You are on page 1of 10

Perspective

https://doi.org/10.1038/s41563-020-0751-3

Towards a better understanding of Lewis acidic


aluminium in zeolites
Manoj Ravi   1, Vitaly L. Sushkevich2 and Jeroen A. van Bokhoven   1,2 ✉

Zeolites are a class of materials that are of great relevance for industrial catalysis. Several fundamental questions relating to
the structure and role of the Lewis acid sites in these materials remain unanswered. Proposals for the origin of such species can
broadly be classified into three categories, which have distinct structures: extra-framework, framework-associated and frame-
work aluminium. In this Perspective, we review each of these proposals and proceed to analyse their suitability to understand
experimental results. Contrary to traditional belief, the number of Lewis acid sites does not always correlate to extra-framework
aluminium content. As a result, we highlight that the terms ‘extra-framework’ and ‘framework-associated’ aluminium should
be used with caution. We propose how the usage of different characterization techniques can enable the closure of knowledge
gaps concerning the strength, multiplicity, localization and structure of catalytically active Lewis acid sites in zeolites.

Z
eolites constitute a class of heterogeneous catalysts for acid-, Hence, zeolites play a vital role in several catalytic applications in
base- and metal-catalysed reactions in several small- to biomass valorization chemistry.
large-scale industrial applications. Their catalytic versatility is Lewis acid sites in zeolites can be generated by introducing
due to the coexistence of aluminium and silicon in the framework1. heteroatoms, such as titanium, tin or zirconium, into the frame-
In a zeolite, framework aluminium is tetrahedrally coordinated with work (Fig. 1). Such acid sites are exploited to catalyse aldol con-
four –O–Si linkages in accordance with Löwenstein’s rule, which densation14 and propylene epoxidation15. Synthesis of tin- and
prohibits the presence of Al–O–Al motifs (Fig. 1)2. The presence zirconium-doped Lewis acidic zeolites is often complicated16,17, as
of aluminium in a framework position introduces a negative obtaining isolated acid sites is not trivial and metal incorporation
charge to the framework that must be charge balanced by an has been proved only for a limited number of zeolite topologies18,19.
extra-framework cationic species. If the cation is a proton, this In contrast, Lewis acidic aluminium in zeolites can be formed with-
results in formation of a hydroxyl group bridging aluminium and out tedious synthesis procedures and is not restricted to specific
silicon, forming a catalytically active Brønsted acid site3 (Fig. 1). zeolite topologies. However, unlike Brønsted acidity, the concept
Consequently, zeolites have become the backbone of catalytic pro- of Lewis acidic aluminium in zeolites is still a grey area. While the
cesses in the petrochemical industry4, with application to some structure of the Brønsted acid site, described as the bridging OH
prominent reactions, including fluid catalytic cracking (FCC)5, group in the Si–(OH)–Al unit (Fig. 1)3, is well defined, the same
methanol to olefins6, methanol to gasoline7 and hydroisomerization cannot be said of Lewis acidic aluminium in zeolites. Evaluating
of hydrocarbons8. the structure and strength of Lewis acidic aluminium remains an
The possibility of hosting diverse active site motifs in zeolites, area of active research and discussion20–23. One of the reasons for
which can be tuned for a particular application, means that the this uncertainty is the plurality of potential Lewis acid sites, aris-
industrial use of zeolites is unrivalled by other catalytic materi- ing from the structural versatility of aluminium that results in
als. They can be used for the production of value-added chemicals many aluminium species. Aluminium with Lewis acidic character
from oil derivatives, but also have the potential to be catalysts for a may be generated through post-synthesis methods, such as steam-
biomass-based economy9. The active sites in biomass valorization ing, calcination at high temperature, and acid or base leaching24–27.
reactions involve both Brønsted and Lewis acidity; however, the These treatments generally result in the hydrolysis of framework
role of these sites is not well understood as of yet. Such coopera- Si–O–Al bonds, removing aluminium from the framework and
tive catalysis featuring both types of acidity often entails a multistep generating ‘extra-framework aluminium’ (EFAl) species with Lewis
conversion of biomolecules to valuable compounds. For example, acidity28. Moreover, Lewis acidic aluminium can also be a property
two such important cascade reactions are the conversion of glucose inherent to the zeolite framework20,21,23. Multiple proposals on the
to 5-hydroxymethylfurfural (HMF)10 and of C3 sugars to lactic acid structure and generation of Lewis acid sites have been advanced,
esters11. The former reaction comprises two steps, namely, isomeri- which can broadly be classified as originating from framework20,29,
zation of glucose to fructose, catalysed by Lewis acid sites, and dehy- framework-associated23,30 and extra-framework aluminium31,32.
dration of fructose to HMF, catalysed by Brønsted acid sites. In this Thus, Lewis acid sites in a zeolite are not necessarily defined by a
regard, aluminium zeolites represent bifunctional catalysts as they single structure, and the total Lewis acidity has contributions from
possess both Lewis and Brønsted acidity, showing high activity and multiple species.
selectivity to HMF12. Elsewhere, Lewis acid sites are paramount in Numerous publications have discussed the concept of Lewis acid
modulating product selectivity in the dehydration of glycerol. Using sites in zeolites33–35. A number of different proposals on the origins
materials with a greater number of Lewis acid sites results in a lower of Lewis acidity have been posited, but the suitability of each in
selectivity to acrolein and a higher selectivity to acetaldehyde13. matching experimental data collected from many studies has not

Institute for Chemical and Bioengineering, Department of Chemistry and Applied Biosciences, ETH Zurich, Zurich, Switzerland. 2Laboratory for Catalysis
1

and Sustainable Chemistry, Paul Scherrer Institute, Villigen, Switzerland. ✉e-mail: jeroen.vanbokhoven@chem.ethz.ch

Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials 1047


Perspective NAture MAterIALS

steaming, wherein water cleaves Si–O–Al linkages at high tempera-


ture to force aluminium into interstitial space27,40–42. There is a pre-
Oil Biomass disposition to associate such EFAl with Lewis acidic property. While
this standpoint is not per se incorrect, there are caveats that need
FCC Glucose to fructose to be considered. One of the first reports that characterized HY
Isomerization Fructose to HMF
Alkylation Glycerol to acrolein zeolites dealuminated by steaming explicitly states that only a small
Zeolites
proportion of EFAl has acidic character43. Steaming leads to the
loss of Brønsted acid sites through irreversible framework dealu-
mination and reversible poisoning by cationic EFAl. In the case of
the latter, acid leaching of the steamed sample extracts EFAl and
? restores some of the Brønsted acid sites32,43. Increasing the severity
of acid treatment increases the number of restored Brønsted acid
M = Sn, Zr, Ti
sites while simultaneously reducing the total number of Lewis acid
Framework aluminium Aluminium Lewis Framework metal sites44. Therefore, EFAl does not refer to a single aluminium species
Brønsted acid sites acid sites Lewis acid sites
with Lewis acid character, but instead refers to a family of species
that are external to the zeolite lattice and have varying degrees of
Fig. 1 | Zeolites, from acid sites to catalytic applications. A sketch showing acidity (Box 1).
the structures of Brønsted and Lewis acid sites in zeolites and their relevant In the case of dealuminated zeolite Y, Al(OH)3 and AlOH2+
applications in oil- and biomass-based processes. have been identified as the extra-framework species38. 27Al
double-quantum magic angle spinning (MAS) NMR experiments
reveal tetrahedral framework aluminium, octahedral Al(OH)3 and
been assessed. Through evaluation of published data and analysis of penta-coordinated extra-framework Al(OH)2+ and that they are all
the different mechanisms that generate Lewis acidic species, a better in close proximity to one another (structures A, B and C, respectively,
understanding of Lewis acid sites in zeolites can be attained. From in Fig. 2)27. 31P solid-state NMR of trimethylphosphine-adsorbed
a materials perspective, this would provide greater control over the zeolite Y and density functional theory calculations suggest
generation of Lewis acidic aluminium, paving a way towards the that the strongest Lewis acid sites in dealuminated zeolites are
currently elusive design of such sites in zeolites. From a process three-coordinated EFAl (structure D in Fig. 2)45. Periodic density
perspective, this leaves a tangible impact on catalytic applications functional theory calculations reveal a thermodynamic preference
based on Lewis acidic aluminium and Brønsted–Lewis acid synergy. for the formation of multinuclear [Al3O4H3]4+ extra-framework
Finally, we proceed to identify knowledge gaps in the field and dis- clusters inside the sodalite cage in faujasite (FAU); however, this
cuss how these may be addressed in the future. Even though there is remains to be experimentally confirmed46,47.
room for development, we believe that the tools necessary to attain Our analysis of published data establishes that there is no cor-
a comprehensive understanding are available, and through a con- relation between the amount of EFAl and the number of Lewis acid
certed effort, breakthroughs are a real possibility. sites in zeolites (Box 1). Therefore, it becomes apparent that trivially
associating EFAl with Lewis acid property is a fallacious argument.
Assessing the proposals for Lewis acidic aluminium in While certain EFAl species undeniably possess Lewis acid character,
zeolites the underlying heterogeneity of EFAl debris must not be overlooked.
Herein, we review the different proposals for Lewis acidity in The classification of aluminium as ‘framework’ or ‘extra-framework’
zeolites, starting with Lewis acid sites originating from EFAl, and is commonly done on the basis of 27Al MAS NMR spectra acquired
progressing to those originating from framework-associated and on wet non-dehydrated samples. In contrast, Lewis acidity is typi-
framework aluminium. cally assessed after dehydrating the zeolite, which often affects
the intrinsic structure of the aluminium sites (Box 1). The term
Lewis acidity of EFAl. The term ‘extra-framework aluminium’ is, by ‘extra-framework aluminium’ constitutes a range of aluminium spe-
convention, reserved to describe aluminium species that are ‘entirely’ cies with largely varying acidic properties. Its structure and acidic
dislodged from the zeolite framework. EFAl exists in several oxide properties are influenced by several factors that include the Si/Al
and hydroxide forms, such as Al(OH)2+, Al(OH)2+, AlOOH, Al2O3, ratio of the zeolite and the method used for dealumination (Box 2).
Al(OH)3 or multinuclear clusters (Box 1)27,36–38. In the ensuing discus-
sion, we refer to such species as ‘extra-framework’, while those that are Lewis acidity of framework-associated aluminium. As opposed
‘partially’ dislodged from the framework will be referred to as ‘frame- to EFAl, the term ‘framework associated’ refers to species that
work associated’ (vide infra). In addition, amorphous aluminosilicate are not entirely dislodged from the zeolite framework (Fig. 2).
can also constitute extra-framework species, although their experi- Framework-associated aluminium adopts a tetrahedral or an
mental observation requires severe steaming/thermal treatment39. octahedral geometry as a function of conditions and can undergo
One of the foremost publications to attribute EFAl debris to Lewis reversible transformation from one coordination to the other
acidic property is the work of Jacobs and Beyer, wherein Lewis acid (structures A, E, F and G in Fig. 2)48–51. On being charge balanced
sites were shown to consist of Al–O species leached from the frame- by cations such as sodium, potassium and ammonium, this alumin-
work of zeolite Y31. As performed in this study, Fourier transform ium species assumes a tetrahedral coordination (27Al chemical shift
infrared (FTIR) spectroscopy of adsorbed bases is a familiar character- around 60 ppm) in the hydrated form of the zeolite, and an octa-
ization technique to study acid sites in zeolites. The pre-treatment of hedral environment (27Al chemical shift around 0 ppm) on being
the sample before the adsorption of the base, which entails the heating charge balanced by protons23,30,48–50. The property of this species to
of the zeolite under vacuum at high temperatures, changes the struc- fit back into a tetrahedral framework coordination after being at
ture of aluminium species. Thus, the structural information retrieved least partially dislodged from the framework suggests that it has a
by FTIR spectroscopy corresponds to that prevailing under activated, memory of its native position, thereby intuitively justifying the epi-
dehydrated conditions and not at ambient conditions (Box 1). thet ‘framework-associated’ aluminium. Specific T-sites are more
prone to exhibit this phenomenon than others41 and the amount of
Generation of EFAl and its experimental characterization. framework-associated aluminium depends on the zeolite structure
Aluminium species external to the lattice are usually generated by and Si/Al ratio23,52.

1048 Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials


NAture MAterIALS Perspective
Box 1 | Correlating EFAl to Lewis acidic property

In the graph below, we collate data from published literature for two tetrahedral EFAl species characterized by chemical shifts of
the correlation between the number of Lewis acid sites and the 44 and 63 ppm in 27Al MAS NMR, the fraction of aluminium spe-
amount of EFAl in different zeolites26,28,85,89,90. Interestingly, a cor- cies associated with the former resonance falls substantially on
relation between the two quantities, one that is often assumed steaming85. These results suggest that besides displacing alumini-
to exist, does not hold for most cases. While some species of the um from the framework, steaming transforms the nature of EFAl,
extra-framework phase are Lewis acidic, not all are. Therefore, thereby influencing its acidity. This serves as another example to
a zeolite with more EFAl content does not necessarily possess a illustrate the pitfall of assuming all of EFAl as Lewis acidic.
greater number of Lewis acidic species. Consider as an example, An important proviso that applies while correlating a particular
the experiments with dealuminated HY91. The zeolite steamed at aluminium species to Lewis acidic property comes from the
973 K has a framework Si/Al ratio of 15 and a bulk ratio of 2.4. The characterization method used to probe Lewis acid sites. The
presence of Lewis acidic species in the sample is evidenced by an plurality of acid sites and their varying strengths are reflected
infrared band at 1,455 cm–1 on being probed by pyridine. Treating differently in different techniques. Consider as an example, the
the zeolite in 1.5 M HCl results in the framework Si/Al ratio re- study of Lewis acidity by a rather exotic method like electron
maining unchanged, but the bulk Si/Al ratio increases to 6.7. Thus, paramagnetic resonance spectroscopy of adsorbed nitric oxide.
a substantial amount of EFAl is leached out by acid treatment, but This technique only probes sites that are strong enough to quench
the framework composition remains unaffected. However, this the orbital magnetic moment of nitric oxide. The number of such
sample with lower EFAl content shows a more intense infrared strong Lewis acid sites increases initially with the amount of EFAl,
feature for Lewis-bound pyridine91. Besides eliminating cationic and then tapers off with further increase in EFAl (grey points)28.
species located in the supercages91, the acid treatment modulates This trend does not hold when probing the same samples by
the number of Lewis acid sites by manipulating EFAl species gen- FTIR spectroscopy of adsorbed carbon monoxide28. Hence, the
erated during steaming, or inducing framework-based Lewis acid- sensitivity to a particular Lewis acid species is fundamentally
ity (vide infra), or a combination of both. However, dealuminat- different from one technique to another.
ing USY using citric acid instead of steaming, generates EFAl that The figure shows the structural models for different aluminium
linearly correlates to the number of sites that show Lewis acidic species: framework2,3, extra-framework27,36–38 and framework-
property (green points)89. The opposite trend is observed while associated aluminium (inspired from refs. 27,50). Experimental
steaming H-BEA at 753 K (orange points)85. Herein, the popula- techniques and typical conditions during measurement of
tion of octahedral EFAl barely increases upon steaming, with the aluminium structure and acidity are listed in blue text boxes. The
more pronounced changes occurring to tetrahedral EFAl. Of the curation of data for the number of Lewis acid sites as a function

Measurements of Al coordination Structural models Measurements of Al acidity


• Three-coordinated Framework aluminium • Lewis acid sites
• Tetrahedral (Td) • Brønsted acid sites
• Penta-Coordinated
• Octahedral (Oh)
Techniques Techniques
• 27Al MAS NMR • FTIR of adsorbed probe
• AI K-edge XAS Extra-framework aluminium molecules
• XPS (Al KLL peak) • Structure-sensitive
reactions

Typical experimental conditions Typical experimental conditions


• In hydrated state • After dehydration in
for 27Al MAS NMR vacuum or dry gas

Framework-associated aluminium

Linking acidity and coordination to structure


Number of Lewis acid sites (a.u.)

