yale University Library] at 01:03 15 August 2013

methylcarbonate [ 153,345,366,3671.

A. Alkylation Reactions

Several reviews have described the alkylation properties of zeolites


[224~,277,278,359,360].Very different selectivities are obtained with acidic
or basic zeolites as one may expect from cationic or anionic catalysis [1,4].
The first reaction described, showing very clearly different products, is the
alkylation of toluene with methanol, giving either xylenes or ethylbenzene
plus styrene with acidic or basic faujasites, respectively [6]. Both kinds of
products are interesting in industry and research is still active in the two fields
of acidic and basic zeolites. Other alkylation reactions are also described in
basic zeolites.

1. Alkylation of Toluene with Methanol


It is well accepted that formaldehyde issued from methanol dehydrog-
enation is the alkylating agent. The role of the zeolite intermediate electro-
negativity Sintis very nicely shown in Fig. 30 after gathering the results from
12 references [194]. At high Sint(acidity range) xylenes are formed, while
side-chain alkylation to ethylbenzene and styrene is favored with basic zeo-
lites (low Sint)[194]. Improvements of the catalytic performances may be
achieved by various modifications of the zeolites. The presence of excess
KOH or CsOH was soon reported as enhancing the activity and the selectivity
to ethylbenzene, particularly for the X form [53,54]. Decomposition of cesium
acetate in CsNaY leads to cesium oxide inside the zeolite [47], increasing the
basic character and hence side-chain alkylation at the expense of formalde-
hyde decomposition [362]. The question may arise of the possible formation
of CsOH with water issued from the reaction. The enhancement is also very
effective in CsX [49]. This is in line with the previous results in Refs. 53
and 54. The hypothesis of a bifunctional character of catalysts containing an
excess of alkaline ions is described in CsX zeolite. One type of site consists
BASIC ZEOLITES 579

40

-e 30

2* 20

10

0
2.8 3.2 3.6 4.0 Sint
BASIC RANGE ACIDIC RANGE
Downloaded by [Yale University Library] at 01:03 15 August 2013

FIG. 30. Dependence of yield of either ethylbenzene and styrene (0)or xy-
lenes (0)upon the intermediate electronegativity of various zeolites (from Ref. 194).

of Cs' ions and the other type of oxide clusters (Cs,O or ZnO) [57]. The
bifunctional character is also evidenced with K,O as added oxide [364].
The activity is enhanced alternatively by adding alkaline metal vapor
using NaN, [55,56] or by heating the zeolites above 600 K [59,368]. The
alkaline metals supported in basic zeolites increase the basicity of close
framework oxygens [72-751. One has nevertheless to consider that the water
formed during the reaction very likely reacts with the alkaline metals giving
the corresponding hydroxides. One would then come back to the case of
hydroxide clusters. Modifications of alkaline faujasites with boron increase
the side-chain alkylation and the overall activity of the catalysts [363,370-
3741. The importance of carbonates in this reaction is emphasized. A decrease
or an increase [369] of activity are both reported. Various other additives are
described (see Ref. 278) such as Cu, for instance [54]. Their role is to de-
hydrogenate methanol to formaldehyde, a very good alkylating agent. An
improvement of catalytic side-chain alkylation is also observed for binary
zeolitic systems. The W K Z S M - 5 catalyst is more active than the separate
zeolites [375].
Independently of any modification of the zeolites or of the global influ-
ence of the catalyst reflected, for instance, in the Sintapproach, it is considered
that not only basic sites but also acid sites are involved in the side-chain
alkylation of alkylbenzenes [320,362,363,376,377]. The two sites may arise
from the presence of two different cations, for instance, Li and Rb, in X
zeolites [377]. This may be related to the description of localized sites in
zeolites with two types of exchanged cations [195,255]. This idea of acid-
base cooperative sites is used by various authors [251,375,3781, toluene and
methanol being adsorbed at different sites [378,379]. Quantum-chemical cal-
culations have indicated that specific configuration of acid and basic sites
580 BARTHOMEUF

with steric restrictions are required [320]. Figure 24(a), proposed in Ref. 320a,
explains that the basic site determines the selectivity of the side-chain alkyl-
ation of toluene by interaction with the methyl group of toluene. Toluene is
adsorbed and stabilized on the acid site by its interaction with the benzenic
ring. Formaldehyde is adsorbed on the protons H' [320a]. The interaction of
aromatics described in Sect. V for basic zeolites indicates that the strength
of the interaction between the cation (Lewis acids) and the IT electron of the
ring is the lowest for the cations with the lowest electronegativity such as K,
Rb, and Cs (Table 4). The observed increase of catalytic alkylation in the
presence of oxide (or hydroxide) clusters entrapped in the zeolite suggests
that the cations of these clusters may also participate in the adsorption of
toluene or formaldehyde. The basic sites 02-interacting with the methyl
Downloaded by [Yale University Library] at 01:03 15 August 2013

group may belong to the zeolite framework and/or the oxide clusters. In order
to explain the activation of the methyl group which generates the side-chain
alkylation, a rather strong basicity is required simultaneously with a weak
cationic Lewis acidity. A balance between the two functions is needed. This
is obtained in the K-, Rb-, or Cs-exchanged zeolites (Table 2). The Li and
Na zeolites give too strong a 7r-electron-cation interaction and too weak a
methyl-weakly basic oxygen interaction (Table 2). The alkylation then occurs
on the ring. Figure 24(b) suggests a scheme to sum up these results.
More information on the reaction mechanism may be obtained experi-
mentally using I3C NMR. It shows a restricted mobility of the aromatic in
the CsX material. This suggests that ring alkylation is hindered by the pres-
ence of cesium cations due to a restriction in the formation of the required
transition states [380]. The same technique distinguishes two alkylating
agents, formaldehyde for the side-chain alkylation and methylcarbocations
for the ring alkylation. In CsX the zeolite plays a crucial role by binding the
highly reactive carbocation as surface methoxy groups. This prevents ring
alkylation from occurring on CsX [381].

2. Alkylation of Alkylbenzenes

The selective alkylation of alkylbenzenes in acid and basic zeolites was


reviewed in Ref. 382. As in the case of toluene, the side-chain alkylation of
p-xylene with methanol requires the presence of an assembly of acid and base
sites which act cooperatively [376,377]. The two types of sites were char-
acterized by an infrared study using probe molecules [383]. The activity is
promoted by addition of Cu" or Ago, which catalyzes the dehydrogenation
of methanol to formaldehyde [383]. The side-chain methylation of ethylben-
zene occurs on KX while the ring alkylation takes place on HX. On KX at
500°C, alkylation and dealkylation proceed, respectively, via a carbanion and
carbonium ion mechanism, whereas dehydrogenation and demethylation oc-
cur via a free-radical mechanism [384]. The side-chain alkylation of alkyl-
benzenes is enhanced by treatment of NaX, Nay, or NaL with NaN, [74].
BASIC ZEOLITES 581

3. Alkylation of Heteroaromatics
The alkylation of N-containing aromatics may occur on N or C atoms
[351]. Examples are given for pyridine or picolines. The side-chain alkylation
of picolines on NaY or CsY occurs simultaneously with that of the nucleus
[350].
The alkylation of aromatic amines, for instance, aniline, may be carried
out with methanol [153,344,345,385,386], dimethylcarbonate [366,367], or
2-propanol [387] on basic zeolites. Basic zeolites appear as environmentally
friendly catalysts since the methylation reactions are usually carried out in
the liquid phase by using methyl halides and dimethylsulfate, which are cor-
rosive and toxic chemicals. A general trend observed when methanol is used
Downloaded by [Yale University Library] at 01:03 15 August 2013

as an alkylating agent is the increased N-alkylation as the basicity of the


zeolite rises [153,345,385,386]. For instance, Table 12 shows the highest
selectivity to N-alkylates for the faujasites X or Y exchanged with K, Rb, or
Cs. As proposed for the alkylation of toluene with methanol [320], both basic
and acid sites could participate in the N-alkylation. The more selective cat-
alysts would require a balance between Lewis cation acidity and framework
oxygen basicity. In addition, a very high aging, due very likely to polymer-
ization reactions, modifies the order of reactivities, particularly for the more
basic zeolites [153,345]. By contrast, the alkylation of aniline with dimethyl-
carbonate is a significantly simpler reaction on basic zeolites. In defined re-
action conditions only methylation on the N atom occurs, with no ring al-

TABLE 12
Alkylation of Aniline with Methanol"": Selectivities to N-Alkylates at Isoconversion
Conversion (%)
Zeolites 10 15 20 27 40 60
c
LiY 53 20
c
NaY 35 41
c c
KY 100
c
RbY 60 66
CSY 100 96 96
c
LiX 51 60 58 30
c
NaX 43 47 46 20
c
Kx 100 85 70 76
c C
RbX 100 88 85
csx 80 73 77 86 92
"From Refs. 153, 345.
bCatalyst pretreatment: 723 K, reaction temperatufe: 673 K, relative pressure aniline/
methanol: 1/3, carrier gas: Nz,N-alkylates: N-methyl-, NJV-dimethylaniline and N 8 , -
dimethylparatoluidine.
'No results at this conversion.
582 BARTHOMEUF

kylation [366,367] and the aging is quite low [367]. Very different zeolites
were tested in different cationic forms (X and Y, L, mordenite, and ZSM-5
[366]; or X, Y, beta and EMT [367]). KY [366] and KEMT [367] are the
most active and selective for the production of N-methylaniline at 408 or 453
K (Table 13). No straight correlation with basicity alone is observed. At high
temperature (503 K), the more basic KX and CsX are very active and selec-
tive for the dimethylation on the nitrogen atom giving N,N-dimethylaniline.
The results suggest that the reaction at high temperature involves mainly the
basic sites for the catalysts [367] while at low temperature both acid and
basic sites are required [366,367].
Dimethylcarbonate appears to be a very good alkylating agent. It selec-
tively gives anisole by 0-methylation of phenol on NaX at 553 K (82%
Downloaded by [Yale University Library] at 01:03 15 August 2013

phenol conversion, 93% selectivity to anisole). A deactivation or lower ac-


tivities are observed with the other X or Y catalysts tested [388]. The results
indicate that the basicity is essential for 0-methylation. The side-chain al-
kylation of phenylacetonitrile (PAN) to a-methylphenylacetonitrile (MPAN)
proceeds on CsX at 623 K to gives 27% yield of MPAN if methanol is used.
With dimethylcarbonate, a 72% yield of MPAN is obtained at 533 K over
NaY [389]. For both phenol and phenylacetonitrile methylation, it is shown
that basicity of the catalyst is required.
In conclusion, basic zeolites give specific selectivities in alkylation dif-
ferent from those obtained with protonic forms.

TABLE 13
Alkylation of Aniline with Dimethylcarbonate
Aniline NMA NNDMA
T conv. ZN alkyl. EN alkyl.
Zeolites (K) W F
~
(%I (%Ib (%>b Ref.
c
KY 453 2.08 99.6 93.5 366
c
Kx 453 2.08 65.4 86.7 366
KY 453 0.206 90.9 64.5 35.5 367
KEMT 453 0.206 90.6 73.6 26.4 367
KY 408 0.206 78.3 91.6 8.4 367
Kx 408 0.206 86.0 81.8 18.2 367
KEMT 408 0.206 81.3 92.5 7.5 367
K Beta 408 0.206 63.4 93.0 7.0 367
KY 503 0.411 95.7 29.2 70.8 367
Kx 503 0.411 97.1 22.6 77.4 367
csx 503 0.411 96.5 26.5 73.5 367
KEMT 503 0.411 92.3 55.7 44.3 367
K Beta 503 0.411 84.2 39.8 60.2 367
ug h mol-’.
bCN alkyl.: N-methyl aniline (NMA) + Nfl-dimethylaniline (NNDMA).
“Not known.
BASIC ZEOLITES 583

Not only basic but also weak acid sites are usually involved.
Some reactions appear to require mainly basic sites (alkylation of
aniline with dimethylcarbonate at 503 K on KX or Cs X [367] or
O-methylation of phenol to anisole at 553 K on NaX [388].
In alkylation reactions basic zeolites are not dangerous, and are easy-
to-use catalysts for the protection of environment.

B. Dehydrocyclization of n-Alkanes

A large number of works have been devoted to the study of zeolite-


supported metals [111,134,390].
Downloaded by [Yale University Library] at 01:03 15 August 2013

Bernard found that some Pt/zeolites in the Na or K forms (L, X, Y,


omega) selectively transform n-parafins to aromatics; the best material is
Pt/KL [137]. This zeolite has a structure with parallel channels (Fig. 3). Only
cations in site D are accessible in the channels. The selectivity to benzene
(when starting from n-hexane) is improved by exchange in L zeolite of the
alkaline cations from Li to Cs in the alkaline series [113]. With PtBaL ex-
changed with different alkali cations, an increase in the selectivity of aro-
matics is shown from Li to Cs [391,392]. The PtBaL zeolite when heat
treated becomes a very active and selective catalyst [171,393]. The Ba ions
move from D sites in the channel (Fig. 3) to sites inaccessible to reactants
[171]. As seen, polyvalent cations may interact with water to generate protons
[Eq. (2)]. When located in inaccessible sites the Ba ions lose this possibility
and no acidity is detected. Nevertheless, most of the framework oxygens
linked to the Ba ions are still accessible from the channel. The basicity is
then preserved. The distribution of Pt particles is said to be more uniform in
Pt/BaKL than in Pt/KL [394]. Several ideas are proposed to understand the
very good selectivity of Pt zeolites in a nonprotonic form. They are still a
matter of discussion [134]. The first one, which is chemical in nature, pro-
poses that the selectivity to aromatics of the Pthasic L zeolites is linked to
the electron-rich character of the Pt particles [113,138,139,141,158,395-3971.
The electron excess in these Pt particles and characteristic of basic zeolites
was evidenced by various techniques [113,138,139,141,142,144,145,158,162].
Since the aromatization is monofunctional on Pt [137,171,178,398], the mod-
ification of the electronic state of Pt strongly suggests specific properties of
the metal particles. In line with this approach is the good activity and selec-
tivity of Pt/Mg(Al)O basic catalyst in the same reaction [399]. In the same
view, the metal/Mg(Al)O interaction is comparable (from infrared CO studies)
to those in Ptibasic zeolites [157]. In addition, the decrease in the hydrogen-
ation activity (a characteristic of metal catalysis) is also observed for Pt
[143,145] or Ru [140] metals supported on basic zeolites (faujasites) as the
basicity increases.
A second approach to understanding the aromatization properties is to
consider geometric effects (such as the geometry of the zeolite channels and
584 BARTHOMEUF

of the Pt clusters). It involves a die effect [400] or preorganization of paraffins


in L channels [401]. The importance of the spatial and geometric character-
istics of various zeolite cages [402-4071 and of Pt particles [408-4101 are
studied with care.
In any case the absence of protons in PtL is a requirement of a high
and selective dehydrocyclization of paraffins. The mode of preparation was
early reported as being important in that it must avoid the presence of protonic
acidity in the catalysts [137]. A detailed study of KCl addition in the prep-
aration and of calcination temperature showed the importance of the two
parameters [396,397,411]. It was also observed that, in Pt/L, reduction of Pt
at temperatures above 300°C leads to a decrease both in the hydrogen chemi-
sorption capacity and in the catalytic activity [412]. Novel Pt incorporation
Downloaded by [Yale University Library] at 01:03 15 August 2013

procedures are described which involve sublimation of Pt(acac), [413] or the


use of hexafluoroacetylacetonate [414]. Both methods give highly dispersed,
nonacidic Pt/L catalysts. Nucleophilic Pt clusters are described in these ma-
terials from DRIFTS and EXAFS studies [414]. Modification of KL-type
zeolite with halocarbons improves the catalytic performances in aromatization
[176,415]. An important role of CF3C1treatment is to keep a high dispersion
of Pt and a low rate of coke deposition [416].
Metals other than Pt, i.e., Ir and Rh supported on BaKL, were shown
to be less active than Pt [409]. For Ir/KL the activity increases up to a max-
imum K:Al ratio of 1.14 [156]. The PtFe/KL catalyst is described as more
active than P t K L in hexane dehydrocyclization and methylcyclopentane de-
hydroisomerization [166]. Pt-Re, Pt-Sn [417], and Pt-Ni [169] supported
L zeolites are also described. Ni prevents the metal particle growth [169].
A synergism is observed for the reforming of n-hexane, methylcyclo-
pentane, and methylcyclohexane when Pt is supported on composites of basic
L zeolite and either H-beta or H USY [418]. Pt KL treated after reduction
with K2C0, is reported to give higher conversion of hexane and benzene
selectivity than the original Pt KL catalyst [419]. Pt supported on supports
other than L (y-A1203)[420] or SAPO-11 [421] made basic by specific treat-
ments, are reported as good catalysts. The question arises as to their de-
activation.
In conclusion, interest in the aromatization reaction is growing in re-
search laboratories. Very relevant and consistent papers have been published.
The main question on the origin of the very good performances of Pt in L
zeolites is still intriguing. Results obtained by various techniques support the
idea that the Pt particles are electron rich or polarized in Pt L zeolite. This
may be the reason for the Pt specific selectivity.

