LDH Catalysts PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

REVIEW

Layered Double Hydroxides www.afm-journal.de

Layered Double Hydroxide-Based Catalysts: Recent


Advances in Preparation, Structure, and Applications
Ming Xu and Min Wei*

performance. Many studies have shown


Layered double hydroxides (LDHs) are a class of functional anionic clays, that active sites can be efficiently modu-
which consist of positively charged host layers (brucite-like M(OH)6 octa- lated by controlling the particle size, com-
hedra) and interlayer anions. By virtue of their unique combination of struc- position, and morphology of catalysts,
regulating their surface/interface defect
tural features (including the tunability of both host layers and interlayer guest
structure (e.g., type and concentration),
anion, exfoliation property, structure topological transformation, confinement and modulating their electronic structures
effect), LDHs have many potential applications in heterogeneous catalysis— or metal–support interactions.[7,8] Thus, in
as catalysts themselves, catalyst supports, or catalyst precursors. In addition, order to substantially improve the activity,
the properties of LDH-based catalysts can be tailored for specific purposes selectivity, and stability of catalysts, it is
by facile modulation of their surface/interface defect structure (e.g., oxygen desirable to develop nanocatalysts with
tailored geometric, electronic, and sur-
vacancy defects or metal defects), controlling the concentration/strength
face/interface defect structures, which
of surface acid/base sites, tuning the geometric/electronic structure of also have a high exposure of active sites.
active sites, or by taking advantage of the confinement effect intrinsic to 2D However, using the traditional methods
materials. In addition, by utilizing the topological structural transformation of for the preparation of heterogeneous cata-
LDH precursors, supported metal catalysts can be obtained (as single metals, lysts, it is difficult to rationally design and
bimetallic alloys or heterostructures, and intermetallic compounds) with tun- precisely modulate active centers at the
atomic scale, and thus attaining simul-
able particle size/morphology and intriguing electronic properties. The main taneously enhanced activity and stability
focus of this review is on recent advances in structure design, preparation, remains a big challenge.
and catalytic applications of LDH-based heterogeneous catalysts. In addition, Layered double hydroxides (LDHs)[9,10]
future challenges and development strategies are discussed from the view- are a class of 2D anionic compounds
points of modulation of intrinsic active sites and establishment of scalable made up of positively charged brucite-like
host layers containing M(OH)6 octahedra,
fabrication processes.
with intercalated anions and water mol-
ecules confined in the interlayer gallery.
LDHs can be represented by the general
1. Introduction formula [M2+1−xM3+x(OH)2]z+An−z/n⋅mH2O, where M2+ and
M3+ are divalent and trivalent metal cations located in the host
Recent advances in the rational design and preparation of high- layers and An− are exchangeable anions in the interlayer region
performance catalysts with excellent stability have attracted compensating for the positive charge on the layers. By virtue
considerable attention in the heterogeneous catalysis field, of the wide range of possible physical and chemical proper-
owing to the extensive applications of the resulting materials in ties, LDH materials have found many applications in the het-
the energy and chemical industries.[1–3] Great efforts have been erogeneous catalysis field,[11,12] including as photocatalysts,
devoted to modulating fine structure of active catalytic sites, electrocatalysts, and thermocatalysts. The composition, particle
since this determines the activity and selectivity of catalysts.[4–6] size, morphology, and surface defect structure, as well as the
In particular, the simultaneous enhancement of activity and electronic properties of LDHs can be modulated by various
stability/lifetime is a key criterion in the evaluation of catalytic synthetic strategies and exfoliation approaches, affording the
coordinatively unsaturated active sites (e.g., low-coordinated
steps, edges, and/or corner atoms) which facilitate heteroge-
Dr. M. Xu, Prof. M. Wei
State Key Laboratory of Chemical Resource Engineering
neous catalysis.[10,13] Moreover, the construction of nanocom-
Beijing Advanced Innovation Center for Soft Matter Science and posites with other materials can also substantially enhance the
Engineering catalytic performance of LDHs, by modulating their geometric
Beijing University of Chemical Technology structure or coupling the electronic properties of the two mate-
Beijing 100029, P. R. China rials.[14,15] In particular, taking advantage of the topological
E-mail: weimin@mail.buct.edu.cn
structural transformation of LDH precursors, a series of high-
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adfm.201802943.
performance layered mixed metal oxide (MMO) catalysts and
supported metal catalysts (monometallic or bimetallic catalysts)
DOI: 10.1002/adfm.201802943 can be fabricated; many of these materials have high catalytic

Adv. Funct. Mater. 2018, 1802943 1802943  (1 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

activity, selectivity, and stability.[8,11,12] Therefore, by virtue of


their unique combination of structural features (layered struc- Ming Xu received his
ture, compositional flexibility, tunable interlayer anions, and Bachelor’s degree in 2012
exfoliation properties) and their topological transformation, from the Hebei University
LDH materials have been widely studied as catalysts, catalyst of Technology. He joined
precursors, and catalyst supports. Professor Wei Min’s group
This Review highlights recent developments in the design as a Ph.D. candidate at
and preparation of LDH-based nanocatalysts (Figure 1) and their Beijing University of Chemical
potential applications in heterogeneous catalysis. New synthetic Technology (BUCT) in 2013.
approaches and exfoliation techniques for the preparation of His research interests cur-
LDH-based catalysts are first reviewed, with a strong emphasis rently focus on the design
on the ways in which the structure of the active sites (e.g., surface and preparation of LDH-
defect structure, interface structure, geometric/electronic struc- based catalysts with new-
ture, and preferential exposure of highly active facets) of LDH- type Strong Metal–Support Interactions (SMSI) and their
based nanocatalysts can be precisely tailored. The identification applications in the heterogeneous catalytic water gas shift
and characterization of active sites as well as structure–property (WGS) reaction.
correlations are discussed in detail, since these provide helpful
guidelines for the rational design and preparation of future high-
performance heterogeneous catalysts. In the final section, future Min Wei received her PhD
opportunities and challenges in the fabrication of LDH-derived from Peking University in
catalysts are discussed in terms of structural design and control 2001, after which she joined
over intrinsic active sites, and some strategies to resolve these the staff of Beijing University
critical issues are proposed. It is hoped that this review will focus of Chemical Technology.
attention on new LDH-based catalysts and encourage future She was promoted as full
work in this exciting and rapidly developing area. professor in 2006. She has
been a visiting scholar in
the Georgia Institute of
2. Catalysts Involving Pristine LDHs or LDH Technology (in 2008). Her
Nanohybrids main research interests focus
on: (1) 2D intercalation
In the past three decades, LDHs prepared by conventional chemistry and functional materials and (2) new catalytic
methods have attracted considerable interest in fields such as materials for energy storage and conversion.
solid base catalysis,[16–19] selective catalytic oxidation,[20,21] and elec-
trocatalysis.[22–27] In order to improve the catalytic performance
of traditional bulk LDHs, an efficient strategy is to decrease the
particle size to the nanoscale level or to exfoliate them to give
ultrathin nanosheets; this can be achieved by both reverse micro- effect” of LDH materials can substantially increase the perfor-
emulsion methods (bottom-up) and delamination approaches mance of the original LDH in base catalysis. Base-catalyzed
(top-down).[10] The fabrication of nanocomposites, which can lead reactions (e.g., aldol,[18,30,36] Knoevenagel,[16] and Claisen–
to synergistic effects between LDHs and other nanomaterials, is Schmidt condensations[37,38]) have considerable commercial
another strategy to enhance the catalytic activity of LDHs.[14,15] applications in the production of high value-added chemicals.
It is recognized that the catalytic performance of an LDH is
mainly determined by the geometric structure and electronic
2.1. Bulk LDH Materials as Catalysts properties of the basic sites. However, methods for identifica-
tion and characterization of solid base sites are rather limited,
In view of the environmental pollution and equipment cor- and finding ways to enhance catalytic activity by modulating the
rosion caused by soluble alkali catalysts, their substitution by microstructure of these basic sites remains a great challenge.
solid base catalysts offers the advantages of environmental Bing et al. recently reported that activated CaxAl-LDHs show
friendliness and facile manipulation.[28,29] LDHs have many excellent activity as solid base catalysts for the aldol condensa-
potential applications in the solid base catalysis area,[30,31] due tion reaction (Figure 2).[39] The catalysts were prepared by a facile
to their abundant surface hydroxyl groups (Brønsted type basic two-step process: CaxAl-mixed metal oxides (CaxAl-MMOs) were
sites located in the brucite-like layers). By taking advantage of obtained by calcination of CaxAl-LDH precursors, followed by a
the versatile elemental composition and structural architecture subsequent rehydration process in basic solution. The optimal
of LDHs, the structure of the base sites can be tailored. In addi- rehydrated re-CaxAl-LDH catalyst showed excellent catalytic
tion, the number and strength of basic sites can be effectively activity for the aldol condensation of isobutyraldehyde and for-
modulated by preparing the materials via calcination of LDHs maldehyde. The maximum yield of hydroxypivaldehyde was
to afford MMOs in a topotactic transformation, followed by ≈61.5%, which is superior to that obtained with conventional
rehydration of the MMOs to regenerate an LDH material,[32–35] solid base catalysts and close to that afforded by soluble alkali
often referred to as an “activated LDH;” this so-called “memory catalysts. The basic catalyst possesses sevenfold CaOH

Adv. Funct. Mater. 2018, 1802943 1802943  (2 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 1.  General overview of the ways in which LDHs can be employed in the heterogeneous catalysis field.

