Characterization of Acidity in ZSM-5 Zeolites: An X-Ray Photoelectron and I R Spectroscopy Study

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

J . Phys. Chem.

1990, 94, 5989-5994

5989

equilibria influence this value as well. However, these latter effects


remain obscured by the redox ones as long as all redox parameters
are not strictly controlled. For any practical application in the
oxidized (conducting) state, PT and its surrounding electrolyte
must be protected from any external redox effect. In other words,
PT and the bathing solution must be sealed tightly, and then its
metallic state is as stable as that of, e.g., polypyrrole under normal
circumstances (for which the redox capacity of the environment
favors the oxidized state).
A significant and instantaneous light effect (==600-mVchange
upon application of white light) has been observed with PT in the
presence of thiophene monomer in the relaxation solution. Further
investigation of this phenomenon and that of the unusually fast
relaxation found in the presence of Ag+ ions are in progress.

limitation prevails while in more dilute solutions mass-transport


limitation becomes more important, which evidently results in very
different polymers. Even the cell geometry can influence the
quality of the films significantly, as was shown in part I . Relaxation processes may also differ accordingly: if, e.g., the supporting electrolyte were more diluted than the monomer, the
uptake of anions would certainly become much slower than in our
experiments where this process happened to be practically instantaneous. Thus, the physical cause of the slow negative shift
in open cell potential during relaxation in this work could not be
the slow uptake of anions. Our relaxation results can rather be
explained in terms of cations leaving the film, like H+, which is
liberated in the reduction of PT by water. This process is interrelated with the decrease in the redox ratio in the second term
of eq 1.
In order to obtain a film with stable properties it is necessary
to equilibrate it in the desired electrolyte solution until its rest
potential has been reached. The redox environment (water, oxygen, monomer, oligomers, and substrate) seems to be the principal
facter determining the final rest potential, though ion-exchange

Acknowledgment. This work was supported in part by a


contract from the Office of Naval Research.
Registry No. PT, 25233-34-5; TBATFB, 429-42-5; ACN, 75-05-8;
7732-1 8-5; Pt, 7440-06-4; thiophene, 1 10-02-1;
PMT, 84928-92-7; H20,
3-methylthiophene,61 6-44-4.

Characterization of Acidity in ZSM-5 Zeolites: An X-ray Photoelectron and I R


Spectroscopy Study
R. Borade, A. Sayari, A. Adnot, and S. Kaliaguine*
Dgpartement de gCnie chimique and CRAPS, Universite Laval, Ste-Foy, QuCbec. C I K 7P4. Canada
(Received: December 4, 1989)

An X-ray photoelectron spectroscopic (XPS) method is proposed for the identification and quantitation of Br~nstedand
Lewis acid sites in ZSM-5 zeolites. The method consists of deconvoluting the N,, XPS level of chemisorbed pyridine and
measuring the relative intensities of the peak components. It was found that pyridine is chemisorbed in three different states
on ZSM-5 zeolites corresponding to N,, binding energy of 398.7,400.0, and 401.8 eV, respectively. The first peak at 398.7
eV was assigned to N,, level of pyridine adsorbed on Lewis sites, while the second and third were assigned to N1, levels of
pyridine adsorbed on relatively weak and strong Brplnsted acid sites, respectively. Comparison of the concentrations of the
various acid sites as determined from the relative intensities of the N,, components with IR spectroscopic data showed that
XPS has potential applications in the identification and the quantitative determination of Br~nstedand Lewis acid sites in
zeolites.

Introduction
The activity of zeolites in catalyzing a great number of hydrocarbon transformation reactions is attributed to their acidic
character wherein Brensted and/or Lewis sites are involved.'
Even though numerous investigations have dealt with zeolite
acidity and its relationship to catalytic activity,* recent s t u d i e ~ ~ - ~
showed that our current knowledge of the nature and distribution
of acid sites in zeolites as well as their involvement in catalytic
processes is far from being thorough. In order to improve our
understanding of this area of highly practical importance, one
needs to develop quantitative methods for studying the acidity
of zeolites. Ideal techniques or combination of techniques should
provide a comprehensive picture of the catalyst and acid sites in
terms of their nature (Brensted or Lewis sites), numbers and
ultimately their strengths.
ZSM-5 zeolite attracts much attention because of its unique
activity and shape selectivity. The hydrogen form of ZSM-5
zeolite also contains both Brensted and Lewis acid sites in proportions that depend on the activation conditions. Dwyer and
(1) Rabo, J. A. Zeolite Chemistry and Catalysis; ACS Monograph 171;
American Chemical Society: Washington, DC, 1979 Chapter 8.
(2) Jacobs, P. A. Carboniogenic Activity of Zeolites; Elsevier: Amsterdam,
1977.
( 3 ) Fritz, P. 0.;Lunsford, J. H. J. Caral. 1989, 118, 85.
(4) Hall, W. K.; Engelhardt, J.; Sill, G. A. Srud. Surf. Sci. Caral. 1989,
42, 1253.
( 5 ) Lombardo, E. A.; Sill, G. A.; Hall, W. K. J . Caral. 1989, 119, 426.