Amount of EFAI (a.u.)

Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials 1049


Perspective NAture MAterIALS

Box 1 | Correlating EFAl to Lewis acidic property (continued)


of EFAl for different zeolite topologies with the colour codes The amount of Lewis acid sites and EFAl are reported in different
and corresponding references is as follows: blue26, grey28, units in these studies: either on a relative or absolute quantitative
green89, orange85 and purple90. Scatter points from different basis. Hence, the plot is presented with both axes being labelled as
studies are offset for clarity and thus cannot be directly compared. arbitrary units.

Extra-framework aluminium D

Extra-framework LAS

High temperature, vacuum

B C M N

EFAI: octahedral EFAI: penta-coordinated EFAI: tetrahedral Other neutral and cationic EFAI

High Very high Very high Steaming,


temperature temperature temperature acid leaching
H2O, RT H2O, RT H2O, RT Calcination

Framework Al
J A L K
High High
temperature, temperature,
Base, vacuum vacuum vacuum

Framework AI expanding coordination Framework AI with BAS Framework LAS Distorted framework AI

Framework-associated Al H2O, RT H2O, RT


E F

High
temperature,
vacuum

Framework-associated AI: octahedral NH4+ NH3 Framework-associated AI: octahedral


exchange adsorption
H2O, RT G H2O, RT

High temperature, vacuum


>383 K –H2O

Framework AI

I H

High temperature, vacuum

Framework-associated AI: tetrahedral Framework-associated LAS

Fig. 2 | A summary of the current understanding of Lewis acidic aluminium in zeolites. Pathways showing the changes in aluminium structure with
changing conditions and classification of Lewis acidic aluminium as originating from extra-framework27,36–38,89, framework-associated23,27,30,50 and framework
aluminium20,30,34,63. All structures in the figure except for structure K have been reported in FAU. Structure K has been reported in chabazite20. Many
transformations sketched above have been observed in zeolites beyond FAU as well. For the specific zeolite types, refer to the references cited above. LAS,
Lewis acid site; RT, room temperature.

We demonstrated with experiments on zeolite mordenite (MOR) the zeolite, is associated with Lewis acidic character when exam-
that framework-associated aluminium, which adopts an octahe- ined after high-temperature dehydration (structure H in Fig. 2)23.
dral coordination under ambient conditions in the proton form of A linear correlation between the number of Lewis acid sites and

1050 Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials


NAture MAterIALS Perspective
Box 2 | Factors regulating the Lewis acidity of EFAl

EFAl in zeolites can be generated by different methods, such as are observed in Al-SBA-3 synthesized via a post-synthesis
high-temperature calcination, steaming, acid treatment or other modification of Si-SBA-3 using sodium aluminate. Analysis of the
post-synthesis approaches. The so-formed EFAl comprises a num- strengths of these sites is realized by evacuating pyridine at different
ber of different aluminium species, all of which are completely dis- temperatures96. Samples produced via a post-synthetic alumination
lodged from the framework. However, not all of these species are procedure of TUD-1 with a higher amount of EFAl does not result
Lewis acidic. By curating examples from published literature, we in a greater number of Lewis acid sites97. While incorporating
illustrate three factors that regulate the Lewis acidic property of more aluminium through post-synthesis modification results in
EFAl. The qualitative dependence of the amount of EFAl on these more five- and six-coordinated EFAl in TUD-1, the total number
factors is shown in the graphs in the figure. of Lewis acid sites decreases97. Therefore, the post-synthesis
modification of TUD-1 using aluminium isopropoxide generates
Si/Al ratio. The Si/Al ratio influences the properties of the non-Lewis acidic EFAl at high aluminium loadings. Modulating
aluminium species displaced from the framework upon steaming. the steaming temperature influences the extent to which neutral
For instance, steaming of a low-aluminium H-ZSM-5 (Si/Al = 45) or cationic aluminium species are produced43,98, with the
results in EFAl with weak acidity and no strong Lewis acid sites92. probability of generating extra-framework species with a high
In contrast, steaming of H-ZSM-5 with greater aluminium content cationic charge decreasing with the severity of the steaming
(Si/Al = 18) generates Lewis acid sites that are strong enough to treatment62. This apart, the structure of extra-framework species
polarize paraffin molecules40. These sites are responsible for the evolves during acid leaching. Aluminium species characterized
dehydrogenation step in the cracking of alkanes93. In more recently by a signal at about 30 ppm in the 27Al MAS NMR spectrum is
discovered mesoporous aluminosilicates, such as AlTUD-1, there the most vulnerable to acid treatment. While the population of
are several different types of EFAl and Lewis acid sites. While this species diminishes strongly, the peak shape and intensity
the extra-framework species include penta- and octahedrally of the feature centred at a chemical shift of around 0 ppm also
coordinated aluminium, FTIR spectroscopy of adsorbed pyridine change markedly44.
does not always reveal the multiplicity in Lewis acid sites94.
For AlTUD-1 with a Si/Al ratio of 25, the infrared feature for Nature of co-cation. High-temperature treatment of ammonium
Lewis-bound pyridine is sharp, while for samples with a Si/Al ratio and proton forms of zeolites with subsequent exposure to ambient
in the range of 50 to 75, the feature is much broader and manifests conditions generally results in a substantial fraction of aluminium
in the form of two partially overlapping peaks94. species adopting an octahedral coordination48. However, such
octahedrally coordinated EFAl is usually not observed with zeolites
Method used for dealumination and post-synthesis treatment. having sodium or potassium as the charge-compensating cation30,99.
EFAl generated by steaming cannot be removed even after
multiple ammonium-exchange treatments24, but dealumination
performed with silicon tetrachloride results in aluminium species Steaming
Si/Al ratio temperature Co-cation
that can be removed by ammonium exchange95. Dealuminating
zeolite BEA by steaming yields a single kind of Lewis acid site
characterized by an infrared feature at 1,456 cm–1 upon interaction
EFAI

EFAI

EFAI
with pyridine, but dealumination using hydrochloric acid or
ammonium hexafluorosilicate yields a second kind of Lewis acid
site, which shows an additional infrared feature at 1,446 cm–1
(ref. 26). Similarly, two kinds of Lewis acid site bound to pyridine Si/Al Temperature Na+, Cs+ NH4+, H+

the amount of framework-associated octahedral aluminium was removal shows absorption bands centred at 3,640 cm–1 and 3,545
attained with two probe molecules of contrasting size and strength: cm–1, assigned to bridging hydroxyl groups in the super cages and
carbon monoxide and pyridine. The amount of octahedral alumin- sodalite cages, respectively54. This sample has all of its aluminium
ium, and consequently, the number of Lewis acid sites, decreases in a framework tetrahedral coordination (boxed structure A in
with increasing Si/Al ratio of the zeolite. The aluminium content Fig. 2). Calcination to the proton form and subsequent exposure
also influences the plurality of Lewis acid sites23. The Lewis acid to ambient conditions transforms a fraction of aluminium into the
character of three-coordinate aluminium sites in protonic zeolites octahedral coordination (structures E and F in Fig. 2). On evacua-
may be masked when the local topology permits such aluminium tion of such HY under vacuum, the infrared features attributed to
to expand its coordination through an extra bond to the framework Brønsted acidity diminish strongly54. Therefore, the generation of
oxygen atom53. While edingtonite is illustrative of a zeolite topol- Lewis acidic framework-associated aluminium is concomitant with
ogy that allows masking to occur rather easily, other zeolites such as the degeneration of Brønsted acid sites, implying a conversion of
beta (BEA), Zeolite Socony Mobil–5 (MFI) and MOR, do not offer a Brønsted to Lewis acid sites.
favourable environment for under-coordinated aluminium to bind
with the aluminosilicate-framework oxygen atom53. Distinguishing framework-associated aluminium from EFAl. The
Zeolite SSZ-33, which possesses octahedral framework-associated concept of octahedral aluminium insertion into a framework
species, shows a prominent infrared signature for Lewis acid coordination was first illustrated with zeolite BEA48 and has been
sites when probed by pyridine30. On average, there is more than extended to many different zeolites since then23,30,50. Distinguishing
one Lewis acid site per unit cell of SSZ-33. This corroborates the framework-associated aluminium from EFAl, with both of them
assignment of framework-associated aluminium to Lewis acidity. being able to adopt an octahedral coordination, can be challeng-
Also, the evolution of the structure of framework-associated alu- ing. Herein, we highlight specific examples where such a rigorous
minium can be monitored through changes in Brønsted acidity. The distinction has been illustrated. As a first example, consider the
infrared spectrum of NH4–Y after heat treatment and ammonia evolution of aluminium structure in SSZ-33 studied by 27Al MAS

Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials 1051


Perspective NAture MAterIALS

NMR30. Calcining the zeolite to the proton form results in a sub- aluminium atom that is part of the framework is not under-
stantial number of octahedral sites under ambient conditions30, coordinated and cannot be intuitively envisioned as a Lewis acid
with the appearance of a sharp and a broad feature in the range of site. However, multiple reports pointing to framework-based Lewis
0 to −20 ppm. Converting this sample into the ammonium form acidity exist20,30,34. The origin of these acid sites can be framework
virtually eliminates all octahedral species30. The enhanced inten- aluminium or defects in the framework, including those created by
sity of the peak associated with tetrahedral aluminium suggests dehydroxylation of bridging hydroxyl groups63.
that most of the octahedral sites were transformed into framework Aluminium ions can take several coordination environments,
aluminium. On calcining this sample and subsequently exposing it ranging from tetrahedral and penta-coordination to octahedral
to ambient conditions, the sharp octahedral feature at 0 ppm reap- geometry. It is hypothesized that framework tetrahedral aluminium
pears30. Thus, this sharp feature represents aluminium species that can act as a Lewis acid site by expanding its coordination to five
undergoes reversible tetrahedral–octahedral coordination, which is on interaction with a base such as pyridine (structure J in Fig. 2)34.
the inherent feature of framework-associated aluminium. Likewise, While this is yet to be experimentally established, the model is built
it has been demonstrated that octahedral aluminium species char- on the premise that the access of a basic molecule to a framework
acterized by a sharp resonance at 0 ppm in zeolite BEA represents aluminium atom is hindered by the strongly competitive interaction
flexible lattice aluminium that can be forced back to a tetrahedral of the base with a nearby proton21,34. However, the interaction of
environment49. Therefore, the fundamental criteria to distinguish the base with framework aluminium can be facilitated if the bridg-
framework-associated aluminium from extra-framework species ing OH group is in one cavity and the base can attack aluminium
must be based on reversible octahedral–tetrahedral coordination, from the opposite side. For example, pyridine can probe aluminium
where the restored tetrahedral structure is comparable to that of through the supercage if the OH group is localized in the sodalite
pristine tetrahedral framework aluminium. While such reversibil- cage in zeolite Y34.
ity in coordination has been mostly demonstrated through NMR, Lewis acidity can also arise from perturbed or distorted frame-
X-ray photoelectron spectroscopy has independently corroborated work aluminium. In ferrierite and chabazite zeolites, perturbed
this phenomenon55. Tetrahedral, octahedral and three-coordinated framework aluminium species are proposed to be (SiO)3AlOH
aluminium have been identified in zeolite BEA using X-ray pho- groups (structure K in Fig. 2), which result in the broadening
toelectron spectroscopy56,57. Furthermore, in situ X-ray absorption of the 27Al MAS NMR line shape at around 40 ppm (ref. 20).
spectroscopy (XAS) reveals that octahedral framework-associated These species dehydroxylate at relatively mild temperatures
aluminium changes into a tetrahedral geometry on heating the zeo- yielding three-coordinated Lewis acidic aluminium (structure L
lite to temperatures above 383 K (structure I in Fig. 2)50,58. in Fig. 2)20.
High-temperature calcination of zeolite Y yields framework-
associated aluminium, which is three-coordinated to the framework Studying Lewis–Brønsted synergy in zeolites. There are many dif-
and coordinated to three water molecules (structure F in Fig. 2)59,60. ferent possible origins for Lewis acidic aluminium in zeolites, and
The adsorption of ammonia can convert this octahedral species to as elaborated above, it is necessary to be scrupulous to differenti-
a tetrahedral geometry (structure G in Fig. 2)27. However, increas- ate between the different types of Lewis acid sites. This is relevant
ing the severity of the thermal treatment results in the formation of not only for reactions that are catalysed only by Lewis acid sites
octahedral extra-framework Al(OH)3 species (equations (1)–(3)), but also for applications that exploit the synergy between Lewis
whose coordination cannot be reversed by ammonia adsorption61. and Brønsted acid sites, which can enhance catalytic properties in,
Likewise, reversible aluminium coordination changes can also be for example, dealuminated zeolites64. The enhanced activity in the
realized in steamed zeolites. The use of a higher pressure of water conversion of hydrocarbons was initially attributed to several fac-
vapour during steaming degrades more than 75% of the bridging tors, namely the stabilization of the negative charge on the lattice
hydroxyl groups, but ammonia treatment restores a substantial frac- due to the EFAl species65, the polarization of reactants by Lewis
tion of these acid sites62. While steaming is traditionally viewed as a acid sites40 and the confinement effects of zeolite pores66. At pres-
technique to generate genuinely EFAl, some of the octahedral alu- ent, the working hypothesis is that the synergy between Brønsted
minium, even in a strongly steamed zeolite can be inserted back and Lewis acid sites is due to the close location of both sites in
into the framework. Therefore, counting all octahedral aluminium the zeolite pore27,67,68. The proximity of Brønsted and Lewis acid
as EFAl is erroneous. sites is experimentally confirmed in qualitative terms, primarily
through two-dimensional double-quantum MAS NMR spectros-
Al3þ þ H2 O þ ðSi2O2AlÞ� ! AlOH2þ þ Si2OðHÞ2Al ð1Þ copy. However, more precise information on the location of dif-
ferent EFAl species (structures B, C, M and N in Fig. 2) has not
been established, which is highly desirable from a fundamental and
AlOH2þ þ H2 O þ ðSi2O2AlÞ� ! AlðOHÞþ
2 þ Si2OðHÞ2Al a practical point of view. Furthermore, the challenges in demar-
ð2Þ cating truly EFAl species from framework-associated aluminium
sites and three-coordinated aluminium with adsorbed water, all of
which take up octahedral geometry, severely hampers the analysis
AlðOHÞþ �
2 þ H2 O þ ðSi2O2AlÞ ! AlðOHÞ3 þ Si2OðHÞ2Al of Lewis–Brønsted synergy41,60.
ð3Þ
Conclusions and outlook
The term ‘extra-framework aluminium’ has often been used in A close analysis of the assortment of proposals for Lewis acidic
a way that does not distinguish species entirely dislodged from the aluminium in zeolites uncovers fundamental differences. While
framework from partially dislodged framework-associated species. attributing Lewis acidic property to EFAl is a widely accepted con-
Consequently, apprehension of trends for the number of Lewis acid jecture, the literature data reveal that in many cases a correlation
sites (Box 1) is not straightforward. This calls for a more judicious between the amount of EFAl and the number of Lewis acid sites
usage of the terms ‘extra-framework’ and ‘framework associated’ does not hold. From this, we identify inconsistencies in the usage
aluminium. of the term ‘extra-framework aluminium’, and note that distinguish-
ing framework-associated aluminium from EFAl is paramount.
Lewis acidity of framework aluminium. Lewis acidity originating The properties of different aluminium species are liable to change
from framework aluminium is poorly understood. A tetrahedral on sample treatment. Publications that discuss how the relative