coordination, and a combination of advanced characterization Modulation of the electronic structure of LDH materials
techniques (X-ray photoelectron spectroscopy (XPS), extended can also influence their performance in other heterogeneous
X-ray absorption fine structure (EXAFS), and deuterated chlo- catalytic reactions. The use of LDH-based nanocatalysts for
roform–Fourier-transform infrared spectroscopy (FTIR) spectra) the splitting of water to give hydrogen and oxygen by photo-
and density functional theory (DFT) calculations confirmed that catalysis[40–43] or electrocatalysis[23–27] has attracted considerable
the sevenfold CaOH coordination site provides a weak Brøn- interest, in view of its ability to reduce our dependence on fossil
sted type basic site. Structure–activity correlation studies demon- energy resources. The oxygen evolution reaction (OER)[44–48] is
strated that such a weak basic site provides an active center for an important step in electrocatalytic water splitting. Since OER
the aldol condensation reaction, which also accelerates desorp- is a multistep reaction involving a four proton/electron-cou-
tion of the initial aldol product (i.e., inhibits further polyconden- pled process with sluggish kinetics, the development of OER
sation), accounting for the significantly enhanced selectivity com- electrocatalysts[49,50] which can decrease the overpotential and
pared with other solid base catalysts. accelerate the reaction rate is necessary. Numerous studies of
By virtue of the synergistic effects between the surface hydroxyl high-performance OER catalysts have focused on tailoring the
groups and the surface defect structure (e.g., oxygen vacancies), chemical composition and the electronic structure of active
LDHs also show excellent catalytic performance in the heteroge- sites. However, identification of the active sites and under-
neous selective oxidation of alcohols. Recently, Li and co-workers standing the catalytic mechanism are necessary in order to
reported that MgMnTi-LDH catalysts displayed good catalytic design improved OER electrocatalysts.
performance in the aerobic oxidation of benzyl alcohol (under NiFe-LDH catalysts have been shown to exhibit excellent
solvent-free conditions).[20] The optimum Mg4Mn2Ti1.0-LDH cata- electrocatalytic activity for OER,[46,48] although the identity
lyst was investigated in detail by XPS and the diffuse reflectance of the active sites remains controversial. Some studies have
infrared Fourier-transform spectroscopy (DRIFTS) combined reported that Ni serves as the active sites, while others have
with kinetic analysis, and the results showed that the excellent claimed that Fe provides the active centers. For instance, evi-
catalytic performance can be attributed to cooperative catalysis dence from in situ Mössbauer spectra, in situ X-ray absorption,
by Ti4+OH species, Mn4+OH species, and Ti3+O vacancy and in situ XPS strongly suggested the importance of Ni spe-
sites. The synergistic effect of Ti4+OH and Mn4+OH species cies in LDH catalysts for OER. Upon anodization under basic
via π-electron induction promotes the adsorption and activation conditions, abundant discrete Ni4+ domains become located in
of the benzyl alcohol molecule, while the oxygen vacancies asso- regular NiIVO6 octahedra (as a result of charge transfer from
ciated with Ti3+ species located in the host layers significantly NiO6 to FeO6 octahedra),[24,27] which serve as the optimal
accelerate the adsorption of O2 molecules. The calculated acti- active sites. Moreover, γ-NiOOH was shown to be the active
vation energy value for the Mg4Mn2Ti1.0LDH catalyst is lower phase, since it is more active than β-NiOOH and dominates
than that of other nonprecious metal catalysts, and is comparable the OER behavior. Recently, operando Mössbauer spectros-
to the values reported for noble metal catalysts. copy studies were performed to gain an insight into changes

Adv. Funct. Mater. 2018, 1802943 1802943  (3 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 2.  Schematic illustration of the structure of re-MgxAl-LDH and re-CaxAl-LDH. A) Normalized XANES spectra at Ca K-edge for (a) re-Ca2Al-LDH,
(b) re-Ca3Al-LDH, (c) re-Ca4Al-LDH, and calcium hydroxide. B) Fourier-transform EXAFS spectra at Ca K-edge for (a) re-Ca2Al-LDH, (b) re-Ca3Al-LDH,
(c) re-Ca4Al-LDH, and calcium hydroxide (solid line, experimental data; dashed line, fitting curve. Reproduced with permission.[39] Copyright 2017,
American Chemical Society.

in the electronic structure of Fe centers in NiFe-LDH catalysts of deprotonation of the OOH species adsorbed on Ni4+ sites.
under actual reaction conditions.[25] The formation of Fe4+ spe- Therefore, the presence of both Ni4+ and Fe4+ species contrib-
cies (probably located at an edge, a corner, or a defect site) was utes to the excellent OER performance of the NiFe-LDH-based
observed during the steady-state water oxidation; in contrast, no material.
Fe4+ species were detected in the reference catalyst (Fe oxide).
In addition, the Fe4+ species were surrounded by second-coor-
dination-sphere of Ni ions within a NiOOH lattice. This clearly 2.2. Ultrathin Nanosheets of LDHs as Catalysts
indicates the advantage of LDHs as a precursor for mixed metal
oxide catalysts, namely, the uniform atomic scale distribution Bulk LDH materials can be delaminated into ultrathin or even
of the metal cations in the LDH layer; this is much harder to monolayer nanosheets by two general approaches, bottom-up
achieve when using mixed precursors. or top-down.[10,53–55] Li et al. developed a facile electrosynthesis
DFT calculations have been carried out to explore the active approach[56] to fabricate ultrathin Fe based LDH nanosheets
sites and the mechanism of OER over the NiFe-LDH-based (lateral length: ≈200–300 nm; mean thickness: ≈8–12 nm)
catalysts.[51,52] Shin et al.[52] found that both Ni4+ and Fe4+ spe- within 300 s. These ultrathin nanosheet catalysts showed excel-
cies play an important role in the enhancement of OER per- lent catalytic performance in OER (overpotential: ≈224 mV at
formance: the high spin d4 Fe sites facilitate the formation of 10.0 mA cm−2) and stability, as a result of the high exposure
O⋅ intermediates, while the low spin d6 Ni centers provide the of active sites and the formation of nanoplatelet arrays. Mono­
active site for OO bond formation. In addition, the calcula- layer MgAl-LDH catalysts[57] have also been prepared by a
tions also suggested the formation of O2− species, as a result delamination approach. Recently, Song and Hu reported that

Adv. Funct. Mater. 2018, 1802943 1802943  (4 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

exfoliated single-layer nanosheets of NiFe and NiCo-LDHs[58] CoMn-LDHs by a facile coprecipitation method,[64] and showed
exhibited significantly better OER performance than the cor- they behaved as a highly active and robust OER catalyst in an
responding bulk LDH counterparts, with the catalytic activity alkaline medium (with the turnover frequency (TOF) being
even exceeding that of a state-of-the-art IrO2 catalyst. The nine times higher than that of an IrO2 catalyst). The good per-
enhancement of activity of the bulk LDHs after delamination is formance was attributed to the formation of amorphous layers
related to the increased electrochemical surface area and better at the surface, with an accumulation of active Co(IV) species.
electron transport properties. Similarly, ultrathin nanosheets of These results show that the precise modulation of the elec-
molybdate-intercalated NiFe-LDHs prepared by an exfoliation- tronic properties of OER catalysts, either by introducing a sur-
free hydrothermal approach exhibited OER current densities face defect structure or simply by decreasing the layer thickness
about three times higher than those of the corresponding bulk of the LDH precursors, can play an important role in improving
materials, resulting from the larger density of electrochemically the catalytic performance.
active sites associated with the ultrathin sheets.[59] In order to The preparation of single-layer nanosheets of NiCo-LDHs by
further improve the exposure of active sites, Zhang et al. devel- a liquid phase exfoliation method has been reported recently.
oped a one-step in situ growth approach to prepare NiFe-LDH The materials showed significantly higher OER activity than
hollow microspheres,[60] which showed high electrocatalytic the corresponding bulk LDHs, resulting from increased
performance for OER. Electrochemical studies demonstrated number of active edge sites and improved electronic conduc-
that the ultrathin hollow structure can also enhance the ion tivity.[62] However, the liquid exfoliation method shows dis-
transport kinetics. More recently, Lou and co-workers reported advantages of long reaction times, low efficiency, and solvent
a facile self-templated approach for the preparation of hollow pollution. Therefore, it is necessary to develop facile, clean,
nanoprisms composed of ultrathin NiFe LDH nanosheets,[61] and efficient approaches to synthesize monolayer nanosheets
which showed high electrocatalytic performance in OER, with LDHs for use as OER catalysts. For example, Wang et al.
a relatively low overpotential and excellent stability: this was reported a dry exfoliation method for the synthesis of mono­
attributed due to the very large surface area and tailored com- layer CoFe-LDH nanosheets involving Ar plasma etching
position. These studies show that the construction of ultrathin (Figure 3).[65] Compared with traditional liquid exfoliation
or single-layer nanosheets is an efficient way to substantially methods, this strategy is clean, creates no toxic waste, and
enhance the catalytic performance of LDHs in OER; this is due avoids the adsorption of solvent molecules. Moreover, the as-
to simultaneous increases in the degree exposure of active sites, synthesized monolayer CoFe-LDH nanosheets exhibited high
the electrochemical surface area, and the ion transport kinetics. activity for electrocatalytic OER, requiring a lower overpotential
In addition, the electrosynthesis approach offers a new path (266 mV) to reach a current density of 10 mV cm−2 than the
to prepare ultrathin LDH materials amenable to large-scale corresponding bulk CoFe-LDHs (321 mV). This high activity
production. was attributed to the large number of exposed surface multiple
Precise modulation of the electronic structure of LDHs can vacancies (including O, Co, and Fe vacancies). Most impor-
significantly influence their catalytic performance. This can be tantly, the introduction of oxygen vacancies and metal vacan-
achieved by introducing surface defect structures (especially cies simultaneously into monolayer LDH nanosheets can mod-
oxygen vacancies) into the host brucite-like layers. Coordina- ulate the electronic structure of the materials. The presence
tively unsaturated metal cations (located in the edge-sharing of oxygen vacancies decreases the adsorption energy of inter-
MO6 octahedra) can be obtained by reducing the layer thick- mediates and improves their stability, while the Co vacancies
ness to atomic length scales of about 1 nm. Recently, Jin and in the ultrathin nanosheets modulate the surface electronic
co-workers reported that ultrathin NiCo-LDH nanosheets pre- structure with change in valence state from Co2+ to Co3+. In
pared by chemical exfoliation exhibited a high catalytic activity order to further improve the activity of OER, N-doped ultrathin
for OER.[62] The enhancement of catalytic activity obtained CoFe-LDH nanosheets with edge-rich sites have been pre-
by chemically exfoliating LDH nanoplates to afford ultrathin pared by N2 plasma etching.[66] The material exhibited excellent
nanosheets was attributed to a change in electronic structure activity for OER (with an ultralow overpotential of 233 mV at
resulting from the introduction of surface defects. Ultrathin 10 mA cm−2) attributed to the introduction of surface defects.
ZnAl-LDH nanosheets have been synthesized by a facile This dry exfoliation approach opens up a new path for the syn-
bottom-up approach (an inverse microemulsion technique).[63] thesis of ultrathin LDH materials.
The resulting materials possess abundant coordinatively In summary, reducing the layer thickness of LDHs to afford
unsaturated Zn ions due to the high density of oxygen-vacancy ultrathin nanosheets or single-layer nanosheets can substan-
defects confined in the host brucite-like layers, resulting in sig- tially improve their catalytic activity. The ultrathin or monolayer
nificantly enhanced photocatalytic activity for the conversion nanosheets can be efficiently prepared by liquid phase exfolia-
of CO2 to CO in the presence of water vapor. EXAFS spectra tion methods, electrosynthesis or dry exfoliation methods. The
and positron annihilation measurements, combined with DFT excellent catalytic performance of ultrathin LDHs can be attrib-
calculations, conclusively demonstrated the formation of coor- uted to the high exposure of active sites and/or defect structures
dinatively unsaturated Zn (Zn+–Vo complexes) after decreasing (oxygen vacancies and metal vacancies) in the brucite-like host
the layer thickness. The defect structure serves as the active site layers. Most importantly, the electronic properties resulting
for the adsorption and activation of CO2 and H2O molecules from the introduction of a defect structure can significantly
and simultaneously promotes photoinduced charge separa- enhance the adsorption of reactant molecules and improve the
tion, which significantly enhances the rate of photocatalytic stability of reaction intermediates, which results in substantial
CO2 reduction. Song and Hu prepared ultrathin nanosheets of enhancements in catalytic performance.

