Sei sulla pagina 1di 12

https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev.

, 2018, 62, (3), 279–290

www.technology.matthey.com

Accessibility and Location of Acid Sites in


Zeolites as Probed by Fourier Transform
Infrared Spectroscopy and Magic Angle
Spinning Nuclear Magnetic Resonance
Understanding heterogeneous catalysts using probe molecule experiments

Cátia Freitas also be quantified using 2,6-di-tert-butyl-pyridine


Birchall Centre, Keele University, Staffordshire, and 2,4,6-trimethylpyridine. It is concluded that
ST5 5BG, UK using a combination of probe molecules, including
co-adsorption experiments, affords differentiation
Nathan S. Barrow between acid sites located in channels and cavities
Johnson Matthey, Blounts Court, Sonning of different sizes and on the external and internal
Common, Reading, RG4 9NH, UK surfaces of various zeolitic structures.

Vladimir Zholobenko* 1. Introduction


Birchall Centre, Keele University, Staffordshire,
ST5 5BG, UK Zeolites are crystalline solids with a well-defined
structure consisting of molecular scale pores and
*Email: v.l.zholobenko@keele.ac.uk channels. Their primary units such as SiO4 and
AlO4 tetrahedra are linked by oxygen atoms at
their vertices, creating a variety of microporous
The understanding of location and accessibility of framework structures (1). The AlO4 units impart
zeolite acid sites is a key issue in heterogeneous a negative charge within the framework that
catalysis. This paper provides a brief overview of must be balanced by cationic species (2). These
Fourier transform infrared (FTIR) spectroscopy cationic species are retained by steric effects and
and nuclear magnetic resonance (NMR) electrostatic interactions and can be exchanged
characterisation of acidity in zeolites based on with other cations, making zeolites highly valuable
the application of test molecules with a diverse as cation-exchangers (2–4).
range of basicity and kinetic diameters. Many Zeolites have been widely used within the chemical
zeolites, including ZSM-5 and BEA, have been and petrochemical industries as heterogeneous
characterised by monitoring the interaction catalysts. This is due to their unique set of
between the zeolite acid sites and test molecules, characteristics such as high adsorption capacity,
such as 1,3,5-triisopropylbenzene, pyridine intrinsic acidity, hydrothermal stability and shape
and alkylpyridines, to probe the location, selectivity.
accessibility and strength of the Brønsted acid Detailed understanding of their acidic properties
sites. 1,3,5-triisopropylbenzene can be used to is important for the design, modification and
distinguish Brønsted acid sites located on the practical application of zeolite based catalysts.
external and internal surface in most medium and Generally, the most important properties for
large pore channel zeolites. Brønsted acid sites on catalytic reactions are the type, strength,
the external surface of medium pore zeolites can distribution, concentration and accessibility of

279 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

acid sites. Whilst the microporous nature of zeolites (20, 21). Other nitriles such as propionitrile,
imparts some of their essential properties, such as isobutyronitrile (22), 2,2-diphenylpropionitrile,
high surface area, adsorption capacity and shape- benzonitrile and ortho-tolunitrile (23) were also
selectivity, the presence of micropores can also lead utilised in order to assess the acid site accessibility
to diffusional limitations, shorter catalyst lifetime in a number of zeolitic structures. When compared
and poor activity (5, 6). with other probe molecules such as pyridines
To achieve the full potential of a zeolite catalyst, and amines, nitriles interact with acid sites less
it is important to maximise the accessibility of strongly, creating a relatively weak bond with the
active sites and transport efficiency for both the BAS and LAS.
feed molecules and products in catalytic reactions. A variety of hydrocarbons have also been used
This can be achieved by employing zeolites with (24); their interaction with acid sites is relatively
different structures and pore systems, and by weak resulting in the formation of a hydrogen
introducing mesoporosity in addition to the bond between the hydrocarbon and a zeolite OH
existing network of micropores. These materials, group. Alkenes and aromatics achieve stronger
the so-called hierarchical zeolites, developed, interactions than saturated hydrocarbons, which
synthesised and modified in numerous ways have are still weak in comparison with other classes of
received considerable attention (7–10). The aim probe molecules. The location and strength of BAS
of this article is to review recent work evaluating in ZSM-5 zeolites were evaluated by comparing
acid site location and accessibility in zeolites with the data obtained for cyclohexane and benzene
different pore systems, focusing on understanding with those for 1,3,5-trimethylbenzene. It was
the interactions between acid sites and probe concluded that the mesoporosity influences only
molecules, particularly using FTIR and solid-state the accessibility of the acid sites by shortening
NMR, and providing examples of characterisation the diffusion pathways, while the strength of the
data for BEA and MFI zeolites using a variety of interaction with the probes, reflecting the strength
probe molecules. of the acid sites, remains unaffected.
Another approach to understanding the site
accessibility in zeolites is co-adsorption of probe
2. Location and Accessibility of Acid
molecules with different sizes. Adsorption of a small
Sites
probe molecule (for example carbon monoxide
The accessibility and location of acid sites in many (CO)) after pre-adsorption of a larger probe
ways determine the catalytic performance of molecule was the subject of several studies in
zeolites. Acid sites hosted on the external surface of mordenite (25, 26). This zeolite exhibits two types
a zeolite are commonly accessible; the accessibility of channels: 12-membered ring main channels
within the microporous system is dependent upon and 8-membered ring channels connected by
the dimensions of the pore space relative to the 8-membered ring side pockets (27). Pyridine can
guest molecule. This relationship is closely linked interact with the acid sites in the main channels
to the local geometry of the acid sites, position of of MOR but not with those inside the smaller side
the Al atoms and chemical environment (11, 12). pockets, whereas CO interacts with all acid sites
in the pore system. Consequently, by co-adsorbing
2.1 Infrared Spectroscopy these two probes, CO provides information about
Infrared (IR) spectroscopy studies using adsorption the strength of Brønsted acid sites in different
of probe molecules is one of the most important locations, and the steric hindrance of pyridine gives
tools for comprehensive characterisation, including evidence for the location of the acid sites (26).
the nature, strength and accessibility of acid Co-adsorption of CO and nonane was also used
sites in zeolite based catalysts. The nature and to examine the spatial distribution of platinum in
the strength of the Brønsted acid sites (BAS) the micropores and mesopores of bi-functional
(i.e. bridging OH groups) and Lewis acid sites PtH-MFI catalysts (28). This technique involves
(LAS) in zeolites has been addressed in detail nonane pre-adsorption between two successive CO
(11, 13–18); this section is focused on evaluating chemisorption experiments.
the location and accessibility of acid sites. Nesterenko et al. (29, 30) and Bleken et al. (31)
Busca and co-authors (19) carried out a range of presented a methodology based on co-adsorption
experiments using pivalonitrile as a probe molecule of alkylpyridines and CO for the analysis of acid
to distinguish acid sites on internal and external site distribution in dealuminated mordenites and
surfaces in MCM-41, FER and MFI type materials MFI zeolites. The use of these probe molecules with