1052 Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials


NAture MAterIALS Perspective
Structure under catalytically Lewis and Brønsted sites: Relative strength and location Rational synthesis of Lewis acid sites
relevant conditions interconversion and synergy of Lewis acid sites

NMR Si Si Si
Reactant Product Al Si Al Al Al Si
FTIR
– – – Si Si Si Si Si Al
– Solvent – Si Si Si

In situ and operando Probe molecules of Controlling aluminium distribution


characterization: NMR and FTIR different sizes and basicity and thermal treatment
FTIR, NMR and XAS Probe molecules desorbed Theory under relevant conditions
at different temperatures and dynamics

Study evolution of aluminium structure from precursor species to aluminium Lewis site

In situ XRD
and XAS. Framework Al with BAS
FTIR and
Temperature

NMR

Ad
at

Co ate

Co bas
hy

so
He

Under-coordinated AI

ol

ol
dr

rb
Framework Al EFAl

e
Framework-associated Al

Fig. 3 | Future research directions on Lewis acidic aluminium in zeolites. A summary of the expected developments, featuring a list of the most pressing
questions that need to be answered and the techniques that could be used to address them. BAS, Brønsted acid site.

quantities of framework-associated and EFAl change on sample analysis of the mechanism of interconversion of extra-framework
treatment are scarce, with the understanding of the changes that and framework-associated aluminium sites.
occur in the number, location and structure of Lewis acidic alumin-
ium on changing synthesis conditions remaining abstruse. Interconversion and synergy between Lewis and Brønsted acid
sites. The precise characterization of diverse aluminium Lewis acid
Structure under catalytically relevant conditions. While the sites, as well as their proximity and interaction with Brønsted acid
chemical structures of Lewis acidic aluminium species have been sites and the zeolite framework, are of primary importance. While
reported in both dehydrated environments and under ambient 27
Al MAS NMR and FTIR spectroscopy can estimate the total num-
conditions, there is a pressing need to study aluminium structure ber of Lewis acid sites21,23,30,72, they are barely sensitive to the strength
under relevant reaction conditions through in situ and operando or structure of the acid sites. Furthermore, both these techniques
characterization (Fig. 3). In this context, in situ XAS and X-ray dif- are not entirely robust under all conditions. For instance, 27Al MAS
fraction (XRD) are two techniques that can track the evolution of NMR is challenging to execute under dehydrated conditions, where
aluminium structure and distinguish between species that are par- highly distorted aluminium local environments occur and are diffi-
tially or entirely dislodged from the framework. XAS can reveal cult to detect. In this context, measuring NMR on other nuclei, such
the presence of tetrahedral, octahedral and under-coordinated as 1H and 29Si, including two-dimensional correlation spectros-
aluminium during all stages of the zeolite treatment58,60, while copy, can shed light on the local aluminium environment. Another
Rietveld refinement of XRD patterns can identify migration of rapidly developing branch of NMR spectroscopy is the analysis of
aluminium to extra-framework positions and reveal occupancy adsorbed probe molecules through 13C, 15N and 31P nuclei, which
of extra-framework cations as a function of temperature42. The directly accesses the acidity of aluminium centres. 27Al–31P or
removal of aluminium from the framework should be reflected by a 27
Al–15N rotational-echo, double-resonance (REDOR) and transfer
decrease in the d-spacing, but since framework-associated alumin- of population in double-resonance (TRAPDOR) experiments can
ium is only partially dislodged from the framework, it is expected be used to identify and quantify Brønsted and Lewis acid sites when
to show a slightly different decrease. However, the difference can trimethylphosphine and pyridine are used as probe molecules,
be very small69,70, making detection and interpretation challenging. respectively73,74. Recently, low-temperature 15N dynamic nuclear
Correlations between the number of Lewis acid sites and activ- polarization surface-enhanced NMR spectroscopy of adsorbed
ity in a reaction are well established, but a deeper understanding of pyridine revealed the plurality of acid sites in ɣ-alumina and has a
the effect of active site structure is missing. Admittedly, elaboration high potential for studying zeolites75. REDOR may also be used to
of this structure-to-performance relationship requires advanced measure Al–N bond distances to evaluate the binding of the probe
characterization methods that enable in situ characterization. MAS molecule to Lewis acid sites76. Likewise, insensitive nuclei enhance-
NMR spectroscopy with double and triple quantum resonance ment by polarization transfer (INEPT) is another two-dimensional
enables the study of proximity and anisotropy of aluminium sites correlation experiment that helps elucidate the coordination num-
that are associated with Brønsted and Lewis acidity71. Furthermore, ber of Lewis acid sites77.
due to the high symmetry of aluminium sites in a solvent-containing
environment, which leads to substantial narrowing of NMR signals, Relative strength and location of Lewis acid sites. Infrared spec-
the use of these techniques can be naturally extended to study active troscopy of adsorbed probe molecules remains an extensively
sites under ‘working’ reaction conditions. This can aid in better used technique to study acid sites in zeolites. However, pyridine,

Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials 1053


Perspective NAture MAterIALS

which is an often-used probe molecule, is a strong and unselec- 4. Martínez, A., Prieto, G., García‐Trenco, A. & Peris, E. in Zeolites and
tive base, making the differentiation of Lewis acid sites of similar Catalysis: Synthesis, Reactions and Applications (eds Čejka, J., Corma, A. &
Zones, S.) 649–685 (Wiley, 2010).
acidity impossible. In contrast, weaker bases, such as nitrogen and 5. Vogt, E. T. C. & Weckhuysen, B. M. Fluid catalytic cracking: recent
carbon monoxide, are capable of differentiating multiple Lewis developments on the grand old lady of zeolite catalysis. Chem. Soc. Rev. 44,
acid sites, including three-coordinated and octahedral aluminium 7342–7370 (2015).
species28,78,79. Tin and zirconium sites in zeolites selectively inter- 6. Tian, P., Wei, Y., Ye, M. & Liu, Z. Methanol to olefins (MTO): from
fundamentals to commercialization. ACS Catal. 5, 1922–1938 (2015).
act with acetonitrile, and likewise, aluminium species can be dif-
7. Bjørgen, M. et al. Methanol to gasoline over zeolite H-ZSM-5: improved
ferentiated using FTIR spectroscopy of adsorbed nitriles80. The catalyst performance by treatment with NaOH. Appl. Catal. A 345,
use of nitriles with alkyl groups of varying basicity and size, such 43–50 (2008).
as acetonitrile, pivalonitrile and benzonitrile, makes it plausible to 8. Minachev, K. M., Garanin, V. I., Kharlamov, V. V., Isakova, T. A. & Senderov,
assess the relative strength and location of aluminium Lewis acid E. E. Catalytic properties of synthetic mordenite in the isomerization,
hydrogenation, and hydroisomerization of certain hydrocarbons. Russ. Chem.
sites (Fig. 3). Similar insights can be achieved by combining infra- Bull. 18, 1611–1615 (1969).
red spectroscopy with temperature-programmed desorption of 9. Jacobs, P. A., Dusselier, M. & Sels, B. F. Will zeolite‐based catalysis be as
probe molecules81. Likewise, ultraviolet–visible diffuse reflectance relevant in future biorefineries as in crude oil refineries? Angew. Chem. Int.
spectroscopy of zeolites adsorbed with molecules, such as naphtha- Ed. 53, 8621–8626 (2014).
lene82, is another tool to study acid sites83. However, this technique 10. Moreno-Recio, M., Santamaría-González, J. & Maireles-Torres, P. Brönsted
and Lewis acid ZSM-5 zeolites for the catalytic dehydration of glucose into
has not been able to reveal the origin of Lewis acid sites, or deter- 5-hydroxymethylfurfural. Chem. Eng. J. 303, 22–30 (2016).
mine whether these sites are inherent to the zeolite or are genuinely 11. Dapsens, P. Y., Mondelli, C. & Pérez-Ramírez, J. Design of Lewis-acid centres
extra-framework82. in zeolitic matrices for the conversion of renewables. Chem. Soc. Rev. 44,
A special emphasis should be placed on the correlation of data 7025–7043 (2015).
obtained using infrared spectroscopy with catalytic data and NMR, 12. Otomo, R., Yokoi, T., Kondo, J. N. & Tatsumi, T. Dealuminated Beta zeolite as
effective bifunctional catalyst for direct transformation of glucose to
as the NMR data are typically obtained for air-exposed samples, 5-hydroxymethylfurfural. Appl. Catal. A 470, 318–326 (2014).
while infrared measurements are done after complete dehydration 13. Suganuma, S., Hisazumi, T., Taruya, K., Tsuji, E. & Katada, N. Influence of
in vacuum. This discrepancy should be mitigated, by, for instance, acidic property on catalytic activity and selectivity in dehydration of glycerol.
measuring multiple-quantum MAS NMR, or related techniques, ChemistrySelect 2, 5524–5531 (2017).
for dehydrated samples84, or infrared spectra for materials with 14. Lewis, J. D., Van de Vyver, S. & Román‐Leshkov, Y. Acid–base pairs in Lewis
acidic zeolites promote direct aldol reactions by soft enolization. Angew.
pre-adsorbed water23. Such experiments bridge the gap between Chem. Int. Ed. 127, 9973–9976 (2015).
Lewis acid sites present under reaction conditions and those 15. Perego, C., Carati, A., Ingallina, P., Mantegazza, M. A. & Bellussi, G.
detected by conventional spectroscopic methods under starkly dif- Production of titanium containing molecular sieves and their application in
ferent conditions. NMR and FTIR spectroscopy are largely focused catalysis. Appl. Catal. A 221, 63–72 (2001).
on studying the final structure of Lewis acidic aluminium bonded 16. Hammond, C., Conrad, S. & Hermans, I. Simple and scalable preparation of
highly active Lewis acidic Sn‐β. Angew. Chem. Int. Ed. 51, 11736–11739 (2012).
to a basic probe molecule. Therefore, one of the limitations of these 17. Sushkevich, V. L., Ivanova, I. I. & Taarning, E. Ethanol conversion into
techniques is that it only provides partial information, often of the butadiene over Zr-containing molecular sieves doped with silver. Green.
initial and final structures, while being largely blind to the evolution Chem. 17, 2552–2559 (2015).
in aluminium structure during sample pre-treatment. Techniques 18. Astorino, E., Peri, J. B., Willey, R. J. & Busca, G. Spectroscopic characterization
of silicalite-1 and titanium silicalite-1. J. Catal. 157, 482–500 (1995).
that are more robust and can be used in an in situ, time-resolved
19. Corma, A., Domine, M. E., Nemeth, L. & Valencia, S. Al-free Sn-beta zeolite
fashion need to be harnessed. as a catalyst for the selective reduction of carbonyl compounds (Meerwein–
Ponndorf–Verley reaction). J. Am. Chem. Soc. 124, 3194–3195 (2002).
Rational synthesis of Lewis acid sites. Steaming at high tem- 20. Brus, J. et al. Structure of framework aluminum Lewis sites and perturbed
perature was one of the first methods used to generate Lewis acid aluminum atoms in zeolites as determined by 27Al{1H} REDOR (3Q) MAS
NMR spectroscopy and DFT/molecular mechanics. Angew. Chem. Int. Ed. 54,
sites in aluminium-containing zeolites, with later on other de- and 541–545 (2015).
re-alumination methods23,30,85–87 being used. These are all simple and 21. Busca, G. Acidity and basicity of zeolites: a fundamental approach.
inexpensive procedures, which can be easily implemented for zeo- Microporous Mesoporous Mater. 254, 3–16 (2017).
lites of any topology and Si/Al ratio. The formation of Lewis acid 22. Zhang, Y. et al. Promotion of protolytic pentane conversion on H-MFI zeolite
sites of a certain strength and specific location in the pore struc- by proximity of extra-framework aluminum oxide and Brønsted acid sites.
J. Catal. 370, 424–433 (2019).
ture of a zeolite is governed by the distribution of aluminium in the 23. Ravi, M., Sushkevich, V. L. & van Bokhoven, J. A. Lewis acidity
framework. Therefore, the control of aluminium distribution, mor- inherent to the framework of zeolite mordenite. J. Phys. Chem. C. 123,
phology and presence of defects in the parent zeolite88 dictates the 15139–15144 (2019).
quality of the final Lewis acid sites (Fig. 3). Headway in the rational 24. Klinowski, J., Fyfe, C. A. & Gobbi, G. C. High-resolution solid-state nuclear
synthesis of Lewis acidic zeolites, complemented by a methodical magnetic resonance studies of dealuminated zeolite Y. J. Chem. Soc. Faraday
Trans. 1 81, 3003–3019 (1985).
characterization of the active sites, will help to resolve unanswered 25. Bevilacqua, M. & Busca, G. A study of the localization and accessibility of
aspects of catalytic applications that exploit Lewis acid sites in zeo- Brønsted and Lewis acid sites of H-mordenite through the FT-IR spectroscopy
lites. Knowledge of structural proximity and structure under reac- of adsorbed branched nitriles. Catal. Commun. 3, 497–502 (2002).
tion conditions will greatly contribute to understanding the role of 26. Marques, J. P. et al. Infrared spectroscopic study of the acid properties of
Lewis acid sites in catalysis by zeolites. dealuminated BEA zeolites. Microporous Mesoporous Mater. 60, 251–262 (2003).
27. Yu, Z. et al. Insights into the dealumination of zeolite HY revealed by
sensitivity‐enhanced 27Al DQ‐MAS NMR spectroscopy at high field. Angew.
Received: 22 October 2019; Accepted: 29 June 2020; Chem. Int. Ed. 49, 8657–8661 (2010).
Published online: 21 September 2020 28. Catana, G., Baetens, D., Mommaerts, T., Schoonheydt, R. A. & Weckhuysen,
B. M. Relating structure and chemical composition with lewis acidity in
References zeolites: a spectroscopic study with probe molecules. J. Phys. Chem. B 105,
1. Barrer, R. M. Hydrothermal Chemistry of Zeolites Vol. 269 (Academic Press, 4904–4911 (2001).
1982). 29. Woolery, G. L., Kuehl, G. H., Timken, H. C., Chester, A. W. & Vartuli, J. C.
2. Fyfe, C. A., Gobbi, G. C., Klinowski, J., Thomas, J. M. & Ramdas, S. Resolving On the nature of framework Brønsted and Lewis acid sites in ZSM-5. Zeolites
crystallographically distinct tetrahedral sites in silicalite and ZSM-5 by 19, 288–296 (1997).
solid-state NMR. Nature 296, 530–533 (1982). 30. Gil, B., Zones, S. I., Hwang, S.-J., Bejblová, M. & Čejka, J. Acidic properties of
3. Haag, W. O., Lago, R. M. & Weisz, P. B. The active site of acidic SSZ-33 and SSZ-35 novel zeolites: a complex infrared and MAS NMR study.
aluminosilicate catalysts. Nature 309, 589–591 (1984). J. Phys. Chem. C. 112, 2997–3007 (2008).