Adv. Funct. Mater. 2018, 1802943 1802943  (5 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 3.  Schematic illustration of CoFe-LDH nanosheets prepared by Ar plasma exfoliation; AFM images of A) bulk CoFe-LDHs and B) ultrathin
CoFe-LDH-Ar nanosheets. C) The corresponding height curves; D) X-ray diffraction (XRD) patterns of bulk CoFe-LDHs and ultrathin CoFe-LDH-Ar
nanosheets. All panels reproduced with permission.[65] Copyright 2017, Wiley.

2.3. LDH-Based Nanohybrids as Catalysts nanocomposites[14,15] with significantly enhanced catalytic perfor-
mance, due to the synergistic effects between the components.
The combination of LDH materials and a variety of other Moreover, this approach offers an opportunity to engineer sophis-
building blocks (e.g., carbon nanotubes (CNTs), graphene, gra- ticated geometric structures with tunable electronic properties.
phene oxide (GO), and metal oxides) can give rise to hierarchical Great effort has been devoted to developing facile and cost-effective

Adv. Funct. Mater. 2018, 1802943 1802943  (6 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

methods for the preparation of such nanocomposites. A series methods or electrosynthesis, have attracted considerable atten-
of creative fabrication methods have been reported, including tion.[15] Recently, Dou et al. reported a facile two-step approach to
grafting LDH nanosheets on nanocarbon,[67–72] forming fabricate TiO2@CoAl-LDH core–shell nanospheres: preparation
core–shell architecture structures,[73–78] and self-assembly of of TiO2 hollow nanospheres via hydrothermal synthesis followed
ultrathin single-layer LDH nanosheets on 3D materials.[79] by in situ growth of a CoAl-LDH shell (Figure 4). The hybrid
Dai and co-workers were the first to report that ultrathin NiFe- material exhibited an extraordinary high activity for photocata-
LDH nanoplatelets with good crystallinity can be grown on mildly lytic water oxidation,[76] with the oxygen gas generation rates
oxidized multiwalled CNTs.[67] The resulting NiFe-LDH/CNT being 2.34 and 2.24 mmol h−1 g−1 under full sunlight (>200 nm)
nanocomposite exhibited a high OER electrocatalytic activity with and visible light (>420 nm), respectively. The observed full utili-
a TOF of 0.56 s−1 at an overpotential of 300 mV in 1 m KOH, which zation of sunlight results from the combination of the visible-
is superior to that of commercial precious metal Ir catalysts. More- light-responsive LDH shell and the UV-light-responsive TiO2
over, the catalytic performance was approximately three times core. Electron transfer from the LDH shell to the TiO2 core
higher than that of previously reported mixed nickel and iron was observed by XPS, and was confirmed by Bader charge
oxide electrocatalysts. The catalyst also exhibited good durability in analysis (with a charge transfer of 0.62 electrons). The photo­
alkaline solutions. The strong interaction between the LDH and luminescence (PL) signal almost disappeared for the TiO2@LDH
CNTs was verified by X-ray photoelectron spectroscopy, with the nanospheres, indicating that the recombination of photoinduced
substantial increase in the carbonyl π∗ peak intensity (located at electron–hole pairs in this heterostructure is significantly inhib-
≈288.5 eV) indicating the formation of MOC bonding via the ited. This gives rise to an enhanced carrier transfer efficiency,
carboxyl group, accompanied by large perturbations of the carbon which facilitates the photocatalytic performance. DFT calcula-
atom. This work provides a new approach to fabricate low-cost tions verified that the formation of the heterostructure leads to
and high-performance oxygen evolution electrocatalysts. an efficient spatial separation of electrons and holes: the strong
Subsequently, Yang and co-workers reported another strategy donor–acceptor coupling and suitable band matching between
to construct NiFe-GO LDH hybrids, by taking advantage of the the TiO2 core and LDH shell play an important role in the
electrostatic attraction between positively charged exfoliated enhancement of photocatalytic activity for O2 generation.
NiFe-LDH nanosheets and negatively charged GO layers,[70] In summary, the fabrication of nanocomposites based on
resulting in the formation of hybrid sheets. The as-prepared LDHs and other 2D layered materials is an efficient way to
hierarchical NiFe-GO LDH catalyst shows an extraordinarily enhance the catalytic performance by taking advantage of
high electrocatalytic activity for OER with an overpotential as the interfacial synergistic effects between the two materials.
low as 210 mV and Tafel slope of 40 mV dec−1, which is supe- Moreover, the strong coupling between LDHs and nanocarbon
rior to the values for NiFe-LDHs, NiFe-LDHs/GO (NiFe-LDHs materials facilitates the charge transport at the interface (i.e.,
nucleated and grown on GO), and NiFe-LDHs+GO (a physical improves the conductivity), which contributes to the high cata-
mixture of NiFe-LDHs and GO). The enhanced electrocata- lytic activity. However, finding ways of effectively inhibiting the
lytic performance was attributed to the high exposure of active restacking of nanosheets and improving the contact between
sites and the synergistic catalysis between NiFe-LDHs and GO. the reactant and active sites are prerequisites if there are to be
Furthermore, NiFe-rGO hybrids were obtained by reduction of further improvements in catalytic performance.
NiFe-GO LDH, and shown to exhibit even better electrocatalytic
activity for OER: the overpotential decreased to 195 mV, and the
TOF reached 0.98 s−1 at an overpotential of 300 mV. The catalytic 3. MMO Based Nanocatalysts
performance shows a significant enhancement compared with
earth-abundant metal hydroxide catalysts and is comparable to As discussed above, MMOs can be obtained by calcination of
that of noble-metal-based electrocatalysts. This was attributed to LDHs in a topotactic transformation.[8,11] A series of studies
the strong interaction between NiFe-LDHs and rGO giving rise have demonstrated that the exposure of crystal facets and defect
to a molecular level assembly, which facilitates charge transport. structure (e.g., oxygen vacancies) of MMOs can be precisely con-
Hierarchical hybrid architectures involving a metal oxide core trolled by varying the conditions for the synthesis of the LDH
and an LDH shell, which can be prepared via hydrothermal precursors and/or the calcination step. In addition to acting as

Figure 4.  Schematic illustration of the fabrication of TiO2@CoAl-LDH core–shell nanospheres for oxygen evolution from water splitting. Reproduced
with permission.[76] Copyright 2015, Wiley.

Adv. Funct. Mater. 2018, 1802943 1802943  (7 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

catalysts in their own right, MMOs offer many advantages as NiTi-MMO, which is closely related to the formation of oxygen
catalyst supports, including a high specific surface area, control- vacancies. Furthermore, the presence of Ti3+ species was revealed
lable composition and defect structures (e.g., oxygen vacancies), by XPS (a 2p3/2 peak at 458.2 eV), consistent with the EXAFS
and tunable quantity and strength of acid/basic sites. spectra. Therefore, the geometric/electronic structure of mono­
layer NiTi-MMO (the predominance of Ni3+ species, Ti3+ species,
and abundant oxygen vacancy defects) gives rise to a high elec-
3.1. MMOs as Catalysts trical conductivity and facilitates charge transfer, accounting for
its enhanced electrocatalytic activity for water splitting.
Metal oxides are extensively employed in heterogeneous
catalysis.[80–85] It has been shown that modulation of crystal facets
and surface defect structures of metal oxides can have a signifi- 3.2. MMOs as Catalyst Supports
cant influence on their catalytic performance. In this respect,
MMOs[8,86] offer considerable advantages, since the exposure of MMOs have been frequently used as supports for hetero-
crystal facets and defect structure can be controlled by varying the geneous catalysts.[87–95] Recent studies have shown that the
topotactic transformation parameters (e.g., heating rate, calcination acidity and basicity of MMO supports can be controlled by
temperature, and calcination atmosphere) of the LDH precursors. tuning the composition of the brucite-like host layers in the
He et al. reported a facile in situ topotactic preparation of LDH precursor. Moreover, the synergistic interactions between
ZnO nanocatalysts from ZnAl-LDH precursors. A ZnAl-LDH supported metal (single metal or alloy) particles or subnano
flower-like shell was grafted onto amorphous Al2O3 micro- clusters and the acid/base sites of the MMO supports can lead
spheres by an in situ growth method, followed by calcination to remarkable enhancements of catalytic performance.
to afford ZnO hexagonal nanoplatelets (diameter: 10–50 nm; Li and co-workers reported that PdAg catalysts supported
thickness: 6–8 nm) immobilized on a hierarchical flower- on various reducible Mg–Ti mixed oxides with tunable acidity/
like Al2O3 matrix.[81] High resolution transmission electron basicity exhibited high performance and selectivity for the
microscopy (HRTEM) and selected area electron diffraction selective hydrogenation of acetylene.[88] The quantity and
ZnO showed that ≈50% of the exposed facets are ZnO(0001). strength of the acid and basic sites can be controlled by simply
The as-obtained ZnO-based MMOs exhibit excellent photo- changing the Mg/Ti ratio, which significantly influences the
catalytic performance in the degradation of Rhodamine B catalytic activity of the resulting PdAg/MgxTi1−xOy catalysts.
under visible light irradiation. This was attributed to the pref- The material with the optimal composition PdAg/Mg0.5Ti0.5Oy
erential exposure of the highly active ZnO(0001) facets, which exhibits excellent hydrogenation performance at relatively low
have abundant oxygen vacancies and a relatively low bandgap temperature (conversion: ≈99%; selectivity: ≈83.8%; 70 °C). In
(≈3.0 eV) in the UV–visible diffuse reflectance spectrum. More- situ DRIFTS (e.g., pyridine/CO2 DRIFTS) combined with
over, the resulting ZnO hexagonal nanoplatelets exhibit a high temperature-programmed desorption (TPD) experiments (e.g.,
degree of dispersion on the substrate, and thus offer a larger CO2-TPD and NH3-TPD) showed that both the medium acid sites
number of surface active sites than other ZnO catalysts. This and weak basic sites promote the hydrogenation of acetylene.
work provides a new approach to fabricate highly dispersed The selectivity was attributed to two factors: i) the high alloying
oxide catalysts with preferential exposure of highly active facets. degree of PdAg increases the number of linearly coordinated Pd
More recently, a single-layer NiTi-MMO with a high per- sites, which facilitates desorption of ethylene and prevents further
centage exposure of (110) facets was successfully prepared via hydrogenation to ethane; ii) the increased Pd electronic density
topotactic transformation of single-layer NiTi-LDH nanosheets arising from the electron transfer from basic sites and Ti3+ spe-
(Figure 5).[85] The monolayer NiTi-LDH nanosheets with cies also promotes the desorption of ethylene.
mean platelet diameter of ≈20 nm and mean thickness Recent studies have mainly focused on the fabrication of
of ≈0.8 nm were synthesized by a reverse microemulsion highly dispersed metallic catalysts (e.g., single-atom, pseudo-
method. The ultrathin NiO nanosheet catalysts derived from single-atom, or subnano clusters), since these materials often
monolayer NiTi-LDH precursors show high electrocatalytic show excellent catalytic performance in a series of industrially
activity for H2O oxidation. X-ray absorption near-edge struc- important reactions.[96–99] Although some successful methods
ture (XANES) spectroscopy shows that the absorption edge have been developed to enhance the dispersion of active
of the monolayer NiTi-MMO shifts to higher energies with metal sites (e.g., a mass-selected soft-loading technique,[100] a
less oscillation compared with NiO and bulk NiTi-MMO, wet-chemistry approach,[101–103] or a support surface induced
resulting from a higher average Ni oxidation state. The pres- method[104]), the preparation of stable single-atom or pseudo-
ence of Ni3+ species were confirmed by electron spin reso- single-atom catalysts remains a considerable challenge. He
nance (ESR) and XPS. Moreover, EXAFS demonstrates that and co-workers developed an efficient approach to fabricate
the first NiO shell of monolayer NiTi-MMO has a distance stable single-atom Pt, pseudo-single-atom Pt, and small Pt
of ≈2.05 Å and coordination number of ≈5.70, both of which cluster catalysts.[92,93] Their method involved the introduction
are smaller than those of bulk NiO (NiO distance: ≈2.08 Å; of Sn4+ cations into the brucite-like layers of an LDH with an
coordination number: 12.00) and bulk NiTi-MMO (NiO atomic-level dispersion. After calcination to form an MMO,
distance: ≈2.08 Å; coordination number: 7.70), due to the single-atom Pt precursors or small Pt clusters (fewer than
presence of nickel vacancies (VNi) in the NiO nanosheets. Nor- 10 atoms) were immobilized via electrostatic attraction between
malized XANES spectra at the Ti K-edge further confirmed negatively charged surface sites and cationic Pt precursors. The
that the TiO6 octahedra are disordered in the monolayer high degree of dispersion of Pt species (single atoms and small