280 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

increasing steric hindrance allows discrimination Recently, this approach was also applied to
between acid sites located on internal and other probe molecules, for example 2,6-di-tert­
external surfaces. Both pyridine and alkylpyridines butylpyridine was used to quantify external BAS
are protonated by BAS. However, due to steric in various parent and modified zeolites indicating
hindrance induced by the bulky substituents, the that the extended mesoporosity and decrease in
alkylpyridine probes do not interact with LAS (32). average length of micropores resulted in increasing
These bulky probe molecules also have limited accessibility of BAS (35). Pivalonitrile adsorption
access to some micropores and consequently are was used to quantify both BAS and LAS, including
suitable to obtain information about the BAS in multivalent transition metal cations hosted in
different locations (14). zeolites, which are considered as active sites in
Many reports have been published on redox reactions (46).
the application of 2,6-di-tert-butylpyridine The quantitative analysis and interpretation
(29, 33–35), 2,6-dimethylpyridine (29, 36–39) of accessibility studies requires careful
and 2,4,6-trimethylpyridine (30, 40–42) for consideration. Indeed, the key to quantitative
zeolite characterisation. For instance, adsorption measurements is the use of molar absorption
of pyridine and 2,4,6-trimethylpyridine was used coefficients (ε). For some test molecules, there
to detect traces of coke in MFI catalysts, and to is a lack of ε values in the literature, and for
determine which acid sites are specifically perturbed most, the values reported show a significant
by coke molecules (43). In the latter study, the degree of variation. Furthermore, the interaction
authors found that coke deposits, resulting from of the probe with the zeolite can be complicated
ortho-xylene isomerisation, do not perturb BAS by pore blockage as some probe molecules can
but perturb non-acidic silanol groups inside the adsorb at the pore mouth of the zeolite and
micropore system. Using the same approach, restrict access to the internal pores that once
Barbera et al. (41) confirmed that the presence were accessible (47). Also, the strength of
of coke could influence catalyst deactivation. Both interaction with some molecules, such as pyridine
studies clearly distinguish between internal and and acetonitrile, could lead to ‘extraction’ of the
external silanol groups and show that silanol defects protons from inaccessible positions, making
play an important role in coke formation over MFI them accessible (48, 49). For this reason, some
catalysts. Corma et al. (34) used the 2,6-di-tert­ molecules with a relatively large kinetic diameter
butylpyridine to investigate the external surface of can access acid sites in small pores of a zeolite.
many zeolitic structures. 2,6-di-tert-butylpyridine For example, Armaroli et al. (39) reported
can enter the 12-membered channels of BEA but different behaviour using probes of similar
not the 10-membered ring channels of ZSM-5 and size (2,6-dimethylpyridine and meta-xylene) in
MCM-22. Therefore, it can be used to identify acid ZSM-5. 2,6-dimethylpyridine enters the pores
sites situated on the external surface of medium of the ZSM-5 zeolite more readily as compared
pore zeolites. to meta-xylene. In addition, Traa et al. (50)
2,4,6-trimethylpyridine and 2,6-dimethylpyridine suggested that the flexibility and the shape of
were used in a novel approach introduced the molecule in relation to the shape of the pore
by Thibault-Starzyk et al. (44), to quantify openings should also be considered.
the accessibility of acid sites in ZSM-5 Overall, the optimisation of the experimental
samples prepared with different degrees of procedures and the application of a combination of
intracrystalline mesoporosity. This approach probe molecules are imperative for the successful
is based on the calculation of the accessibility evaluation of the location and strength of acid sites
index (ACI), the ratio between the number of in different zeolite based materials.
BAS detected by substituted pyridines and the
total number of BAS in the zeolite detected 2.2 Magic Angle Spinning Nuclear
by pyridine (Py). The results showed that the Magnetic Resonance
formation of mesoporosity reduces the average Both FTIR and 1H NMR can directly probe the acidic
length of micropores and leads to an increase in proton, which can be used to differentiate BAS and
the availability of acid sites at the pore mouths. terminal hydroxyls (51). However, the accessibility
This methodology has been successfully used to (location), strength and distribution of acid sites
evaluate the accessibility of acid sites in both cannot be easily measured by NMR. Basic probe
nanocrystalline zeolites and in zeolites with molecules such as pyridine-d5 and 13C-acetone have
relatively large crystal size (44, 45). been used to adsorb on the acid sites and reveal