0022-3654/90/2094-5989$02.50/0

OMalley6 reviewed various methods that have been employed to


determine the acidity of zeolites, the most important of which are
microcalorimetry,' temperature-programmed desorption of amm ~ n i a ? and
- ~ infrared spectroscopy.lwl2
IR spectroscopy seems to be one of the most widely used
techniques for determining the Brensted and Lewis sites concentrations in zeolites. The only drawback of this technique is
that it only leads to relative concentrations of Bransted and Lewis
acid sites. In order to determine the absolute concentration of
each type of acidic centers, assumptions regarding the mechanism
by which Lewis acid sites are generated during dehydroxylation
have to be made. Such a mechanism would provide an additional
independent relationship between the numbers of Brmsted (B)
and Lewis (L) acid sites per unit cell.
A dehydroxylation mechanism involving the disappearance of
two Br~nstedsites for each Lewis site generated has been adopted
by several authors. On the basis of this mechanism, Datka and
Tuznik'O used IR spectroscopy to obtain quantitative information
(6) Dwyer, J.; O'Malley, P.J. Srud. Surf. Sci. Catal. 1988, 35, 5 .
(7) Vedrine, J. C.; Auroux, A.; Dejaifve, P.; Ducarme, V.; Hoser, H.; Zhou,
S . J . Catal. 1982, 73, 147.
(8) Tops%, N. Y.;Pedersen, K.; Derouane, E. G.J . Caral. 1981, 70,41.
(9) Anderson, J. R.; Foger, K.; Mole, T.; Rajadhyaksha, R. A.; Sanders,
J . V. J . Caral. 1979, 58, 114.
(IO) Datka, J.; Tuznik, E. J. Caral. 1986, 102, 43.
( 1 I ) Jentys, A.; Wareca, G.; Lercher, J. A. J . Mol. Caral. 1989,51,309.
(12) Rhee, K. H.; Rao, U. S.;Stencel, J. M.; Melson, G. A.; Crawford,
J. E. Zeolites 1983, 3, 337.

0 1990 American Chemical Society

5990

The Journal of Physical Chemistry, Vol. 94, No. 15, 1990

about the concentration of Br~nstedand Lewis sites in ZSM-5


zeolites. Similarly, earlier work from this laboratory has shown
the usefulness of IR techniques to determine the Bransted and
Lewis sites concentrations in ZSM-5 zeolite modified by phosphorus and transition and noble metal c o m p l e x e ~ . ' ~ -However,
'~
the aforementioned mechanism has been challenged recently.
KazanskyI6 showed that the dehydroxylation of zeolites can
proceed by two mechanisms in which two Brsnsted acid sites are
destroyed generating either one or two Lewis acid sites depending
on the Si/AI ratio and the method of pretreatment of the sample.
It is obvious that the relationships between B and L stemming
from these two mechanisms will be different from each other.
Therefore, the values of B and L calculated from IR spectra will
depend on the actual dehydroxylation mechanism.
This work was undertaken with the aim of exploring the possible
applicability of XPS technique in combination with IR spectroscopy to determine the concentrations of Bronsted and Lewis
sites present in zeolites.
Experimental Section
Catalyst Preparation. The ZSM-5 zeolites were synthesized
according to the procedure described by Gabelica et a1.I' To
convert them into protonic form, the as-synthesized zeolites were
first calcined in air for 10 h at 500 OC to remove the organic
materials from the zeolite pores. The calcined form was next
converted to ammonium form by repeated ion exchange with 1
M ammonium nitrate solution. The protonic form was obtained
by air calcination of the ammonium form at 500 OC for 10 h.
The crystallinity of these samples was verified by X-ray diffraction using a Philips spectrometer fitted with a Ni-filtered Cu
Kcu radiation source. Elemental analyses were performed by
atomic absorption using a Perkin-Elmer (Model 1100B) spectrometer.
IR Measurements. Infrared measurements of adsorbed pyridine
were performed on all samples. Thin self-supporting wafers of
about 9 mg were prepared by placing powdered zeolite in a
stainless steel die having a diameter of 13 mm and compressing
under a pressure of 160 MPa for 2 min. The wafers were mounted
on moveable stainless steel sample holders and introduced into
a Pyrex cell fitted with calcium fluoride windows and designed
to accommodate four different samples at a time. Finally, the
cell was connected to the vacuum system and the wafers were
outgassed at 4 0 0 OC for about 16 h.
A first IR spectrum was recorded after the samples were cooled
down to room temperature. Then pyridine vapor was introduced
into the cell and allowed to react with the wafers for about 15
min. Afterward, the cell was evacuated at 200 OC for 16 h to
eliminate physisorbed pyridine. The cell was then cooled down
to room temperature and a second spectrum recorded. All the
spectra were recorded from 4000 to 1000 cm-' with a 2 cm-'
resolution, using a Digilab FTS-60 spectrometer. When the
experiment was over, the wafers were taken out of the cell and
analyzed by XPS.
XPS Measurements. The XPS spectra of all the zeolites were
recorded with a V.G. Scientific Escalab Mark I1 system with an
hemispherical analyzer operated in the constant pass energy mode
(20 eV). A Mg K a X-ray source (hv = 1253.6 eV) was used.
The spectrometer was operated at 20 mA and 15 kV. The residual
gas pressure in the spectrometer chamber during data acquisition
was less than IO-' Torr.
Zeolite wafers previously used for pyridine adsorption followed
by IR measurements were mounted in stainless steel cups and
introduced into the spectrometer chamber for analysis. The
measurements were performed in the following sequence: Si(2p),
( I 3) Rahman, A.: Lemay, G.: Adnot, A.; Kaliaguine, S. Appl. Catal. 1989,
50, 131.
(14) Rahman. A.; Lemay, G.; Adnot, A.; Kaliaguine, S. J . Caral. 1988,
112, 453.
(15) Mahay, A.; Lemay, G.; Adnot, A,; Szoghy, I. M.; Kaliaguine, S . J .
Coral. 1987, 103, 480.
(16) Kazansky, V. 9. Catal. Today 1988, 3, 367.
(17) Gabelica, 2.; Derouane, E. G.; Blom, N . Appl. Catal. 1983, 5, 227.