C. Hydrogenation-Dehydrogenation Reactions

Nonprotonic zeolites are reported as active catalysts for the hydroge-


nation of unsaturated hydrocarbons or CO and for the dehydrogenation of
saturated hydrocarbons or alcohols.
BASIC ZEOLITES 585

1. Hydrocarbons
Some basic zeolites containing no metal phase were shown to catalyze
hydrogenation-dehydrogenation reactions. Ethylene is hydrogenated at
180°C on basic Na- and K-erionite activated by spillover [422]. In ZSM-5
exchanged with the series of alkaline cations, the K, Rb, or Cs forms are the
most active for the dehydrogenation of cyclohexane to benzene and cyclo-
hexene [95]. Highly dispersed oxides (such as aFe,O,, NiO, CuO) in NaY
give high activity and selectivity for the dehydrogenation of cyclohexane
[423]. Similarly, ZnO or Ga,O, oxides dispersed in alkaline zeolites induce
hydrogenation-dehydrogenation properties [57]. All these examples refer to
modified zeolites.
Downloaded by [Yale University Library] at 01:03 15 August 2013

Metals supported on zeolites usually involve acidic materials [lll].Nev-


ertheless, metals dispersed in basic zeolites may generate interesting and dif-
ferent selectivities. The main example described separately in Sect. VI1.B
concerns Pt/alkaline L, which is monofunctional on Pt and very selective for
ring closure of alkanes. It was observed that the hydrogenation-dehydroge-
nation activity of metals supported on zeolites strongly depends on the aci-
dobasicity of the solids [ 111,145,424,4251. The cyclohexane dehydrogenation
to benzene is higher with K or Na/PtL than with the acid form [424]. The
benzene hydrogenation on Pt/faujasite varies in the opposite direction and
decreases in the X and Y series as the basicity of the zeolite increases, as
seen in Fig. 9, curve C. The highest apparent activation energy is observed
for the most basic catalyst PtCsX [145]. It increases from 48 kJ/mol for PtHY
to 120 kJ/mol for Pf36CS326Na462 x.
Similarly, the hydrogenation activity of Ru supported on Y zeolites de-
creases from NaY to KY and then CsY, i.e., as the zeolite basicity increases
[1401. The selective hydrogenation of 1,3-butadiene to cis-2-butene is high
on basic zeolites containing Mo(CO), carbonyls [117,119,120].

2. Oxygenated Compounds
a. Alcohol Dehydrogenation. This reaction is very often associated
with dehydration and two types of products arise from the transformation of
alcohols. Nevertheless, dehydrogenation sites may be characterized separately.
The dehydrogenation of 2-propanol to acetone, depressed by an acidic
reagent (phenol.) was shown to occur on the basic sites (framework oxygens)
of X and Y exchanged with alkali cations [5]. The possible role of cationic
iron impurities in this reaction was stressed in Ref. 426. The reaction occurs
at 300°C in CsZSM-5 while dehydration to propylene appears easier, taking
place at 200°C [427]. Cesium oxide clusters trapped in the cages of CsNaY
increase the acetone activity of the parent zeolite by an order of magnitude
[47,360]. The selectivity of acetone is above 97%, and on a surface area basis
the activity is comparable to that found for MgO [46]. The dehydrogenation
of methanol to formaldehyde was examined by infrared [428]. This reaction
586 BARTHOMEUF

was proposed to occur in the side-chain alkylation of toluene with methanol


on basic zeolites (see Sect. VII.A.l). Formaldehyde is also produced from
methanol with a high selectivity at 750 K over silicalite (Al content < 0.03%)
containing an excess of Na' ions (up to 1%).It was proposed that Na' ions
play a major role [365]. This suggests that oxide clusters may be the active
sites even if the zeolite itself (i.e., silicalite) is not basic. This is a very good
and early example of a zeolite system made active by postsynthesis inclusion
of basic sites. In acidic zeolites dehydration occurs alone or in addition to
dehydrogenation. It is suggested that even in the case of protonic catalysts
(H mordenite) [429] or HY [430], acid and basic sites are involved in the
dehydration of methanol [429,430] or ethanol [430] to the corresponding
ether. The conversion of cyclic alcohols on H, Na, or Cs-ZSM-5 suggests
Downloaded by [Yale University Library] at 01:03 15 August 2013

that protons favor olefins production while carbonyl compounds are formed
on basic dehydrogenation sites [431]. The Lewis acid-base pairs are also
needed for the transformation of methylbutynol in a variety of alkali-ex-
changed zeolites (X, Y, L, mordenite, ZSM-5) [96].
All these results show that both acid and basic sites participate in the
dehydration-dehydrogenation of alcohols as was proposed a long time ago
for oxides [l].When the balance between the two functions is in favor of
basicity (high Al content, cations with low electronegativity, presence of basic
oxide clusters), the dehydrogenation activity is preferred.
The addition of a metal function (Pt cluster) improves the dehydroge-
nation of isopropanol to acetone in PtKL [432,433]. No obvious dehydro-
genation property was seen in PtHL, the bifunctional catalyst which enhances
the conversion to H,O and propene [433]. This is associated with the electron-
deficient character of Pt in PtHL [433]. This suggests that in PtKL the elec-
tronic state of Pt favors the dehydrogenation of isopropanol while it does not
in PtHL. This is in line with the better dehydrogenation of cyclohexane to
benzene in PtKL or PtNaL rather than in PtHL [424]. Following the discus-
sion in Sect. VI1.B on the origin of the aromatization selectivity in Pt alkaline
L zeolites, it is tempting to suggest that the proposed hypothesis of electron-
rich or polarized metal clusters [113,138~139~141~142~143~158~395-397] may
very nicely explain the high dehydrogenation activity of Pt supported on basic
zeolites. The adsorption coefficient of the reaction products, unsaturated hy-
drocarbons or acetone, would be small on electron-rich particles favoring a
fast desorption and a high turnover. By contrast, the poor hydrogenation of
benzene on Pt in basic faujasites (Fig. 9) would result from a low adsorption
coefficient of the aromatic decreasing the reaction rate [145]. This is in line
with competitive hydrogenation of benzene and toluene on PtL zeolites hav-
ing various acid-base properties and with the simultaneous changes in ad-
sorption equilibrium constants [1411.
b. Other Compounds. CO hydrogenation on modified basic zeolites
with metal carbonyls [58,122,123] or various cations [167,434-4361 gives
specific selectivities. With bulky molecules steric effect may be important.
For instance, in the selective hydrogenation of cinnamaldehyde to cinnamyl
BASIC ZEOLITES 587

alcohol, geometric constraints may direct the selectivity more than the
strength of basicity in Pt, Ru, or Rh supported on NaY or KY [437,438] and
in PtNa- or PtCs-beta zeolites [439].
As a conclusion to the properties of metals in basic zeolites, it may be
noted that bifunctional catalysis often implies the idea of a metal and an acid
function. One may envisage other reactions which may be bifunctional with
a metal and a basic function. This is almost an unexplored field. An example
may be the CO hydrogenation on Pt/L exchanged with Mg, Ca, or Ba ions
[1481. Furthermore, the dependence of hydrogenation-dehydrogenation re-
actions on the acidobasicity of the zeolite is fully in line with a modification
of the electron distribution in Pt clusters encaged in zeolites.
Downloaded by [Yale University Library] at 01:03 15 August 2013

D. Condensation Reactions

Condensation reactions, which may be catalyzed by bases in solution,


occur in basic zeolites [440]. The Knoevenagel condensation (formation of a
C-C bond through the reaction of a carbonyl with a methylene group) or
aldol condensation (involving two C-0 bonds) are described and compared
to results in solution. The Knoevenagel condensation of benzaldehyde and
substituted benzaldehydes with ethylcyanoacetate [26,61,279,280], ethylma-
lonate [26,211,279,280], and ethylacetoacetate [279,280] are described for
basic faujasites [26,61,279,280] and basic beta [211]. The catalytic activity
increases with the zeolitic basicity. The most active of the X-type zeolites are
CsX [279,280] and NaGeX [26,280], i.e., a zeolite where Si is replaced by
Ge. These zeolites are more active than pyridine and less active than piper-
idine showing basic sites with pK, < 13.3 [26,279,280]. The behavior of the
NaGeX sample is quite in line with the experimental measurement of basicity
in NaGeX zeolite. This zeolite is more basic than NaX [25] (see Sect.
IV.B.2.c). Despite its low Al content, Na-beta zeolite is more active in this
reaction than NaY or NaX. A high basicity of Na or Cs-beta is already re-
ported from benzene [241] and pyrrole adsorption results (Fig. 15). It may
reflect a specific structural influence in beta. The NaGa-beta (part of Al re-
placed by Ga) and NaGe-beta are more active than the Si-Al form [211],
indicating that both Ga (instead of Al) and Ge (instead of Si) increase the
basicity. As indicated in Sect. IV.A, the theoretical calculation of basicity on
Ge or Ga zeolites depends on the actual values of electronegativities of
these two elements in the zeolites [192]. It is not known for Ge. The aldol
condensation of acetone gives mesityloxide and isophorone on faujasites or
L zeolites containing Naj neutral clusters (see Sect. 1II.D). The very likely
formation of extraframework Na oxide or hydroxide upon interaction with
water resulting from the reaction was pointed out, along with steric effects
[74]. Aldol condensation is reported in X [96,441], Y [441,442], ZSM-5
[441,443], and ZSM-11 [443] zeolites which were made basic. In the con-
densation of n-butanal to 2-ethyl-2-hexena1, the presence of both acid and
588 BARTHOMEUF

basic sites was suggested to produce a higher activity with alkali-exchanged


Y than with strongly basic MgO [442]. The bimolecular condensation of
ethanol to 1-butanol is the highest for RbLiX compared to RbNaX or RbKX.
Basic sites of the zeolite are involved in the reaction. It was shown that the
reaction does not proceed through aldol condensation [444,445].

E. Miscellaneous Reactions

A large variety of reactions have been carried out on nonacidic zeolites


[277,278,346,358,360]. The basic character of these materials was pointed
out rather recently in order to explain the selectivities obtained.
Downloaded by [Yale University Library] at 01:03 15 August 2013

The isomerization of olefins, known to occur on basic catalysts, is easy


on basic zeolites [60,72-74,446-4481. The activity is enhanced by addition
of NaN, [72-741 or an excess of alkali ions [60,446]. The occlusion of Cs,O
in CsNaX and CsNaY favors the isomerization of 1-butene to cis-2-butene
[601.
Metal carbonyls stabilized in basic zeolites show specific catalytic prop-
erties. Pt carbonyls similar to those described in Refs. 112 and 114 are very
active in the reduction of NO to CO, N,, and N,O at 473 K [131].
Mo(CO),(OZ),, (OZ) being the framework oxygen (Fig. 8), gives a high se-
lectivity in the hydrogenation of 173-butadieneto cis-2-butene [117,119,120].
Iron carbonyls in NaY are described as catalysts precursors in the Fischer-
Tropsch synthesis [ 1101. Osmium carbonyls [122], rhodium carbonyls [123],
and iridium carbonyls [58] in basic faujasites give non-Schulz-Flory distri-
butions in CO hydrogenation with high yield of C, hydrocarbons. Rh carbonyl
clusters entrapped in NaY are active for the liquid-phase hydroformylation
of olefins [449]. Palladium trimethylphosphine carbonyl clusters encaged in
NaY are active for the hydroformylation of propene [450].
The activation of CO in hydrogenation or carbonylation reaction is fa-
vored on iron-manganese oxide alkali modified ZSM-5 [436], CoNaY [434],
Pd-Fe/NaX [167] or Nait clusters trapped in X zeolite [435]. Conversely,
the decomposition of methylformate to CO and methanol is base-catalyzed
by NaX and CaA doped with NaOH or KOH [451].
Basic sites in alkali faujasite are involved in the reaction of H,S with
0, [271a], alcohols [271b7452], CO, [453], or methylfuran [454]. The gen-
eration of protons in the decomposition of H,S in basic zeolites (Le., Na
forms of Y, mordenite, ZSM-5, and A) promotes the cracking of n-hexane,
isopropylbenzene, and ethylbenzene; and the isomerization of 1-butene and
cyclopropane [272]. The electronegativity of the alkali metal exchanged
forms of X, Y, and ZSM-5 mainly influences the formation of thiophene in
transformation of ethanethiol and that of ethene in decomposition of diethy-
lene sulfide [455]. Preparation of 4-methyl thiazole is carried out on ZSM-5
containing Cs oxide clusters. The presence of water vapor may very likely
assist in the formation of an hydroxide [456].
BASIC ZEOLITES 589

The conversion of methylhalides to hydrocarbons [457]on CsX, the


bromination of methylbenzene on NaY [458], and the dehydrochlorination of
1,1,2-tricloroethane to 1,l-dichloroethene on Cs-ZSM-5 were carried out
[459].The basicity of the zeolite and the Lewis acidity of the cation were
proposed to be involved in the reaction [457].The participation of acid-base
pairs is also involved in the reaction of methylbutynol, the condensation of
acetone, the reaction of monomethylamine 1961, and propene aromatization
[255].The cyclization of acetonylacetone over ZSM-5 containing an excess
of Na cations gives 3-methyl-2-cyclopenten-1-oneselectively while the acid
zeolite yields 2,5-dimethylfuran [63].As in the case of methanol dehydro-
genation [365],this reaction shows how occluded basic species change the
selectivity of a nonbasic zeolite (ZSM-5).
Downloaded by [Yale University Library] at 01:03 15 August 2013

The formation of ketenes from acetic acid, propanoic acid, and isobu-
tanoic acid is catalyzed on alkali zeolites [460].Reverse oxidation to hydro-
peroxides, ketones, or aldehydes of olefins, aromatics, or alkanes occurs very
selectively upon visible irradiation at room temperature or below on Y zeo-
lites exchanged with alkaline or alkaline earth cations [461].
In conclusion, the above section describes a large variety of reactions
and methods of improving zeolite properties by different modifications (e.g.,
incorporation of occluded oxides, alkaline metals, alkaline ions, metal clus-
ters, etc.). It can be seen that very specific applications exist for basic zeolites
in the preparation of high-value chemicals, such as pharmaceuticals, or other
industrial chemicals used in the manufacture of perfumes, soaps, or deter-
gents. Several methods of zeolite improvement have still to be applied to
these reactions. In addition, basic catalysis with zeolites may offer a new
route to find processes for environmental protection.