Adv. Funct. Mater. 2018, 1802943 1802943  (8 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 5. A) TEM image of monolayer NiTi-LDH nanosheets. B,C) TEM and HRTEM images of monolayer NiTi-MMO. D) FFT pattern of the
corresponding NiO area. E) Geometrical model of the hexagonal NiO nanocrystal. F) AFM image, and G) the corresponding height profiles of
monolayer NiTi-MMO. The numbers 1 to 3 in part (G) correspond to the line scan numbers in part (F). Reproduced with permission.[85] Copyright
2016, American Chemical Society.

Adv. Funct. Mater. 2018, 1802943 1802943  (9 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

clusters with a Pt dispersion of 96%) was demonstrated by high- metal–metal bond during the transformation of LDH precur-
angle annular dark-field scanning transmission electron micros- sors can lead to the production of metal nanoparticles with
copy (HAADF-STEM) combined with EXAFS. The optimal selective exposure of specific facets and defects. Finally, using a
0.3 wt% Pt/Mg(Sn)(Al)O@Al2O3 catalyst showed high selectivity single precursor leads to the formation of strong metal–support
for cyclization during the n-heptane reforming reaction. By interactions which should increase catalyst stability. Recently,
contrast, a 2 wt% Pt/Al2O3 catalyst with large Pt particles gave He et al. developed a facile approach to prepare a surface defect-
preferential selectivity for C1 products rather than C7. There- promoted Ni-based nanocatalyst using a flowerlike NiAl-LDH
fore, the size of the Pt particles clearly plays an important role shell formed on an amorphous Al2O3 core via an in situ hydro-
in determining the performance and selectivity of catalysts for thermal method, followed by reduction in hydrogen.[114] X-ray
the n-heptane reforming reaction. This work shows that intro- diffraction (XRD) combined with STEM images reveals that the
ducing high-valent metal ions such as Sn4+ into the brucite-like supported Ni nanoparticles have a high exposure of (111) planes
host layers of LDH precursors provides a facile way to pre- with an average particle size of ≈5.0 nm and extremely high par-
pare highly dispersed (even single-atom) noble metals with ticle density (≈2.4 × 1016 particles m−2), but retain a high degree
promising catalytic behavior. of dispersion. When used as a catalyst for the CO2 methanation
reaction, the optimal Ni/HAl2O3(400) material exhibits excel-
lent performance at a relatively low temperature and pressure
4. Supported Metal Catalysts Derived from LDH (conversion: ≈99% at 300 °C, TOF value: 2.4 × 10−3 s−1 at 220 °C,
selectivity >99%) and the performance is superior to that of
Precursors
supported noble metal (e.g., Ru or Pd) catalysts. An obvious
Supported metal catalysts are extensively employed in many distortion of the lattice fringes of the Ni nanoparticles was
industrial reactions including ammonia synthesis,[105] selec- demonstrated by HAADF-STEM and HRTEM (e.g., bending
tive hydrogenation,[106–108] selective oxidation,[109] and steam and dislocation), indicating the presence of abundant surface
reforming.[110–113] Great effort has been focused on tailoring defects. Positron annihilation spectroscopy (PAS) showed that
the metal active sites so as to simultaneously enhance the cata- the optimal Ni/HAl2O3(400) catalyst had the highest relative
lytic performance, selectivity, and stability. Previous studies concentration of Ni vacancy clusters, consistent with its excel-
have verified that control over composition, particle size, mor- lent low-temperature performance. EXAFS spectroscopy was
phology, dispersion, electronic property, and defect structure of carried out to reveal the structure–activity relationship between
supported metal nanoparticles can tune the adsorption and acti- Ni vacancy clusters and the catalytic performance, and the
vation of reactant molecules, which improves catalytic activity results demonstrated that the Ni/HAl2O3(400) catalyst had
and selectivity.[8,11] Furthermore, the construction of supported the lowest average NiNi coordination number, consistent with
bimetallic catalysts is an efficient strategy for simultaneously the PAS results. In addition, the Ni/HAl2O3(400) catalyst shows
tailoring the geometric and electronic structure of catalytic sites. long-term catalytic stability: CO2 conversion remained constant
In addition to control over the metal sites themselves, tuning over 252 h and no obvious sintering or aggregation of Ni nano-
the metal–support interfacial structure also plays a crucial role particles was observed. The high stability was attributed to the
in improving catalytic performance, by virtue of the synergistic strong anchoring of the Ni nanoparticles on the Al2O3 support.
interactions between metal particles and acid/base sites or Since the first report[118] by Tauster et al. in 1978, the strong
oxygen vacancies on the supports. In particular, the preparation metal–support interaction (SMSI) has attracted considerable
of supported metal catalysts with tunable strong metal–support interest and has been shown to play a crucial role in heteroge-
interactions (involving electron transfer) has been recognized neous catalysis.[5,6,115–117] Tauster showed that supporting VIII
as an important way to enhance catalytic properties. The calci- group metals on TiO2 markedly inhibited their ability to adsorb
nation and subsequent reduction of LDH precursors has been hydrogen and carbon monoxide. The SMSI effect was attributed
widely used to prepare metal oxide-supported monometallic to the migration of partially reduced titania on the surface of the
catalysts, bimetallic alloy catalysts, and intermetallic compound metal at an elevated reduction temperature.[119,120] Recent studies
catalysts with tunable particle size, morphology, and geometric/ have focused on new types of SMSI,[121–125] which differ from the
electronic structure. In this section, we summarize methods for original. Rodriguez and co-workers reported a new type of SMSI
the rational design and preparation of supported metal catalysts in Pt/CeOx/TiO2(110) catalysts, which involved electronic pertur-
using LDH precursors and review our current understanding bation of small Pt particles in contact with ceria, leading to a sig-
of structure–activity relationships. nificantly enhanced dissociation of the OH bonds in water.[122]
Campbell used DFT calculations to study the electronic inter-
action between platinum and reduced ceria, and proposed the
4.1. LDH-Based Monometallic Catalysts new concept of electronic metal–support interactions (EMSI).[126]
Therefore studies of how to effectively regulate SMSI and thus
Employing the calcination of LDH precursors followed by optimize the properties of heterogeneous catalysts are of great
reduction (e.g., under a hydrogen atmosphere) to prepare sup- interest. Calcination of LDH precursors followed by reduction
ported monometallic catalysts offers three advantages:[7,8,11,12] offers one promising way to achieve this goal.
First, due to the atomic-level dispersion of divalent and trivalent Xu et al. used a NiTiLDH precursor to prepare TiO2−x-
metal cations in the brucite-like layers, the reduction process modified Ni nanocatalysts. HRTEM images demonstrate
generally results in a high dispersion of metal nanoparticles. that the highly dispersed Ni particles were partially encapsu-
Second, the cleavage of metal–oxygen bonds and formation of lated by a TiO2−x overlayer (Figure 6). They showed that the

Adv. Funct. Mater. 2018, 1802943 1802943  (10 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 6. A1–D1) TEM images of four Ni@TiO2−x samples (inset: the histogram of size distribution of Ni nanoparticles). A2–D2) HRTEM images of a
single Ni nanoparticle selected from (A1)–(D1), respectively. All panels reproduced with permission.[127] Copyright 2017, American Chemical Society.