Borade et at.
TABLE I: Bulk and Surface Properties of Various ZSM-5Samples
sample
(Si/Al)n (N/AI). (N/Al), SO, m2/g
A
B

D
E
a

19.9
24.4
39.0
45.9
65.0

14.3
21.4
37.2
47.4
36.6

0.37
0.62
0.92
1.07

0.83

0.52
0.53
0.85
0.9 1
0.92

362
392
432
422
445

Nitrogen BET surface area.

N(ls), O(ls), A1(2p), and Si(2p). A binding energy of 103.3 eV


for the Si, level was used as an internal reference for all samples.
The accuracy of the binding energy as determined with respect
to this standard value was within f 0 . 2 eV.
The intensity of the various XPS bands was determined by using
linear background subtraction and integration of the peak areas.
N( 1s) peaks were deconvoluted into two or three components by
keeping full width at half-maximum (fwhm) constant for all
components in a particular spectrum and assuming that the peak
has Gaussian-Lorenzian shape.
Gravimetric Adsorption Measurements. Gravimetric pyridine
adsorption measurements were carried out using a microbalance
MTB 8 (Setaram, France) connected to a conventional highvacuum system. About 100 mg of the sample was placed in a
Pyrex bucket and degassed at 400 OC for 16 h. Then, the sample
was cooled down to 50 "C and pyridine vapor was admitted to
the system for about 15 min. Finally, physisorbed pyridine was
desorbed at 200 OC under vacuum till no change in weight of the
sample was observed.

Results and Discussion


X-ray Photoelectron Spectroscopy. Ratios of the atomic
concentrations in the outer surface layers of the samples were
estimated from the corresponding XPS peak area ratios by using
the relation

where M stands for silicon or nitrogen and A , u, A, and EK are


the normalized XPS peak area, the cross section of the photoelectron emission, the escape depth, and the photoelectron kinetic
energy, respectively. Scofield's cross-section values1*were used
in the calculations whereas escape depths were determined from
an empirical relation given by Vulli and Stark.19
Table I shows the surface
and (N/AI), atomic ratios
as well as their bulk counterparts. The
and (N/Al)b
ratios were determined by chemical analysis and thermogravimetry
of pyridine chemisorption, respectively. As far as the sample's
purity is concerned, some comments may be made. Sample A
showed different surface and bulk compositions, indicating that
aluminum is not homogeneously distributed throughout the zeolite
grains. Moreover, this sample had a relatively low BET surface
area and a very low chemisorbed pyridine uptake [(N/AI)J,
suggesting that it contains a significant amount of extraframework
amorphous nonacidic material. This conclusion was further evidenced by acid leaching of sample A in 0.1 N HCI that led to
sample B with higher bulk and surface Si/A1 ratios. However,
as sample B also had a low pyridine uptake, it is inferred that,
under our conditions, extraframework aluminum species are only
partly extracted by acid leaching and that the remaining amount
is homogeneously distributed throughout the zeolite particles. As
for samples C and D, they exhibited similar bulk and surface Si/AI
ratios, in addition to relatively high N/A1 ratios and high BET
surface areas. This indicates that both samples are clean (no
extraframework aluminum species) and homogeneous (no surface
segregation or depletion of aluminum). On similar grounds, it
may be inferred that sample E is likely to be free of extraframework material, however; in contrast to samples C and D, it
exhibits significant surface aluminum enrichment.
(1 8) Scofield, J. H.J . Electron Spectrosc. Relat. Phenom. 1976, 8. 129.
(19) Vulli. M.: Stark, K.J . Phys. E 1978, 10, 158.