VIII. ACIDIC VERSUS BASIC ZEOLITES

Zeolites receive most of their industrial applications in catalysis in their


protonic forms while the basic solids are used preferentially in adsorption
processes where protonic catalytic activity is unwanted. Section VII of the
present paper describes the catalytic properties of basic zeolites with regards
to some specific reactions. As can be seen, the basic catalysts may be greatly
improved by specific modification procedures. The interest in such materials
varies widely from adsorbents to catalysts. It is then appropriate to consider
which parameters are important in generating an acidic, neutral, or basic
character in a zeolite structure.
It turns out that as an almost unique case in the chemistry of materials
(except some clays or resins), a given zeolite with a well-defined crystalline
structure may have acidic or basic properties depending only on its chemical
composition, i.e., on framework and nonframework atoms. In the case of
faujasite, which is well documented, it is shown that the range of properties
may even extend from superacids to superbases. One may try to rationalize
590 BARTHOMEUF

the parameters of importance to direct the chemical properties on one side or


the other.
Table 14 summarizes the present knowledge on faujasite zeolites (ex-
changed with protons or alkaline cations). The properties move gradually and
continuously from the limit of superacids to that of superbases from the top
to the bottom of Table 14. No discontinuity is known. The framework Al
content must be low in the very acidic forms and high in the basic zeolites.
The progressively increased exchange of protons by alkaline ions moves the
faujasite from the medium and weak acid range (HNaY, HNaX) to that of
basicity (RbX, CsX). The more basic zeolites contain the cations (K, Rb, Cs)
with the lowest electronegativity (Figs. 11, 15, 19).
In fact, cation M and framework oxygen atoms are the acid and basic
Downloaded by [Yale University Library] at 01:03 15 August 2013

sites, respectively. They form acid-base pairs in M-0 species [24,95,96,


162,183,265,4621. For a given Al content, the acidic character prevails when
cations are strongly acidic (Li, Na). The charge on the adjacent oxygen is
then low. The zeolites show basic properties for a high charge on oxygen,
which corresponds to cations with low electronegativity (Rb, Cs), i.e., weak
acidity. This acid-base character exists in any zeolite.
The intrinsic framework acidobasicity may be changed by postsynthesis

TABLE 14
Change in Acidobasic Character of Faujasites with Chemical Composition
Acidobasic
character, j. Characteristics and examples
Acid
Superacid Ultrastable Y Acidity strongly depends on
Dealuminated Y c- the dealumination procedure
Strong Low framework Al and on the chemical form of
highly exchanged extraframework Al
Medium Partially neutralized
(ex: HNaY)
Weak (ex: HNaX)
LiY

Neutral NaY

Basic
Weak KY
RbY
Medium CsY, .1 LiX
NaX Basicity strongly depends on the possible
Kx addition of occluded species [oxides,
Strong RbX, CsX hydroxides, alkali metal clusters, . . .
Superbase X + basic species (nature, amount, basicity of these species)]
BASIC ZEOLITES 591

modifications of the zeolites. It is well known that a large variety of very


acidic faujasite exists. This arises from the distribution of a few Al atoms in
the highly silicic framework, from the presence or not of extraframework Al,
and from the chemical state, size, or location of these extraframework clus-
ters. This is well documented [463] and of major importance for the industrial
use of zeolites. A n analogy exists with the case of basic zeolites as to the
possible modifications of the solids. The Al content must be the highest (i.e.,
Si/Al close to 1) in order to generate a high structural (or intrinsic framework)
basicity. A large variety of strongly basic faujasites may be obtained by the
correct choice of the exchangeable cation and even more by the addition of
basic clusters (such as oxides or hydroxides) or occluded species (such as
alkaline metal or metal clusters) enhancing the basic character of framework
Downloaded by [Yale University Library] at 01:03 15 August 2013

oxygen (Table 14). This domain is still largely unexplored.


More generally, the concepts presented for faujasites appear to apply to
other zeolite structures. A high protonic acid strength is usually related to a
low Al content. On the other hand, the basicity increases with the Al content
(Fig. 11). The basic character of mordenite (Fig. 15) or beta (Fig. 25) in-
creases as expected when the protons are exchanged with the series of al-
kaline ions from Li to Cs. Nevertheless, the range of their basicity is higher
than that expected from Al content and the Sintcorresponding to mean elec-
tronegativity of the zeolite. As detailed in Sects. IVB.2.c and VI.B, a struc-
tural effect should be considered in addition to the influence of chemical
composition. This shifts the range of acidobasic character as compared to that
of faujasite with similar Al content or exchangeable cations. A second feature
must also be pointed out. Some examples are given for ZSM-5 [63] or sili-
calite [365] which show basic properties in catalysis. These zeolites have a
very low Al content and do not possess structural basicity (Figs. 11 and 15;
Refs. 196 and 219). The reported selectivities, typical of basic catalysis, are
found for these zeolites having Na ions in excess and then very likely con-
taining basic clusters of Na oxide or hydroxide. This shows that these zeolites
and probably any zeolite, even with no intrinsic framework basicity, may be
transformed to a basic catalyst by addition of highly dispersed basic species
in the porous system. The protons are exchanged by the cations. The excess
ions generate the basic character. Table 15 proposes a general scheme for the
preparation of basic catalysts based on zeolites. The zeolite behaves mainly
as the supporting matrix for highly dispersed and very active basic clusters
located in the cages and channels of the zeolite itself. These supported basic
species may be the active phase for some specific catalystic reactions. From
Table 15, one may envisage that, in addition to basic selectivity, shape selec-
tivity may also be used to improve the catalytic performances of the supported
basic species. Almost no work has been done to date in this field.
As a conclusion, acidic and basic zeolites belong to a continuum of
materials. The potential chemical properties may be adjusted by the choice
of the correct framework and nonframework composition. Within this broad
592 BARTHOMEUF

TABLE 15
Preparation of Basic Catalysts Based on Zeolites

[ ]
Zeolite" + adduct + catalytic system

Clusters of
Strong
to acid"] + Oxide
Hydroxide + Basic catalysts supported
Strong base' Alkaline metal on zeolites
Ions in excess
Other . . .
"Any structure type possible.
"For instance, silicalite or ZSM-5 [63,365].
Downloaded by [Yale University Library] at 01:03 15 August 2013

'For instance, faujasite (see Table 14).

domain the field of basic zeolites is still not very well understood and has
not yet been considered for any specific application.
The examples reported above give general rules on how to obtain basic
zeolites. The very usual one is to generate structural (or intrinsic framework)
basicity by lowering the mean zeolite electronegativity. A very powerful rule
is to add new species into the zeolite pores. These species may have their
own basic character which may be eventually changed by an interaction with
the zeolite, or they may modify the framework basicity. This gives rise to
supported basic catalysts-or adsorbents. As in the case of acidic zeolites,
shape selectivity may in addition be used to direct catalysis or adsorption.

IX. CONCLUSIONS AND PERSPECTIVES

Zeolites exchanged with alkaline cations have been known since the
beginning of zeolite history, because they are the first form of the as-synthe-
sized materials. Considered from the viewpoint of basic solids, it appears that
the properties of these zeolites are adjustable to specific applications, and the
solids may therefore be seen in a sense as new materials. Up until now no
systematic way of improvement has been applied. The present attempts to
rationalize the concepts of basicity in zeolites opens up a new route for the
scientific search for adsorbents or catalysts with specified properties.
The two sources of basicity known to exist in zeolites are the structural
(or intrinsic framework) basicity and that arising from occlusion of species
which either have their own basic character (like hydroxide or oxide clusters)
or, alternatively, increase the basicity of framework oxygens (like Na; neutral
metal clusters). For a given zeolite structure, the structural basicity is in-
creased by lowering the zeolite intermediate electronegativity Sint.This may
be achieved either by incorporating in the framework atoms different from
Al or Si and/or by using exchangeable cations with a low electronegativity.
BASIC ZEOLITES 593

Examples are given in this paper for the framework approach with incorpo-
rated Ga or Ge, and for exchangeable cations with Rb or Cs.
One may expect that other framework atoms should also generate
different basicities. For example, nothing is known at the present time
on the basicity of zeolites containing B, Fe, or other incorporated
framework atoms and exchanged with alkaline cations.
Mordenite and beta zeolite exhibit basic properties which are greater
than expected from the chemical composition alone (low Al content). The
effect of the structure evidenced in these two cases may well exist for other
zeolites.
In other words, some zeolites which have not yet been investigated
Downloaded by [Yale University Library] at 01:03 15 August 2013

with regards to their basicity may show interesting basic character.


The increase of the basicity in a zeolite system by incorporating new
species (hydroxides, oxides, alkaline metal clusters) has been used for many
years but remains poorly understood. A better understanding on the role of
the species is required, and particularly information on the influence of the
identity of the alkaline cation, the size, chemical form, and location of the
clusters are all required. This will allow one to master the preparation of
these modified zeolites and to obtain tailor-made basic adsorbents or catalysts.
Therefore, other oxides (or hydroxides) or metal clusters have to be
tested, for instance, transition metal oxide clusters, which would gen-
erate oxidation properties leading to bifunctional materials.
More and more evidence is being reported for an effect of the zeolite
basicity on the electronic properties of supported transition metals, mainly Pt.
The consequence of this on the n-alkane aromatization is still under investi-
gation. Results on hydrogenation-dehydrogenation activity are consistent, for
different reactions and various zeolites, with a directing effect of the elec-
tronic properties of Pt on the selectivity. The properties of transition metal
carbonyls are comparable to those obtained in basic media. The influence of
the “basic media” in the zeolite cages is observed for any highly dispersed
occluded species (metal, metal carbonyl, oxide, or hydroxide).
Transition metals supported on basic zeolites, other than Pt, should
be considered. They offer a new route to transition metal clusters
highly dispersed in zeolites. In order to generate neutral, cationic, or
anionic metal carbonyls, attention should be paid to the compatible
acidobasic character of the zeolite. The new properties of the oc-
cluded oxide or hydroxide clusters compared to bulk materials appear
important to consider.
Characterization of the basicity itself is making great progress. The mo-
lecular approach of the zeolite-probe interaction is either theoretical or ex-
perimental. It gives a deep understanding on the mode and energy of the
594 BARTHOMEUF

interaction. Up until now it has been applied only to a small number of


molecules interacting with the framework, and only a few studies consider
the basicity arising from occluded species.
Progress in basicity characterization requires more work on the
search of probes molecules and on their interaction with the basic
sites. Both molecular modeling and experiments are required. The
intrinsic framework basicity and that arising from occluded species
must be clearly distinguished and evaluated.
The main uses of basic zeolites are in adsorption and catalysis. In the
very large amount of work devoted to adsorption of pure components or of
mixtures and to the industrial separations by adsorption on zeolites, very little
Downloaded by [Yale University Library] at 01:03 15 August 2013

is concerned with the chemical properties of the adsorbent. The zeolite is


most usually considered as an inert material whose main characteristic is the
porosity. The present paper shows that the acidobasic properties of the zeolite
cannot be ignored, and that they may direct the adsorption selectivities. They
also explain the efficiency of desorbents in the practical separation of
mixtures due to the competition of adsorption of the various adsorbates hav-
ing different chemical (acid-base) properties.
It appears that consideration of the acidobasic character of zeolites
in adsorption processes is an emerging field which gives some keys
for the understanding of selectivities in adsorption and for the uses
in separation processes.
Catalysis with basic zeolites until now has been mainly concerned with
the search for new reactions using the available materials. At the present time
new ways are offered to direct the catalytic properties by adjusting the zeolites
to the level and amount of basicity required by a given reaction. As said
above, this may be achieved by changes in the zeolite electronegativity (both
of the framework atoms and exchangeable cations) and by postsynthesis mod-
ifications (addition of encaged species). In addition, the zeolite basicity mod-
ifies the catalytic properties of oxide and hydroxide clusters, supported met-
als, or transition metal carbonyls in many reactions.
The zeolite cages are microreactors, the walls of which bear the basic
sites whose strength and number can be adjusted.
In addition, the modification of zeolite basicity as a field is still very
large and not much explored. This opens new routes for the improve-
ment of activity and selectivities in many reactions catalyzed by ba-
ses or in basic media. The combined effects of basicity and shape
selectivity have also to be considered in specific reactions. The do-
main of bifunctional zeolites will be successfully extended to the case
where one function is basicity and the other one is, for instance,
hydrogenation-dehydrogenation (supported metal) or oxidation
(transition metal oxide).
BASIC ZEOLITES 595

ACKNOWLEDGMENTS

T h e author thanks Professor S . Kaliaguine for helpful comments, Dr. D.


Murphy for thorough correction of the manuscript, and Mrs. R. Frank for
interesting discussions.