281 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

their relative strength (52, 53). However, 1H has a As NMR is inherently quantitative, by knowing the
small chemical shift range giving poor differentiation, quantity of TMPO added, a deconvolution of spectral
and 13C (1.1% naturally abundant) is an insensitive peaks directly gives the concentration of acid sites.
nucleus requiring expensive enriched reagents. Furthermore, the 31P chemical shift of the peaks is
Thus, an approach involving 31P was sought, which linear with Brønsted acid strength (proton affinity),
has both a large chemical shift range and high allowing both relative and absolute acid strength to
sensitivity being 100% naturally abundant. be known (71).
Trimethylphosphine (TMP) was first used in 1985
by Lunsford et al. and has since been used as a
3. Experimental Method
probe molecule for BAS and LAS in a variety of solid
acid catalysts (54). TMP can chemisorb on BAS and Ammonium forms of zeolites BEA (CP814E, BEA
LAS or physisorb on weakly acidic hydroxyls, each framework, Si:Al = 12.5) and ZSM-5 (CBV8014,
giving a distinct chemical shift range for each type MFI framework, Si:Al = 40) were obtained from
of interaction. Although TMP is a sensitive probe Zeolyst International, USA. Prior to FTIR studies,
for Lewis acid characterisation, a small Brønsted the zeolites were pressed into self-supporting
acid chemical shift range and volatility of the discs (~8–10 mg) and pretreated in situ in an IR
probe molecule have limited the popularity of this cell at 450°C under vacuum (10–5 Torr) for 5 h.
approach. The adsorption experiments with different probe
Trimethylphosphine oxide (TMPO), which is a molecules were monitored by Thermo ScientificTM
solid at room temperature not susceptible to NicoletTM iSTM10 FT-IR spectrometer equipped with
oxidation, retains the benefits of using 31P NMR a deuterated triglycine sulfate (DTGS) detector,
and offers a greater chemical shift range for BAS at a spectral resolution of 4 cm–1. An excess of
characterisation. The applicability of TMPO probe probe molecules was admitted by injection of
molecules was first shown for zeolites by Rakiewicz 1.0 μl into the IR cell. Physisorbed molecules
et al. in 1998 (55). Over the past 20 years there were subsequently removed by evacuation
have been many 31P NMR studies of phosphorus at the adsorption temperature. Adsorption of
containing probe molecules, especially for the 1,3,5-triisopropylbenzene (C15H24, Acros Organics,
industrially relevant zeolites Beta and ZSM-5 Belgium, 95%) was performed at room temperature.
shown in our examples (56–64). Pyridine (C5H5N, Acros Organics, 99.5%),
A typical procedure for loading a zeolite with 2,6-di-tert-butyl-pyridine (C13H21N, Sigma-Aldrich,
TMPO is to first dehydrate the zeolite, then add USA, 97%), 2,6-dimethylpyridine (C7H9N, Sigma-
TMPO dissolved in dichloromethane (CH2Cl2) Aldrich, 99%) and 2,4,6-trimethylpyridine (C8H11N,
under an inert atmosphere, before heating the BDH reagents, UK, 95%) were adsorbed at 150°C.
sample above the melting point of TMPO (140°C) Desorption profiles were obtained by evacuating
(65). However, we and others (66) have also had the sample at increasing temperatures in 50°C
success in a solvent-free method, by just adding steps.
TMPO at elevated temperatures, which has the The obtained IR spectra were analysed (including
added advantage of avoiding any solvent-zeolite integration, subtraction and determination of peak
interactions. It has been noted that a slight excess positions) using Thermo ScientificTM OMNICTM
of TMPO is crucial in ensuring a complete coverage Series Software. All the spectra presented in this
of acid sites during the experiment (67). work were normalised to 10 mg sample mass.
By varying the length of the alkyl chain, internal For MAS NMR experiments, the two zeolite
and external acid sites can be discriminated (68). samples were dehydrated at 350°C under vacuum
The kinetic diameter of TMPO is 0.55 nm (55), (10–5 Torr) overnight. A slight excess of TMPO
whereas the butyl equivalent, tributylphosphine was added to the zeolites in an argon glovebox,
oxide (TBPO), is 0.82 nm (69), which is larger than followed by a treatment at 165°C for a few hours
the pore size in ZSM-5. to melt the TMPO and distribute it throughout the
Unlike TMP, the chemical shift ranges for TMPO sample. Solid-state NMR spectra were acquired
interacting with Brønsted and Lewis sites overlap. at a static magnetic field strength of 9.4 T on a
However, the TMPO adsorption to Lewis sites is Bruker AvanceTM III console using TopSpin® 3.1
weak and it can be readily displaced by water. software. A widebore Bruker 4 mm BB/1H WVT
Thus, acquiring two spectra with and without MAS probe was used, tuned to 161.98 MHz and
hydration allows one to determine Brønsted versus referenced to ammonium dihydrogen phosphate
Lewis acidity (70). at 0.9 ppm. The samples were packed into