The Journal of Physical Chemistry, Vol. 94, No. 15, 1990 5991

Acidity in ZSM-5 Zeolites


TABLE 11: XPS Data after Pyridine Adsorption
WP)
(31s)
samole BE fwhm BE fwhm BE fwhm
A
103.3 2.4 532.3 2.5 74.3 2.8
B
103.3 2.4 532.6 2.3 74.4 2.4
C
103.3 2.4 533.1 2.4 74.8 2.5
D
103.3 2.4 532.9 2.4 74.5 2.4
E
103.3 2.4 532.6 2.4 74.1 2.4
r

N(ls)
BE fwhm
401.1 4.2
399.7 3.4
400.2 4.6
399.5 2.6
399.3 4.2

As for the origin of the nonacidic extraframework material,


it is assumed that during the ZSM-5 synthesis aluminum hydroxide/oxide species were formed and, upon calcination, these
species might have undergone transformation into an aluminum
oxide phase which does not exhibit acid properties. It may be
mentioned here that in a detailed XPS investigation of aluminum
species present on the ZSM-5 zeolite surfaces, Barr and Lishkam
have also observed that the extralattice aluminum constitute more
than 50% of the total aluminum present on the surface. However,
these extralattice aluminum species were still less than 10% of
the bulk value.
The binding energy and fwhm values of Si(2p), O(ls), A1(2p),
and N( Is) XPS lines are given in Table 11. A binding energy
of 103.3 eV for Si(2p) level was used as an internal reference.
Both the BE and fwhm values for Si(2p), O(ls), and AI(2p) levels
in our samples agree well with those of Vedrine et aL2' and
Ratnasamy et a1.22for H-ZSM-5. It is important to note that
all N( Is) peaks were found to be substantially (40-8Wo) broader
and less symmetric than their corresponding Si(2p), A1(2p), and
O(1 s) peaks. This behavior is a strong indication that, in all cases,
the N( Is) peak is actually a composite peak that has to be deconvoluted. As a result, the N,, BE values shown in Table I1 may
not have a real physical meaning for they were simply the values
corresponding to the maxima of N Is envelopes.
Defosse and CanessonZ3have studied by XPS the adsorption
of pyridine on NHIY zeolite activated at various temperatures.
They also found N,, peaks with fwhm values close to 5 eV which
were much higher than those of Si(2p), A1(2p), and O(1s) peaks.
They attributed this broadening of N(ls) peak to a distribution
of acidity over a wide range of acid strengths. Therefore, we
concur that the broadening of N( 1s) peaks reported here is due
to the overlapping of several N( Is) peaks from pyridine interacting
with various acid sites.
In all cases, the N( Is) envelopes could be deconvoluted into
three components with the same fwhm (2.4 eV) and reproducible
binding energies a t 398.7 rt 0.3, 400.0 f 0.3, and 401.8 f 0.3
eV (Table 111). Figure 1 gives the curve fitted N(ls) level spectra
for the various ZSM-5 zeolites presented in Table I. The only
significant difference in these spectra is the relative intensity of
the three components (Table 111). Since the binding energy values
of these components are reproducible it is contended that the
ZSM-5 sample on which pyridine is adsorbed at room temperature
and desorbed at 200 OC always exhibits three N( Is) components.
The most important question about these results concerns peak
assignment. As mentioned above, Defosse and CanessonZ3also
found a shift of 2 eV toward lower BE when the activation temperature exceeded 500 O C . They indicated that this change in
BE was related, as shown by IR spectroscopy, to a decrease in
the concentration of Bransted sites and increase in the concentration of Lewis sites by dehydroxylation at high temperature.
Based on these observations they attributed the peak a t lower
binding energy to nitrogen associated with Lewis sites and the
peak at higher binding energy to nitrogen in pyridinium ions. On
similar grounds, the first N( Is) peak which occurs at 398.7 0.3
eV for our samples is assigned to nitrogen atoms of pyridine

(20) Barr, T. L.; Lishka, M. A. J . Am. Chem. SOC.1986, 108, 3178.