REFERENCES

1. H. Pines and J . Manassen, Adv. Catal., 16, 49 (1966).


2. K. Tanabe, Solid Acids and Bases, Kodansha, Tokyo, 1970.
3. M. Ai, Bull. Jpn. Petrol. Inst., 18, 50 (1976).
Downloaded by [Yale University Library] at 01:03 15 August 2013

4. H. Pines and W. M. Stalick, Base Catalyzed Reactions of Hydrocarbons and


Related Compounds, Academic Press, New York, 1977.
5. T. Yashima, H. Suzuki, and N. Hara, J. Catal., 33, 486 (1974).
6. Y. N. Sidorenko, P. N. Galich, V. S. Gutyrya, V. G. Il’in, and I . E. Neimark,
Dokl. Akad. Nuuk, SSSR, 173, 132 (1967).
7. T. Yashima, K. Sato, H. T. Hayasaka, and N. Hara, J. Cutul., 26, 303 (1972).
8. J. Schemer, J. L. Bass, and F. B. Hunter, J . Phys. Chem., 79, 1194 (1975).
9. J. M. Newsam, M. M. J. Treacy, W. T. Koetsier, and C. B. de Gruyter, Proc.
R. SOC.Lond., A420, 375 (1988).
10. J . B. Higgins, R. B. La Pierre, J. L. Schlenker, A. C. Rohrman, J. D. Wood,
G. T. Kerr, and W. J. Rohrbaugh, Zeolites, 8, 446 (1988).
11. R. M. Barrer and H. Villiger, Z. Kristallogr., 128, 352 (1969).
12. W. J. Mortier, Compilation of Extraframework Sites in Zeolites, Butterworth,
GuiIford, 1982, p. 54.
13. D. W. Breck, Zeolite Molecular Sieves, Wiley, New York, 1974.
14. S. T. Wilson, B. M. Lok, C. A. Messina, T. R. Cannan, and E. M. Flanigen,
J . Am. Chem. SOC.,104, 1146 (1982).
15. B. M. Lok, C. A. Messina, R. L. Patton, R. T. Gajek, T. R. Cannan, and E.
M. Flanigen, J. Am. Chem. Soc., 106, 6092 (1984).
16. E. M. Flanigen, B. M. Lok, R. L. Patton, and S. T. Wilson, Proc. 7th Znt.
Zeol. Conf:, Tokyo (Y. Murakami, A. Iijima, and J. W. Ward, eds.), Kodanska,
Tokyo, and Elsevier, Amsterdam, 1986, p. 103.
17. E. M. Flanigen, R. L. Patton, and S. T. Wilson, Stud. Surf: Sci. Catal., 37, 13
(1988).
18. G. V. Gibbs, E. P. Meagher, J. V. Smith, and J. J. Pluth, ACS Symp. Ser, 40,
19 (1977).
19. J. A. Rabo, Catal. Rev., 31, 385 (1991).
20. K. A. Van Genechten and W. J. Mortier, Zeolites, 8, 273 (1988).
21. J. V. Smith, Molecular sieve I , Adv. Chem. Ser, 101, 171 (1971).
22. C. Mirodatos and D. Barthomeuf, J. Catal., 57, 136 (1979); 93, 246 (1985);
114, 121 (1988); and C. Mirodatos, A. Billoul, and D. Barthomeuf, J. Chem.
SOC.Chem. Commun., 1987, p. 149.
23. W. M. Meier and D. H. Olson, Atlas of Zeolite Structure Types, 3rd ed.,
Butterworth-Heinemann, London, 1992.
24. D. Barthomeuf, J. Phys. Chem., 88, 42 (1984).
25. P. 0. Scokart and P. G. Rouxhet, Bull. SOC.Chim. Belge, 90, 983 (1981).
596 BARTHOMEUF

26. A. Corma, R. M. Martin-Aranda, and F. Sanchez, J . Catal., 126, 192 (1990).


27. T. L. Barr, Zeolites, 10, 760 (1990).
28. C. Mirodatos, P. Pichat, and D. Barthomeuf, J. Phys. Chem., 80, 1335 (1976).
29. C. Mirodatos, A. Abou Kais, J. C. Vedrine, P. Pichat, and D. Barthomeuf, J.
Phys. Chem., 80, 2366 (1976).
30. C. J. Planck, Proc. Int. Congr. Catal., 3rd, Amsterdam, 1964, Vol. 1, p. 727.
31. W. T. Reichle, J. Catal., 101, 352 (1986).
32. M. E. Davis, C. Montes, P. E. Hathaway, J. P. Arhancet, D. L. Hasha, and J.
M. Garces, J . Am. Chem. SOC., 111, 3919 (1989).
33. M. Estermann, L. B. McCusker, C. Baerlocher, A. Merrouche, and H. Kessler,
Nature, 352, 320 (1991).
34. R. A. Van Santen, B. W. H. Van Beest, and A. J. M. de Man, in Guidelines
for Mastering the Properties of Molecular Sieves (D. Barthomeuf, E. G. Der-
Downloaded by [Yale University Library] at 01:03 15 August 2013

ouane, and W. Holderich, eds) NATO AS1 Series, Plenum Press, New York,
1990; Ser. B: Physics, Vol. 221, p. 201.
35. I. N. Senchenya, V. B. Kazansky, and S . Beran, J. Phys. Chem., 90, 4857
(1986).
36. A. G. Pelmenshchikov, V. I. Pavlov, G. M. Zhidomirov, and S . Beran, J. Phys.
Chem., 91, 3325 (1987).
37. P. J. O’Malley and J. Dwyer, J. Phys. Chem., 92, 3005 (1988).
38. M. Che and A. J. Tench, Adv. Catal., 31, 77 (1982).
39. H. Hattori, Muter. Chem. Phys., 18, 533 (1988).
40. P. Gallezot and B. Imelik, J. Phys. Chem., 77, 2364 (1973).
41. Ch. Peuker, K. Moeller, and D. Kunath, J. Mol. Struct., 114, 215 (1984).
42. M. D. Baker, G. A. Ozin, and J. Godber, Catal. Rev.-Sci. Eng., 27, 591
(1985).
43. R. A. Van Santen, A. J. M. de Man, W. P. J. H. Jacobs, E. H. Teunissen, and
G. J. Kramer, Catal. Lett., 9, 273 (1991).
44. D. B. Akolekar, M. Huang, and S . Kaliaguine, Zeolites, 14, 519 (1994).
45. A. Abou-Kais, C. Mirodatos, J. Massardier, D. Barthomeuf, and J. C. Vedrine,
J. Phys. Chern., 81, 397 (1977).
46. P. E. Hathaway and M. E. Davis, J. Catal., 116, 263 (1989).
47. P. E. Hathaway and M. E. Davis, J. Catal., 116, 279 (1989).
48. T. Turk, F. Sabin, and A. Vogler, Muter. Res. Bull., 27, 1003 (1992).
49. M. Lasperas, H. Cambon, D. Brunel, I. Rodriguez, and P. Geneste, Microp.
Muter., 1, 343 (1993).
50. H. Tsuji, F. Yagi, H. Hattori, and K. Kita, Proc. 10th Znt. Congr. Catal.,
Budapest (L. Guczi, F. Solymosi, and P. TCtenyi, eds.), Akademiai Kiado,
Budapest, 1993, B, p. 1171.
51. J. Ward, in Zeolite Chemistry and Catalysis (J. Rabo, ed.), ACS Monograph
171, 1976, p. 233.
52. D. W. Breck, Zeolite Molecular Sieves, Wiley, New York, 1974, p. 460.
53. J. Engelhardt, J. Szanyi, and B. Jover, Acta Symp. Ibero-Amer. Catal., 9th,
1984, Vol. 2, p. 1435.
54. C. Lacroix, A. Deluzarche, A. Kiennemann, and A. Boyer, J . Chem. Phys.,
81, 473, 481, 486 (1984).
55. V. Barbarin, Thesis, Paris (1987).
56. D. Barthomeuf and V. Barbarin, French patent 2,623,423 (1989).
BASIC ZEOLITES 597

57. D. Archier, G. Coudurier, and C. Naccache, Proc. 9th Int. Zeol. Con$, Mon-
treal (R. Van Ballmoos, J. B. Higgins, and M. M. J. Treacy, eds.),
Butterworth-Heinemann, Boston, 1993, Vol. 11, p. 525; D. Archier, Thesis,
Lyon (1989).
58. S. Kawi, J. R. Chang, and B. C. Gates, J. Catal., 142, 585 (1993).
59. J. M. Garces, G. E. Vrieland, S. I. Bates, and F. M. Scheidt, Stud. Sur$ Sci.
Catal., 20, 67 (1985).
60. J. C. Kim, H. X. Li, and M. E. Davis, Symp. New Catal. Chem., Div. Petrol.
Chem., 206th National Meeting, Am. Chem. SOC.,Chicago, 1993, Vol. 38, p.
776; J. C. Kim, H. X. Li, C. Y. Chen, and M. E. Davis, Microp. Muter., 2,
413 (1994).
61. I. Rodriguez, H. Cambon, D. Brunel, M. Lasperas, and P. Geneste, Stud. Surf:
Sci. Catal., 78, 623 (1993).
Downloaded by [Yale University Library] at 01:03 15 August 2013

62. (a) I. Rodriguez, H. Cambon, D. Brunel, M. Lasperas, and P. Geneste, Eu-


ropacat, Montpellier, 1993, Book of Abstracts, I, p. 164. (b) M. Lasperas,
personal communication.
63. R. M. Dessau, Zeolites, 10, 205 (1990).
64. S. Malinowski and S. Szczepanska, J. Catal., 2, 310 (1963).
65. J. Kijenski, K. Brzozka, and S . Malinowski, Bull. Acad. Pol. Sci., Ser. Sci.
Chim., 25, 20 (1977).
66. G. Suzukamo, M. Fukao, T. Hibi, and K. Chikaishi, in Acid-Base Catalysis
(K. Tanabe, H. Hattori, T. Yamaguchi, and T. Tanaka, eds.), Kodanska-VCH,
1988, p. 45.
67. A. Z. Khan and E. Ruckenstein, J . Catal., 143, 1 (1993).
68. T. Ushikubo, H. Hattori, and K. Tanabe, Chem. Lett., 1984, p. 649.
69. J. A. Rabo, C. L. Angell, P. H. Kasai, and V. Schomaker, Disc.F'araday SOC.,
4, 328 (1966); P. H. Kasai and R. J. Bishop, Jr., J. Phys. Chem., 77, 2308
(1973); P. H. Kasai, J. Chem. Phys., 43, 3322 (1965).
70. P. P. Edwards, M. R. Harrison, J. Klinowski, S. Ramdas, J. M. Thomas, D.
C. Johnson, and C. J. Page, J . Chem. SOC.,Chem. Commun., 1984, p. 982.
71. M. R. Harrison, P. P. Edwards, J. Klinowski, J. M. Thomas, D. C . Johnson,
and C. J. Page, J. Solid State Chem., 54, 330 (1984).
72. L. R. Martens, P. J. Grobet, W. J. M. Vermeiren, and P. A. Jacobs, Stud. Surf:
Sci. Catal., 28, 935 (1986).
73. L. R. M. Martens, W. J. M. Vermeiren, P. J. Grobet, and P. A. Jacobs, Stud.
Surf: Sci. Catal., 31, 531 (1987).
74. L. R. M. Martens, W. J. M. Vermeiren, D. R. Huybrechts, P. J. Grobet, and
P. A. Jacobs, Proc. 9th Int. Congr. Catal., Calgary (M. J. Phillips and M.
Ternan, eds.), Chem. Institute of Canada, Ottawa, 1988, Vol. 1, p. 420.
75. P. J. Grobet, L. R. M. Martens, W. J. M. Vermeiren, D. R. C. Huybrechts,
and P. A. Jacobs, Z. Phys. D Atoms Molecules Clusters, 12, 37 (1989).
76. G. A. Ozin, Adv. Muter., 4, 612 (1992).
77. E. Trescos, F. Rachdi, L. C. de Menorval, F. Fajula, T. Nunes, and G . Feio,
J. Phys. Chem., 97, 11855 (1993).
78. L. R. M. Martens, P. J. Grobet, and P. A. Jacobs, Nature, 315, 568 (1985).
79. Bo Xu and L. Kevan, J . Phys. Chem., 96, 2642 (1992).
80. Bo Xu and L. Kevan, J. Chem. SOC. Faraday Trans., 88, 1695 (1992).
81. T. Sun and K. Seff, J . Phys. Chem., 98, 10156 (1994).
598 BARTHOMEUF

82. H. Nakayama, D. D. Mug, C. I. Ratcliffe, and J. A. Ripmeester, J. Am. Chem.


SOC.,116, 9777 (1994).
83. S . H. Song, Y. Kim, and K. J. Seff, J. Phys. Chem., 95, 9919 (1991); J. Phys.
Chem., 96, 10937 (1992).
84. N. H. Heo and K. Seff, J. Am. Chem. SOC., 109, 7986 (1987); and M. S.
Jeong, Y. Kim, and K. Seff, J. Phys. Chem., 97, 10139 (1993).
85. T. Sun, K. Seff, N. H. Heo, and V. P. Petranovskii, J. Phys. Chem., 98, 5768
(1994).
86. P. A. Anderson, A. R. Armstrong, and P. P. Edwards, Angew. Chem., 106, 669
(1994).
87. H. D. Lee, U. S. Kim, J. Y. Park, and Y. Kim, J. Korean Chem. SOC., 38, 186
(1994).
88. J. B. A. F. Smeulders, M. A. Hefni, A. A. K. Klaasen, E. de Boer, U. Westphal,
Downloaded by [Yale University Library] at 01:03 15 August 2013

and G. Geismar, Zeolites, 7, 347 (1987).