1054 Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials


NAture MAterIALS Perspective
31. Jacobs, P. A. & Beyer, H. K. Evidence for the nature of true Lewis sites in 59. Altwasser, S., Jiao, J., Steuernagel, S., Weitkamp, J. & Hunger, M. in Studies in
faujasite-type zeolites. J. Phys. Chem. 83, 1174–1177 (1979). Surface Science and Catalysis Vol. 154 (eds van Steen, E., Claeys, M. &
32. Lohse, U., Löffler, E., Hunger, M., Stöckner, J. & Patzelova, V. Hydroxyl Callanan, L. H.) 3098–3105 (Elsevier, 2004).
groups of the non-framework aluminium species in dealuminated Y zeolites. 60. van Bokhoven, J. A., Van der Eerden, A. M. J. & Koningsberger, D. C.
Zeolites 7, 11–13 (1987). Three-coordinate aluminum in zeolites observed with in situ X-ray
33. Weitkamp, J. & Puppe, L. Catalysis and Zeolites: Fundamentals and absorption near-edge spectroscopy at the Al K-edge: flexibility
Applications (Springer Science & Business Media, 2013). of aluminum coordinations in zeolites. J. Am. Chem. Soc. 125,
34. Phung, T. K. & Busca, G. On the Lewis acidity of protonic zeolites. 7435–7442 (2003).
Appl. Catal. A 504, 151–157 (2015). 61. Rybakov, A. A., Larin, A. V. & Zhidomirov, G. M. Influence of alkali cations
35. Derouane, E. G. et al. The acidity of zeolites: concepts, measurements and on the inter-conversion of extra-framework aluminium species in
relation to catalysis: a review on experimental and theoretical methods for dealuminated zeolites. Microporous Mesoporous Mater. 189, 173–180 (2014).
the study of zeolite acidity. Catal. Rev. 55, 454–515 (2013). 62. Jiao, J., Altwasser, S., Wang, W., Weitkamp, J. & Hunger, M. State of
36. Shannon, R. D. et al. The nature of the nonframework aluminum species aluminum in dealuminated, nonhydrated zeolites Y investigated by
formed during the dehydroxylation of HY. J. Phys. Chem. 89, 4778–4788 (1985). multinuclear solid-state NMR spectroscopy. J. Phys. Chem. B 108,
37. Bhering, D. L., Ramírez-Solís, A. & Mota, C. J. A. A density functional theory 14305–14310 (2004).
based approach to extraframework aluminum species in zeolites. J. Phys. 63. Ward, J. W. The nature of active sites on zeolites: I. The decationated Y
Chem. B 107, 4342–4347 (2003). zeolite. J. Catal. 9, 225–236 (1967).
38. Li, S. et al. Brønsted/Lewis acid synergy in dealuminated HY zeolite: a 64. Mirodatos, C. & Barthomeuf, D. Superacid sites in zeolites. J. Chem. Soc.
combined solid-state NMR and theoretical calculation study. J. Am. Chem. Chem. Commun. 39–40 (1981).
Soc. 129, 11161–11171 (2007). 65. Lunsford, J. H. Surface interactions of NaY and decationated Y zeolites with
39. Dimitrijevic, R., Lutz, W. & Ritzmann, A. Hydrothermal stability of zeolites: nitric oxide as determined by electron paramagnetic resonance spectroscopy.
determination of extra-framework species of HY faujasite-type steamed J. Phys. Chem. 72, 4163–4168 (1968).
zeolite. J. Phys. Chem. Solids 67, 1741–1748 (2006). 66. van Bokhoven, J. A. et al. An explanation for the enhanced activity for light
40. Zholobenko, V. L. et al. On the possible nature of sites responsible for the alkane conversion in mildly steam dealuminated mordenite: the dominant
enhancement of cracking activity of HZSM-5 zeolites dealuminated under role of adsorption. J. Catal. 202, 129–140 (2001).
mild steaming conditions. Zeolites 10, 304–306 (1990). 67. Yu, Z. et al. Brønsted/lewis acid synergy in H-ZSM-5 and H–MOR zeolites
41. van Bokhoven, J. A., Koningsberger, D. C., Kunkeler, P., Van Bekkum, H. & studied by 1H and 27Al DQ-MAS solid-state NMR spectroscopy. J. Phys.
Kentgens, A. P. M. Stepwise dealumination of zeolite βeta at specific T-sites Chem. C. 115, 22320–22327 (2011).
observed with 27Al MAS and 27Al MQ MAS NMR. J. Am. Chem. Soc. 122, 68. Li, S. et al. Extra-framework aluminium species in hydrated faujasite zeolite
12842–12847 (2000). as investigated by two-dimensional solid-state NMR spectroscopy and
42. Agostini, G. et al. In situ XAS and XRPD parametric Rietveld refinement to theoretical calculations. Phys. Chem. Chem. Phys. 12, 3895–3903 (2010).
understand dealumination of Y zeolite catalyst. J. Am. Chem. Soc. 132, 69. Dzwigaj, S. & Che, M. Incorporation of Co(ii) in dealuminated BEA zeolite at
667–678 (2009). lattice tetrahedral sites evidenced by XRD, FTIR, diffuse reflectance UV–vis,
43. Macedo, A., Raatz, F., Boulet, R., Janin, A. & Lavalley, J. C. in Studies in Surface EPR, and TPR. J. Phys. Chem. B 110, 12490–12493 (2006).
Science and Catalysis Vol. 37 (eds Grobet, P. J. et al.) 375–383 (Elsevier, 1988). 70. Hajjar, R., Millot, Y., Man, P. P., Che, M. & Dzwigaj, S. Two kinds of
44. Menezes, S. M. C. et al. Characterization of extra-framework species of framework Al sites studied in BEA zeolite by X-ray diffraction, Fourier
steamed and acid washed faujasite by MQMAS NMR and IR measurements. transform infrared spectroscopy, NMR techniques, and V probe. J. Phys.
Appl. Catal. A 207, 367–377 (2001). Chem. C. 112, 20167–20175 (2008).
45. Yi, X. et al. Origin and structural characteristics of tri-coordinated 71. Zheng, A., Li, S., Liu, S.-B. & Deng, F. Acidic properties and structure–
extra-framework aluminum species in dealuminated zeolites. J. Am. Chem. activity correlations of solid acid catalysts revealed by solid-state nmr
Soc. 140, 10764–10774 (2018). spectroscopy. Acc. Chem. Res. 49, 655–663 (2016).
46. Liu, C., Li, G., Hensen, E. J. M. & Pidko, E. A. Nature and catalytic role of 72. Farneth, W. E. & Gorte, R. J. Methods for characterizing zeolite acidity.
extraframework aluminum in faujasite zeolite: a theoretical perspective. Chem. Rev. 95, 615–635 (1995).
ACS Catal. 5, 7024–7033 (2015). 73. Kao, H.-M. & Grey, C. P. Probing the Brønsted and Lewis acidity of zeolite
47. Li, G. & Pidko, E. A. The nature and catalytic function of cation sites in HY: a 1H/27Al and 15N/27Al TRAPDOR NMR study of monomethylamine
zeolites: a computational perspective. ChemCatChem 11, 134–156 (2019). adsorbed on HY. J. Phys. Chem. 100, 5105–5117 (1996).
48. Bourgeat-Lami, E. et al. Study of the state of aluminium in zeolite-β. Appl. 74. Kao, H.-M., Liu, H., Jiang, J.-C., Lin, S.-H. & Grey, C. P. Determining the
Catal. 72, 139–152 (1991). structure of trimethylphosphine bound to the Brønsted acid site in zeolite
49. Haouas, M., Kogelbauer, A. & Prins, R. The effect of flexible lattice HY: double-resonance NMR and ab initio studies. J. Phys. Chem. B 104,
aluminium in zeolite beta during the nitration of toluene with nitric acid and 4923–4933 (2000).
acetic anhydride. Catal. Lett. 70, 61–65 (2000). 75. Moroz, I. B., Larmier, K., Liao, W.-C. & Copéret, C. Discerning γ-alumina
50. Omegna, A., van Bokhoven, J. A. & Prins, R. Flexible aluminum coordination surface sites with nitrogen-15 dynamic nuclear polarization surface
in alumino–silicates. Structure of zeolite H–USY and amorphous silica– enhanced NMR spectroscopy of adsorbed pyridine. J. Phys. Chem. C. 122,
alumina. J. Phys. Chem. B 107, 8854–8860 (2003). 10871–10882 (2018).
51. Wouters, B. H., Chen, T. H. & Grobet, P. J. Reversible tetrahedral–octahedral 76. Grey, C. P. & Kumar, B. S. A. 15N/27Al double resonance NMR study of
framework aluminum transformation in zeolite Y. J. Am. Chem. Soc. 120, monomethylamine adsorbed on zeolite HY. J. Am. Chem. Soc. 117,
11419–11425 (1998). 9071–9072 (1995).
52. Abraham, A. et al. Influence of framework silicon to aluminium ratio on 77. Kao, H.-M. & Grey, C. P. Determination of the 31P–27Al J-coupling constant
aluminium coordination and distribution in zeolite Beta investigated by 27Al for trimethylphosphine bound to the Lewis acid site of zeolite HY. J. Am.
MAS and 27Al MQ MAS NMR. Phys. Chem. Chem. Phys. 6, 3031–3036 (2004). Chem. Soc. 119, 627–628 (1997).
53. Busco, C., Bolis, V. & Ugliengo, P. Masked Lewis sites in proton-exchanged 78. Hadjiivanov, K. I. & Vayssilov, G. N. Characterization of oxide surfaces and
zeolites: a computational and microcalorimetric investigation. J. Phys. Chem. zeolites by carbon monoxide as an IR probe molecule. Adv. Catal. 47,
C. 111, 5561–5567 (2007). 307–511 (2002).
54. Xu, B., Rotunno, F., Bordiga, S., Prins, R. & van Bokhoven, J. A. Reversibility 79. Wischert, R., Copéret, C., Delbecq, F. & Sautet, P. Dinitrogen: a selective
of structural collapse in zeolite Y: alkane cracking and characterization. probe for tri-coordinate Al “defect” sites on alumina. Chem. Commun. 47,
J. Catal. 241, 66–73 (2006). 4890–4892 (2011).
55. Remy, M. J., Genet, M. J., Poncelet, G., Lardinois, P. F. & Notté, P. P. 80. Boronat, M., Concepción, P., Corma, A., Renz, M. & Valencia, S.
Investigation of dealuminated mordenites by X-ray photoelectron Determination of the catalytically active oxidation Lewis acid sites in Sn-beta
spectroscopy. J. Phys. Chem. 96, 2614–2617 (1992). zeolites, and their optimisation by the combination of theoretical and
56. Collignon, F., Jacobs, P. A., Grobet, P. & Poncelet, G. Investigation of the experimental studies. J. Catal. 234, 111–118 (2005).
coordination state of aluminum in β zeolites by X-ray photoelectron 81. Jin, F. & Li, Y. A FTIR and TPD examination of the distributive properties of
spectroscopy. J. Phys. Chem. B 105, 6812–6816 (2001). acid sites on ZSM-5 zeolite with pyridine as a probe molecule. Catal. Today
57. Esquivel, D., Cruz-Cabeza, A. J., Jiménez-Sanchidrián, C. & Romero-Salguero, 145, 101–107 (2009).
F. J. Local environment and acidity in alkaline and alkaline-earth exchanged β 82. Moissette, A., Vezin, H., Gener, I. & Brémard, C. Generation and migration
zeolite: structural analysis and catalytic properties. Microporous Mesoporous of electrons and holes during naphthalene sorption in acidic Al-ZSM-5
Mater. 142, 672–679 (2011). zeolites. J. Phys. Chem. B 107, 8935–8945 (2003).
58. van Bokhoven, J. A. In-situ Al K-edge spectroscopy on zeolites: 83. Marquis, S., Moissette, A., Vezin, H. & Brémard, C. Long-lived radical
instrumentation, data-interpretation and catalytic consequences. Phys. Scr. cation–electron pairs generated by anthracene sorption in non Brønsted
2005, 76–79 (2005). acidic zeolites. J. Phys. Chem. B 109, 3723–3726 (2005).

Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials 1055


Perspective NAture MAterIALS
84. Jiao, J. et al. Characterization of framework and extra-framework aluminum 95. DeCanio, S. J., Sohn, J. R., Fritz, P. O. & Lunsford, J. H. Acid catalysis by
species in non-hydrated zeolites Y by 27Al spin-echo, high-speed MAS, and dealuminated zeolite-Y: I. Methanol dehydration and cumene dealkylation.
MQMAS NMR spectroscopy at B0 = 9.4 to 17.6 T. Phys. Chem. Chem. Phys. J. Catal. 101, 132–141 (1986).
7, 3221–3226 (2005). 96. Anunziata, O. A., Martínez, M. L. & Costa, M. G. Characterization and acidic
85. Maier, S. M., Jentys, A. & Lercher, J. A. Steaming of zeolite BEA and its effect properties of Al-SBA-3 mesoporous material. Mater. Lett. 64, 545–548 (2010).
on acidity: a comparative NMR and IR spectroscopic study. J. Phys. Chem. C. 97. Xia, Z. et al. Post synthesis of aluminum modified mesoporous TUD-1
115, 8005–8013 (2011). materials and their application for FCC diesel hydrodesulfurization catalysts.
86. Oumi, Y. et al. Effect of the framework structure on the dealumination– Catalysts 7, 141–160 (2017).
realumination behavior of zeolite. Mater. Chem. Phys. 78, 551–557 (2003). 98. Raatz, F., Freund, E. & Marcilly, C. Study of small-port and large-port
87. Jia, C., Massiani, P. & Barthomeuf, D. Characterization by infrared and mordenite modifications. Part 2.—Ion-exchange properties of thermally
nuclear magnetic resonance spectroscopies of calcined beta zeolite. J. Chem. treated ammonium forms. J. Chem. Soc. Faraday Trans. 1 81, 299–310 (1985).
Soc. Faraday Trans. 89, 3659–3665 (1993). 99. Kunkeler, P. J. et al. Zeolite Beta: the relationship between calcination procedure,
88. Dědeček, J., Tabor, E. & Sklenak, S. Tuning the aluminum distribution in aluminum configuration, and Lewis acidity. J. Catal. 180, 234–244 (1998).
zeolites to increase their performance in acid‐catalyzed reactions.
ChemSusChem 12, 556–576 (2019).
89. Pu, X., Liu, N.-w & Shi, L. Acid properties and catalysis of USY zeolite with Acknowledgements
different extra-framework aluminum concentration. Microporous Mesoporous The authors gratefully acknowledge the ESI Platform of the Paul Scherrer Institute and
Mater. 201, 17–23 (2015). ETH Zurich for financial support.
90. Gruver, V. & Fripiat, J. J. Lewis acid sites and surface aluminum in aluminas
and mordenites: an infrared study of CO chemisorption. J. Phys. Chem. 98, Competing interests
8549–8554 (1994). The authors declare no competing interests.
91. Janin, A. et al. FT IR study of the silanol groups in dealuminated HY zeolites:
nature of the extraframework debris. Zeolites 11, 391–396 (1991).
92. Niwa, M., Sota, S. & Katada, N. Strong Brønsted acid site in HZSM-5 created Additional information
by mild steaming. Catal. Today 185, 17–24 (2012). Correspondence should be addressed to J.A.v.
93. Zholobenko, V. L. et al. On the nature of the sites responsible for the Reprints and permissions information is available at www.nature.com/reprints.
enhancement of the cracking activity of HZSM-5 zeolites dealuminated under
mild steaming conditions: part 2. Zeolites 11, 132–134 (1991). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
94. Anand, R., Maheswari, R. & Hanefeld, U. Catalytic properties of the novel published maps and institutional affiliations.
mesoporous aluminosilicate AlTUD-1. J. Catal. 242, 82–91 (2006). © Springer Nature Limited 2020

1056 Nature Materials | VOL 19 | October 2020 | 1047–1056 | www.nature.com/naturematerials

You might also like