You are on page 1of 11

Trends in Chemistry

Review

Creating Hierarchical Pores in Zeolite Catalysts


Risheng Bai,1 Yue Song,1 Yi Li,1,2 and Jihong Yu1,2,*

Hierarchical zeolites containing both microporosity and meso/macroporosity Highlights


have emerged as a new and important branch of zeolites due to their potential Hierarchical zeolites with two (or more)
for improving mass transfer and molecular accessibility, which are key to over- levels of pore size are advantageous for
many industrially important catalytic
coming steric, diffusional, and coke-formation limitations in catalytic reactions.
reactions.
In this review, we present significant recent progress in the fabrication of hierar-
chical zeolites via mesoporogen, mesoporogen-free, and demetallization strate- In addition to the conventional
gies and discuss the advantages and disadvantages of each method. The impact demetallization methods and

of the hierarchical structures of zeolites (e.g., CHA, FAU) on their enhanced mesoporogen methods, there has been
increasing interest in the development
catalytic activity and product selectivity is highlighted. Finally, we address the of the cost-efficient mesoporogen-free
current challenges and future perspectives for the designed construction of hier- methods for generating zeolite hierarchi-
archical zeolite catalysts to meet the increasing demands for industrial catalytic cal structures.

applications. Hierarchical zeolites have been applied in


various important catalytic reactions,
Zeolites and Hierarchical Structures which show remarkably improved cata-
lytic activity and selectivity. However,
Zeolites are an important class of inorganic microporous crystalline materials widely used in ion
the inherent relationship between the
exchange, gas adsorption, and heterogeneous catalysis (Box 1). Zeolites are among the most hierarchical structures and the catalytic
important solid catalysts in the chemical industry because of their unique shape selectivity, strong properties remains not fully understood.
intrinsic acidity, and high stability [1–7]. Shape selectivity in particular renders zeolites of
paramount importance in heterogeneous catalysis. This molecular handle affords screening of
reactants and/or products that diffuse into or out of the zeolites according to their molecular
sizes [8], and only the stable and pertained transition states can be formed under the steric
constraints of the pores, cages, and zeolite channels [9]. The philosophy of shape selectivity in
zeolites has been well accepted in the petroleum and petrochemical industries and in fine-
chemical catalytic processes.

However, the shortcomings of single-sized zeolite micropores cannot be neglected. On the one
hand, the narrow zeolite pore sizes that afford high selectivity to catalytic reactions fail to affect
molecules with molecular dimensions greater than that of the characteristic pore size. On the
other hand, the small pore sizes and long diffusion path lengths not only reduce transport effi-
ciency but also cause poor catalyst utilization and decreased catalytic rates [10]. Therefore, it is
highly desirable to enhance the accessibility of bulky molecules to the active sites located in the
zeolite channels and to reduce the impact of diffusional limitations while maintaining high selectiv-
ity and catalytic activity [11,12]. Besides exploring extra-large pore zeolites and nanosized
zeolites [13–20], significant effort has been expended towards fabricating hierarchical zeolites
1
to overcome these transport limitations. These hierarchical zeolites feature at least two levels State Key Laboratory of Inorganic
Synthesis and Preparative Chemistry,
of pore systems integrating intrinsic zeolite micropores with mesopores and/or macropores
College of Chemistry, Jilin University,
[4,21–24]. Changchun 130012, People’s Republic
of China
2
Hierarchical zeolites offer the potential to: (i) reduce steric limitations for converting bulky mole- International Center of Future Science,
Jilin University, Changchun 130012,
cules; (ii) increase the rate of intracrystalline diffusion; (iii) inhibit deactivation due to coking;
People’s Republic of China
(iv) maximize catalyst utilization; (v) and modulate selectivity towards target products [25–28].
These advantages endow hierarchical zeolites with superior catalytic performance compared
with their microporous counterparts, especially in the case of reactions where coke deposits *Correspondence:
readily or those involving relatively large molecules. In this review, we present recent progress jihong@jlu.edu.cn (J. Yu).

Trends in Chemistry, September 2019, Vol. 1, No. 6 https://doi.org/10.1016/j.trechm.2019.05.010 601