89. K . B. Yoon and J. K. Kochi, J . Chem. SOC. Chem. Commun., 1988, p. 510.
90. (a) Y. S. Park, Y. S. Lee, and K. B. Yoon, J . Am. Chem. SOC., 115, 12220
(1993). (b) Y. S. Park, Y. S. Lee, and K. B. Yoon, Stud. Surf: Sci. Catal., 84B,
901 (1994).
91. S. Bordiga, A. Ferrero, E. Giamello, G. Spoto, and A. Zecchina, Catal. Lett.,
8, 375 (1991).
92. I. Hannus, I. Kiricsi, K. Varga, and P. Fejes, React. Kinet. Catal. Lett., 12,
309 (1979); P. Fejes, I. Kiricsi, I. Hannus, T. Tihanyi, and A. Kiss, Stud. Surf:
Sci. Catal., 5, 135 (1980); I. Kiricsi, I. Hannus, A. Kiss, and P. Fejes, Zeolites,
2, 247 (1982).
93. W. E. Garner and D. J. B. Marke, J. Chem. SOC.,1936, p. 657.
94. J. W. Ward, J. Catal., 10, 34 (1968); and Zeolite Chem. Catal. (J. Rabo ed.),
ACS Monograph 171, 1976, p. 226.
95. M. Huang and S. Kaliaguine, J. Mol. Catal., 81, 37 (1993).
96. N. Huang and S . Kaliaguine, Stud. Surf: Sci. Catal., 78, 559 (1993).
97. D . N. Stamires and J. Turkevich, J. Am. Chem. Soc., 86, 749 (1964).
98. J. Bandiera, Y. Ben Taarit, and C. Naccache, Bull, SOC. Chim. Fr., 10, 3419
(1969).
99. Y. Ben Taarit, J. Bandiera, M. V. Mathieu, and C. Naccache, J. Chim. Phys.,
67, 37 (1970).
100. Y. Ben Taarit, C. Naccache, and B. Imelik, J. Chim. Phys., 67, 389 (1970).
101. J. Turkevich and Y. Ono, Adv. Catal., 20, 135 (1969).
102. B. D. Flockhart, L. Mc Loughlin, and R. C . Pink, J. Catal., 25, 305 (1972).
103. J. H. Lunsford, Catal. Rev., 12, 137 (1975).
104. C. Naccache and Y. Ben Taarit, Proc. 5th Znt. Conf: Zeol., Naples (L. V. C .
Rees, ed.), Heyden, London, 1980, p. 592.
105. Y. Ben Taarit and M. Che, Stud. Surf: Sci. Catal., 5, 167 (1980).
106. P. A. Jacobs, N. I. Jaeger, P. Jiru, and G. Schulz-Ekloff, Stud. Surf: Sci. Catal.,
18, 1982.
107. F. Lefebvre, P. Gelin, C. Naccache, and Y. Ben Taarit, Proc. 6th Znt. Zeol.
Conf:,Reno (D. Olson and A. Bisio, eds.), Buttenvorths, Guildford, 1984, p.
435.
108. G. A. Ozin and C. Gil, Chem. Rev., 89, 1749 (1989).
109. M. Ichikawa, Adv. Catal., 38, 283 (1992).
BASIC ZEOLITES 599

110. D. Ballivet-Tkatchenko, G. Coudurier, and H. Mozzanega, Stud. Surf: Sci.


Catal., 5, 309 (1980).
111. W. M. H. Sachtler and Z. Zhang,Adv. Catal., 39, 129 (1993).
112. A. de Mallmann and D. Barthomeuf, Catal. Lett., 5, 293 (1990).
113. C. Besoukhanova, J. Guidot, D. Barthomeuf, M. Breysse, and J. R. Bernard,
J. Chem. Soc. Faraday Trans. I, 77, 1595 (1981).
114. J. R. Chang, Z. Xu, S. K. Purnell, and B. C . Gates, J. Mol. Catal., 80, 49
(1993).
115. G. Schulz-Ekloff, R. J. Lipski, N. I. Jaeger, P. Hiilstede, and L. Kubelkova,
Catal. Lett., 30, 65 (1995); H. Bischoff, N. I. Jaeger, G. Schulz-Ekloff, and
L. Kubelkova, J. Mol. Catal., 80, 95 (1993).
116. Y. Okamoto, A. Maezawa, H. Kane, and T. Imanaka, J. Catal., 112, 585
(1988).
Downloaded by [Yale University Library] at 01:03 15 August 2013

117. A. Maezawa, H. Kane, Y. Okamoto, and T. Imanaka, Chem. Lett., 1988, p.


241.
118. Y. Okamoto, A. Maezawa, H. Kane, I. Mitsushina, and T. Imanaka, J. Chem.
SOC. Faraday Trans. I, 84, 851 (1988).
119. Y. Okamoto, A. Maezawa, H. Kane, and T. Imanaka, Proc. 9th Znt. Congr.
Catal. (M. J. Phillips and M. Ternan, eds.), Chem. Institute of Canada, Ottawa,
1988, Vol. I, p. 11.
120. Y. Okamoto, T. Imanaka, K. Asakura, and Y. Iwasawa, J . Phys. Chem., 95,
3700 (1991).
121. P. L. Zhou and B. C . Gates, J . Chem. Soc. Chem. Commun., 1989, p. 347.
122. P. L. Zhou, S. D. Maloney, and B. C. Gates, J. Catal., 129, 315 (1991).
123. T. J. Lee and B. C. Gates, Catal. Lett., 8, 15 (1991); J. Mol. Catal., 71, 335
(1992).
124. L. Basini, R. Patrini, A. Aragno, and B. C . Gates, J. Mol. Catal., 70, 29
(1991).
125. T. J. Lee, J. H. Kim, W. C . Chang, D. J. Lee, Y. K. Kim, Hwahak Konghak,
31, 593 (1993).
126. Z. Zhang and W. M. H. Sachtler, J. Mol. Catal., 67, 349 (1991).
127. H. M. Ziethen and A. X. Trautwein, Stud. Surf:Sci. Catal., 46, 789 (1989).
128. C. Bremard, C. Depecker, H. des Grousilliers, and P. Legrand, Appl. Spec-
troscop., 45,1278 (1991).
129. Th. Bein, F. Schmidt, W. Gunsser, and G. Schmiester,. Surf: Sci., 156, 57
(1985).
130. Y. Wada, Y. Yoshizawa, and A. Morikawa, Shokubai, 32, 328 (1990).
131. G. J. Li, T. Fujimoto, A. Fukuoka, and M. Ichikawa, Stud. Surf: Sci. Catal.,
75, 1607 (1993); Catal. Lett., 12, 171 (1992).
132. L. Dixit, G. Lu, and L. Guczi, Zeolites, 14, 588 (1994).
133. P. Gallezot, Catal. Rev.-Sci. Eng., 20, 121 (1979).
134. R. Davis, Heterogeneous Chem., Rev., 1 , 41 (1994).
135. R. A. Della Betta and M. Boudart, Proc. 5th Znt. Congr. Catal., Miami (J. H.
Hightower, ed.), 1973, p. 1329.
136. P. Gallezot, R. Weber, R. A. Della Betta, and M. Boudart, 2. Naturforsch.,
43a, 40 (1979).
137. J. R. Bernard, Proc. 5th Int. Zeol. Conf:,Naples (L. V. C. Rees, ed.), Heyden,
London, 1980, p. 686.
600 BARTHOMEUF

138. C. Besoukhanova, M. Breysse, J. R. Bernard, and D. Barthomeuf, Stud. Surf.


Sci. Catal., 6 , 201 (1980).
139. C. Besoukhanova, M. Breysse, J. R. Bernard, and D. Barthomeuf, Proc. 7th
Int. Congr. Catal., Tokyo (T. Seiyama and K. Tanabe, eds.), Kodanska, Tokyo;
Elsevier, Amsterdam, 1980, Vol. B, p. 1410.
140. I. R. Leith, J. Chem. SOC. Chem. Commun., 1983, p, 93.
141. G. Larsen and G. L. Haller, Catal. Lett., 3, 103 (1989).
142. B. J. Mc Hugh, G. Larsen, and G. L. Haller, J . Phys. Chem., 94, 8621 (1990).
143. A. de Mallmann, Thesis 3" cycle, Paris (1986), and Thesis, Paris (1989); A.
de Mallmann and D. Barthomeuf, to be published.
144. A. de Mallmann and D. Barthomeuf, Stud. Surf: Sci. Catal., 46, 429 (1989).
145. A. de Mallmann and D. Barthomeuf, J. Chim. Phys., 87, 535 (1990); K. J.
Chao, A. de Mallmann, and D. Barthomeuf, J. Chim. Phys., 85, 449 (1988).
Downloaded by [Yale University Library] at 01:03 15 August 2013

146. F. Stoop, F. J. C. M. Toolenaar, and V. Ponec, J. Catal., 73, 50 (1982).


147. M. Primet, J. Catal., 88, 273 (1984).
148. G. Larsen and G. H. Haller, 1st Conf.Adv. Catal. Sci. Technol., Tokyo, 1991,
Vol. 1, p. 135.
149. W. J. Han, A. B. Kooh, and R. F. Hicks, Catal. Lett., 18, 193 (1993).
150. L. M. Kustov, D. Ostgard, and W. M. H. Sachtler, Catal. Lett., 9, 121 (1991).
151. G. S. Lane, J. T. Miller, F. S. Modica, and M. K. Barr, J . Catal., 141, 465
(1993).
152. M. J. Kappers, M. Vaarkamp, J. T. Miller, F. S. Modica, M. K. Barr, J. H.
Van der Maas, and D. C. Koningsberger, Catal. Lett., 21, 235 (1993).
153. B. L. Su, Thesis, Paris (1992).
154. B. L. Su, D. Barthomeuf, and M. Che, to be published.
155. D. Barthomeuf, in Acidity and Basicity of Solids (J. Fraissard and L. Petrakis,
eds.), NATO AS1 Series C, 444, 181 (1994).
156. N. D. Triantafillou, J. T. Miller, and B. C. Gates, Div. Petrol. Chem. Am.
Chem. SOC.,206th Natl. Meeting, Chicago, 1993, 812.
157. (a) V. B. Kazansky, V. Yu. Borovkov, N. Sokolova, N. I. Jaeger, and G .
Schulz-Ekloff, Catal. Lett., 23, 263 (1994). (b) V. B. Kazansky, V. Yu. Bo-
rovkov, and E. G. Derouane, Catal. Lett., 19, 327 (1993).
158. X. Jiang, Z. N. Gao, Z. K. Ruan, S. S. Sheng, J. S. Huang, X. H. Luo, Y. D.
Xu, and X. X. Guo, Div. Petrol. Chern., Am. Chem. SOC.,206th Natl. Meeting,
Chicago, 1993, 55.
159. M. Vaarkamp, F. S. Modica, J. T. Miller, and D. C. Koningsberger, J. Catal.,
144, 611 (1993).
160. A. Yu. Stakheev and W. M. H. Sachtler, J. Chem. SOC.Faraday Trans., 87,
3703 (1991).
161. S . T. Homeyer, Z. Karpinski, and W. M. H. Sachtler, J. Catal., 123, 60 (1990).
162. D. Barthomeuf, Stud. Surf. Sci. Catal., 65, 157 (1991).
163. A. Sauvage, P. Massiani, M. Briend, D. Barthomeuf, and F. Bozon-Verduraz,
J. Chem. SOC.Faraday Trans., 91, 3291 (1995).
164. P. Gallezot, J. Datka, J. Massardier, M. Primet, and B. Imelik, Proc. 6th Int.
Congr. Catal., London (G. C. Bond, P. B. Wells, and E. C. Tompkins, eds.),
Chemical Society, 2, 1977, Vol. 2, p. 696.
165. G. Larsen and G. H. Haller, Proc. 9th Int. Zeol. Conf:,Montreal (R. Von
BASIC ZEOLITES 601

Ballmoss, J. B. Higgins, and M. M. J. Treacy, eds.), Butterworth-Heinemann,


Boston, 1993, Vol. 11, p. 441.
166. G. Xu and J. Wang, Shiyou Xuebao, Shiyou Jiagong, 8, 29 (1992).
167. B. M. Choudary, K. Lazar, I. Bogyay, and L. Guczi, J. Chem. SOC. Faraday
Trans., 86, 419 (1990).
168. S. K. Purnell, J. R. Chang, and B. C. Gates, J. Phys. Chem., 97, 4196 (1993).
169. G. Larsen, D. E. Resasco, V. A. Durante, J. Kim, and G. L. Haller, Stud. Surf:
Sci. Catal., 83, 321 (1994).
170. J. A. Rabo, V. Schomaker, and P. E. Pickert, Proc. 3rd Int. Congr. Catal.,
1965, Vol. 2, p. 1264.
171. T. R. Hughes, W. C. Buss, P. W. Tamm, and R. L. Jacobson, Stud. Surf: Sci.
Catal., 28, 725 (1986).
172. M. Vaarkamp, J. T. Miller, F. S. Modica, G. S. Lane, and D. C. Koningsberger,
Downloaded by [Yale University Library] at 01:03 15 August 2013

J. Catal., 138, 675 (1992).


173. G. B. Mc Vicker, J. L. Kao, J. J. Ziemak, W. E. Gates, J. L. Robbins, M. M.
J. Treacy, S. B. Rice, T. H. Vanderspurt, V. R. Cross, and A. K. Ghosh, J.
Catal., 139, 48 (1993).
174. J. L. Kao, G. B. Mc Vicker, M. M. J. Treacy, S. B. Rice, J. L. Robbins, W.
E. Gates, J. J. Ziemiak, V. R. Cross, and T. H. Vanderspurt, Stud. Surf: Sci.
Catal., 75, 1019 (1993).
175. G. Larsen and G. H. Haller, Catal. Lett., 17, 127 (1993).
176. M. Sugimoto, H. Katsuno, T. Hayasaka, N. Ishikawa, and K. I. Hirasawa,
Appl. Catal. A General, 102, 167 (1993); N. Ishikawa, K. Hirasawa, M. Sug-
imoto, H. Katsuno, T. Hayasaka, and T. Takyu, Jpn. J. Appl. Phys., Part 1,
1993, p. 32; M. Sugimoto, Sekiyu Gakkaishi, 37, 10 (1994).
177. M. Vaarkamp, J. V. Grondelle, J. T. Miller, D. J. Sajkowski, F. S. Modica, G.
S. Lane, B. G. Gates, and D. C. Koningsberger, Catal. Lett., 6, 369 (1990).
178. E. Mielczarski, S. B. Hong, R. J. Davis, and M. E. Davis, J. Catal., 134, 359
(1992); R. J. Davis and E. Mielczarski, in Selectivity in Catalysis (M. E. Davis
and E. Mielczarski, eds.), ACS Symp. Ser., 517, 327 (1993).
179. J. Zheng, J. L. Dong, and Q . H. Xu, Chin. Chem. Lett., 5, 339 (1994).
180. S. J. Cho, S. M. Jung, Y. G. Shul, and R. Ryoo, J. Phys. Chem., 96, 9922
(1992).
181. R. T. Sanderson, Chemical Bonds and Bond Energy, Academic Press, New
York, 1976.
182. R. T. Sanderson, J. Am. Chem. SOC.,105, 2259 (1983).
183. D. Barthomeuf and A. de Mallmann, Stud. Surf: Sci. Catal., 37, 365 (1988).
184. S. Beran and J. Dubsky, J. Phys. Chem., 83, 2538 (1979).
185. G. M. Zhidomirov and V. B. Kazansky, Adv. Catal., 34, 131 (1986).
186. E. G. Derouane and J. G. Fripiat, J. Phys. Chem., 91, 145 (1987).
187. E. Kassab, K. Seiti, and M. Allavena, J. Phys. Chem., 92, 67 (1988).
188. P. D. O’Malley and J. Dwyer, Chem. Phys. Lett., 143, 97 (1988).
189. A. Goursot, F. Fajula, C. Daul, and J. Weber, J . Phys. Chem., 92,4456 (1988).
190. K. P. Schroder and 3. Sauer, Proc. 9th In?. Zeol. Conf:, Montreal (R. Von
Ballmoos, J. B. Higgins, and M. M. J. Treacy, eds.), Butterworth-Heinemann,
Boston, 1993, Vol. I, p. 687.
191. W. J. Mortier, J. Catal., 55, 138 (1978).
602 BARTHOMEUF

192. D. H. Dompas, W. J. Mortier, 0. C. H. Kenter, M. J. G. Janssen, and J. P.


Verduijn, J . Cutul., 129, 19 (1991).
193. W. J. Mortier, Stud. Surf: Sci. Cutul., 37, 253 (1988).
194. N. Giordano, L. Pino, S. Cavallaro, P. Vitarelli, and B. S . Rao, Zeolites, 7,
131 (1987).
195. M. Huang and S. Kaliaguine, J. Chem. SOC.Faruduy Trans., 88, 751 (1992).
196. M. Huang, A. Adnot, and S. Kaliaguine, J. Am. Chem. SOC., 114, 10005
(1992).
197. J. Xie, M. Huang, and S . Kaliaguine, Cutul. Lett., 29, 281 (1994).
198. L. Uytterhoeven, D. Dompas, and W. J. Mortier, J . Chem. SOC. Furuduy
Trans., 88, 2753 (1992).
199. W. A. Wachter, Proc. 6th Znt. Conf: Zeol., Reno (D. Olson and H. Bisio, eds.),
Butterworth, Guilford, 1984, p. 141.
Downloaded by [Yale University Library] at 01:03 15 August 2013