metal–support interaction between Ni and the TiO2−x support (denoted as Ni/NiAl-MMO) using calcination and reduction of
could be controlled by varying the synthesis parameters (e.g., a NiAl-LDH precursor.[128] The quantity and strength of acid/
heating rate, calcination and reduction temperature, and reduc- basic sites could be efficiently controlled via precisely tuning the
tion atmosphere).[127] In situ EXAFS, in situ DRIFTS, and XPS synthesis parameters (e.g., reduction temperature, atmosphere,
confirmed the formation of electron-rich Ni species (Niδ−), heating rate, flow rate). A cubic NiO-like phase (Al3+-doped
resulting from electron transfer from the TiO2−x support to NiO structure) of Ni/NiAl-MMO catalysts was identified by in
interfacial Ni by SMSI during the reduction of NiTi-MMO. The situ EXAFS, in situ Raman spectroscopy, and in situ XRD. The
as-obtained optimized Ni@TiO2−x(450) catalyst exhibited excel- presence of medium-strong acid–base sites (Niδ+Oδ− pairs)
lent catalytic activity (reaction rate: ≈356.8 µmolCO gcat−1 s−1; was demonstrated by CO2-TPD and NH3-TPD. In situ EXAFS
TOF value: ≈3.8 s−1) in the water gas shift (WGS) reaction; spectra demonstrate that the first NiO shell distance of the
these values are much higher than those for Ni-based catalysts NiO-like phase (≈2.05 Å) is smaller than that of the NiO refer-
prepared by conventional methods. Moreover, XPS combined ence material (2.08 Å); while the second NiNi shell (NiOAl)
with XANES spectra at the Ti–L edge demonstrated that the bond is significantly larger than that of the NiO reference owing
strong interaction between Ni particles and the TiO2−x sup- to a severe distortion of the octahedral NiO coordination envi-
port gives rise to the formation of abundant oxygen vacancy ronment, resulting from Al3+ doping and the formation of
(OvTi3+) sites and Niδ−OvTi3+ interface sites, which serve as NiOAl bonds. When used as catalyst for the dehydrogena-
the dual-active-site for the WGS reaction (Ni metal sites activate tion of 2-octanol to 2-octanone under oxidant-free conditions,
CO, while H2O molecules dissociate at oxygen vacancies). Fur- the optimal Ni/NiAl-MMO(400) material shows much higher
thermore, in situ DRIFTS spectra (using CO as a probe mol- activity than a conventional Ni/Al2O3 catalyst (with the TOF
ecule) were carried out to identify the strength of the SMSI at values being 174.4 ± 1.59 h−1 and 55.2 ± 1.11 h−1, respectively).
Niδ−OvTi3+ interfacial sites in different materials. The rela- Operando time-resolved EXAFS spectra combined with kinetic
tionship between the strength of the SMSI at Niδ−OvTi3+ isotope effect mesurements were carried out to study the struc-
interfacial sites and the catalytic performance was revealed by ture–property correlation: the presence of Ni0 sites accelerates
quantitative analysis of reaction kinetics models. The results α-CH bond cleavage, while the presence of medium-strong
demonstrate that interfacial sites (Niδ−OvTi3+) with mod- acid–base sites facilitates OH bond cleavage. This work clearly
erate SMSI afford the maximum activity in the WGS reaction. shows careful choice of the reaction conditions during calcina-
The synergistic interactions between metal sites and acid– tion and reduction of LDH precursors and offers an effective
base sites on supports have been shown to play an important way to tailor the synergistic interactions between the metal and
role in many heterogeneously catalyzed reactions, including acid/base sites in the resulting supported metal catalysts.
hydrogenation of α,β-unsaturated aldehydes, partial hydrogena- Generally, the surface defect structure (e.g., oxygen vacancies)
tion of acetylene, and dehydrogenation of alcohols. Recently, of a support determines the nature of surface acid/base sites,
Chen et al. reported the preparation of acid–base-promoted and thus has a significant influence on catalytic performance.
Ni nanocatalysts supported on NiAl-mixed metal oxides Hu et al. prepared Mn-containing spinel-supported copper

Adv. Funct. Mater. 2018, 1802943 1802943  (11 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

nanocatalysts via calcination and reduction of a CuMnAl-LDH Cu to Co, indicating a strong electronic interaction between two
precursor.[129] XPS and H2-temperature-programmed reduc- metallic species. In situ CO-FTIR spectra demonstrated that the
tion (TPR) experiments confirmed the presence of highly adsorbed CO molecule gradually migrated from a bridge-type
dispersed Cu0 nanoparticles, together with abundant surface to linear-type adsorption as the Cu content increased; this was
oxygen vacancies and Mn2+ species. The optimal catalyst dis- attributed to electron transfer and active-site isolation effects. It
played exceptional performance in the catalytic hydrogenation was found that Cu-rich catalysts with predominantly bridge-type
of various biomass-derived compounds containing carbonyl CO adsorption give higher selectivity for hydrocarbons, while
groups (e.g., gas-phase hydrogenation of dimethyl succinate Co-rich catalysts favor the formation of oxygenated compounds.
(DMS) to γ-butyrolactone). The correlation between the catalytic The following reaction mechanism was proposed: negatively
reaction rate and parameters such as the Cu+/(Cu++Cu0) ratio charged Co sites facilitate the dissociative adsorption of CO,
and number of oxygen vacancies was investigated in detail. The CC chain growth, and the formation of *CnHz groups, while
results showed that surface oxygen vacancies (defect structure) the electron-deficient Cu sites serve as the active site for asso-
contribute to catalytic activity, and in situ FTIR measurements ciative adsorption of CO to afford CO* species. The construc-
of DMS adsorbed on different catalysts demonstrate that the tion of core–shell structures via the calcination and subsequent
surface oxygen vacancies adjacent to metallic copper species reduction of LDH precursors is thus clearly a promising way
substantially promote the adsorption of DMS molecules. Simul- to maximize the synergistic effects in bimetallic alloy catalysts.
taneously, Mn2+ species are oxidized to Mn3+ by the CO group The rational design and modulation of the surface/inter-
with concomitant formation of CO σ bonds, which deceases face microstructure of bimetallic catalysts is another effective
the barrier for DMS hydrogenation. This work shows that the strategy to enhance catalytic performance, and a core–shell
synergistic interactions between active metal sites and the sur- nanostructure containing a transition metal core encapsu-
face defect structure of the support can significantly enhance lated by a noble metal based alloy shell is a good demonstra-
the activity and selectivity of catalytic hydrogenation reactions. tion of this strategy. Recently, Li et al. reported the prepara-
tion of a Ni@(RhNi-alloy) nanocatalyst with a Ni metal core
and a RhNi-alloy shell (≈2 nm) using an LDH precursor.[138]
4.2. LDH-Based Bimetallic Catalysts The synthesis involved the following three steps: 1) the for-
mation of highly dispersed Ni nanoparticles supported on an
The geometric and electronic effects generated in supported Al2O3 substrate obtained by calcination and hydrogenation of
bimetallic catalysts can lead to them having significantly a NiAl-LDHs precursor; 2) preparation of a Ni@Rh/Al2O3 cata-
improved catalytic performance when compared with the indi- lyst by chemical etching of this material with RhCl3 solution; 3)
vidual component metals alone. Examples include selective calcination of Ni@Rh/Al2O3 to afford Ni@(NiRh-alloy)/Al2O3
hydrogenation,[130–133] steam reforming,[112,134] and the conversion catalysts. HRTEM images reveal the surface defect structures
of synthesis gas.[135,136] In particular, the selectivity of catalytic of the Ni@(NiRh-alloy) nanocatalyst (e.g., the lattice disorder
reactions can be significantly enhanced as a result of selective at the core–shell interface and the contraction/expansion of the
adsorption and activation of functional groups in the reactant lattice fringe in the shell). The material has excellent catalytic
molecules. Bimetallic catalysts can be divided into bimetallic performance in hydrogen generation from N2H4BH3 (a rate of
alloys, bimetallic heterostructures, and intermetallic compounds 1.2 mol H2 per min⋅mol⋅metal). The utilization efficiency of the
(IMCs). Since LDHs can be prepared containing two (or more) noble metal is substantially enhanced, since it is only present in
transition metal ions in the layers, calcination and reduction of the alloy shell rather than the whole nanoparticle, and this method
such LDH precursors should be an attractive way to prepare uni- can be extended to the preparation of other noble metal catalysts.
formly and highly dispersed supported bimetallic catalysts.[8,11,12] In addition to achieving high dispersion of supported alloy
nanoparticles, another challenging goal is increasing cata-
lyst stability by inhibiting the sintering of active species. He
4.2.1. Bimetallic Alloy Catalysts and co-workers prepared a CoGa-alloy catalyst by calcina-
tion of a CoGaZnAl-LDH precursor in a reducing atmosphere
Gao et al. prepared core–shell structure Cu@(CuCo-alloy) (Figure 7).[139] HRTEM images showed a mean particle size of
nanocatalysts embedded on a Al2O3 matrix via a facile two-step 12.3 nm with a high degree of exposure of (111) facets. Interest-
process: an in situ growth of a CuCoAl-LDHs precursor on an ingly, HAADF images with element mapping and linescan pro-
aluminum substrate followed by calcination and reduction.[137] files demonstrated that CoGa nanoparticles become trapped in
HRTEM and HAADDF-STEM images demonstrated that the the MMO support formed during the topotactic transformation
as-obtained CuCo-alloy catalysts (nanoparticles: 15–20 nm of the CoGaZnAl-LDH precursor. The as-obtained CoGa-alloy
in diameter) show a core–shell structure with a uniform and catalyst exhibits excellent catalytic performance in syngas conver-
homogeneous distribution of Cu and Co (Cu in the core with sion to higher alcohols (43.5% CO conversion and 59% selectivity
a CuCo-alloy as the exterior shell). When compared with for alcohols). Moreover, the formation of the trapped structure
powdered-CuCo/Al2O3, Co/Al2O3, and Cu/Al2O3 catalysts, the helps to increase catalyst stability by inhibiting sintering (the CO
core–shell Cu@(CuCo-alloy)/Al2O3 catalysts showed signifi- conversion and selectivity for alcohols remained stable during
cantly better catalytic performance in the hydrogenation of reaction for 100 h). Co K-edge EXAFS, CO-FTIR, and CO-TPD
CO to higher alcohols The C6+ 1-alcohol slate selectivity of the measurements demonstrated that the formation of the CoGa
optimal catalyst (Cu/Co = 1/2) reached 48.9% at a CO conver- alloy promotes the electron transfer from Ga to Co, and the
sion of 21.5%. XPS provided evidence of electron transfer from resulting electron-enriched Co serves as the site for dissociative

Adv. Funct. Mater. 2018, 1802943 1802943  (12 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 7.  Schematic illustration of the preparation of trapped CoGa particles using LDH precursors. A) XRD patterns of (a) γ-Al2O3, (b) CoZnGaAl-
LDHs/γ-Al2O3, (c) CoZnAl-LDHs/γ-Al2O3, (d) ZnGaAl-LDHs/γ-Al2O3, and (e) ZnAl-LDHs/γ-Al2O3. B) SEM image of CoZnGaAl-LDHs/γ-Al2O3
with cross-sectional image inset. C) SEM–EDS elemental mapping images for Co, Ga, and Zn of the corresponding region (yellow frame in (B)) of
CoZnGaAl-LDHs/γ-Al2O3. All panels reproduced with permission.[139] Copyright 2017, Elsevier.

adsorption of CO, which facilitates the CC bond coupling form a new bimetallic heterostructure. Recently, Chen et al.
needed for carbon-chain growth. In addition, the isolated Co reported a defect-induced synthesis of heterogeneous bime-
site (Co atom separated by contiguous Ga atoms) is responsible tallic RuNi catalysts, in which Ru clusters (1.1 ± 0.4 nm) were
for the linear nondissociative adsorption of CO, promoting the immobilized on the surface of Ni nanoparticles, as confirmed by
CO insertion step and the formation of alcohols. Therefore, the HRTEM (Figure 8).[140] PAS further confirmed that Ru clusters
trapped structure leads to the maximum matching of CO insertion are anchored on the surface of Ni vacancy cluster defects, which
and CC coupling, accounting for the very high catalytic activity enhances the average electron density of Ni, as revealed by
and exceptional selectivity for alcohols in syngas conversion. EXAFS. Moreover, HRTEM combined with Fourier-transform
EXAFS spectra verified that the formation of a RuNi inter-
face induces an interfacial strain due to the lattice expansion
4.2.2. Bimetallic Heterostructure Catalysts of Ni nanoparticles. The as-obtained Ru0.8–Ni/H–Al2O3 catalyst
showed excellent catalytic performance in the ethanol steam
Precise modulation of the surface defect structure of supported reforming reaction, with a H2 yield of 4.2 molH2/molEtOH at
metal nanoparticles should have a significant influence on cata- 350  °C. In situ FTIR measurements and catalytic evaluation
lytic selectivity. A facile and effective approach is to immobilize confirmed that the RuNi interface substantially promotes
clusters of a second metal onto the surface of nanoparticles of a CC bond cleavage and efficiently inhibits the CO bond rup-
different metal, so as to modify the original surface defects and ture during the reaction, accounting for the very high H2 yield.