282 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

zirconia MAS rotors with Kel-F® caps in an argon 1545 cm–1 and 1637 cm–1 is due to pyridinium ion
glovebox. (PyH+), two bands assigned to pyridine coordinated
to Lewis acid sites (PyL) at 1456 cm–1 and 1622
cm–1 and the superposition of signals of Lewis and
4. Results and Discussion
Brønsted acid sites at 1491 cm–1 (Figure 1(b)).
4.1 Acidity Measurements on ZSM-5 The concentrations of Brønsted and Lewis acid
and BEA Zeolites sites (Figure 2) have been calculated from the
MFI structure is characterised by two types of intensities of peaks at 1545 cm–1 for BAS and 1456
10-membered ring channels: straight channels cm–1 for LAS. The total concentrations of acid sites
with a nearly circular opening of 5.3 Å × 5.6 Å and for ZSM-5 and BEA parent zeolites are about 350
sinusoidal channels with an elliptical opening of μmol g–1 and 760 μmol g–1. ZSM-5 zeolite with
5.1 Å × 5.5 Å. BEA is a large pore size zeolite with Si:Al = 40 has less Al in the structure compared to
smaller 12-membered ring channels with a cross- BEA with Si:Al = 12.5, and consequently, a lower
section of 5.6 Å × 5.6 Å and larger 12-membered total concentration of acid sites. The concentration
ring channels with a cross-section of 7.7 Å × 6.6 Å. of LAS in ZSM-5 is ~12% of the total number of
FTIR spectra of zeolites BEA and ZSM-5 show acid sites. BEA zeolite presents similar amounts
two major peaks at 3745 cm–1, with a shoulder at of BAS and LAS, 340 μmol g–1 and 420 μmol g–1,
~3735 cm–1, and 3610 cm–1 (Figure 1(a)). The respectively. These data are corroborated by 27Al
band at 3610 cm–1 is assigned to acidic bridging MAS NMR experiments indicating a higher amount
Si-OH-Al groups and the bands at 3745 cm–1 of extra-framework aluminium in BEA compared to
and 3735 cm–1 are attributed to external and the ZSM-5 zeolite.
internal silanol groups (Si-OH), respectively. The The type and concentration of acid sites can
separation of external and internal silanol groups is be readily determined from FTIR spectra of
more noticeable in the spectra of the BEA zeolite. pyridine adsorbed on zeolites. Their strength can
The interaction of pyridine with ZSM-5 and BEA be evaluated by the temperature programmed
zeolites results in a complete disappearance of desorption of pyridine, ammonia or other probe
the band at 3610 cm–1 corresponding to bridging molecules. However, such measurements give
Si-OH-Al groups and a decrease in the intensity of only an effective strength as the probe molecules
the band assigned to Si-OH groups. This means can re-adsorb on available acid sites during the
that pyridine can access all the acid sites of BEA desorption process. Clearly, the observed apparent
and ZSM-5 providing an overall concentration of strength would be affected by the concentration of
acid sites. acid sites, the size of the micropores as well as by
In the range of 1400–1700 cm–1, chemisorbed a number of experimental parameters. In contrast,
31
pyridine is revealed by the following sets of bands: P MAS NMR spectra of TMPO-loaded zeolites

(a) (b)

0.7
0.7
0.7
Absorbance, arbitrary units
Absorbance, arbitrary units

0.6 Si-OH
0.6
0.5
0.5
0.4
0.4 PyL
Si-OH-Al PyH+
0.3
A 0.3
0.2
0.2
B A
0.1 0.1 B
0 0
3800 3750 3700 3650 3600 3550 3500 1700 1650 1600 1550 1500 1450 1400
–1
Wavenumber, cm Wavenumber, cm –1

Fig. 1. (a) Infrared spectra of the hydroxyl region of BEA, A, and ZSM-5, B, activated at 450°C; (b) infrared
spectra of the pyridine region following pyridine adsorption on BEA, A, and ZSM-5, B

283 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

where TMPO was added directly. Both methods


800
ZSM-5 produce similar results, whereas the solvent-
700 BEA free method avoids any potential solvent-zeolite
600 interaction. A small excess of physisorbed TMPO is
Acid sites, μmol g–1

visible around 46 ppm, giving confidence that the


500 accessible acid sites are completely covered. Two
400 types of strong Brønsted site are observed, around
77 ppm and 69 ppm, along with a weaker Brønsted
300
site at 54 ppm. Lewis acid sites from the extra-
200 framework aluminium appeared around 65 ppm.
100
There also appeared to be a very strong Lewis acid
site at 84 ppm, but at a very low concentration.
0
Total Brønsted Lewis The assignment of BAS and LAS was based on
amount acid sites acid sites literature data (70).
Fig. 2. Concentration of acid sites in ZSM-5 Figure 3(b) presents 31P MAS NMR spectra of
and BEA zeolites in quantitative experiments using TMPO dosed BEA zeolite, obtained via the solvent-
pyridine adsorption monitored by FTIR free method. A small quantity of mobile TMPO is
visible around 31 ppm, suggesting physisorbed
TMPO is also present around 46 ppm, but overlapping
provide a direct measure of the strength of acid with Brønsted or Lewis acid sites. Strong BAS can
sites given by the chemical shift of the phosphorus be clearly seen around 75 ppm, along with very
signal. These data are also quantitative, but they strong LAS at 85 ppm. Additional LAS appeared
do require complete coverage of the zeolite by around 65 ppm, as seen for ZSM-5. For BEA, the
TMPO, and therefore rely on the appropriate dosing determination of Lewis versus Brønsted was based
procedure being followed. on hydrating the sample to observe which peaks
Figure 3(a) shows 31P MAS NMR spectra of TMPO disappear. The greater quantity of LAS in BEA over
dosed ZSM-5 by two different methods, one where ZSM-5 agrees with the pyridine-FTIR and 27Al NMR
TMPO was first dissolved in CH2Cl2 and the other results.

(a) (b)
Brønsted sites Lewis sites
77.4 64.5
69.3

53.5
84.0 46.3 Residual TMPO
Sum of fit H-ZSM-5
Spectrum with DCM Brønsted sites
75.5

Lewis sites
65.2 46.7

84.7 Residual TMPO


31.1
H-ZSM-5 Sum of fit H-Beta
solvent free Spectrum

120 100 80 60 40 20 0 120 100 80 60 40 20 0


δ(31P), ppm δ(31P), ppm

Fig. 3. 31P Solid-state MAS NMR spectra of: (a) ZSM-5 dosed with TMPO, with and without using CH2Cl2 as a
solvent; (b) BEA dosed with TMPO without solvent