© 2019 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Trends in Chemistry

Box 1. Nanoporous Materials and Zeolites


Glossary
In terms of the International Union of Pure and Applied Chemistry (IUPAC) classification, nanoporous materials such as
*BEA, CHA, FAU, MFI, and MOR:
zeolites, carbons, and oxides can be classified into microporous materials (with pore diameter less than 2 nm,
zeolite framework type codes assigned
d b2 nm), mesoporous materials (2b d b50 nm), and macroporous materials (d N50 nm) according to their pore diameters.
by the International Zeolite Association:
*BEA, with channels: b100N 12** ↔
Zeolites are an important class of inorganic crystalline materials. The frameworks of zeolites possess ordered distributed
[001] 12*; CHA, with channels: ⊥[001]
pores with diameters typically less than 2 nm. Typically, the frameworks of zeolites are strictly built up from tetrahedral TO4
8***; FAU, with channels: b111N 12***;
units (T = Si, Al, P, etc.) by sharing their vertex oxygen atoms, and different ways of linkage lead to over 200 distinct
MFI, with channels: {[100] 10 ↔ [010]
framework types. Although many zeolites are formed in nature, most of the currently known zeolites are synthesized in
10}***; MOR, with channels: [001] 12* ↔
autoclaves under hydrothermal/solvothermal conditions.
[001] 8*** (http://www.iza-structure.org).
Notation: * preceding BEA indicates that
BEA is partially disordered. Each
in the synthesis of hierarchical zeolites (Figure 1, Key Figure), highlighting the advantages of hier- channel system is characterized by the
archical structures for several important catalytic reactions, including methanol-to-olefin (MTO) channel direction relative to the lattice
axes of the corresponding zeolite
conversion (see Glossary) and Fischer–Tropsch (FT) synthesis. Finally, we address current framework (angle brackets means that
challenges and future perspectives with respect to the fabrication of hierarchical zeolites to there are channels parallel to all crystal-
meet the requirements of industrial catalytic applications. lographically equivalent axes, ⊥ indicates
the direction perpendicular to the
succeeding lattice axes); the channel
Synthetic Strategies for Hierarchical Zeolites aperture presented by number of
Mesoporogen (i.e., mesopore-generating agent) and mesoporogen-free approaches are based T-atoms (in bold type); the number of
asterisks succeeding the channel aper-
on the direct crystallization method widely used to create hierarchical structures in microporous
ture indicates that the channel system is
zeolites. The former includes the utilization of hard or soft templates [29,30], while the latter in- 1D, 2D, or 3D; interconnecting channel
volves seed assistance [31], kinetic control of crystallization [32], and steam-assisted conversion systems are separated by a double
arrow (↔).
3D electron diffraction tomography
(3D-EDT): a state-of-the-art technique
Key Figure allowing 3D reciprocal space recon-
struction of a crystal based on a series of
A Schematic Depicting Synthetic Methods Employed to Fabricate selected-area electron diffraction
Hierarchical Zeolite patterns.
Anderson–Schulz–Flory (ASF)
distribution model: predicts the
hydrocarbon product distributions in FT
synthesis.
Fischer–Tropsch (FT) synthesis: a
reaction converting syngas (CO and H2)
into hydrocarbon fuels and chemicals.
Methanol-to-olefin (MTO)
conversion: a process for the
production of ethylene and propylene
from methanol feedstock, which can be
derived from non-oil raw materials, in-
cluding natural gas, coal, and biomass.
Temperature-programmed
desorption of ammonia (NH3-TPD):
a widely used technique to characterize
the amount and strength of acid sites in
the catalyst by measuring the desorption
of ammonia in the catalyst at different
temperatures.

Trends in Chemistry

Figure 1.

602 Trends in Chemistry, September 2019, Vol. 1, No. 6


Trends in Chemistry

[33]. Demetallization of Al or Si from microporous zeolites through post-treatment (e.g., acid


leaching, base leaching, steaming, irradiation) has also been utilized as an efficient method of
generating zeolite hierarchical structures [34,35]. In this section, we discuss these representative
methods in detail.

Mesoporogen-Directed Hierarchical Zeolites


Mesoporogen-based strategies are based on the introduction of secondary templates (hard or
soft) to the synthetic gels of microporous zeolites [4,21,36,37]. Hard templates, a kind of solid
material with a relatively rigid structure (e.g., carbonaceous materials, polymers, biological mate-
rials, inorganic materials), have been used to produce hierarchical structures in several zeolites
[33,38]. Typically, the resulting hierarchical materials obtained from the hard-template method
feature intrinsic zeolitic characteristics and controllable mesopore size distributions, but suffer
from low hydrothermal/mechanical stability, low pore interconnectivity, and high production cost.

Conversely, the soft-template method offers several advantages over the hard-template method.
However, the mesoporogens (e.g., surfactants, polymers, organosilanes) usually produce
composites of both zeolite crystals and mesoporous materials instead of a crystalline hierarchical
zeolite phase [36,39–42]. To avoid phase segregation, Ryoo and coworkers developed several
‘two-in-one’ templates comprising a hydrophobic long-chain alkyl group and two hydrophilic
quaternary ammonium groups spaced by a short alkyl linkage; the head groups of the multiple
quaternary ammoniums were favorable for the formation of a microporous MFI structure, while
the hydrophobic alkyl chains were assembled into a mesoscale micellar structure that blocked
excessive zeolite growth, culturing the zeolite into a multi/unilamellar mesostructure [43,44].

On the basis of the two-in-one template strategy, Che and coworkers further introduced biphenyl
and naphthyl into the amphiphilic templates together with alkyl tail and quaternary head groups.
The introduced rigid aromatic groups enabled interaction of the tails through aromatic–aromatic
or π–π stacking [30]. The single ammonium groups of the template molecules directed the nucle-
ation and growth of microporous structures that were located in the straight channels of the MFI
framework; the aromatic groups in the hydrophobic tails formed a stable network with sturdy π–π
stacking that geometrically matched the MFI topology and gave rise to single-crystalline
mesostructured zeolite nanosheets with a lamellar structure (Figure 2A). The single-crystalline
nature of the ZSM-5 nanosheets (Figure 2B–C) was supported via 3D electron diffraction
tomography (3D-EDT).

Besides surfactant templates, nonsurfactant templates [e.g., the cationic polymer


polydiallyldimethylammonium chloride (PDADMA)] have been utilized by Xiao and coworkers as
bifunctional templates in directing both the microporous structure and the mesoporous structure
of the zeolite Beta (*BEA, named Beta-MS) [45]. Distinct from the surfactants that can construct
periodically ordered mesostructures through the formation of regular micelles, the missing hydro-
phobic segments allow PDADMA to act as a ‘porogen’ to generate only disordered mesopores.
Selected-area electron diffraction (SAED) patterns and high-resolution transmission electron
microscopy (HR-TEM) images of the individual zeolite particles confirmed that the highly intercon-
nected, hierarchical Beta-MS comprised single crystals rather than randomly aggregated
nanocrystals.