(21) Vedrine, J. C.; Auroux, A.; Bolis, V.; Dejaifve, P.; Naccache, C.;
Wierzchowski, P.; Derouane. E. G.; Nagy, J. B.; Gilson, J.; Van Hoff, J. H.
C.; Vander Berg, J. P.; Wolthuizen, J. J . Cazal. 1979, 59, 248.
(22) Badrinarayana, S.; Hegde, R. I.; Balakrishnan, 1.; Kulkarni, S. B.;
Ratnasamy, P. J . Catal. 1981, 71. 439.
(23) Defosse, C.; Canesson, P. 1. Chem. Soc., Faraday Tr4ns.1 1976.11,
2565.

I\

Binding energy (eV)


Figure 1. N(1s) XPS profiles of adsorbed pyridine on various ZSM-5
samples recorded at room temperature. Samples designation as in Table
I.

associated with Lewis sites, and the second and third peaks
corresponding to BE values 400.0 f 0.3 eV and 401.8 f 0.3 eV
are assigned to nitrogen atoms of pyridine associated with relatively
weak and strong Bransted acid sites, respectively.
The fact that ZSM-5zeolite contains two types of Bransted
acidic sites is confirmed by carrying out additional experiments.
In these experiments all conditions were kept the same as stated
above except that pyridine desorption was carried out in the
temperature range of 250-400 O C instead of 200 OC. Under these
conditions one should expect faster desorption of pyridine from
Lewis sites than from Br~nstedsites, and as a consequence the
decrease in the relative intensity of the corresponding N( 1s) peak
in the XPS spectra. Figure 2 shows that indeed the desorption
of pyridine from Lewis site does occur upon evacuation between
250 and 400 O C as the relative intensity of the low BE component
decreases substantially. The simultaneous decrease in the intensity
of the 400.0-eV band relative to the 401.8-eV band indicates that,
parallel to the desorption of pyridine from Lewis sites, a partial
desorption from weak Bransted sites also takes place. The presence
of N(ls) XPS component even at the pyridine desorption temperature of 400 OC suggests that pyridine molecules associated
with Lewis sites do not desorb completely. This observation is
in agreement with the result of Kucherov et al.,' who studied the
(24) Kucherw, A. V.; Slinkin, A A.: Khamn, M. S.;Bondarenko,T. N.;
Minachev. Kh. M. Kinez. Kurd 1989, 30, 165.

5992

The Journal of Physical Chemistry, Vol. 94, No. 15, 1990

Borade et al.

TABLE 111: Deconvolution of XFS N(1s) Peaks


Deak comwnents

(1)

(2)

samDle

BE

fwhm

BE

398.6
398.7
398.7
398.8
398.5

2.3
2.3
2.4
2.3
2.3

400.3
400.0
400.3
399.8
399.8

B
C
D
E

fwhm
2.4
2.3
2.4
2.4
2.4

BE

50
I40
200
300
400

(Si/AI)s
36.7
33.3
37.2
38.4
39.6

(N/AI),
1.10
1.19
0.93
0.56
0.55

N( Is) components
(1)

19.5
21.0
24.4
9.5
7.7

(2)
46.5
48.0
47.6
26.0
29.5

(2)
46.4
52.6
47.6
65.4
48.5

(1)

14.7
26.9
24.4
18.3
26.7

(3)
38.9
20.5
28.0
16.3
24.8

'-i

re1 intensity of
XPS

fwhm
2.4
2.4
2.5
2.4
2.4

402.0
402.0
402.1
401.6
401.8

TABLE IV: Effect of Pyridine Desorption Temperature on the


Relative Intensities of N(ls) Components for Sample C

pyridine desorpn
..
temp, 'C

relative intensity

(3)

/D

(3)
34.0
31.0
28.0
64.5
62.8

Temperature ("C)

Figure 3. Variation of (a) (TB/L) and (b) (SB/WB) ratios for sample
C as a function of pyridine desorption temperature.
TABLE V Comparison of Bulk and Surface Properties of Various
Zeolites

B/L
Si/AI
N/AI
____
surface
sample bulk surface bulk' surface bulk (IR)
(XPS)
A

B
C
D
E

19.9
24.4
39.0
45.9
65.0

14.3
21.4
37.2
47.4
36.6

0.52
0.53
0.85
0.91
0.92

0.37
0.62
0.92
1.07
0.83

4.5
2.7
3.7
4.7
4.7

5.8
2.7
3.1
4.6
2.7

'(N/Al)bulkis determined by the gravimetric adsorption method.