284 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

Further work would be required to obtain the intensity of the Si-OH band at 3745 cm–1. In the
accurate data for the concentration of each type case of BEA (Figure 4(c)) there is a clear separation
of the detected sites. This would involve precise between external SiOH groups at ~3745 cm–1, which
weighing of the samples and TMPO dose that are interacting with the hydrocarbon molecules, and
was not undertaken for this study. Additionally, a internal silanols at ~3735 cm–1, which are not. At
quantitative analysis of the acid sites present on the the same time, the Si-OH-Al band at ~3610 cm–1
internal and external surfaces could be performed appears to be almost unchanged. However, the
by using bulkier probe molecules such as TBPO. difference spectra (Figures 4(b) and 4(d)) show a
low intensity negative peak at ~3610 cm–1 detected
4.2 Accessibility of Acid Sites in ZSM-5 for both zeolites. This negative peak corresponds to
and BEA Zeolites the acidic Si-OH-Al groups on the external surface
Adsorption of 1,3,5-triisopropylbenzene (kinetic interacting with the probe molecule with the formation
diameter of ~8.5 Å) at 30°C on ZSM-5 and BEA of a hydrogen bond. These data demonstrate that
zeolites (Figure 4) leads to a significant reduction in 1,3,5-triisopropylbenzene, which is too large to enter

(a) (b)

0.40
Si-OH-Al 0.04
Absorbance, arbitrary units

Absorbance, arbitrary units


0.35 Si-OH
0.02

External Si-OH-Al
0.30 0

B –0.02
0.25
A
–0.04
0.20 Si-OH

3800 3750 3700 3650 3600 3550 3500 3800 3750 3700 3650 3600 3550 3500
–1 –1
Wavenumber, cm Wavenumber, cm

(c) (d)

1.2 0.4
Absorbance, arbitrary units
Absorbance, arbitrary units

1.0
Si-OH 0.2

0.8
0 External Si-OH-Al
0.6 Si-OH-Al

B –0.2
0.4
A Si-OH
0.2 –0.4
3800 3750 3700 3650 3600 3550 3500 3800 3750 3700 3650 3600 3550 3500
–1 –1
Wavenumber, cm Wavenumber, cm

Fig. 4. (a) FTIR spectra of ZSM-5 before, A, and after, B, 1,3,5-triisopropylbenzene adsorption at 30°C; (b)
difference spectrum of ZSM-5 before and after adsorption of the probe; (c) FTIR spectra of BEA before, A,
and after, B, 1,3,5-triisopropylbenzene adsorption at 30°C; (d) difference spectrum of BEA before and after
adsorption of the probe

285 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

the 10- and 12-membered channels of ZSM-5 and in the OH region show a low intensity negative
BEA, respectively, can be used to quantify the BAS peak at ~3610 cm–1 after 2,4,6-trimethylpyridine
located on the external surface. and 2,6-di-tert-butyl-pyridine adsorption, which
FTIR spectra of ZSM-5 zeolite following adsorption corresponds to the small fraction of bridging
of 2,6-dimethylpyridine, 2,4,6-trimethylpyridine OH groups located on the external surface of the
and 2,6-di-tert-butyl-pyridine at 250oC are zeolite or near the pore mouths.
presented in Figure 5. All the substituted Adsorption of 2,6-dimethylpyridine (Figure 5(c),
pyridines interact with terminal SiOH groups on the Spectrum A) leads to the appearance of two bands
external surface reducing the intensity of the band around 1600–1680 cm–1 (72). 2,4,6-trimethylpyridine
at 3745 cm–1. The band of Si-OH-Al is virtually adsorption (Figure 5(c), Spectrum B) gives rise
unaffected in the case of 2,4,6-trimethylpyridine to the band at ~1634 cm–1 with a shoulder
and 2,6-di-tert-butyl-pyridine adsorption, but at ~1649 cm–1 resulting from the interaction with
does decrease noticeably following adsorption of BAS; two low intensity bands at 1619 cm–1 and
2,6-dimethylpyridine, indicating that the latter 1575 cm–1 are assigned to the probe adsorbed on
probe can access some of the BAS inside the Si-OH groups (42). The spectra of 2,6-di-tert-butyl­
micropore system. 2,6-di-tert-butyl-pyridine pyridine (Figure 5(c), Spectrum C) show a band at
and 2,4,6-trimethylpyridine are not able to enter 1615 cm–1 attributed to the probe bonded to BAS
the micropores of ZSM-5. The difference spectra (35). Based on the assignment of these bands and

(a) (b)

1.1 0.3
Si-OH-Al
Si-OH
1.0 A
Absorbance, arbitrary units
Absorbance, arbitrary units

0.2
0.9 A
0.8 0.1
B
0.7 B
0
0.6
C –0.1
0.5 C

0.4 –0.2
D Si-OH-Al
Si-OH
0.3 –0.3
3800 3750 3700 3650 3600 3550 3500 3800 3750 3700 3650 3600 3550 3500
–1
Wavenumber, cm–1 Wavenumber, cm

(c)

0.4 Fig. 5. (a) FTIR spectra of ZSM-5 before,


A, and after adsorption of alkylpyridines:
2,6-dimethylpyridine, B, 2,4,6-trimethylpyridine,
Absorbance, arbitrary units

C, and 2,6-di-tert-butyl-pyridine, D; (b) difference


0.3
spectra in the OH region after adsorption
of alkylpyridines: 2,6-dimethylpyridine, A,
2,4,6-trimethylpyridine, B, and 2,6-di-tert-butyl­
0.2 pyridine, C; (c) difference spectra in the region
A
of the aromatic ring vibrations of alkylpyridines:
2,6-dimethylpyridine, A, 2,4,6-trimethylpyridine,
0.1 B B, and 2,6-di-tert-butyl-pyridine, C