In brief, both well-designed surfactants and nonsurfactant polymers can act as dual-functional
templates to construct hierarchical zeolites with abundant and highly interconnected mesopores.
However, defective sites may be formed during the introduction of hierarchical structures via
mesoporogens, which can act as the Achilles’ heel to their hydrothermal stability. Furthermore,
the high synthetic cost, complex preparation process, and generation of harmful gases during

Trends in Chemistry, September 2019, Vol. 1, No. 6 603


Trends in Chemistry

(A) (B) (C)

5 nm

(D) (E)
(seed assistance)

(F) (G) (H)


(kinetic regulation)

(I) (J)

Trends in Chemistry

Figure 2. Hierarchical Zeolites Constructed via Mesoporogen, Mesoporogen-Free, and Demetallization Methods. (A) Schematic of the position and interaction of
quaternary ammoniums in the zeolite channel, which serve as a template to direct the formation of the lamellar hierarchical zeolite structure. (B) Transmission electron microscopy
(TEM) image and (C) 3D electron diffraction tomography (3D-EDT) projections with multiple unit-cell meshes along different directions illustrating the structural details in 3D
reciprocal space. Adapted, with permission, from [30]. (D) Schematic of the formation process and (E) TEM images of the resulting nanosized hierarchical SAPO-34 zeolite via
the seed-assisted method. Adapted, with permission, from [46]. (F) Schematic of the crystal evolution process, (G) TEM image, and (H) spherical aberration-corrected
scanning TEM annular dark field (Cs-corrected STEM-ADF) image, as well as the corresponding electron diffraction pattern of the single-crystalline hierarchical ZSM-5 zeolite
nanocrystal. Adapted, with permission, from [49]. (I) Scanning electron microscopy (SEM) image of the coupled dealumination–desilication-treated ZSM-5 zeolite. Adapted,
with permission, from [34]. (J) TEM images of the dissolution of aggregates (ZSM-5-I) and apparent single crystals (ZSM-5-II) treated with NH4F for 30 and 60 min,
respectively. Adapted, with permission, from [35].

604 Trends in Chemistry, September 2019, Vol. 1, No. 6


Trends in Chemistry

the removal of excess organic mesoporogens may also hinder practical industrial application of
these hierarchical zeolites.

Seed-Assisted Synthesis of Hierarchical Zeolites


Mesoporogen-free synthetic methods have proven to be a cost-effective and green synthetic
way to fabricate hierarchical zeolites. Some effective strategies have been developed, including
self-pillaring, steaming assisting, seed assisting, and kinetically regulated crystallization.

Seed-assisted synthesis in particular has demonstrated advantages in the preparation of hierar-


chical zeolites with improved crystallinity, decreased crystallization time, and adjustable mesopo-
rous structures. With the assistance of zeolite seeds, hierarchical zeolites such as SAPO-34
(CHA), ZSM-5 (MFI), and TS-1 (MFI) have been prepared by regulating the synthesis gel compo-
sitions without adding extra mesoporogens [31,46,47]. For example, Tang and coworkers
prepared a series of hierarchical ZSM-5 zeolites with tunable intracrystalline/intercrystalline
mesopores that could be finely tuned through additives such as KF and secondary organic
structure-directing agents [47]. The dissolution degree of zeolite seeds increased with the
amount of KF additive, facilitating the formation of intracrystalline mesoporous structures. Adding
organic templates with a significant structure-directing effect or increasing the amount of
tetrapropylammonium bromide (TPABr) or H2O in the seed-assisted system enabled the forma-
tion of oriented self-assembled nanocrystalline hierarchical zeolites.

In addition to the aluminosilicate zeolites, silicoaluminophosphate (SAPO-34) and titanosilicate


(TS-1) zeolites with hierarchical structures have been prepared using the seed-assisted method
[31,46,48]. For example, Yu and coworkers introduced small zeolite seeds in the preparation of
SAPO-34 [46,48]. The seeds were partially dissolved into fragments and SAPO-34 crystallized
epitaxially on the surface of the fragments, resulting in interconnected meso- and macroporous
hierarchical structures (Figure 2D,E).

As a facile and efficient strategy for generating hierarchical structures in various zeolites, the seed-
assisted method has aroused growing interest. However, there remain challenges in the rational
construction of hierarchical zeolites with controllable meso/macropores by this method, partly
due to the current lack of mechanistic understanding (e.g., how exactly the seeds direct the hier-
archical structures).

Kinetically Regulated Hierarchical Zeolites


In addition to the seed-assisted method, regulation of crystallization kinetics has also been
explored as an efficient mesoporogen-free method for the preparation of hierarchical zeolites.
For example, Stucky and coworkers prepared a hierarchical ZSM-5 zeolite catalyst with a
core–shell structure using tetrapropylammonium hydroxide (TPAOH) as the sole structure-
directing agent by regulating the nucleation and growth kinetics of zeolite crystals [32]. The key
point in constructing the hierarchical porous structure was the precisely controlled epitaxial
growth of aluminum, achieved by a two-step crystallization process. During the first step at low
temperature (20–80°C), most of the raw materials hydrolyzed and condensed to form
precrystallized clusters that further transformed into hierarchical zeolites with Al-rich branches
at the second crystallization step occurring at a higher temperature (100–140°C). Notably, the
branched portion of the hierarchical Al-zoned zeolites exhibited oriented attachment, showing
the single-crystalline nature of the sample.

Recently, an amino acid-assisted synthesis was proposed by Yu and coworkers whereby defect-
free single-crystalline hierarchical nano-ZSM-5 zeolites were constructed via a two-step crystal-
lization strategy [49]. At low temperature (90°C), L-lysine added to the concentrated reaction

Trends in Chemistry, September 2019, Vol. 1, No. 6 605


Trends in Chemistry

gel inhibited excessive growth of the initially formed protozeolite nanoparticles. As the crystalliza-
tion process proceeded, the protozeolites converted into integrated nanocrystals by oriented
attachment generating intracrystalline isolated mesopores. At the second crystallization stage
(at high temperature, 170°C), the isolated mesopores in the orientedly attached zeolites evolved
into highly interconnected mesopores via intraparticle ripening (Figure 2F). Electron diffraction
patterns, 27Al magic-angle spinning (MAS) NMR spectra, and 29Si MAS NMR spectra revealed
the single-crystalline and defect-free characteristics of the final hierarchical zeolite nanocrystals
(Figure 2G,H). Furthermore, the size of the secondary pores in the hierarchical ZSM-5 zeolites
could be adjusted by varying the temperature of the second crystallization stage.