200. D. Barthomeuf, Muter. Chem. Phys., 17, 49 (1987); Stud. Surf: Sci. Catul.,
38, 177 (1988).
201. M. Iwamoto, M. Tajima, and S . Kagawa, J. Chem. SOC. Chem. Commun.,
1986, p. 598.
202. M. Sawa, M. Niwa, and Y. Murakami, Zeolites, 10, 532 (1990).
203. A. L. Klyachko and I. V. Michin, Neft. Khim., 30(3), 339 (1990).
204. H. Stach, J. Jaenchen, and H. Lohse, Cutul. Lett., 13, 389 (1992).
205. H. Stach and J. Jaenchen, Zeolites, 12, 152 (1992).
206. H. Stach, J. Jaenchen, H. G. Jerschekewitz, U. Lohse, B. Parlitz, and M.
Hunger, J. Phys. Chem., 96, 8480 (1992).
207. I. V. Mishin, A. L. Klyachko, T. R. Brueva, 0. P. Tkachenko, and H. K.
Beyer, Kin. Cut., 34, 562 (1993) (Kin. Cat., 34, 1993, p. 502).
208. G . Engelhardt and D. Michel, High Resolution Solid-state NMR of Silicates
and Zeolites, Wiley, Chichester, 1987; G. Engelhardt, Stud. Surf. Sci. Cutul.,
52, 151 (1991).
209. H. K. C . Timken, G . L. Turner, J. P. Gilson, L. B. Welsch, and E. Oldfield,
J. Am. Chem. SOC., 108, 7231 (1986).
210. H. K. C. Timken, N. Janes, G. L. Turner, S. L. Lambert, L. B. Welsch, and
E. Oldfied, J. Am. Chem. SOC., 108, 7236 (1986).
211. M. A. Camblor, A. Corma, R. M. Martin-Aranda, and J. Perez-Pariente, Proc.
9th Int. Zeol. Conf:, Montreal (R. Von Ballmoos, J. B. Higgins, and M. M. J.
Treacy, eds.), Butterworth-Heinemann, Boston, 1993, Vol. 11, p. 647.
212. G. Engelhardt and R. Radeglia, Chem. Phys. Lett., 108, 271 (1984).
213. R. Radeglia and G. Engelhardt, Chem. Phys. Lett., 114, 28 (1985).
214. J. F. Tempere, D. Delafosse, and J. P. Contour, Chem. Phys. Lett., 33, 95
(1975).
215. T. L. Barr, Appl. Surf:Sci., 15, 1 (1983).
216. T. L. Barr and M. A. Lishka, J. Am. Chem. SOC., 108, 3178 (1986).
217. Y. Okamoto, M. Ogawa, A. Maezawa, and T. Imanaka, J. Catal., 112, 427
(1988).
218. Y. Okamoto, Zeoraito, 10, 195 (1993).
219. M. Huang, A. Adnot, and S . Kaliaguine, J. Cutal., 137, 322 (1992).
220. S. J. Kulkarni, Indian J. Chem. Sect. A: Inorg. Phys., Theor. Anal., 29A(11),
1125 (1990).
BASIC ZEOLITES 603

221. M. J. Edgell, R. W. Paynter, S. C. Mugford, and J. E. Castle, Zeolites, 10, 51


(1990).
222. J. Stoch, J. Lercher, and S . Ceckiewicz, Zeolites, 12, 81 (1992).
223. K. Tanabe, in Catalysis Science and Technology (J. R. Anderson and M. Bou-
dart, eds.), Springler-Verlag, Berlin, 1981, Vol. 2, Chap. 5.
224. K. Tanabe, M. Misono, Y. Ono, and H. Hattori, Stud. Surf. Sci. Catal., 51,
1989: a, p. 14; b, p. 215; c, p. 225.
225. J. C . Lavalley, Trends Phys. Chem., 2, 305 (1991).
226. C. Lahousse, J. Bachelier, J. C. Lavalley, H. Lauron-Pernot, and A. M. Le
Govic, J. Mol. Catal., 87, 329 (1994).
227. E. A. Paukshtis, N. S. Kotsarenko, and L. G. Karakchiev, React. Kin. Kat.
Lett., 12, 315 (1979).
228. P. 0. Scokart and P. G. Rouxhet, J. Chem. SOC. Faraday Trans. I, 76, 1476
Downloaded by [Yale University Library] at 01:03 15 August 2013

(1980).
229. A. Gervasini and A. Auroux, J. Therm. Anal., 37, 1737 (1991).
230. A. Auroux and A. Gervasini, J. Phys. Chem., 94, 6371 (1990).
231. H. G. Karge, M. Laniecki, and M. Ziolek, Proc. 7th Int. Zeol. Con$, Tokyo
(Y. Murakami, A. Iijima, and J. W. Ward, eds.), Kodansha, Tokyo; Elsevier,
Amsterdam, 1986, p. 617.
232. M. Waqif, A. M. Saad, M. Bensitel, J. Bachelier, 0. Saur, and J. C . Lavalley,
J. Chem. SOC. Faraday Trans., 88, 2931 (1992).
233. H. Lauron-Pernot, F. Luck, and J. M. Popa, Appl. Catal., 78, 213 (1991).
234. M. Huang and S . Kaliaguine, Catal. Lett., 18, 373 (1993).
235. P. F. Rossi, G. Busca, V. Lorenzelli, M. Lion, and J. C. Lavalley, J . Catal.,
109, 378 (1988).
236. M. Berkani, J. L. Lemberton, M. Marczewski, and G. Perot, Catal. Lett., 31,
405 (1995).
237. W. Przystajko, R. Fiedorow, and I. G. Dalla Lana, Zeolites, 7, 477 (1987).
238. N. Nagaraju, S. P. Walvekar, and N. M. Nanje Gowda, Ind. J. Technol., 28,
59 (1990).
239. Can Li, Shi-fu Fu, Hui Zhang, and Qin Xin, J. Chem. SOC. Chem. Commun.,
1, 17 (1994).
240. 0. V. Krylov and E. A. Fokina, Problemy KinCt. Katal., Akad. Nauk, Moskva,
8, 246 (1955).
241. S. Dzwigaj, A. de Mallmann, and D. Barthomeuf, J. Chem. SOC. Faraday
Trans., 86, 431 (1990).
242. T. Yamanaka and K. Tanabe, J. Phys. Chem., 79, 2409 (1975).
243. T. R. Brueva, A. L. Klachko-Gurvich, and A. M. Rubinstein, Im.Akad. Nauk
SSSR, Ser. Khim., 12, 2807 (1972).
244. D. Barthomeuf and B. H. Ha, J. Chem. SOC.Faraday Trans. I, 69,2147,2158
(1973).
245. G. Zhang, H. Hattori, and K. Tanabe, Appl. Catal., 36, 189 (1988).
246. A. M. Maitra, I. Campbell, and R. J. Tyler, Appl. Catal. A General, 85, 27
(1992).
247. P. Cartraud, B. Chauveau, M. Bernard, and A. Cointot, J. Therm. Anal., 11,
51 (1977).
248. S. S. Khvoshchev, V.E. Skazyvaev, and E. A. Vasiljeva, Proc. 5th Int. Zeolite
Conf., Naples (L. V. C. Rees, ed.), Heyden, London, 1980, p. 476.
604 BARTHOMEUF

249. D. Amari, J. M. Lopez Cuesta, N. P. Nguyen, R. Jerrentrup, and J. L. Ginoux,


J , Therm. Anal., 38, 1005 (1992).
250. P. N. Joshi and V. P. Shiralkar, J. Phys. Chem., 97, 619 (1993).
251. S. A. Zheng, J. J. Cai, and D. C. Liu, Proc. 9th Int. Congr. Catal. (M. J.
Phillips and M. Ternan, eds.), Chem. Institute of Canada, Ottawa, 1988, Vol.
1, p. 476.
252. H. Knozinger, to be published.
253. H. Tsuji, F. Yagi, and H. Hattori, Chem. Lett., 1991, p. 1881.
254. J. C. Kim, H. X. Li, C. Y. Chen, and M. E. Davis, Microp. Muter., 2, 413
(1994).
255. M. Huang and S . Kaliaguine, Stud. Surf: Sci. Catal., 73, 291 (1992).
256. C. L. Angel1 and M. V. Howell, Canad. J. Chem., 47, 3811 (1969).
Downloaded by [Yale University Library] at 01:03 15 August 2013

257. P. A. Jacobs, F. H. Van Cauwelaert, E. F. Vansant, and J. B. Uytterhoeven, J.


Chem. SOC. Faraday Trans. Z, 69, 1056 (1973).
258. N. Kh. Tas-001, A. A. Kubasov, and K. V. Topchieva, Kin. Cat., 19, 200
(1978).
259. Y. Delaval and E. Cohen de Lara, J. Chem. SOC. Faraday Trans. I, 77, 869
(1981).
260. H. Forster and M. Schumann, J. Chem. SOC. Faraday Trans. I, 85, 1149 and
references therein (1989).
261. L. J. Bellamy and H. E. Hallam, Trans. Faraday SOC.,54, 1120 (1958).
262. M. S. Nozari and R. S. Drago, J. Am. Chem. SOC., 92, 7086 (1970).
263. G. Poncelet and M. L. Dubru, J. Catal., 52, 321 (1978).
264. J. L. Schlenker, J. J. Pluth, and J. V. Smith, Muter. Res. Bull., 14, 751 (1979).
265. A. de Mallmann and D. Barthomeuf, Proc. 7th Znt. Zeol. Conf:, Tokyo (Y.
Murakami, A. Iijima, and J. W. Ward, eds.), Kodanska, Tokyo; Elsevier, Am-
sterdam, 1986, p. 609.
266. A. de Mallmann and D. Barthomeuf, Zeolites, 8, 292 (1988).
267. A. de Mallmann and D. Barthomeuf, J. Chem. SOC. Chem. Commun., 1989,
p. 129.
268. D. Barthomeuf and A. de Mallmann, Ind. Eng. Chem. Res., 29, 1435 (1990).
269. B. L. Su and D. Barthomeuf, J. Catal., 139, 81 (1993).
270. B. L. Su and D. Barthomeuf, Zeolites, 13, 626 (1993).
271. (a) M. Ziolek, D. Szuba, and R. Leksowski, Stud. Surf: Sci. Catal., 37, 427
(1988); (b) M. Ziolek and Z. Dudzik, Zeolites, 1, 117 (1981).
272. M. Sugioka, M. Amisawa, H. Abe, and K. Sato, Stud. Surf: Sci. Catal., 84C,
1603 (1994).
273. C. J. Blower and T. D. Smith, J. Chem. SOC. Faraday Trans., 90, 931 (1994).
274. S . Bordiga, E. Garrone, C. Lamberti, A. Zecchina, C. 0. Arean, V. B. Kazan-
sky, and L. M. Kustov, J. Chem. SOC.Faraday Trans., 90, 3367 (1994).
275. R. B. Borade, M. Huang, A. Adnot, A. Sayari, and S . Kaliaguine, Proc. 10th
Znt. Congr. Catal., Budapest (L. Guczi, F. Solymosi, and P. Tetenyi, eds.),
Akamediai Kiado, Budapest, 1993, Vol. B, p. 1625.
276. C. Bremard and M. Le Maire, J. Phys. Chem., 97, 9695 (1993).
277. Y. Ono, Stud. Surf. Sci. Catal., 5, 19 (1980).
278. D. Barthomeuf, G. Coudurier, and J. Vedrine, Muter. Chem. Phys., 18, 553
(1988).
BASIC ZEOLITES 605

279. A. Corma, V. Fornes, R. M. Martin-Aranda, H. Garcia, and J. Primo, Appl.


Catal., 59, 237 (1990).
280. A. Corma, R. M. Martin-Aranda, and F. Sanchez, Stud. SurJ Sci. Catal., 59,
503 (1991).
281. C. L. Angel1 and M. V. Howell, J . Colloid Znterfac. Sci., 28, 279 (1968).
282. A. G. Bezus, A. V. Kiselev, Z. Sedlacek, and P. Q . Du, Trans. Faraday SOC.,
67, 468 (1971).
283. H. Pfeifer, W. Schirmer, and H. Winkler, Adv. Chem. Ser., 121, 430 (1973).
284. D. Michel, W. Meiler, and H. Pfeifer, J. Mol. Catal., 1 , 85 (1975).
285. V. Yu. Borovkov, W. K. Hall, and V. B. Kazansky, J. Catal., 51, 437 (1978).
286. B. Coughlan, W. Caroll, P. O’Malley, and J. Nunan, J. Chem. SOC.Faraday
Trans., 77, 3037 (1981).
287. J. Sauer and D. Deininger, Zeolites, 2, 114 (1982).
Downloaded by [Yale University Library] at 01:03 15 August 2013

288. H. Jobic, M. Bee, and A. Renouprez, Surf. Sci., 140, 307 (1984).
289. J. M. Newsam, J. Phys. Chem., 93, 7689 (1989).
290. B. G. Silbernagel, A. R. Garcia, and J. M. Newsam, Colloids Surf., A72, 71
(1993).
291. J. P. Shen, J. Ma, T. Sun, D. Z. Jiang, and E. Z. Min, J. Chem. SOC.Faraday
Trans., 90, 1351 (1994).
292. B. L. Su, J. M. Manoli, C. Potvin, and D. Barthomeuf, J. Chem. SOC.Faraday
Trans., 89, 857 (1993).
293. J. J. Freeman and M. L. Unland, J. Catal., 54, 183 (1978).
294. M. L. Unland and J. J. Freeman, J. Phys. Chem., 82, 1036 (1978).
295. H. Lechert and K. P. Wittern, Ber. Bunsenges. Phys. Chem., 82, 1054 (1978).
296. M. Primet, E. Garbowski, M. V. Mathieu, and B. Imelik, J . Chem. SOC. Far-
aday Trans. I, 76, 1942 (1980).
297. A. N. Fitch, H. Jobic, and A. J. Renouprez, J. Chem. SOC.Chem. Commun.,
1985, p. 284.
298. A. N. Fitch, H. Jobic, and A. J. Renouprez, J. Phys. Chem., 90, 1311 (1986).
299. H. Jobic, A. Renouprez, A. N. Fitch, and H. J. Lauter, J. Chem. SOC.Faraday
Trans., 83, 3199 (1987).
300. B. G. Silbernargel, A. R. Garcia, J. M. Newsam, and R. Hulme, J . Phys.
Chem., 93, 6506 (1989).
301. P. Demontis, S. Yashonath, and M. L. Klein, J . Phys. Chem., 93,5016 (1989).
302. B. Coughlan and M. A. Keane, J. Chem. SOC. Faraday Trans., 86, 3961
(1990).
303. J. G. Pearson, B. F. Chmelka, D. N. Shykind, and A. Pines, J. Phys. Chem.,
96, 8517 (1992).
304. S. B. Liu, L. J. Ma, M. W. Lin, J. F. Wu, and T. L. Chen, J. Phys. Chem.,
96, 8120 (1992).
305. G. Schrimpf, M. Schlenkrich, J. Brickmann, and P. Bopp, J. Phys. Chem., 96,
7404 (1992).
306. J. A. Hriljac, M. M. Eddy, A. K. Cheetham, J. A. Donohue, and G . J. Ray,
J. Sol. State Chem., 106, 66 (1993).
307. S. B. Liu, J. F. Wu, L. J. Ma, M. W. Lin, and T. L. Chen, ACS Symp. Ser.,
517, 272 (1993).
308. L. M. Bull, S. J. Heyes, and K. Cheetham, Proc. 9th Znt. Zeol. Conf. (R. Von
606 BARTHOMEUF