Adv. Funct. Mater. 2018, 1802943 1802943  (13 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 8.  HRTEM images of Ni/H–Al2O3: A) an overall image, B) one nanoflake with embedded Ni nanoparticles (inset: lattice fringes of Ni nanopar-
ticle), C–E) a single Ni nanoparticle (the scale bar is 1 nm). HRTEM images of Ru0.8–Ni/H–Al2O3: F) an overall image, G) a single RuNi nanoparticle
(inset: a large magnification image), H) EDS mapping of Ni (green, NiK) and Ru (red, RuK) elements for two selected RuNi nanoparticles, I) the
cross-sectional image of one RuNi nanoparticle (inset: a large magnification image), J) the image of the RuNi interface, K) the top-view image of
one Ru cluster embedded on the surface of a Ni nanoparticle. All panels reproduced with permission.[140] Copyright 2016, American Chemical Society.

DFT calculations were carried out to investigate the d-band pro- hydrogen atmosphere; ii) rehydration of the resulting RuxCuy/
jected densities of states for Ni atoms located at a Ni terrace MgAl-MMO to obtain RuxCuy/MgAl-LDH).[141] HRTEM images
(111), a Ni step (311), and the perimeter of the RuNi interface. showed the presence of uniform and highly dispersed Ru
It was found that Ni vacancy cluster defects with step-edge sites nanoparticles (mean particle size: ≈2.1 nm) anchored to the
have a strong ability to simultaneously break CC and CO surface of the platelet-like matrix. Moreover, HRTEM combined
bonds, while the RuNi interface sites with a moderate d-band with CO-TPD and in situ FTIR demonstrates that Cu preferen-
center position facilitate CC bond cleavage but inhibit CO tially occupies Ru sites with a low coordination number. XPS
bond cleavage. Therefore, modulation of metal surface defects showed that there was no electron transfer between Ru and
via the construction of a bimetallic heterostructure can shift the Cu atoms, indicating the formation of a RuCu bimetallic het-
metal d-band center position to a moderate energy state, which erostructure rather than a RuCu alloy. The optimal Ru1.0Cu0.5/
gives rise to the desired bond-breaking selectivity. MgAl-LDH catalyst exhibited excellent catalytic activity in the
Liu et al. reported the preparation of Cu-decorated Ru bime- selective hydrogenation of benzene (with a maximum
tallic catalysts supported on MgAl-LDH via a facile two-step cyclohexene yield of 44.0% at 150 °C and 5.0 MPa without any
procedure: i) the concomitant calcination and reduction of additives). Kinetic investigations demonstrated that the reaction
RuCl3-impregated CuMgAl-LDH by heating at 400 °C in a rate of benzene hydrogenation over Cu-decorated Ru catalysts

Adv. Funct. Mater. 2018, 1802943 1802943  (14 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

decreased to some extent with increasing Cu content, while the particle size (≈20 nm). The as-prepared CoSn IMC catalysts
selectivity for cyclohexene increased significantly. DFT calcu- show much higher selectivity than comparable monometallic
lations confirmed that low-coordination Ru sites facilitate the Co catalysts in the hydrogenation of α,β-unsaturated aldehydes.
dissociation of hydrogen, but are unfavorable for the desorp- EXAFS and DFT calculations showed that this can be attributed
tion of cyclohexene, while Cu-decorated Ru sites slightly depress to electron transfer from Sn to Co, and effective isolation of the
hydrogen dissociation, but significantly promote the desorption active sites by Sn. According to H2-TPD measurements, four
of cyclohexene leading to a substantial improvement in selectivity different Co sites (labeled as α, β, γ, and σ) can be identified on
for cyclohexene. Thus, the rational design and effective construc- the surface of CoSn IMCs, and DFT calculations demonstrate
tion of bimetallic heterostructures based on the topotactic trans- that the β-adsorption site (CoSn coordination) serves as the
formation of LDH precursors provides a feasible approach for active center for hydrogenation of the CO bond.
the preparation of high-performance heterogeneous catalysts for
use in systems where consecutive reactions can occur.
4.3. Metal Phosphide Catalysts

4.2.3. Intermetallic Compound Catalysts Supported metal phosphide catalysts have attracted consider-
able interest as heterogeneous catalysts, due to their versatile
IMCs with a definite composition and crystal structure have chemical composition, flexible structural architecture, and unu-
attracted considerable attention as heterogeneous catalysts in sual electronic properties. Traditional methods for the synthesis
reactions such as selective hydrogenation, methanol steam of metal phosphide catalysts (e.g., Fe2P, CoP, Ni2P, and MoP),
reforming, and electrocatalytic oxidation.[8,11,12] Although great including solvothermal reactions,[144] solid-state reactions,[145]
effort has been devoted to developing syntheses of IMCs with and high temperature reduction of oxide precursors,[146] have
controlled structure and morphology, many of the methods been developed, but these methods normally involve organic rea-
employed suffer from the need to use toxic solvents and/or gents and rigorous reaction conditions and the products suffer
complicated procedures. Therefore, the development of facile from self-agglomeration. Chen et al. have reported an alternative
and green synthetic routes is a key goal. Recently, various IMC way to prepare uniform and highly dispersed nickel phosphide
catalysts supported on MMO supports have been prepared by (Ni12P5, Ni2P, and NiP2) catalysts anchored to an Al2O3 matrix
taking advantage of the topotactic calcination and reduction of based on a modified LDH precursor approach[147] involving the
LDH precursors. Li et al. developed a method to prepare different following two steps: 1) highly dispersed Ni nanoparticles sup-
NiIn IMC catalysts (Ni3In, Ni2In, NiIn, and Ni2In3) with tun- ported on Al2O3 were prepared by an in situ topotactic calcination
able particle size and morphology by in situ reduction of NixIny- and reduction of a NiAl-LDHs precursor; 2) various supported
LDH precursors.[142] The atom-scale interspersion of Ni2+ and nickel phosphides with different crystal phases were obtained
In3+ in the brucite-like host layers of the LDH precursor facili- via a solid–gas reaction between the Ni nanoparticles and red
tates the preparation of highly dispersed IMCs under moderate phosphorus. The chemical composition, structure, and particle
reduction conditions. Moreover, the particle size and crystal size can be controlled by precisely modulating the calcination/
phase can be controlled by simply adjusting the Ni/In ratio and reduction reaction parameters (e.g., heating rate and reduction
reduction temperature. The as-obtained NiIn IMC catalysts temperature) and molar ratio of metal/red phosphorous. A Ni2P/
exhibited excellent catalytic performance in the hydrogenation of Al2O3 catalyst shows excellent catalytic activity and selectivity
α,β-unsaturated aldehydes (e.g., furfural, 1-phenylethanol, and for the selective hydrogenation of phenylacetylene (conversion:
crotonaldehyde). EXAFS and DFT calculations showed that the 98.6%, styrene selectivity: 88.2%; 100 °C, 0.3 MPa, 3 h). In situ
formation of NiIn IMCs induces electron transfer from In to Ni FTIR, EXAFS, XPS combined with DFT calculations showed that
due to the polarization of the NiIn bonds, leading to electron- the Ni active sites are isolated by P atoms and electron transfer
rich Ni species (Niδ−). This favors the hydrogenation of the nucle- occurs from Ni to P, resulting in the formation of Niδ+ centers
ophilic CO group at Niδ− sites and simultaneously inhibits the with reduced electron density. EXAFS shows that the NiNi
hydrogenation of the electrophilic alkene (CC) moiety, resulting bond length increased to 0.264 nm. Both experimental studies
in the observed high selectivity for unsaturated alcohols. and DFT calculations demonstrated that the active site isolation
Although the in situ topotactic calcination and reduction effect and the electron-deficient nature of the Ni site facilitates
of LDH precursors offers a way to prepare IMC catalysts with di-π(CC) adsorption, but decreases the desorption energy of
controllable composition and particle size, as well as high dis- the alkene initially formed and thus improves the selectivity for
persion, it requires that both metallic components of the IMC this species. In addition, due to the strong anchoring on the sup-
can be simultaneously incorporated into the host layers of LDH port, the as-obtained catalyst exhibits good stability without any
precursors. However, some large metal ions (e.g., Sn, Pb, Sb, obvious sintering/aggregation after long-term use. This approach
and Bi) as well as noble metal elements (e.g., Pt, Ir, Rh, Au, can also be extended to the preparation of other supported metal
and Pd) are difficult or impossible to incorporate into the host phosphides (e.g., CoP and Fe2P).
layers. Recently, Zhou et al. reported a modified procedure to
overcome this problem. Supported CoSn IMC catalysts with
various compositions (Co2.9Sn2, CoSn, CoSn2) were prepared 5. LDHs as Catalysts Supports
via a two-step procedure[143] (Figure 9): CoZnAl-LDHs were
homogeneously impregnated with Na2SnO4, followed by a core- LDH materials have been widely used as catalyst supports.
duction process to obtain supported CoSn IMCs with uniform The structure LDH materials (e.g., 2D surface area, flexible