Yu and coworkers also demonstrated an intermediate-crystallization strategy to prepare hierar-


chical zeolites via regulating crystallization time [31]. In the seed-assisted synthesis of
titanosilicate TS-1, the intermediate-crystallized TS-1 zeolites possessed both intracrystalline
and intercrystalline mesopores stemming from the dissolution of zeolite seed fragments and
aggregation of initially formed nanozeolite crystals, respectively. In addition, this intermediate-
crystallization strategy can also afford abundant hexacoordinated Ti species that are active
sites for several oxidation reactions (e.g., the oxidative desulfurization reactions).

It should be noted that the kinetics-regulation method relies on the control of zeolite nucleation
and growth, which not only affords the construction of hierarchical structures with tunable
mesopore sizes but also regulates the distribution and state of the active sites (e.g., Al, Ti).
Extending this efficient method to the preparation of many other types of zeolites is highly desired.
Such mesoporogen-free methods may facilitate the large-scale production of industrially impor-
tant zeolite catalysts for catalytic applications.

Demetallization in Constructing Hierarchical Zeolites


Demetallization processes (e.g., dealumination, degermanation, desilication) represent a class of
destructive methods for constructing hierarchical structures in crystalline microporous zeolitic
materials [50–54]. Dealumination (involving acid and steam treatments) has been traditionally
used as an industrial technology in modifying and optimizing microporous zeolites; removing
alumina species from the zeolite framework alters the acidity and Si/Al ratio, which contribute to
the catalytic activity and stability [21]. However, dealumination usually generates isolated cavities
or mesovoids rather than the interconnected mesopores, limiting the improvement of the
substrate mass transport involved in the catalytic reactions [55,56]. By contrast, desilication,
which is applicable to high-silica microporous zeolites, has an advantage in producing intercon-
nected mesopores in zeolites but fails to generate abundant mesopores in large-particle-size
crystals [52,57].

To overcome the shortcomings mentioned above, especially for Al-rich zeolites, Huang and
coworkers bridged the dealumination and desilication processes to construct hierarchical
structures in ZSM-5 zeolites with high Al content (Si/Al = 10–20) [34]. The consecutive
steaming–alkaline treatment generated uniform mesopores in both large and small ZSM-5 crystals;
this cannot be achieved by the steaming or alkaline treatment independently (Figure 2I). During the
demetallization process, the reconstruction of Al sites and Si sites occurred inside the microporous
zeolites and framework Al species were elaborated as the pore-directing agent [58]. Notably, ver-
satile interconversions among framework Al and extraframework Al species were observed during
the coupled dealumination–desilication process, which provided an understanding of the role of
(extra)framework Al and the zeolite structures from different perspectives.

In addition to traditional acid leaching, base leaching, and steaming, Valtchev and coworkers
utilized NH4F as a nondiscriminate extractor to extract both framework Si and Al from zeolites

606 Trends in Chemistry, September 2019, Vol. 1, No. 6


Trends in Chemistry

such as ZSM-5, silicalite-1, and mordenite (MOR), resulting in a mosaic pattern of rectangular
and aligned hierarchical pores [35]. The HF and HF2− species formed by the double hydrolysis
of NH4F endowed a controllable aspect to the etching process, which extracted both Si and Al
at the same rate. Defect zones in zeolites (including the interfaces between intergrown crystals
and grain boundaries) were sensitive to the F species, giving the mosaic patterns. With the
etching time prolonged, these mosaic-like structures in zeolites transformed into highly intercon-
nected mesopores (Figure 2J).

Although demetallization has been widely applied to various zeolites to construct hierarchical
structures, the low interconnectivity of the secondary meso/macropores, the decreased crystal-
linity, and the mass loss of zeolite crystals during the demetallization process should be improved
and overcome.

Catalytic Reactions Exploiting Hierarchical Zeolite Structures


Industrially important zeolites (e.g., *BEA, CHA, FAU, and MFI zeotype frameworks) with unique
porous architectures, strong Brønsted acidity, and high stability show excellent catalytic perfor-
mance in catalytic processes such as (hydro)cracking, isomerization, and aromatization reactions
[21,25,59,60]. The introduction of hierarchical structures in these zeolites can greatly improve the
catalytic activity and the product selectivity in bulky-molecule-involved and coke-prone reactions,
which could be attributed to the modified acidity and the enhanced accessibility to active sites as
well as the improved mass-transfer efficiency. Herein, taking zeolites SAPO-34 and Y as exam-
ples, we demonstrate the impact of hierarchical structures on the enhanced catalytic perfor-
mance in MTO conversion and FT synthesis.

MTO
MTO conversion has proved to be an effective nonpetrochemical method for the production of
light olefins [61–63]. Silicoaluminophosphate SAPO-34 (CHA) zeolite is an ideal catalyst for the
MTO reaction to produce ethylene and propylene with the complete conversion of methanol,
but suffers from rapid deactivation due to the inherent diffusion bottleneck induced by the
deposited coke species. Many synthetic efforts have been devoted to introducing hierarchical
structures in SAPO-34 catalysts to improve their MTO performance [27,64].

With organosilane dimethyloctadecyl-(3-trimethoxysilylpropyl)-ammonium chloride (TPOAC) as


the mesoporogen along with the microporous template morpholine, Yu and coworkers obtained
micron-sized hierarchical SAPO-34 crystals comprising aggregated nanocrystallites [65]. The
resultant catalyst showed superior catalytic performance in the MTO reaction with about four-
times-prolonged catalytic lifetime (up to 466 min) and nearly 10% improvement of selectivity for
ethylene and propylene compared with the microporous SAPO-34 catalyst of micron size. The
coke formation rate of the hierarchical SAPO-34 zeolite was 0.132 mg/min (much lower than
that of the microporous SAPO-34 catalyst at 0.334 mg/min), demonstrating the benefit of hierar-
chical structures in decreasing the coke formation rate and thus prolonging the catalyst lifetime.