390

395
400
405
Binding energy (eV)

410

Figure 2. N( Is) XPS profiles of pyridine adsorbed on sample C recorded


at room temperature. Figures indicate pyridine desorption temperature.

adsorption/desorption of pyridine on ZSM-5 zeolites at various


temperatures by IR and found that, even at pyridine desorption
temperature of 365 O C , there still exist weak pyridine bands
associated with Lewis sites.
Now, assuming that at the pyridine desorption temperature of
400 O C the contribution of N( 1s) component associated with Lewis
sites to the main N ( Is) peak is negligible, then the fwhm of the
N( Is) peak should be comparable to those of Si(2p), A1(2p), and
O( 1s) peaks if all Bronsted sites were of the same strength. Yet,
the N ( Is) peak was found substantialy broader than the abovementioned peaks (3.4 eV as compared to 2.4 eV). Also, from
Figure 3 and the data presented in Table IV, it is seen that the
desorption of pyridine from sample C in the temperature range
of 50-200 OC has almost no effect either on (N/AI), or on the
total Bronsted to Lewis (TB/L) or the strong to weak Bronsted
(SB/WB) sites ratios. Above this temperature, the (N/AI), ratio
decreases and the (TB/L) and (SB/WB) ratios increase. As

expected, the ratio (TB/L) increases much faster than the


(SB/WB) ratio. The change in (SB/WB) ratio at temperature
higher than 200 O C is yet another indication of the existence of
two different types of Bronsted acid sites. At the pyridine desorption temperature of 400 O C , the presence of two main N( 1s)
components with fwhm of 2.4 eV and the increase in the relative
intensity ratio of strong to weak Brsnsted sites clearly demonstrate
the existence of two types of Brsnsted acid sites.
The presence of two types of Bransted sites in the ZSM-5 zeolite
has been reported by Datka and Tuznik.'O In their work, the
assignment of strong and weak Bronsted acid sites was based on
the adsorption of excess pyridine on the zeolite followed by desorption at various temperatures while monitoring the changes
in the intensities of the 3610-cm-I band and pyridine associated
with Brsnsted sites at the 1545-cm-I band and Lewis sites at the
1455-cm-' band. The amount of pyridinium ions decomposed after
the disappearance of the 1455-cm-' band and before the reappearance of the 3610-cm-' OH band was taken as the amount
of weak Bronsted acid sites. By this method, the concentration
of strong Bronsted acid sites relative to all Bronsted sites was found
to be 77% for a H-ZSM-5 sample with Si/AI = 47. In contrast,
we found by XPS that, depending on the zeolite, the strong
Brsnsted acid sites account for only 20-37% of all Bronsted sites.
Our results seem to be in line with recent NH, poisoning studies4
that indicate that strong Bronsted acid sites in ZSM-5 zeolites
are likely to be in very limited numbers.
Assuming that the stoichiometry of pyridine chemisorption is
one pyridine molecule adsorbed per acid site regardless of its
nature, the relative concentrations of Bronsted and Lewis acid
sites (B/L) in the outer layers of our zeolites can be estimated
from the relative intensities of N( 1s) XPS bands corresponding

The Journal of Physical Chemistry, Vol. 94, No. 15, 1990 5993

Acidity in ZSM-5 Zeolites


TABLE VI: Infrared h t a for Samples A-E

frequency of OH groupsa
sample
A

C
D
E

1
19.9
24.4
39.0
45.9
65.0

3742
3741
3741
3742

3728

NV

NV

3680

3725

NV
NV
NV

NV
NV

pyridine region

UR

361 1
361 1
3612
361 1
3612

3515
3515
3512
351 1
3512

1547
1547
1541
1547
1547

VR

1455
1456
1456
1455
1455

VI

1491
1491
1491
1491
1491

ARIAI
3.00
1.80
2.46
3.15
3.15

Before pyridine adsorption; NV, not visible. bRatio of absorbance of pyridine chemisorbed on Bransted and Lewis acid sites.

n
l

0)
0

e
P

l! 5

1545

1515

1485

1455

Wovenumber
4000

3850

3700
3550
3400
Wavenumber Ccm? 1
in the region 3400-4000 cm-I: (1)

Figure 4. Infrared spectra


pyridine adsorption; (2) after pyridine adsorption.

1 15

(cm-)

Figure 5. Infrared spectra in the region 1425-1575 cm-I after pyridine


adsorption.

before

to each type of acidity. These values are gathered in Table V.