Adv. Funct. Mater. 2018, 1802943 1802943  (15 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

shows excellent catalytic performance in the


selective hydrogenation of CO2 to formic acid
under mild conditions (2.0 MPa, 100 °C).
The strong Brønsted base type OH− groups
(as electron donors) boost the adsorption and
activation of CO2 molecules and facilitate
the subsequent isomerization step in basic
media. The catalytic activity of single-atom
Ru/MgAl-LDH can be tuned by modulating
the ratio of Mg2+/Al3+, since this changes the
base strength. This work shows how LDH
materials can be a very effective support for
the preparation of single-atom metal catalysts
with specific electronic structure.
By taking advantage of the 2D confinement
effect and anion exchange properties of LDH
supports, a series of supported noble-based
catalysts have been prepared by intercalation
of appropriate precursor anions. Wang et al.
reported 2D gold nanosheet catalysts sup-
ported on MgAl-LDH,[152] which exhibited
excellent catalytic activity and stability in the
solvent-free selective oxidation of carbon–
hydrogen bonds (various phenylic alkanes
with primary and secondary CH bonds) with
molecular oxygen. The Au precursor (AuCl4−)
was first intercalated into the interlayer gallery
of the MgAl-LDH precursor, followed by chem-
ical reduction to afford ultrathin Au nanosheet
catalysts anchored on the MgAl-LDH support.
HRTEM, XPS, and EXAFS demonstrated that
Figure 9.  A) Normalized intensity of Co K-edge XANES spectra for the CoSn IMCs and Co
the as-obtained ultrathin 2D gold nanosheets
foil. B) corresponding Fourier transform k3-weighted EXAFS spectra in R space compared with
Co foil. Atomic arrangement and chemical bonding of the preferential crystal facet of Co and are negatively charged with a thickness of
CoSn IMCs (based on XRD diffraction): C) Co (111) face, D) Co2.9Sn2 (004) face, E) CoSn a few atomic layers; moreover, preferential
(201) face, F) CoSn2 (211) face. All panels reproduced with permission.[143] Copyright 2016, exposure of (001) facets and a low gold–gold
Royal Society of Chemistry. coordination number are observed. The high
catalytic activity of the materials was attributed
composition, confinement effect) allow a series of supported to the ultrathin properties and high degree of unsaturation of the
noble metal (Ru, Pd, Pt, and Au) catalysts[148–150] to be prepared; surface Au atoms. After treatment at high temperature, the Au/
the metal particles in these materials have a high degree of LDH catalysts show no obvious sintering or serious aggregation,
dispersity, as well as structural stability resulting from the suggesting that this is inhibited by the confinement effect of the
strong anchoring effect of the support. The presence of surface brucite-like host layers of the LDH support.
hydroxyl groups and defect structures (e.g., oxygen vacancies) In addition to the immobilization of noble metal particles,
located in the brucite-like host layers can significantly influ- other catalytically active anions can be intercalated into the
ence the particle size, morphology, and electronic structure of interlayer galleries by anion exchange. Wang et al. reported the
the supported metal catalysts. An additional advantage of LDHs immobilization of α-amino acids onto the surface of delami-
is that a wide range of anionic metal complexes can be inter- nated ZnAl-LDH nanosheets,[153] taking advantage of the elec-
calated into the interlayer galleries in a regular arrangement, trostatic interaction between the carboxylate groups and the
offering an alternative to conventional impregnation as a way to positively charged LDH nanosheets. The as-obtained catalysts
introduce precursors to catalytically active metal particles. exhibit high enantioselectivity for the vanadium-catalyzed
The nature of the exposed crystal facets and surface epoxidation of allylic alcohols (yield of epoxy alcohol of ≈93%
hydroxyl groups in different LDH materials play an impor- (after 520 min), and ee values for trans isomers of ≈96% in
tant role in determining the particle size and electronic struc- the case of L-glutamate-intercalated LDHs). This is due to the
ture of supported metal nanoparticles. Mori et al. reported a highly dispersed active species confined on the surface of LDH
stable and well-defined single-atom Ru catalyst anchored to nanosheets (the rigid inorganic layers), together with a syner-
MgAl-LDH.[151] EXAFS showed that the strongly basic surface gistic effect between the active center and the support. By virtue
hydroxyl groups promote the formation of electron-rich Ru of this type of immobilization, the catalytic reaction occurs
centers, due to the strong electronic metal–support interactions. under pseudo-homogeneous conditions, accounting for both
The obtained single-atom catalyst immobilized on MgAl-LDH the fast reaction rate and high enantioselectivity. Therefore,

Adv. Funct. Mater. 2018, 1802943 1802943  (16 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

using delaminated LDH nanosheets as support to immobilize Second, the fabrication of supported metal catalysts with SMSI
homogeneous active sites (i.e., forming a pseudo-homogeneous is an efficient strategy to precisely modulate the geometric/elec-
catalyst) is clearly an efficient approach to simultaneously tronic structure of interfacial sites. If reducible metal cations
improve activity and enantioselectivity in asymmetric catalysis. (e.g., Fe2+/Fe3+, Co2+/Co3+, Ni2+, and Cu2+) and reducible support
species (e.g., Ti3+/Ti4+, Mn3+/Mn4+, Zr4+, and Ce4+) can be simul-
taneously introduced into the host layers of LDH materials, this
6. Summary and Perspectives would greatly increase the opportunities to prepare this type of
catalyst. The strength of the SMSI can be efficiently controlled by
This Review has summarized recent progress in the design and simply regulating the calcination and reduction conditions (e.g.,
fabrication of LDH-based nanostructured catalysts as well as calcination/reduction temperature and atmosphere), since these
highlighting structure–activity relationships in a variety of impor- affect the geometric/electronic structure of the interfacial sites
tant catalytic reactions. By taking advantage of the numerous due to the electron transfer associated with the SMSI and the
advantages of LDH precursors and their topotactic transforma- formation of oxygen defects located at the interface. Therefore,
tion conversion to MMOs, the surface/interface defect structures detailed studies of the mechanism of the topotactic transfor-
(e.g., oxygen vacancy defects or metal defects), type, quantity and mation during the calcination of LDH precursors should allow
strength of surface acid/base sites, geometric and electronic struc- the metal–support interactions to be more finely tuned. In situ
ture (e.g., particle size, composition, morphology, and electron time-resolved characterization techniques (e.g., in situ EXAFS
transfer), can be strategically engineered to give a catalyst with and TG–MS) should be carried out to monitor the structure
high activity and excellent selectivity for a specific target reaction. evolution process during the whole process of calcination and
Moreover, the confinement effect of the host layers and anchoring reduction of LDH precursors, including monitoring changes in
effect of the support contribute to the high stability of the cata- the electronic state, coordination number, and bond distance.
lytically active sites. The geometric/electronic structure and sur- Third, based on reaction kinetics models, the establishment of
face defect structure can be characterized by a range of advanced quantitative relationships between reaction rate and structure
techniques (e.g., HRTEM, STEM, XPS, DRIFTS, EXAFS, ESR, (e.g., metal sites, defect sites, or interfacial sites) should allow a
Mössbauer spectra, and positron annihilation), which leads to a better understanding of the structure of the active sites. In order
fundamental understanding of structure–activity relationships. to acquire direct evidence about the actual active sites, in situ
Although great progress has been made, critical issues and characterization techniques are necessary to track catalytic reac-
great challenges still remain to be addressed in the preparation tion process under in situ or even operando conditions. In situ/
of high-performance LDH-based nanocatalysts, with the fol- operando XPS, in situ/operando EXAFS combined with in situ/
lowing three problems being the most important: i) More effi- operando XRD can provide information about the structural
cient and simple ways of fabricating monolayer LDH nanosheets evolution and variation in electronic state of active sites under
are required. Although the preparation of monolayer LDH practical reaction conditions. In the case of high-valent transi-
nanosheets has been reported by traditional liquid phase exfo- tion Fe based catalysts, in situ/operando Mössbauer spectros-
liation or dry delamination methods (Ar plasma etching), these copy is an especially efficient approach to observe the change
synthetic methods are either time-consuming, environmen- of electronic structure and to identify intermediate active phases
tally unfriendly, or require rigorous conditions and/or sophis- of catalyst. In situ/operando DRIFTS combined with mass spec-
ticated techniques. This is a considerable barrier to developing trometry can be used as a powerful tool to explore the adsorp-
large-scale practical applications of these materials. ii) Based tion and activation of a reactant molecule and to monitor reac-
on the topotactic transformation of LDH precursors, supported tion intermediates. Combined studies of the surface geometric/
monometallic or bimetallic catalysts can be prepared with a electronic structure of a catalyst and the surface chemisorbed
high degree of dispersion. However, more precise methods of species should provide valuable evidence about the mechanism
engineering the geometric/electronic structure (e.g., metal and of a catalytic reaction. However, the timescales of the structural
oxygen vacancies) of interfacial sites in the as-obtained catalysts evolution of a catalyst and the formation of active intermediates
are needed in order to optimize the catalytic performance of the are normally transient and, therefore, the development of in situ
resulting materials. iii) The characterization and identification of time-resolved characterization techniques is a necessary prereq-
active sites are rather difficult at an atomic level, due to the com- uisite for obtaining such information. In addition, DFT calcula-
plexity of a heterogeneous reaction and the fact that the nature tions serve as a valuable supplementary tool to study activation
of the active site may change under practical reaction conditions. energy barriers, transition states, and reaction intermediates,
In the light of these challenges, future studies should focus and to simulate the whole reaction process. As well as providing
on developing new synthetic methods, particularly those which a reasonable explanation for experimental observations, DFT
are amenable to being scaled up, and a detailed and thorough calculations can also be used to make predictions about future
investigation of the mechanism of the topotactic transformation. reactions. In summary, the combination of advanced in situ
The development of improved synthetic approaches (bottom-up) characterization methodology and DFT calculations is neces-
and delamination techniques (top-down) is necessary in order sary to reveal details about the intrinsic structure of catalytically
to prepare monolayer LDH nanosheets. The reverse micro­ active sites, and to provide a fundamental understanding of a
emulsion method can be tuned by precisely controlling the oil/ catalytic mechanism at the molecular/atomic level. To overcome
water ratio and the property of the oil phase so as to synthesize the challenges mentioned above, more creative investigations
uniform and single-layered LDH nanosheets. Scale-up of these should be carried out to boost the development of new LDH
procedures would then allow practical industrial applications. materials for heterogeneous catalysis.