Yu and coworkers also fabricated hierarchical SAPO-34 catalysts with micro-meso-macropores


via the seed-directed method, which showed a prolonged catalytic lifetime (366 min) and a higher
selectivity for ethylene and propylene of up to 85.4% in the MTO reaction compared with that of
the microporous counterpart (126 min and 79.7%, respectively) [48]. A larger number of coke
species (e.g., methyl-substituted benzene and pyrene) were detected inside the hierarchical
SAPO-34 catalysts compared with their microporous counterparts. The hierarchical structures
reduced the mass-transfer resistance imposed by the narrow micropores, increased active-site
utilization, and accommodated more coke, all leading to improved catalytic performance
(Figure 3).

Trends in Chemistry, September 2019, Vol. 1, No. 6 607


Trends in Chemistry

Trends in Chemistry

Figure 3. Benefits of Hierarchical Porosity in SAPO-34 Catalysts for the Synthesis of Ethylene and Propylene
via the Methanol-to-Olefin Reaction.

FT Synthesis
Currently, FT synthesis has received renewed attention due to the growing global demand for
liquid fuels and the oil shortage [66–73]. Extensive research effort has been expended towards
the development of efficient FT synthesis catalysts, but the effective control of product selectivity
remains a great challenge.

Significantly, Bao and coworkers developed composite bifunctional FT-synthesis catalysts com-
prising a partially reduced oxide (ZnCrOx) and hierarchical SAPO-34 zeolite [74]. The composite
catalyst (OX-ZEO) showed C2−C4 selectivity of up to 94% (including 80% C2=–C4= and 14%
C2°–C4°) for all hydrocarbons (carbon atom-based), with only 2% methane at a CO conversion
of 17%, far beyond the value predicted by the Anderson–Schulz–Flory (ASF) distribution
model (58%) [75–78]. The activation of CO and H2 occurred at the oxide surfaces, while
the C–C coupling took place at the acid sites located within zeolite pores, which controlled the
C–C coupling process precisely by suppressing over-hydrogenation and methane formation.
The composite catalyst with the hierarchical SAPO-34 zeolite presented an olefin/paraffin ratio
(C2=–C4=/C2°–C4°) of 4.7, superior to that observed for the microporous counterpart (0.9).
The enhanced olefin selectivity of hierarchical SAPO-34 zeolite is potentially attributable to the
weaker acid strength than that of the microporous zeolite, as evidenced by the temperature-
programmed desorption of ammonia (NH3-TPD) measurements (NH3 desorption peak at
350°C vs 394°C). The catalytic performance manifested the key role of zeolite acid strength in de-
termining the product distributions in FT synthesis, which can be tuned by introducing hierarchical
structures into the zeolites.

Other than the hierarchical SAPO-34 zeolites, aluminosilicate faujasite Y (FAU) zeolites have also
demonstrated their unique catalytic ability in FT synthesis [79]. Recently, Wang and coworkers
proposed a new strategy of controlling hydrogenolysis in FT synthesis by using mesoporous
zeolite Y-supported cobalt nanoparticles as a composite catalyst to improve the selectivity of
C10–C20 hydrocarbons up to 60%, far exceeding the expected ASF distribution (39%) [80].

608 Trends in Chemistry, September 2019, Vol. 1, No. 6


Trends in Chemistry

Notably, the variation of the size of the mesopores did not change the conversion of the original Outstanding Questions
substrates, but exerted significant influence on product selectivity. The selectivities for C10– Will the structures (dimension, shape,
C20 increased with mesopore size, while the selectivities for CH4 decreased. The hierarchical and connectivity) of the generated sec-
structure not only improved the mass-transport efficiency but also improved the product selectiv- ondary meso/macropores in hierarchical
zeolites be precisely fabricated?
ity by promoting the hydrogenolysis process and suppressing the successive demethylation in
the hierarchical zeolite structures [80,81]. How can the distribution of mesopores
be rationally controlled to achieve
As demonstrated above, several important catalytic reactions could benefit significantly from use ‘molecular traffic control’ in hierarchical
of hierarchical structures in zeolite catalysts, which afford more easily accessed diffusion path- zeolites?

ways, greater coke accommodation space, appropriately modulated acid strength, and greatly
What is the formation mechanism of
improved active-site accessibility. hierarchical zeolite structures regulated
by kinetic control of the nucleation and
Concluding Remarks crystal growth?
Compared with conventional microporous zeolites, hierarchical zeolites with two (or more) levels
Is it possible to construct the long-range
of pore size are advantageous for many industrially important catalytic reactions. The highly inter-
ordered mesoporous structures inside
connected porosity of hierarchical zeolite structures is favorable to overcome the mass-transfer zeolites, especially zeolites with a struc-
issues associated with microporous zeolites. Thus, the diffusion and steric limitations of bulky turally hierarchical system of connected
molecules and the pore-blocking effect of deposited coke species can be reduced, affording micropores and mesopores?

high utilization efficiency, prolonged catalytic lifetime, and improved catalytic performance of
Can the mesopore-afforded shape
zeolite catalysts. So far, a variety of synthetic methods based on mesoporogen-directed, selectivity of hierarchical zeolites be im-
mesoporogen-free synthetic methods, and demetallization strategies have been employed to proved, in order to allow the discriminate
create hierarchical structures in zeolites. These synthetic strategies afford hierarchical zeolite screening of bulky reactants and prod-
ucts as well as transition states?
characteristics (e.g., distribution, dimensionality, connectivity, even ordering of secondary
pores) to be tuned to enhance the diffusivity of bulky molecules and the accessibility of the
substrates to the active sites of the catalysts. However, constructing hierarchical zeolites with
precisely controlled distributions of secondary meso/macropores (e.g., location, dimension,
shape, connectivity, ordering) and adjustable active sites remains a significant challenge (see
Outstanding Questions).