Ballmoos, J. B. Higgins, and M. M. J. Treacy, eds.), Buttenvorth-Heinemann,


Boston, 1993, Vol. 11, p. 63.
309. L. M. Bull, N. J. Henson, A. K. Cheetham, J. M. Newsam, and S. J. Heyes,
J. Phys. Chem., 97, 11176 (1993).
310. V. Voss and B. Boddenberg, Surf: Sci., 298, 241 (1993).
311. D. Hasha, V. Miner, J. Garces, and S. Rocke, in Syrnp. New Surface Science
in Catalysis, ACS Meeting, Philadelphia, Aug. 26-31, 1984, p. 953.
312. B. Zibrowius, J. Caro, and H. Pfeifer, J. Chem. SOC.Faraday Trans. I, 84,
2347 (1988).
313. D. A. Fischer, J. L. Gland, and S . M. Davis, Catal. Lett., 6, 99 (1990).
314. B. L. Su and D. Barthomeuf, Appl. Catal. A General, 24, 73, 81 (1995).
315. B. L. Su and D. Barthomeuf, Zeolites, 15, 470 (1995).
316. J. F. Wu, T. L. Chen, L. J. Ma, H. W. Lin, and S . B. Liu, Zeolites, 12, 86
Downloaded by [Yale University Library] at 01:03 15 August 2013

(1992).
317. Suk Bong Hang, H. M. Cho, and M. E. Davis, J. Phys. Chem., 97, 1622,
1629 (1993).
318. B. F. Chmelka, J. G. Pearson, S. B. Liu, R. Ryoo, L. C. de Menorval, and A.
Pines, J. Phys. Chem., 95, 303 (1991).
319. C. Mellot, Thesis, Paris (1993); C. Mellot, D. Espinat, B. Rebours, Ch. Baer-
locher, and P. Fischer, Catal. Lett., 27, 159 (1994); C . Mellot and D. Espinat,
Bull. SOC. Chim. Fr., 131, 742 (1994) and Rev. Inst. Fr. Petr., 49, 667 (1994).
320. (a) H. Itoh, A. Miyamoto, and Y. Murakami, J. Catal., 64, 284 (1980). (b) A.
Corma, G. Sastre, and P. Viruela, Stud. Surf: Sci. Catal., 84C, 2171 (1994).
321. 0. M. Dzhigit, A. V. Kiselev, K. N. Milos, and G. G. Muttik, J. Chem. SOC.
Faraday Trans. I, 67, 458 (1971).
322. Si-Yang Zhang, 0. Talu, and D. T. Hayhurst, J. Phys. Chem., 95, 1722 (1991);
97, 12894 (1993).
323. Chen Lifeng, Cuihua Xuebao, 11(5), 386 (1990); Da Ming Sun, Vacuum, 45,
1175 (1994).
324. R. Goyal, A. N. Fitch, and H. Jobic, J. Chem. SOC., Chern. Comrnun., 1990,
p. 1152.
325. P. A. Wright, J. M. Thomas, A. K. Cheetham, and A. K. Nowak, Nature, 318,
611 (1985).
326. M. Czjzek, T. Vogt, and H. Fuess, Zeolites, 11, 832 (1991).
327. R. M. Barrer, Zeolites and Clay Minerals us Sorbents and Molecular Sieves,
Academic Press, London, 1976.
328. D. M. Ruthven, Principles of Adsorption and Adsorption Processes, John Wi-
ley and Sons, New York, 1984.
329. D. M. Ruthven and M. Goddard, Zeolites, 6, 275 (1986).
330. L. 0. Stine and D. B. Broughton, U.S. Patent 3,636,121 to U P 0 (1969); R.
Bearden, Jr., and R. J. de Foe, U.S. Patent 3,686,343 to Esso (1972).
331. R. W. Neuzil, U.S. Patent 3,734,974 to UOP (1973).
332. A. J. de Rosset, U.S. Patent 3,917,734 to UOP (1975).
333. (a) R. W. Neuzil and D. H. Rosback, U.S. Patent 3,943,182 to UOP (1976).
(b) P. R. Geissler, U.S. Patent 4,175,099 to Exxon (1979). (c) D. Barthomeuf,
U S . Patent 4,593,149 and 4,613,725 to Exxon (1986).
334. (a) R. W. Neuzil, U.S. Patent 4,326,092 to UOP (1982). (b) D. Barthomeuf,
U.S. Patent 4,593,150 to Exxon (1986).
BASIC ZEOLITES 607

335. (a) D. Barthomeuf, U.S. Patent 4,584,424 to Exxon (1986). (b) D. Barthomeuf
and L. G. Daniel, U.S. Patent 4,751,346 to Exxon (1988).
336. V. Solinas, R. Monaci, E. Rombi, and M. Morbidelli, Stud. Surf: Sci. Catal.,
46, 595 (1989).
337. Maomi Seko, Oil Gas J . , July 2, 1979, p. 81.
338. K. Iwayama and M. Suzuki, Stud. SurK Sci. Catal., 83, 243 (1994).
339. G. A. Olah, Acc. Chern. Res., 4, 240 (1971).
340. A. de Mallmann and D. Barthomeuf, Stud. Surf: Sci. Catal., 49B, 935 (1989).
341. J. A. Rabo, in Zeolite Chemistry and Catalysis (J. A. Rabo, ed.), ACS Mon-
ograph 171, 1976, p. 332; J. A. Rabo and P. H. Kasai, Progr. Solid State
Chern., 9, 1 (1975).
342. A. Haas, K. E. Finger, and U. Alkemade, Appl. Catal. A: General, 115, 103
Downloaded by [Yale University Library] at 01:03 15 August 2013

(1994).
343. D. Barthomeuf, Appl. Catal. A: General, 126, 187 (1995).
344. P. R. Hari Prasad Rao, P. Massiani, and D. Barthomeuf, Stud. Sure Sci. Catal.,
84B, 1449 (1994).
345. B. L. Su and D. Barthomeuf, Appl. Catal. A: General, 124, 73, 81 (1995).
346. P. B. Venuto and P. S . Landis, Adv. Catal., 18, 331 (1968).
347. J. A. Rabo and M. L. Poutsma, Adv. Chern. Ser., 102, 284 (1971).
348. P. A. Jacobs, Carboniogenic Activity of Zeolites, Elsevier, Amsterdam, 1977.
349. I. E. Maxwel1,Adv. Catal., 31, 1 (1982).
350. W. F. Holderich, M. Hesse, and F. Naumann, Angew. Chern., 27, 226 (1988).
351. W. F. Holderich, Stud. Surf: Sci. Catal., 46, 193 (1989).
352. G . Perot and M. Guisnet, J. Mol. Catal., 61, 173 (1990).
353. J. Cornier, J. Popa, and M. Gubelmann, Actual. Chim., 6, 405 (1992).
354. W. F. Holderich, Proc. 10th Int. Congr. Catal., 1992, p. 127.
355. Y. Izumi and M. Onaka,Adv. Catal., 38, 245 (1992).
356. J. Weitkamp, Proc. 9th Innt. Zeol. Con$ (R. Von Ballmoos, J. B. Higgins, and
M. M. J. Treacy, eds.), Butterworth-Heinemann, Boston, 1993, Vol. I, p. 13.
357. P. B. Venuto, Microp. Muter., 2, 297 (1994).
358. W. F. Holderich, Proc. Int. Syrnp. Acid-Base Catalysis, Sapporo (K. Tanabe,
H. Hattori, T. Yamaguchi, and T. Tanaka, eds.), Kodanska, Tokyo, 1989, p. 1.
359. Y. Xu, S. Yao, X. Liu, and W. Pang, Gaodeng Xuexiao Huaxue Xuebao, 13,
281 (1992).
360. C . B. Dartt and M. E. Davis, Catal. Today, 19, 151 (1994).
361. J. Engelhardt, J. Szani, and 3. Valyon, J . Catal., 107, 296 (1987).
362. P. E. Hathaway and M. E. Davis, J. Catal., 119, 497 (1989).
363. S . H. Park, M. H. Kim, H. J. Kim, and S . K. Moon, Hwahak Konghak, 31,
369 (1993).
364. A. Malecka and D. Barthomeuf, to be published.
365. Y. Matsumura, K. Hashimoto, and S . Yoshida, J. Catal., 100, 392 (1986).
366. Z. H. Fu and Y. Ono, Catal. Lett., 22, 277 (1993).
367. P. R. Hari Prasad Rao, P. Massiani, and D. Barthomeuf, Catal. Lett., 31, 115
(1995).
368. S . T. King and J. M. Garces, J. Catal., 104, 59 (1987).
369. N . Ya. Usachev, A. L. Lyapidus, 0. N. Usacheva, M. M. Savelev, L. L. Kras-
nova, and Kh. M. Minachev, Nepkhimiya, 33, 305 (1993).
370. M . L. Unland and G. E. Barker, U.S. Patent 4,115,424 to Monsanto (1978).
608 BARTHOMEUF

371. M. L. Unland and G. E. Barker, U S . Patent 4,140,726 to Monsanto (1979).


372. M. L. Unland and G. E. Barker, in Catalysis of Organic Reactions ( M . R.
Moser, ed.), Dekker, New York, 1981, p. 51 and references therein.
373. W. G. Guo, G. Y. Cai, J. W, Zhang, J. Liang, and G. Q. Chen, Chin. Chem.
Lett., 4, 873 (1993).
374. S. K. Moon, H. J. Kim, and K. T. Seo, Hwahak Konghak, 27, 422 (1989).
375. X . Wang, G. Wang, D. Shen, C. Fu, and M. Wei, Zeolites, 11, 254 (1991).
376. H. Itoh, T. Hattori, K. Suzuki, A. Miyamoto, and Y. Murakami, J. Catal., 72,
170 (1981).
377. H. Itoh, T. Hattori, K. Suzuki, and Y. Murakami, J. Catal., 79, 21 (1983).
378. W. Guo, Z. Zhang, J. Liang, G. Cai, and G. Che, Proc. Int. Conf.Pet. Refin.
Petrochem. Process, 1991, Vol. 3, p. 1459.
379. E. Mielczarski and M. E. Davis, Ind. Eng. Chem. Res., 29, 1579 (1990).
Downloaded by [Yale University Library] at 01:03 15 August 2013

380. M. D. Sefcik, J. Am. Chem. SOC.,101, 2164 (1979).


381. A. Philippou and M. W. Anderson, J. Am. Chem. SOC.,116, 5774 (1994).
382. Y. N. Sidorenko and P. N. Galich, Neftkhimia, 31, 54 (1991).
383. C. Yang, Z. Meng., Appl. Catal., 71, 45 (1991).
384. C. S. Huang and A. N. KO, Catal. Lett., 19, 319 (1993).
385. P. Y. Chen, M. C. Chen, H. Y. Chu, N. S. Chang, and T. K. Chuang, Proc.
7th Int. Zeol. Conf. (Y. Murakami, A. Iijima, and J. W. Ward, eds.) Kodansha,
Tokyo; Elsevier, Amsterdam, 1986, p. 739.
386. P. Y. Chen, S. J. Chu, N. S. Chang, and T. K. Chuang, Stud. Surf. Sci. Catal.,
49B, 1105 (1989).
387. R. B. C. Pillai and C. N. Pillai, Ind. J. Chem., 32B, 592 (1993).
388. Z. H. Fu and Y. Ono, Catal. Lett., 21, 43 (1993).
389. Z. H. Fu and Y. Ono, J. Catal., 145, 166 (1994).
390. D. Delafosse, J. Chim. Phys. Chim. Biol., 83, 791 (1986).
391. W. J. Han, A. B. Kooh, and R. F. Hicks, Catal. Lett., 18, 219 (1993).
392. R. F. Hicks, W. J. Han, and A. B. Kooh, Stud. Surf. Sci. Catal., 75, 1043
(1993).
393. P. W. Tamm, D. H. Mohr, and C. R. Wilson, Stud. Surf. Sci. Catal., 38, 335
(1987).
394. J. Dong, C. Jin, and Q. Hu, Ranliao Huaxue Xuebao, 20, 244 (1992).
395. Z. K. Ruan, X. Jiang, Z. Gao, Y. Zhang, and J. Huang, Shiyou Huagong, 22,
711 (1993); X . Jiang, Z. Gao, Z. Ruan, J. Huang, Y. Xu, X. Guo, X. Luo,
and J. He, Ranliao Huaxue Xuebao, 22, 9 (1994); Z. Ruan, X. Jiang, Z. Gao,
Y. Zhang, and Y. Xu, Shiyou Huagong, 23, 281 (1994).
396. L. X. Dai, H. Sakashita, and T. Tatsumi, Bull. Chem. SOC. Jpn., 67, 1553
(1994).
397. T. Tatsumi, L. X. Dai, and H. Sakashita, Catal. Lett., 27, 289 (1994).
398. A. B. Kooh, W. J. Han, and R. F. Hicks, Catal. Lett., 18, 209 (1993).
399. R. J. Davis and E. G. Derouane, Nature, 349, 313 (1991).
400. S. J. Tauster and J. J. Steger, J. Catal., 125, 387 (1990); Muter. Res. SOC.
Symp. Proc., 111, 419 (1988).
401. E. G. Derouane and D. J. Vanderveken, Appl. Catal., 45, 11 (1988).
402. Z. Zhan, I. Manninger, Z. Paal, and D. Barthomeuf, J. Catal., 147,333 (1994).
403. W. E. Alvarez and D. E. Resasco, Catal. Lett., 8, 53 (1991).
404. I. Manninger, Z. Zhan, X. L. Xu, and Z. Paal, J. Mol. Catal., 66, 223 (1991).
BASIC ZEOLITES 609