Infrared Characterization. The hydroxyl region of the infrared
spectra for all the samples before and after pyridine adsorption
is shown in Figure 4. The IR spectra contain two main bands
at around 3612 and 3740 cm-I. The 3612-cm-l band is due to
the vibration of bridged OH groups linking aluminum and silicon
framework atoms and the 3740-cm- band is associated with the
terminal O H groups bound to silicon framework atom^.^^^^^ In
a non-dehydroxylated zeolite, the intensity of the former band
is proportional to the framework aluminum content but this is
not the case for the 3740-cm- band. The broad band at 3500
cm-I assigned to hydrogen-bonded Si-OH groups at dealumination
or lattice imperfection site^^',^^ is also detectable in some of our
samples (Figure 4).
In addition, two more O H stretching bands at 3720 and 3680
cm-I have been reported.25*29-30The band at 3720 cm-l was
attributed to the extralattice amorphous materialZ5and the band
at 3680 cm- was assigned by Ison and GorteZ9to water adsorbed
in molecular form, whereas Kazansky et aL3O assigned it to the
extralattice A1-OH species formed by dehydroxylation. It is
difficult to detect the presence of the 3720- and 3680-cm-l bands
in the spectra of our ZSM-5 samples. Therefore, it is hard to draw
(25) Jacobs, P. A,; von Ballmoos, R. J . Phys. Chem. 1982, 86, 3050.
(26) Chu, C. T. W.;Chang, C. D. J . Phys, Chem. 1985.89, 1569.
(27) Woolery, G. L.; Almany, L. B.; Dessau. R. M.;Chester, A. W.
Zeolites 1986, 6 , 14.
(28) Moser, W.R.; Chiang, C. C.; Thompson, R. W . J . Catal. 1989,115,
532.
(29) Ison, A.; Gorte, R. J. J . Coral. 1984, 89, 150.
(30) Zholobenko, V . L.; Kustov, L. M.; Borovkov, V. Y.; Kazansky, V. B.
Zeolites 1988, 8, 175.

from IR results any concrete conclusion about samples purity.


The IR bands at 1545 and 1455 cm- assigned to pyridinium
ions and coordinately bound pyridine, respectively, were used to
determine the relative bulk concentrations of Br~nstedand Lewis
acid sites (Figure 5). This ratio was determined by using the
following equation

where AB/AL stands for IR absorbance ratio and tJtg is the


extinction coefficient ratio. Rhee et al.lz found that tL/tg can
be taken as 1.5 for zeolites having Si02/A1203ratio higher than
15. The IR frequencies in the hydroxyl as well as pyridine regions
and the calculated AB/AL ratios are presented in Table VI.
In order to check the validity of XPS quantitative determination
of acid sites, we have made an attempt to compare the estimates
of B/L ratios obtained from the component peak area ratios in
N( 1s) XPS spectra with IR data (Table V). It is interesting to
note that the B/L values obtained by XPS method are in close
agreement with IR data for samples B, C, and D. Since, as pointed
out earlier, sample A contains a significant amount of amorphous
material and exhibits a nonhomogeneous AI distribution, it is not
surprising that a large difference in B/L ratios determined by these
two techniques was found. The lack of agreement for sample E
is also likely to be attributable to its inhomogeneous A1 distribution.
It is interesting, however, to note the good agreement obtained
by both techniques for sample B, even though this sample was
not free of extraframework nonacidic material. Therefore, it may
be concluded that XPS can be used as a reliable technique for
the identification of acid sites and the determination of B/L ratios
in zeolites provided that these zeolites have homogeneous spatial
A1 distributions.

5994 The Journal of Physical Chemistry, Vol. 94, No. 15, I990
TABLE VII: Effect of Dehydroxylation on B/L and N/AI Ratios for
Sample D

calcin

"AI),

temp, OC (Si/Al)b
45.9
500
675

(Si/Al),

(B/L)b

(B/L),

47.4
40.1
43.4
41.9

4.71
1.17
0.46
0.49

4.60
0.65
0.46
0.44

45.9
45.9
45.9

800
950

XPS calc"
1.07
1.04
1.18
0.87

0.85
0.62
0.59
0.59

"Calculated from eq 4 using the experimental values of (B/L),.


As pointed out earlier, the determination of absolute densities
of Brernsted and Lewis acid sites requires an additional relationship
between the numbers of these two types of sites. Such a relationship can be established from the dehydroxylation mechanism.
KazanskyI6 reported that, in the case of zeolites with high silica
content, dehydroxylation occurs via a mechanism proposed earlier
by Uytterhoeven et aL3' According to this mechanism, two
Bronsted sites are destroyed to generate two Lewis sites consisting
of trigonal atoms of AI and Si:
H
I

H
I

\st0\i0f
\si

\-/ 0\sf 0\AI


+si
/A1\
/ \ / \ / \

\-I 0
/A?