The production cost and potential health, safety, and environmental issues related to the use of
expensive and occasionally toxic organic mesoporogen templates (and the release of harmful
gases on template removal) may limit the large-scale industrial applications of hierarchical zeolites
created using this approach. Further, the demetallization approach suffers from difficulties
with controlling interconnectivity and pore depth, damage of microporous and crystalline proper-
ties, and material loss. The development of the cost-efficient mesoporogen-free methods
(e.g., seed-assisted and kinetically regulated crystallization methods) may promise scalable
production of hierarchical zeolites. However, critical economic analyses of the developed
approaches to the hierarchical zeolites must be conducted to make the industrial implementation
realistic.

To date, the catalytic mechanisms associated with hierarchical zeolite catalysts remain unclearly
established. To this end, deepened understanding of the relationship between the multilevel pore
system of specific zeolite structures and catalytic performance is needed. The innovation of facile
and cost-effective synthetic strategies, the development of characterization techniques to probe
the diffusion of molecules inside the hierarchical pore system, and understanding of the
structure–catalytic relationships in hierarchical zeolites promise to boost the rational design and
industrial application of hierarchical zeolite catalysts.

Acknowledgments
The authors thank the National Key Research and Development Program of China (2016YFB0701100), the National Natural
Science Foundation of China (21835002 and 21621001), and the 111 Project, China (B17020) for supporting this work. R.B.
also acknowledges the Graduate Interdisciplinary Research Funding of Jilin University (10183201815).