405. Z. Gao, X. Jiang, Z. Ruan, and Y. Xu, Catal. Lett., 19, 81 (1993).
406. Y. Z. Shi, X. Jiang, Y. Zhang, Z. K. Ruan, Y. Zhang, X. X. Guo, Cuihua
Xuebao, 14, 312 (1993).
407. S. B. Sharma, P. P. Ouraipryvan, H. A. Nair, P. Balaraman, T. W. Root, and
J. A. Dumesic, J. Catal., 150, 234 (1994).
408. E. Iglesia and J. E. Baumgartner, Stud. Surf: Sci. Catal., 75, 993 (1993).
409. E. Iglesia and J. E. Baumgartner, Proc. 9th Int. Zeol. Con& (R. Von Ballmoos,
J. B. Higgins, and M. M. J. Treacy, eds.), Buttenvorth-Heinemann, Boston,
1993, Vol. 11, p. 421.
410. P. Meriaudeau, A. Thangaraj, C. Naccache, and S. Narayanan, J. Catal., 146,
579 (1994).
411. L. X. Dai, H. Sakashita, and T. Tatsumi, J. Catal., 147, 311 (1994).
412. M. Vaarkamp, J. Van Grondelle, R. A. Van Santen, J. T. Miller, B. L. Meyers,
Downloaded by [Yale University Library] at 01:03 15 August 2013

F. S. Modica, G. S. Lane, and D. C. Koningsberger, Proc. 9th Int. Zeol. Conf:


(R. Von Ballmoos, J. B. Higgins, and M. M. J. Treacy, eds.), Buttenvorth-
Heinemann, Boston, 1993, Vol. 11, p. 433.
413. S. B. Hong, E. Mielczarski, and M. E. Davis, J. Catal., 134, 349 (1992).
414. C. Dossi, R. Psaro, A. Bartsch, A. Fusi, L. Sordelli, R. Ugo, M. Bellatreccia,
R. Zanoni, and G. Vlaic, J. Catal., 145, 377 (1994).
415. H. Katsuno, T. Fukunaga, and M. Sugimoto, Stud. Surf: Sci. Catal., 75, 2419
(1993).
416. M. Sugimoto, T. Murakawa, T. Hirano, and H. Ohashi, Appl. Catal. A: Gen-
eral, 95, 257 (1993).
417. J. Y. Wang, Q. P. Zhang, X. F. Yang, and G. C. Xu, Shiyou Xuebao, Shiyou
Jiagong, 9, 26 (1993).
418. E. Ruckenstein and P. G. Smirniotis, Catal. Lett., 14, 123 (1994).
419. T. Tatsumi, M. Taniguchi, L. X. Dai, and H. Tominaga, Sekiyu Gakkaishi, 37,
553 (1994).
420. J. L. Dong, J. H. Zhu, and Q. H. Xu, Shiyou Huagong, 22, 225 (1993).
421. M. A. Chaar and J. B. Butt, Appl. Catal., A : General, 114, 287 (1994).
422. G. M. Pajonk and A. El-Tanany, Catal. Lett., 6, 173 (1990).
423. Z. Gao, J. Cui, and S. Pu, Chin. J. Chem., 10, 320 (1992).
424. L. P. Chirinskaia, V. S. Komarov, P. P. Oubranovitch, M. Th. Roussak, and
N. S. Kozlov, Neftkhim, 14, 568 (1974).
425. B. Coughlan and M. A. Keane, Catal. Lett., 5, 89 (1990); Zeolites, 11, 483
(1991).
426. P. A. Jacobs and J. B. Uytterhoeven, J. Catal., 50, 109 (1977).
427. J. B. Nagy, J. P. Lange, A. Gourgue, P. Bodart, and Z. Gabelica, Stud. Surf:
Sci. Catal., 20, 127 (1985).
428. M. L. Unland, J. Phys. Chem., 82, 580 (1978).
429. J. Bandiera and C. Naccache, Appl. Catal., 69, 139 (1991).
430. E. Santacesaria, D. Gelosa, E. Giorgi, and S. Carra, J. Catal., 90, 1 (1984).
431. Ts. Bezukhanova, Yu. Kalvachev, and H. Lechert, Chem. Znd. (Dekker), 47,
185 (1992); Stud. Surf: Sci. Catal., 75, 1739 (1993).
432. J. L. Dong, J. H. Zhu, and Q. H. Xu, Appl. Catal., A: General, 112, 105
(1994).
433. J. H. Zhu, J. I. Dong, Q. H. Xu, and J. Zhang, CuihuaXuebao, 15,268 (1994).
610 BARTHOMEUF

434. M. M. Sung, M. Y. Youn, Y. Kim, and H. N. Paik, Tae Hun Hwahakhoe Chi,
32, 501 (1988).
435. K. B. Yoon and J. K. Kochi, Catal. Lett., 1, 183 (1988).
436. L. Xu, G. Chen, G. Cai, and Q. Wang, Cuihua Xuebao, 11, 442 (1990).
437. P. Gallezot, A. Giroir-Fendler, and D. Richard, Catal. Lett., 5, 169 (1990).
438. D. G. Blackmond, R. Oukaci, B. Blanc, and P. Gallezot, J. Catal., 131, 401
(1991).
439. P. Gallezot, B. Blanc, D. Barthomeuf, and M. I. Pais da Silva, Stud. Surf:
Catal. Sci., 84B,1433 (1994).
440. A. Corma, Muter. Res. SOC. Symp. Proc., 233, 17 (1991).
441. P. T. Wierzchowski and L. W. Zatorski, Catal. Lett., 9, 411 (1991).
442. E. J. Rode, P. E. Gee, L. N. Marquez, T. Uemura, and M. Bazargani, Catal.
Lett., 9, 103 (1991).
Downloaded by [Yale University Library] at 01:03 15 August 2013

443. R. M. Dessau, U.S. Patent 5,026,919 to Mobil (1991).


444. C. Yang and Z. Y. Meng, J. Catal., 142, 37 (1993).
445. C. Yang and Z. Y. Meng, Cuihua Xuebao, 15, 28 (1994).
446. H. Tsuji, F. Yagi, and H. Hattori, Chem. Lett., Chem. SOC.Jpn., 11, 1881
(1991).
447. T. F. Brownscombe, U.S. Patent 4,992,613 to Shell (1991).
448. A. M. Efstathiou, S. L. Suib, and C. 0. Bennett, J. Catal., 135, 223 (1992).
449. E. Mantovani, N. Palladino, and A. Zanobi, J . Mol. Catal., 3, 285 (1977/78).
450. Z. Karpinski, Z. Zhang, and W. M. H. Sachtler, J. Mol. Catal., 77, 181 (1992).
451. F. Q. Ma, D. S. Lu, and Z. Y. Guo, J. Mol. Catal., 78, 309 (1993).
452. A. V. Mashkina and V. N. Yakovleva, Kin. Cat., 32, 566 (1991).
453. P. Fellmuth, W. Lutz, and M. Bulow, Zeolites, 1, 367 (1987).
454. I. Ferino, R. Monaci, B. V. Solinas, C. Oliva, I. Peiri, and L. Forni, J . Chem.
SOC.Faraday Trans., 86, 193 (1990).
455. M. Ziolek and P. Decyk, Stud. Surf: Sci. Catal., 84, 1579 (1994).
456. F. P. Gortsema, B. Beshty, J. J. Friedman, D. Matsumoto, J. J. Sharkey, G.
Wildman, T. J. Blacklock, and S. H. Pan, 14th Conf. Catal. Organ. React.,
Albuquerque, April 1992.
457. D. K. Murray, J. W. Chang, and J. F. Haw, J. Am. Chem. SOC.,115,4732 (1993).
458. S . M. Nad, E. A. Zubkov, V. G. Shubin, A. G. Pelmenshchikov, L. A. Vos-
trikova, and K. G. Ione, Mendeleev Commun., 3, 94 (1991).
459. I. Mochida, Y. Yasumoto, H. Fujitsu, and Y. Kojima, Sekiyu Gakkaishi, 36,
448 (1993); I. Mochida, Y. Kojima, and H. Tejima, Jpn. Kokai, Tokyo koho
J.P. 05,132,435 (93,132,435) (1993).
460. L. M. Parker, D. M. Bibby, and I. J. Miller, J. Catal., 129, 438 (1991).
461. F. Blatter and H. Frei, J . Am. Chem. SOC.,115, 7501 (1993); 116, 1812 (1994);
H. Sun, F. Blatter, and H. Frei, J. Am. Chem. SOC.,116, 7951 (1994).
462. V. B. Kazansky, Stud. Sure Sci. Catal., 65, 117 (1991).
463. J. Scherzer, in Catalytic Materials: Relationship Between Structure and Reac-
tivity (T. E. Whyte, Jr., R. A. Della Betta, E. G. Derouane, and R. T. K. Baker,
eds.), ACS Symp. Ser., 1984, p. 157.

Note Added in Proof. Since June 1995, when this paper was submitted, sev-
eral publications have appeared bringing new information on basic zeolites.
BASIC ZEOLITES 611

With regards to the basicity characterization using pyrrole [24,25,183,


195,197,228,2551, a detailed infrared study distinguishes various types of interac-
tion between pyrrole and oxide surfaces. In NaX the low oxygen basicity does not
dissociate pyrrole [464]. More generally, the conjugated acid-base pair species in
basic zeolites (metal cation and framework oxygen [24]) were characterized using
infrared and XPS applied to adsorbed pyrrole [195,219]. The generation of two
different basic sites when oxygen is conjugated to two different cations-for in-
stance, Na and K, Rb, or Cs-is confirmed by microcalorimetry [465]. The results
show that in addition to the local basic property a long-range effect may modify the
overall basic strength [465]. A computer deconvolution of the infrared spectra of
pyrrole shows in EMT [466a] and X and Y [466b] exchanged with alkali cations
that the framework oxygens acquire a basic strength specific to each type of location
of the close cation (SI, S:, S,, . . .) in addition to the effect of the identity of the
Downloaded by [Yale University Library] at 01:03 15 August 2013

cation itself.
The interaction of Pd2+ ions with framework oxygen in NaX studied by MAS
NMR of "Si [467] is the highest in the dehydrated zeolite when Pd" ions are located
close to strongly basic sites exhibiting the Al-richest environment.
For reduced metal clusters, the benzene hydrogenation is higher in PdNaY than
in PdNaX [468]. This order, in agreement with that of the basicity of the support, is
comparable to the one obtained in the case of Pt metal clusters ([145]; Fig. 9). For
Pt supported on NaX or hydrotalcite the strong Pt-C interaction when CO is ad-
sorbed on the metal is studied by diffuse reflectance IR [157,469]. It is proposed to
arise from a negative charging of the platinum surface which originates from the
polarization of very small Pt particles by negatively charged surface oxygen of the
support (so-called Schottky barrier at the metal-oxide interfaces) [469]. Such an
explanation defines precisely the effect of the electric field in the cages [143].
The study of the adsorption (and coadsorption) of methanol and ammonia,
carried out by infrared spectroscopy on mordenite and erionite, indicates that the
interaction with the cations (Na, K) and the reactivity to give dimethylether or mon-
omethylamine strongly depend on the Al content [470]. Dimethylether would result
from a high interaction (hydrogen bonding) between the hydroxyls of methanol and
the charged lattice oxygen atoms (case of high Al content).
New basic materials may be formed by addition of hydroxides or oxides, not
only in basic or neutral Si-Al zeolites [28,29,45-50,51-631, but also in mesoporous
molecular sieves. Besides the basic properties of Na-MCM-41 or Cs-MCM-41
(prepared by ion exchange) in the base-catalyzed Knoevenagel condensation, the
cesium acetate impregnated MCM-41 (Cs in excess) shows strong base activity in
the Michael addition and seems promising for superbase catalysis [471]. More gen-
erally one may expect that the chemistry of highly dispersed basic oxides in various
porous supports (neutral or basic) will open an expanding field for the generation of
tailor-made basic catalysts. Both the support and the dispersed oxide may gain new
properties. For instance, the impregnation with Cs acetate of previously exchanged
CsNaX and CsNaY generates occluded species which are more strongly basic when
located in CsNaX than in CsNaY [472]. Simultaneously, the modified CsNaX zeolites
are more thermally stable than the modified CsNaY.
A different family of basic catalysts consists of amorphous aluminophosphate
oxynitrides (AIPON) [473]. The nitridation by pure ammonia of amorphous AlPO,
gives rise to materials active in the Knoevenagel condensation. The acid-base prop-
612 BARTHOMEUF

erties of AIPON may be tuned by adjusting the O/N ratio but the basicity range
cannot yet be determined [473]. Considering these results, one may wonder whether
the interaction of ammonia in crystalline solids like AlP0,-5 [474] or SAPO-34 [475]
would not also generate basic properties.

464. C. Binet, A. Jadi, J. Lamotte, and J. C. Lavalley, J. Chem. SOC. Faraday


Trans., 92, 123 (1996).
465. M. Huang, S. Kaliaguine, and A. Auroux, Stud. SurF Sci. Catal., 97, 311
(1995).
466. D. Murphy, P. Massiani, R. Franck, and D. Barthomeuf, (a) J . Phys. Chem.,
100, 6731 (1996); (b) Stud. Sur$ Sci. Catal., 11th Int. Zeol. Conf. Seoul,
1996, in press.
467. A. Sauvage, M. Oberson de Souza, M. J. Peltre, P. Massiani, and D. Bartho-
Downloaded by [Yale University Library] at 01:03 15 August 2013

meuf, J. Chem. SOC. Chem. Commun., 1996, in press.


468. A. Sauvage, M. Oberson de Souza, P. Massiani, and D. Barthomeuf, in Proc.
DGMK Con6 Catalysis on Solid Acids and Bases (J. Weitkamp and B. Lucke,
eds.), DGMK, Hamburg, 1996, p. 295.
469. V. B. Kazansky and V. Yu. Borovkov, Stud. Surf: Sci. Catal., 92, 275 (1995).
470. A. Kogelbauer, C. Grundling, and J. A. Lercher, J . Phys. Chem., 100, 1852
(1996).
471. K. R. Kloestra and H. Van Bekkum, J . Chem. SOC.,Chem. Commun., 1995,
p. 1005.
472. M. Lasperas, H. Cambon, D. Brunel, I. Rodriguez, and P. Geneste, Microp.
Muter., 1996, in press.
473. J. J. Benitez, M. A. Centeno, J. A. Odriozola, R. Conanec, R. Marchand, and
Y. Laurent, Catal. Lett., 34, 379 (1995); A. Massinon, J. A. Odriozola, Ph.
Bastians, R. Conanec, R. Marchand, Y. Laurent, and P. Grange, Appl. Catal.
A : General, 137, 9 (1996); A. Massinon, E. Guegen, R. Conanec, R. Mar-
chand, Y. Laurent, and P. Grange, 11th Int. Congx Catal., Baltimore, 1996,
in press.
474. A. Stein, B. Wehrle, and M. Jansen, Zeolites, 13, 291 (1993).
475. R. Vomscheid, M. Briend, M. J. Peltre, D. Barthomeuf, and P. P. Man, J .
Chem. SOC.Faraday Trans., 91, 3281 (1995).

You might also like