C
0
1700 1650 1600 1550 1500
–1
Wavenumber, cm

286 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

the extinction coefficient values available in the 5. Conclusions


literature (30, 35, 37), the number of BAS accessible
to these probe molecules and the corresponding Characterisation of the acidic properties of zeolites
accessibility indices have been calculated assuming has received a great deal of attention in recent
1:1 interaction with the BAS (Table I). The total decades. FTIR and MAS NMR are now established
amount of BAS is obtained by probing the zeolites as major analytical techniques providing detailed
with pyridine (44). 2,6-dimethylpyridine, being information on the type, concentration, accessibility
bigger than pyridine, probes 47% of the total and location of acid sites. This work demonstrates
amount of BAS; 2,4,6-trimethylpyridine can access several examples of FTIR and NMR evaluation of the
8% and 2,6-di-tert-butyl-pyridine only 5%. These acidic properties of ZSM-5 and BEA zeolites using
results agree with previously published reports on a range of test molecules under in situ conditions.
the adsorption of alkylpyridines on ZSM-5 zeolites For instance, the accessibility and the number of
(35, 44). The diffusion of 2,6-dimethylpyridine, with acid sites on the internal and external surfaces has
the kinetic diameter of 6.7 Å, in the micropores of been determined using adsorption of bulky probe
ZSM-5 zeolite (maximum pore size of 5.6 Å) is molecules monitored by FTIR, hence providing a
restricted, depending on the temperature and duration clear method for the detailed examination of the
of the experiment it can access up to about 50% of acid sites in MFI and BEA structures. The application
the BAS. These data confirm that the accessibility of of 31P MAS NMR to the analysis of the interactions
the acid sites in the zeolite micropores is controlled between TMPO as a probe molecule and the zeolite
by both molecular sieving and strength of interaction has provided in-depth information about the type
between the probe molecule and the acid site. and the strength of the acid sites.
The relatively large size of 2,6-di-tert-butyl-pyridine This work can be further extended to include
(7.9 Å) and 2,4,6-trimethylpyridine (7.4 Å) prevents detailed characterisation of new and modified
their access to BAS in the micropores of ZSM-5. zeolite-based catalysts, particularly utilising a
In the case of BEA zeolite, the FTIR spectra combination of several techniques. In addition,
demonstrate that all three substituted pyridines are the experimental method should be optimised to
protonated on all Si-OH-Al and some Si-OH groups, improve the accuracy of the quantitative analysis
as they interact with all BAS on the external surface under in situ and realistic reaction conditions and
and in the micropores of zeolite BEA. Indeed, the for cross validation of the data obtained from
size of the substituted pyridines are similar to the different techniques.
dimensions of the larger pores in the BEA structure
(7.7 Å × 6.6 Å) allowing their access to the BAS in
Acknowledgements
the micropore system.
We thank Laura Powell and Peter J. Ellis for
performing initial TMPO experiments as well as
Loredana Mantarosie and Markus Knaebbeler-
Buss for helping with sample preparation. We
Table I Concentration of Brønsted gratefully acknowledge Johnson Matthey, UK and
Acid Sites and Accessibility Indices for Keele University, UK for their support and funding
ZSM-5 Zeolite Determined using provided for this work.
Adsorption of Alkylpyridines
Py a Lu b Coll c DTBPy d References
e
Concentration 305 143 25 14
of accessible 1. J. Weitkamp, Solid State Ionics, 2000, 131,
BAS, μmol g–1 (1–2), 175
Accessibility 100 47 8 5 2. A. Corma, J. Catal., 2003, 216, (1–2), 298
index, % 3. J. Čejka, G. Centi, J. Perez-Pariente and W. J.
Roth, Catal. Today, 2012, 179, (1), 2
a
Py = pyridine
b
Lu = 2,6-dimethylpyridine 4. R. Xu, W. Pang, J. Yu, Q. Huo and J. Chen,
c
Coll = 2,4,6-trimethylpyridine “Chemistry of Zeolites and Related Porous
d
DTBPy = 2,6-di-tert-butylpyridine
e Material: Synthesis and Structure”, John Wiley &
Depends on the temperature and duration of the adsorption
experiment Sons (Asia) Pte Ltd, Singapore, 2007, 616 pp

287 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

5. A. Corma, Chem. Rev., 1997, 97, (6), 2373


The Netherlands, 2007, pp. 218–219
6. S. van Donk, A. H. Janssen, J. H. Bitter and K. P.
28. L. M. Chua, I. Hitchcock, R. S. Fletcher, E. M. Holt,
de Jong, Catal. Rev. Sci. Eng., 2003, 45, (2), 297
J. Lowe and S. P. Rigby, J. Catal., 2012, 286, 260

7. J. Pérez-Ramírez, C. H. Christensen, K. Egeblad, 29. N. S. Nesterenko, F. Thibault-Starzyk, V. Montouillout,
C. H. Christensen and J. C. Groen, Chem. Soc.
V. V. Yuschenko, C. Fernandez, J.-P. Gilson, F. Fajula

Rev., 2008, 37, (11), 2530


and I. I. Ivanova, Micro. Meso. Mater., 2004, 71,

8. Y. Wei, T. E. Parmentier, K. P. de Jong and J.


(1–3), 157

Zečević, Chem. Soc. Rev., 2015, 44, (20), 7234


30. N. S. Nesterenko, F. Thibault-Starzyk, V.
9. M. Milina, S. Mitchell, P. Crivelli, D. Cooke and J.
Montouilliout, V. V. Yushchenko, C. Fernandez, J.­
Pérez-Ramírez, Nat. Commun., 2014, 5, 3922
P. Gilson, F. Fajula and I. I. Ivanova, Kinet. Katal.,

2006, 47, (1), 45; translated into English in Kinet.