Trends in Chemistry, September 2019, Vol. 1, No. 6 609


Trends in Chemistry

References
1. Li, Y. and Yu, J.H. (2014) New stories of zeolite structures: their 28. Yang, X-Y. et al. (2017) Hierarchically porous materials: synthesis
descriptions, determinations, predictions, and evaluations. strategies and structure design. Chem. Soc. Rev. 46, 481–558
Chem. Rev. 114, 7268–7316 29. Serrano, D.P. et al. (2013) Synthesis strategies in the search for
2. Martinez, C. and Corma, A. (2011) Inorganic molecular sieves: hierarchical zeolites. Chem. Soc. Rev. 42, 4004–4035
preparation, modification and industrial application in catalytic 30. Xu, D. et al. (2014) π–π Interaction of aromatic groups in amphi-
processes. Coord. Chem. Rev. 255, 1558–1580 philic molecules directing for single-crystalline mesostructured
3. Li, C. et al. (2018) Building zeolites from precrystallized units: zeolite nanosheets. Nat. Commun. 5, 4262–4270
nanoscale architecture. Angew. Chem. Int. Ed. Engl. 57, 31. Bai, R. et al. (2018) Intermediate-crystallization promoted cata-
15330–15353 lytic activity of titanosilicate zeolites. J. Mater. Chem. A 6,
4. Sachse, A. and Garcia-Martinez, J. (2017) Surfactant-templating of 8757–8762
zeolites: from design to application. Chem. Mater. 29, 3827–3853 32. Ding, K. et al. (2015) Constructing hierarchical porous zeolites
5. Dusselier, M. and Davis, M.E. (2018) Small-pore zeolites: via kinetic regulation. J. Am. Chem. Soc. 137, 11238–11241
synthesis and catalysis. Chem. Rev. 118, 5265–5329 33. Machoke, A.G. et al. (2015) Micro/macroporous system: MFI-
6. Li, J. et al. (2015) Synthesis of new zeolite structures. Chem. type zeolite crystals with embedded macropores. Adv. Mater.
Soc. Rev. 44, 7112–7127 27, 1066–1070
7. Prech, J. et al. (2018) From 3D to 2D zeolite catalytic materials. 34. Yang, S. et al. (2017) Bridging dealumination and desilication for
Chem. Soc. Rev. 47, 8263–8306 the synthesis of hierarchical MFI zeolites. Angew. Chem. Int. Ed.
8. Kaerger, J. and Valiullin, R. (2013) Mass transfer in mesoporous Engl. 56, 12553–12556
materials: the benefit of microscopic diffusion measurement. 35. Qin, Z. et al. (2016) The mosaic structure of zeolite crystals.
Chem. Soc. Rev. 42, 4172–4197 Angew. Chem. Int. Ed. Engl. 55, 15049–15052
9. Csicsery, S.M. (1986) Catalysis by shape selective zeolites – 36. Luo, H.Y. et al. (2015) One-pot synthesis of MWW zeolite nano-
science and technology. Pure Appl. Chem. 58, 841–856 sheets using a rationally designed organic structure-directing
10. Li, K. et al. (2014) Realizing the commercial potential of hier- agent. Chem. Sci. 6, 6320–6324
archical zeolites: new opportunities in catalytic cracking. 37. Shen, X. et al. (2018) A hierarchical MFI zeolite with a two-
ChemCatChem 6, 46–66 dimensional square mesostructure. Angew. Chem. Int. Ed.
11. Schneider, D. et al. (2016) Transport properties of hierarchical Engl. 57, 724–728
micro-mesoporous materials. Chem. Soc. Rev. 45, 3439–3467 38. Schmidt, I. et al. (2000) Catalytic epoxidation of alkenes with
12. Bingre, R. et al. (2018) An overview on zeolite shaping technol- hydrogen peroxide over first mesoporous titanium-containing
ogy and solutions to overcome diffusion limitations. Catalysts zeolite. Chem. Commun. 21, 2157–2158
8, 163–180 39. Cheneviere, Y. et al. (2010) Synthesis and catalytic properties of
13. Awala, H. et al. (2015) Template-free nanosized faujasite-type TS-1 with mesoporous/microporous hierarchical structures ob-
zeolites. Nat. Mater. 14, 447–451 tained in the presence of amphiphilic organosilanes. J. Catal.
14. Mintova, S. et al. (2015) Nanosized microporous crystals: 269, 161–168
emerging applications. Chem. Soc. Rev. 44, 7207–7233 40. Wu, Y. et al. (2015) Spatial distribution and catalytic perfor-
15. Zaarour, M. et al. (2014) Progress in zeolite synthesis promotes mance of metal-acid sites in Mo/MFI catalysts with tunable
advanced applications. Microporous Mesoporous Mater. 189, meso-/microporous lamellar zeolite structures. J. Catal. 323,
11–21 100–111
16. Bai, R. et al. (2016) Simple quaternary ammonium cations- 41. Jin, D. et al. (2017) Hierarchical silicoaluminophosphate catalysts
templated syntheses of extra-large pore germanosilicate zeo- with enhanced hydroisomerization selectivity by directing the ori-
lites. Chem. Mater. 28, 6455–6458 entated assembly of premanufactured building blocks. ACS
17. Jiang, J. et al. (2015) ITQ-54: a multi-dimensional extra-large Catal. 7, 5887–5902
pore zeolite with 20 × 14 × 12-ring channels. Chem. Sci. 6, 42. Castro, M. et al. (2018) Unraveling direct formation of hierarchical
480–485 zeolite Beta by dynamic light scattering, small angle X-ray scat-
18. Yang, J. et al. (2017) Principles of designing extra-large pore tering, and liquid and solid state NMR: insights at the supramo-
openings and cages in zeolitic imidazolate frameworks. J. Am. lecular level. Chem. Mater. 30, 2676–2686
Chem. Soc. 139, 6448–6455 43. Choi, M. et al. (2009) Stable single-unit-cell nanosheets of zeolite
19. Smeets, S. et al. (2014) High-silica zeolite SSZ-61 with MFI as active and long-lived catalysts. Nature 461, 246–249
dumbbell-shaped extra-large-pore channels. Angew. Chem. 44. Na, K. et al. (2011) Directing zeolite structures into hierarchically
Int. Ed. Engl. 53, 10398–10402 nanoporous architectures. Science 333, 328–332
20. Kang, J.H. et al. (2016) Synthesis and characterization of CIT- 45. Zhu, J. et al. (2014) Highly mesoporous single-crystalline
13, a germanosilicate molecular sieve with extra-large pore zeolite Beta synthesized using a nonsurfactant cationic poly-
openings. Chem. Mater. 28, 6250–6259 mer as a dual-function template. J. Am. Chem. Soc. 136,
21. Schwieger, W. et al. (2016) Hierarchy concepts: classification 2503–2510
and preparation strategies for zeolite containing materials with 46. Sun, Q. et al. (2016) Seeding induced nano-sized hierarchical
hierarchical porosity. Chem. Soc. Rev. 45, 3353–3376 SAPO-34 zeolites: cost-effective synthesis and superior MTO
22. Chen, L-H. et al. (2012) Hierarchically structured zeolites: syn- performance. J. Mater. Chem. A 4, 14978–14982
thesis, mass transport properties and applications. J. Mater. 47. Zhang, H. et al. (2016) Tailoring zeolite ZSM-5 crystal morphol-
Chem. 22, 17381–17403 ogy/porosity through flexible utilization of silicalite-1 seeds as
23. Sun, M-H. et al. (2016) Applications of hierarchically structured templates: unusual crystallization pathways in a heterogeneous
porous materials from energy storage and conversion, catalysis, system. Chem. Eur. J. 22, 7141–7151
photocatalysis, adsorption, separation, and sensing to biomedi- 48. Sun, Q. et al. (2018) Mesoporogen-free synthesis of hierarchical
cine. Chem. Soc. Rev. 45, 3479–3563 SAPO-34 with low template consumption and excellent
24. Perez-Ramirez, J. et al. (2008) Hierarchical zeolites: enhanced methanol-to-olefin conversion. ChemSusChem 11, 3812–3820
utilisation of microporous crystals in catalysis by advances in 49. Zhang, Q. et al. (2019) Amino acid-assisted construction of
materials design. Chem. Soc. Rev. 37, 2530–2542 single-crystalline hierarchical nanozeolites via oriented-
25. Hartmann, M. et al. (2016) Catalytic test reactions for the evalu- aggregation and intraparticle ripening. J. Am. Chem. Soc. 141,
ation of hierarchical zeolites. Chem. Soc. Rev. 45, 3313–3330 3772–3776
26. Verboekend, D. et al. (2016) Synthesis, characterisation, and cata- 50. Qin, Z. et al. (2014) Comparative study of nano-ZSM-5 catalysts
lytic evaluation of hierarchical faujasite zeolites: milestones, chal- synthesized in OH– and F– media. Adv. Funct. Mater. 24,
lenges, and future directions. Chem. Soc. Rev. 45, 3331–3352 257–264
27. Sun, Q. et al. (2018) The state-of-the-art synthetic strategies for 51. Qin, Z. et al. (2013) Chemical equilibrium controlled etching of
SAPO-34 zeolite catalysts in methanol-to-olefin conversion. Natl. MFI-type zeolite and its influence on zeolite structure, acidity,
Sci. Rev. 5, 542–558 and catalytic activity. Chem. Mater. 25, 2759–2766