/ \ / \ /

If such a mechanism is operative, the number of A1 atoms in the


zeolite framework would be related to B and L as follows:
n ~ =l B L
(2)

However, according to KuhI?* dehydroxylation of the hydrogen


form of zeolites is followed by a framework dealumination involving the expulsion of the trigonal aluminum as an extralattice
AIO+ species and simultaneous saturation of the coordinatively
unsaturated silicon:
0\si/0\A,
\-/

'Si
/ \ / \ / \

\-/o\si/o\s(
/A?

/ \ /

A,O+

The final outcome of this mechanism corresponds to the destruction of two Bronsted sites for each Lewis site generated. In
this case, the relationship between the number of the original
framework aluminum and the numbers B and L becomes
nAl = B 2L
(3)

Let us assume that the hydrogen form of the zeolite does not
contain any extralattice nonacidic aluminum species. Then if the
first dehydroxylation mechanism is the only one to apply, the N/AI
ratio should be equal to 1 regardless of the extent of dehydroxylation. This, of course, assumes also that the stoichiometry of
chemisorption is of one pyridine molecule adsorbed per Brernsted
and one per Lewis acid site. If, however, the dehydroxylation
proceeds by the second mechanism with complete conversion, the
N/AI ratio will depend on the extent of dehydroxylation as follows:
(31) Uytterhoeven, L. B.; Cristner, L. G.; Hall, W. K. J . Phys. Chem.
1965. 69, 21 17.
(32) Kiihl, G. H. J . Phys. Chem. Solids 1977, 38, 1259.

Borade et al.
1 + B/L
N/Al= 2 + B/L

(4)

Equation 4 will apply to both ( N / M h and ( N / d ) , if no migration


of extraframework species occurs during dehydroxylation.
Assuming that (N/AI), atomic ratios are determined with a
maximum uncertainty of 15%,33XPS technique cannot discriminate between eqs 2 and 3 as long as B/L is higher than ca. 4.67,
a value for which the calculated (N/AI), ratio is equal to 0.85.
However, for samples with (B/L), ratios lower than 4.67, surface
(N/AI), ratios systematically lower than unity should be expected.
Table VI1 summarizes the results obtained for sample D which
was found to satisfy the above-mentioned conditions of purity and
homogeneity fairly well. It is seen that, within experimental errors,
the (si/Al)b and @/AI), remain practically constant and very
close to each other after calcination at temperatures as high as
950 O C . As the extent of dehydroxylation increases, the (B/L),
ratio decreases to reach a value of 0.44 for a calcination temperature of 950 OC. It is also seen that the (B/L)b and (B/L),
ratios as determined by IR and XPS, respectively, remain fairly
close to each other except at the calcination temperature of 675
"C, suggesting that in early stages dehydroxylation proceeds more
readily in the outer layers of the zeolite than in the bulk. The
(B/L), ratios were used to calculate the corresponding (N/AI),
according to eq 4. As expected, these were found to be significantly lower than unity for calcination temperatures higher than
500 O C . The experimental XPS value, however, remain very close
to 1 which seems to validate eq 2 over eq 3.

Conclusions
The XPS results presented above demonstrate that on an activated H-ZSM-5 zeolite pyridine is chemisorbed in three different
states: namely, on Lewis sites, weak Brernsted sites, and strong
Br~nstedsites corresponding to BEs of 398.8 f 0.3,400.0 f 0.3,
and 401.8 f 0.3 eV, respectively. Upon pyridine adsorption, the
IR bands at 3680 and 3500 cm-' remain unchanged, indicating
that weak Br~nstedacid sites identified in the present work are
not due to extralattice AlOH groups or SiOH groups of hydroxyl
nests that may be generated during the calcination steps. The
relative concentration of Bronsted and Lewis acid sites as determined by XPS were found to be in close agreement with the
data obtained from IR measurements for samples that exhibit a
homogeneous AI distribution, even if a fraction of the AI is
comprised of extraframework material. Moreover, the XPS
technique has the ability of providing the relative concentrations
of strong and weak Brernsted sites. Usually such measurements
cannot be made by IR spectroscopy.
XPS measurements, of pyridine chemisorbed on a clean sample
previously calcined at different temperatures seem to indicate that
the dehydroxylation of ZSM-5 zeolites occurs via a mechanism
that generates one Lewis site for each Bronsted site destroyed.
Using this conclusion, it is possible to calculate from N( 1s) bands
of chemisorbed pyridine the absolute densities of Lewis as well
as weak and strong Bronsted sites in ZSM-5 zeolite samples having
identical surface and bulk compositions.
(33) Penn, D.R. J . Electron. Spect. Relat. Phenom. 1916, 9, 29.

You might also like