10. “Mesoporous Zeolites: Preparation, Characterization


Catal., 2006, 47, (1), 40

and Applications”, eds. J. García-Martínez and


K. Li, Wiley-VCH Verlag GmbH & Co, Weinheim, 31. F. L. Bleken, K. Barbera, F. Bonino, U. Olsbye,
Germany, 2015, 608 pp
K. P. Lillerud, S. Bordiga, P. Beato, T. V. W. Janssens

and S. Svelle, J. Catal., 2013, 307, 62



11. K. Hadjiivanov, Adv. Catal., 2014, 57, 99

32. P. A. Jacobs and C. F. Heylen, J. Catal., 1974, 34,

12. D. Zhai, Y. Li, H. Zheng, L. Zhao, J. Gao, C. Xu and


(2), 267

B. Shen, J. Catal., 2017, 352, 627

33. V. V. Ordomsky, V. Y. Murzin, Yu. V. Monakhova,
13. C. Lamberti, A. Zecchina, E. Groppo and S. Bordiga,

Y. V. Zubavichus, E. E. Knyazeva, N. S. Nesterenko

Chem. Soc. Rev., 2010, 39, (12), 4951

and I. I. Ivanova, Micro. Meso. Mater., 2007, 105,

14. S. Bordiga, C. Lamberti, F. Bonino, A. Travert and (1–2), 101

F. Thibault-Starzyk, Chem. Soc. Rev., 2015, 44,

34. A. Corma, V. F ornés, L. Forni, F. Márquez, J.

(20), 7262

Martínez-Triguero and D. Moscotti, J. Catal., 1998,

15. A. V
imont, F. Thibault-Starzyk and M. Daturi,
179, (2), 451

Chem. Soc. Rev., 2010, 39, (12), 4928

35. K. Góra-Marek, K. Tarach and M. Choi, J. Phys.

16. M. Niwa, N. Katada and K. Okumura, “Characterisation


Chem. C, 2014, 118, (23), 12266

and Design of Zeolite Catalysts: Solid Acidity, Shape

36. L. Oliviero, A. Vimont, J.-C. Lavalley, F. R. Sarria,


Selectivity and Loading Properties”, Springer-Verlag,

M. Gaillard and F. Maugé, Phys. Chem. Chem.

Berlin, Germany, 2010, 184 pp

Phys., 2005, 7, (8), 1861



17. G. Busca, Micro. Meso. Mater., 2017, 254, 3

37. T. Onfroy, G. Clet and M. Houalla, Micro. Meso.

18. L.-E. Sandoval-Díaz, J.-A. González-Amaya a nd


Mater., 2005, 82, (1–2), 99

C.-A. Trujillo, Micro. Meso. Mater., 2015, 215, 229

38. A. Corma, C. Rodellas and V. Fornes, J. Catal.,

19. M. Trombetta, G. Busca, M. Lenarda, L. Storaro and 1984, 88, (2), 374

M. Pavan, Appl. Catal. A: Gen., 1999, 182, (2), 225


39. T. Armaroli, M. Bevilacqua, M. Trombetta, A. G.
20. T. Armaroli, M. Bevilacqua, M. Trombetta, Alejandre, J. Ramirez and G. Busca, Appl. Catal.
F. Milella, A. G. Alejandre, J. Ramírez, B. Notari, A: Gen., 2001, 220, (1–2), 181

R. J. Willey and G. Busca, Appl. Catal. A: Gen.,


40. K. Barbera, F. Bonino, S. Bordiga, T. V. W. Janssens

2001, 216, (1–2), 59


and P. Beato, J. Catal., 2011, 280, (2), 196

21. M. Trombetta, T. Armaroli, A. G. Alejandre, J. R.
41. S. M. T. Almutairi, B. Mezari, E. A. Pidko, P. C. M.
Solis and G. Busca, Appl. Catal. A: Gen., 2000,
M. Magusin and E. J. M. Hensen, J. Catal., 2013,

192, (1), 125


307, 194

22. M. Bevilacqua and G. Busca, Catal. Commum.,


42. M. S. Holm, S. Svelle, F. Joensen, P. Beato, C. H.

2002, 3, (11), 497


Christensen, S. Bordiga and M. Bjørgen, Appl.

23. T. Montanari, M. Bevilacqua and G. Busca, Appl.


Catal. A: Gen., 2009, 356, (1), 23

Catal. A: Gen., 2006, 307, (1), 21


43. F. Thibault-Starzyk, A. Vimont and J.-P. Gilson,

24. D. Tzoulaki, A. Jentys, J. Pérez-Ramírez, K. Egeblad


Catal. Today, 2001, 70, (1–3), 227

and J. A. Lercher, Catal. Today, 2012, 198, (1), 3


44. F. Thibault-Starzyk, I. Stan, S. Abelló, A. Bonilla,
25. O. Marie, P. Massiani and F. Thibault-Starzyk, J.
K. Thomas, C. Fernandez, J.-P. Gilson and J. Pérez-

Phys. Chem. B, 2004, 108 (16), 5073


Ramírez, J. Catal., 2009, 264, (1), 11

26. M. Maache, A. Janin, J. C. Lavalley and E. Benazzi,


45. K. Mlekodaj, K. Tarach, J. Datka, K. Góra-Marek

Zeolites, 1995, 15, (6), 507


and W. Makowski, Micro. Meso. Mater., 2014, 183,

27. C. Baerlocher, L. B. McCusker and D. H. Olson, 54

‘MOR: Cmcm’, in “Atlas of Zeolite Framework 46. K. Sadowska, K. Góra-Marek and J. Datka, J. Phys.

Types”, 6th Edn., Elsevier Science BV, Amsterdam, Chem. C, 2013, 117, (18), 9237

288 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

47. A. Zecchina, S. Bordiga, G. Spoto, D. Scarano, G.


59. A. Zheng, S.-J. H
uang, Q. Wang, H. Zhang, F.

Spanò and F. Geobaldo, J. Chem. Soc., Faraday


Deng and S.-B. Liu, J. Catal., 2013, 34, (3), 436

Trans., 1996, 92, (23), 4863
60. A. Zheng, S.-B. Liu and F. Deng, Solid State Nucl.