610 Trends in Chemistry, September 2019, Vol. 1, No. 6


Trends in Chemistry

52. Verboekend, D. and Perez-Ramirez, J. (2011) Design of hierar- 68. Zhang, Q. et al. (2014) Fischer–Tropsch catalysts for the produc-
chical zeolite catalysts by desilication. Catal. Sci. Technol. 1, tion of hydrocarbon fuels with high selectivity. ChemSusChem 7,
879–890 1251–1264
53. Mazur, M. et al. (2015) Synthesis of ‘unfeasible’ zeolites. Nat. 69. Su, J. et al. (2019) Syngas to light olefins conversion with high
Chem. 8, 58–62 olefin/paraffin ratio using ZnCrOx/AlPO-18 bifunctional catalysts.
54. Roth, W.J. et al. (2013) A family of zeolites with controlled pore Nat. Commun. 10, 1297
size prepared using a top-down method. Nat. Chem. 5, 70. Cheng, K. et al. (2012) Mesoporous Beta zeolite-supported ru-
628–633 thenium nanoparticles for selective conversion of synthesis gas
55. Janssen, A.H. et al. (2001) Three-dimensional transmission elec- to C5–C11 isoparaffins. ACS Catal. 2, 441–449
tron microscopic observations of mesopores in dealuminated 71. Cheng, K. et al. (2016) Direct and highly selective conversion of
zeolite Y. Angew. Chem. Int. Ed. Engl. 40, 1102–1104 synthesis gas into lower olefins: design of a bifunctional catalyst
56. Zecevic, J. et al. (2012) Mesoporosity of zeolite Y: quantitative combining methanol synthesis and carbon–carbon coupling.
three-dimensional study by image analysis of electron tomo- Angew. Chem. Int. Ed. Engl. 128, 4803–4806
grams. Angew. Chem. Int. Ed. Engl. 51, 4213–4217 72. Zhou, W. et al. (2018) Direct conversion of syngas into
57. Luk, H.T. et al. (2019) Impact of carrier acidity on the conversion methyl acetate, ethanol, and ethylene by relay catalysis via the
of syngas to higher alcohols over zeolite-supported copper-iron intermediate dimethyl ether. Angew. Chem. Int. Ed. Engl. 57,
catalysts. J. Catal. 371, 116–125 12012–12016
58. Groen, J.C. et al. (2005) Mechanism of hierarchical porosity 73. Liu, X. et al. (2018) Design of efficient bifunctional catalysts for
development in MFI zeolites by desilication: the role of aluminium direct conversion of syngas into lower olefins via methanol/
as a pore-directing agent. Chem. Eur. J. 11, 4983–4994 dimethyl ether intermediates. Chem. Sci. 9, 4708–4718
59. Mitchell, S. et al. (2013) From powder to technical body: the 74. Jiao, F. et al. (2016) Selective conversion of syngas to light ole-
undervalued science of catalyst scale up. Chem. Soc. Rev. 42, fins. Science 351, 1065–1068
6094–6112 75. Torres Galvis, H.M. and de Jong, K.P. (2013) Catalysts for pro-
60. Shi, J. et al. (2015) Recent advances of pore system construc- duction of lower olefins from synthesis gas: a review. ACS
tion in zeolite-catalyzed chemical industry processes. Chem. Catal. 3, 2130–2149
Soc. Rev. 44, 8877–8903 76. Friedel, R. and Anderson, R.B. (1950) Composition of synthetic
61. Tian, P. et al. (2015) Methanol to olefins (MTO): from fundamen- liquid fuels. I. Product distribution and analysis of C5–C8 paraffin
tals to commercialization. ACS Catal. 5, 1922–1938 isomers from cobalt catalyst 1. J. Am. Chem. Soc. 72,
62. Yarulina, I. et al. (2018) Recent trends and fundamental insights 1212–1215
in the methanol-to-hydrocarbons process. Nat. Catal. 1, 77. Puskas, I. and Hurlbut, R.S. (2003) Comments about the causes
398–411 of deviations from the Anderson–Schulz–Flory distribution of the
63. Li, C. et al. (2018) Synthesis of reaction-adapted zeolites as Fischer–Tropsch reaction products. Catal. Today 84, 99–109
methanol-to-olefins catalysts with mimics of reaction intermedi- 78. Cheng, Q. et al. (2018) Confined small-sized cobalt catalysts
ates as organic structure-directing agents. Nat. Catal. 1, stimulate carbon-chain growth reversely by modifying ASF law
547–554 of Fischer–Tropsch synthesis. Nat. Commun. 9, 3250
64. Zhong, J. et al. (2017) Recent advances of the nano-hierarchical 79. Zhang, Q. et al. (2010) Development of novel catalysts for
SAPO-34 in the methanol-to-olefin (MTO) reaction and other Fischer–Tropsch synthesis: tuning the product selectivity.
applications. Catal. Sci. Technol. 7, 4905–4923 ChemCatChem 2, 1030–1058
65. Sun, Q. et al. (2014) Organosilane surfactant-directed synthesis 80. Peng, X. et al. (2015) Impact of hydrogenolysis on the selectivity
of hierarchical porous SAPO-34 catalysts with excellent MTO of the Fischer–Tropsch synthesis: diesel fuel production over
performance. Chem. Commun. 50, 6502–6505 mesoporous zeolite-Y-supported cobalt nanoparticles. Angew.
66. Sartipi, S. et al. (2014) Catalysis engineering of bifunctional solids Chem. Int. Ed. Engl. 54, 4553–4556
for the one-step synthesis of liquid fuels from syngas: a review. 81. Li, J. et al. (2018) Integrated tuneable synthesis of liquid fuels via
Catal. Sci. Technol. 4, 893–907 Fischer–Tropsch technology. Nat. Catal. 1, 787–793
67. Li, N. et al. (2019) High‐quality gasoline directly from syngas by
dual metal oxide–zeolite (OX‐ZEO) catalysis. Angew. Chem. Int.
Ed. Engl. 58, 7400–7404

Trends in Chemistry, September 2019, Vol. 1, No. 6 611

You might also like