48. G. Crépeau, V. Montouillout, A. Vimont, L. Mariey, Magn. Reson., 2013, 55–56, 12

T. Cseri and F. Maugé, J. Phys. Chem. B, 2006,


61. A. Zheng, F. Deng and S.-B. Liu, Ann. Rep. NMR

110, (31), 15172


Spectrosc., 2014, 81, 47

49. W. Daniell, N.-Y. Topsøe and H. Knözinger,


62. C. E. Hernandez-Tamargo, A. Roldan and N. H. de

Langmuir, 2001, 17, (20), 6233


Leeuw, J. Phys. Chem. C, 2016, 120, (34), 19097

50. Y. Traa, S. Sealy and J. Weitkamp, ‘Characterization 63. A. Zheng, S. Li, S.-B. Liu and F. Deng, Acc. Chem.

of the Pore Size of Molecular Sieves Using Molecular Res., 2016, 49, (4), 655

Probes’, in “Molecular Sieves: Characterization II”, 64. R. Zhao, Z. Zhao, S. Li, and W. Zhang, J. Phys.

eds. H. G. Karge and J. Weitkamp, Vol. 5, Springer- Chem. Lett., 2017, 8, (10), 2323

Verlag, Berlin, Germany, 2007, pp. 103–154


65. A. Zheng, S.-J. Huang, S.-B. Liu and F. Deng, Phys.

51. M. Müller, G. Harvey and R. Prins, Micro. Meso.


Chem. Chem. Phys., 2011, 13, (33), 14889

Mater., 2000, 34, (3), 281

66. S. Hayashi, K. Jimura and N. Kojima, Micro. Meso.

52. Y. J iang, J, Huang, W. Dai and M. Hunger, Solid


Mater., 2014, 186, 101

State Nucl. Magn. Reson., 2011, 39, (3–4), 116



67. A. Zheng, S.-B. Liu and F. Deng, Chem. Rev.,

53. L.-E. Sandoval-Díaz, J.-A. González-Amaya and C.­ 2017, 117, (19), 12475

A. Trujillo, Micro. Meso. Mater., 2015, 215, 229

68. Q. Zhao, W.-H. Chen, S.-J. Huang, Y.-C. Wu, H.-K.

54. J. H. Lunsford, W. P. Rothwell and W. Shen, J. Am.


Lee and S.-B. Liu, J. Phys. Chem. B, 2002, 106,

Chem. Soc., 1985, 107, (6), 1540


(17), 4462

55. E. F. Rakiewicz, A. W. Peters, R. F. Wormsbecher, 69. C. E. Webster, R. S. Drago and M. C. Zerner, J.

K. J. Sutovich and K. T. Mueller, J. Phys. Chem. B,


Phys. Chem. B, 1999, 103, (8), 1242

1998, 102, (16), 2890


70. P. V. Wiper, J. Amelse and L. Mafra, J. Catal., 2014,

56. H.-M. Kao, C.-Y. Yu and M.-C. Yeh, Micro. Meso.


316, 240

Mater., 2002, 53, (1–2), 1


71. A. Zheng, L. Chen, J. Yang, M. Zhang, Y. Su, Y.

57. J. Guan, X. Li, G. Yang, W. Zhang, X. Liu, X. Han


Yue, C. Ye and F. Deng, J. Phys. Chem. B, 2005,

and X. Bao, J. Mol. Catal. A: Chem., 2009, 310,


109, (51), 24273

(1–2), 113
72. F. Leydier, C. Chizallet, A. Chaumonnot, M.

58. Y. Seo, K. Cho, Y. Jung and R. Ryoo, ACS Catal.,


Digne, E. Soyer, A.-A. Quoineaud, D. Costa and P.

2013, 3, (4), 713


Raybaud, J. Catal., 2011, 284, (2), 215

The Authors
Cátia Freitas graduated from the University of Porto, Portugal, in 2012 with a Bachelor’s
degree in Chemistry, and a Master’s degree in Chemical Analysis and Characterisation
Techniques from the University of Minho, Portugal, in 2014. She worked as a researcher in
the Heterogeneous Catalysis and Catalytic Processes (CATHPRO) in Lisbon, Portugal. Cátia
is now in the third year of a PhD project under the supervision of Vladimir Zholobenko at
the Birchall Centre, Keele University, UK, sponsored by Johnson Matthey, UK. Her research
focuses on the FTIR characterisation of acid sites in zeolite-based catalysts in terms of
nature, location, concentration and strength.

289 © 2018 Johnson Matthey


https://doi.org/10.1595/205651318X696792 Johnson Matthey Technol. Rev., 2018, 62, (3)

Nathan Barrow is currently a Principal Scientist in the Advanced Characterisation department


at Johnson Matthey’s Technology Centre, Sonning Common, UK. He graduated with an
MPhys in 2006 from the University of Warwick, UK, where he remained to gain a PhD in
solid-state NMR. In 2010 Barrow was a Knowledge Transfer Partnership associate between
the University of Warwick and Johnson Matthey, helping to install and run a solid-state NMR
service. His current research focuses on applying advanced characterisation to materials
such as zeolites, alumina, glasses and batteries.

Vladimir Zholobenko is a senior lecturer at Keele University, UK. He graduated with a


MSc Diploma in Chemistry from the Moscow State University, Russian Federation, later
obtaining his PhD in Chemistry from N. D. Zelinsky Institute of Organic Chemistry, Moscow,
Russian Federation. This was followed by postdoctoral research at N. D. Zelinsky Institute
of Organic Chemistry; UMIST, Manchester, UK and Northwestern University, Evanston,
USA. His current research is concerned with catalysis by zeolites and nanostructured
materials and spectroscopic methods for their characterisation.

290 © 2018 Johnson Matthey

Potrebbero piacerti anche