Adv. Funct. Mater. 2018, 1802943 1802943  (17 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Acknowledgements [23] Y. Li, L. Zhang, X. Xiang, D. Yan, F. Li, J. Mater. Chem. A 2014, 2,
13250.
This work was supported by the National Key Research and [24] L. Trotochaud, S. L. Young, J. K. Ranney, S. W. Boettcher, J. Am.
Development Programme (Grant No. 2017YFA0206804), the National Chem. Soc. 2014, 136, 6744.
Natural Science Foundation of China (NSFC), and the Fundamental [25] J. Y. C. Chen, L. Dang, H. Liang, W. Bi, J. B. Gerken, S. Jin, E. E. Alp,
Research Funds for the Central Universities (buctylkxj01). The authors S. S. Stahl, J. Am. Chem. Soc. 2015, 137, 15090.
are thankful for the support of the BSRF (Beijing Synchrotron Radiation
[26] B. J. Trzesniewski, O. Diaz-Morales, D. A. Vermaas, A. Longo,
Facility) during the XAFS measurements at the beamline 1W1B, 1W2B
W. Bras, M. T. Koper, W. A. Smith, J. Am. Chem. Soc. 2015, 137,
and 4B9B.
15112.
[27] D. K. Bediako, B. Lassalle-Kaiser, Y. Surendranath, J. Yano,
V. K. Yachandra, D. G. Nocera, J. Am. Chem. Soc. 2012, 134, 6801.
Conflict of Interest [28] G. Busca, Chem. Rev. 2010, 110, 2217.
[29] O. Kikhtyanin, R. Bulánek, K. Frolich, J. Čejka, D. Kubička, J. Mol.
The authors declare no conflict of interest. Catal. A: Chem. 2016, 424, 358.
[30] Z. Wang, P. Fongarland, G. Lu, N. Essayem, J. Catal. 2014, 318,
108.
[31] I. Delidovich, R. Palkovits, J. Catal. 2015, 327, 1.
Keywords [32] J. Han, Y. Dou, M. Wei, D. G. Evans, X. Duan, Angew. Chem., Int.
catalyst support, delamination, heterogeneous catalysis, layered double Ed. 2010, 49, 2171.
hydroxides, topotactic transformation [33] J. S. Valente, H. Pfeiffer, E. Lima, J. Prince, J. Flores, J. Catal. 2011,
279, 196.
Received: April 30, 2018 [34] Y. Dou, T. Pan, S. Xu, H. Yan, J. Han, M. Wei, D. G. Evans,
Revised: July 11, 2018 X. Duan, Angew. Chem., Int. Ed. 2015, 54, 9673.
Published online: [35] O. Kikhtyanin, Z. Tišler, R. Velvarská, D. Kubička, Appl. Catal.,
A 2017, 536, 85.
[36] L. Smoláková, K. Frolich, J. Kocík, O. Kikhtyanin, L. Čapek, Ind.
Eng. Chem. Res. 2017, 56, 4638.
[1] L. Dai, Y. Xue, L. Qu, H. J. Choi, J. B. Baek, Chem. Rev. 2015, 115, 4823. [37] S. Abelló, S. Dhir, G. Colet, J. Pérez-Ramírez, Appl. Catal., A 2007,
[2] J. A. Rodriguez, D. C. Grinter, Z. Liu, R. M. Palomino, 325, 121.
S. D. Senanayake, Chem. Soc. Rev. 2017, 46, 1824. [38] S. Abelló, F. Medina, D. Tichit, J. Pérez-Ramírez, J. E. Sueiras,
[3] Q. Fu, X. Bao, Chem. Soc. Rev. 2017, 46, 1842. P. Salagre, Y. Cesteros, Appl. Catal., B 2007, 70, 577.
[4] J. K. Norskov, T. Bligaard, B. Hvolbaek, F. Abild-Pedersen, [39] W. Bing, L. Zheng, S. He, D. Rao, M. Xu, L. Zheng, B. Wang,
I. Chorkendorff, C. H. Christensen, Chem. Soc. Rev. 2008, 37, 2163. Y. Wang, M. Wei, ACS Catal. 2017, 8, 656.
[5] S. Yao, X. Zhang, W. Zhou, R. Gao, W. Xu, Y. Ye, L. Lin, X. Wen, [40] J. L. Gunjakar, T. W. Kim, H. N. Kim, I. Y. Kim, S. J. Hwang, J. Am.
P. Liu, B. Chen, E. Crumlin, J. Guo, Z. Zuo, W. Li, J. Xie, L. Lu, Chem. Soc. 2011, 133, 14998.
C. J. Kiely, L. Gu, C. Shi, J. A. Rodriguez, D. Ma, Science 2017, 357, [41] C. G. Silva, Y. Bouizi, V. Fornés, H. García, J. Am. Chem. Soc. 2009,
389. 131, 13833.
[6] J. C. Matsubu, S. Zhang, L. DeRita, N. S. Marinkovic, J. G. Chen, [42] J. Luo, J. H. Im, M. T. Mayer, M. Schreier, M. K. Nazeeruddin,
G. W. Graham, X. Pan, P. Christopher, Nat. Chem. 2017, 9, 120. N. G. Park, S. D. Tilley, H. J. Fan, M. Grätzel, Science 2014, 345,
[7] S. He, Z. An, M. Wei, D. G. Evans, X. Duan, Chem. Commun. 2013, 1593.
49, 5912. [43] Y. Zhao, B. Li, Q. Wang, W. Gao, C. J. Wang, M. Wei, D. G. Evans,
[8] C. Li, M. Wei, D. G. Evans, X. Duan, Small 2014, 10, 4469. X. Duan, D. O’Hare, Chem. Sci. 2014, 5, 951.
[9] A. M. Fogg, V. M. Green, H. G. Harvey, D. O’Hare, Adv. Mater. [44] T. R. Cook, D. K. Dogutan, S. Y. Reece, Y. Surendranath, T. S. Teets,
1999, 11, 1466. D. G. Nocera, Chem. Rev. 2010, 110, 6474.
[10] Q. Wang, D. O’Hare, Chem. Rev. 2012, 112, 4124. [45] G. Wu, P. Zelenay, Acc. Chem. Res. 2013, 46, 1878.
[11] G. Fan, F. Li, D. G. Evans, X. Duan, Chem. Soc. Rev. 2014, 43, 7040. [46] M. Gong, H. Dai, Nano Res. 2014, 8, 23.
[12] J. Feng, Y. He, Y. Liu, Y. Du, D. Li, Chem. Soc. Rev. 2015, 44, 5291. [47] N. T. Suen, S. F. Hung, Q. Quan, N. Zhang, Y. J. Xu, H. M. Chen,
[13] Y. Zhao, X. Jia, G. I. N. Waterhouse, L. Z. Wu, C. H. Tung, Chem. Soc. Rev. 2017, 46, 337.
D. O’Hare, T. Zhang, Adv. Energy Mater. 2016, 6, 1501974. [48] J. Li, S. Jiang, M. Shao, M. Wei, Catalysts 2018, 8, 214.
[14] M. Q. Zhao, Q. Zhang, J. Q. Huang, F. Wei, Adv. Funct. Mater. [49] C. Zhang, J. Zhao, L. Zhou, Z. Li, M. Shao, M. Wei, J. Mater. Chem.
2012, 22, 675. A 2016, 4, 11516.
[15] M. Shao, M. Wei, D. G. Evans, X. Duan, Chem. - Eur. J. 2013, 19, 4100. [50] J. Zhang, J. Liu, L. Xi, Y. Yu, N. Chen, S. Sun, W. Wang, K. M. Lange,
[16] M. L. Kantam, B. M. Choudary, C. V. Reddy, K. K. Rao, F. Figueras, B. Zhang, J. Am. Chem. Soc. 2018, 140, 3876.
Chem. Commun. 1998, 9, 1033. [51] H. Xiao, H. Shin, W. A. Goddard III, Proc. Natl. Acad. Sci. USA
[17] B. M. Choudary, M. Lakshmi Kantam, V. Neeraja, K. Koteswara 2018, 115, 5872.
Rao, F. Figueras, L. Delmotte, Green Chem. 2001, 3, 257. [52] H. Shin, H. Xiao, W. A. Goddard III, J. Am. Chem. Soc. 2018, 140,
[18] D. Tichit, D. Lutic, B. Coq, R. Durand, R. Teissier, J. Catal. 2003, 6745.
219, 167. [53] G. Hu, N. Wang, D. O’Hare, J. Davis, Chem. Commun. 2006,
[19] J. Cueto, L. Faba, E. Díaz, S. Ordóñez, Appl. Catal., B 2017, 201, 287.
221. [54] C. J. Wang, Y. A. Wu, R. M. J. Jacobs, J. H. Warner, G. R. Williams,
[20] Y. Du, Q. Wang, X. Liang, Y. He, J. Feng, D. Li, J. Catal. 2015, 331, 154. D. O’Hare, Chem. Mater. 2011, 23, 171.
[21] Y. Du, Q. Wang, X. Liang, P. Yang, Y. He, J. Feng, D. Li, Catal. Sci. [55] G. Hu, N. Wang, D. O’Hare, J. Davis, J. Mater. Chem. 2007, 17,
Technol. 2017, 7, 4361. 2257.
[22] X. Zou, A. Goswami, T. Asefa, J. Am. Chem. Soc. 2013, 135, [56] Z. Li, M. Shao, H. An, Z. Wang, S. Xu, M. Wei, D. G. Evans,
17242. X. Duan, Chem. Sci. 2015, 6, 6624.

Adv. Funct. Mater. 2018, 1802943 1802943  (18 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[57] F. Wypych, K. G. Satyanarayana, J. Colloid Interface Sci. 2005, 285, [90] L. Wang, J. Zhang, Y. Zhu, S. Xu, C. Wang, C. Bian, X. Meng,
532. F.-S. Xiao, ACS Catal. 2017, 7, 7461.
[58] F. Song, X. Hu, Nat. Commun. 2014, 5, 4477. [91] P. Sun, G. Siddiqi, W. C. Vining, M. Chi, A. T. Bell, J. Catal. 2011,
[59] N. Han, F. Zhao, Y. Li, J. Mater. Chem. A 2015, 3, 16348. 282, 165.
[60] C. Zhang, M. Shao, L. Zhou, Z. Li, K. Xiao, M. Wei, ACS Appl. [92] Y. Zhu, Z. An, J. He, J. Catal. 2016, 341, 44.
Mater. Interfaces 2016, 8, 33697. [93] Y. Zhu, Z. An, H. Song, X. Xiang, W. Yan, J. He, ACS Catal. 2017,
[61] L. Yu, J. F. Yang, B. Y. Guan, Y. Lu, X. W. D. Lou, Angew. Chem., Int. 7, 6973.
Ed. 2018, 57, 172. [94] X. Ma, Z. An, Y. Zhu, W. Wang, J. He, ChemCatChem 2016, 8, 1773.
[62] H. Liang, F. Meng, M. Caban-Acevedo, L. Li, A. Forticaux, L. Xiu, [95] Z. Tian, Q. Li, J. Hou, Y. Li, S. Ai, Catal. Sci. Technol. 2016, 6, 703.
Z. Wang, S. Jin, Nano Lett. 2015, 15, 1421. [96] B. Qiao, A. Wang, X. Yang, L. F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li,
[63] Y. Zhao, G. Chen, T. Bian, C. Zhou, G. I. Waterhouse, L. Z. Wu, T. Zhang, Nat. Chem. 2011, 3, 634.
C. H. Tung, L. J. Smith, D. O’Hare, T. Zhang, Adv. Mater. 2015, 27, [97] J. Lin, A. Wang, B. Qiao, X. Liu, X. Yang, X. Wang, J. Liang, J. Li,
7824. J. Liu, T. Zhang, J. Am. Chem. Soc. 2013, 135, 15314.
[64] F. Song, X. Hu, J. Am. Chem. Soc. 2014, 136, 16481. [98] M. Yang, S. Li, Y. Wang, J. A. Herron, Y. Xu, L. F. Allard, S. Lee,
[65] Y. Wang, Y. Zhang, Z. Liu, C. Xie, S. Feng, D. Liu, M. Shao, J. Huang, M. Mavrikakis, M. Flytzanistephanopoulos, Science
S. Wang, Angew. Chem., Int. Ed. 2017, 56, 5867. 2014, 346, 1498.
[66] Y. Wang, C. Xie, Z. Zhang, D. Liu, R. Chen, S. Wang, Adv. Funct. [99] P. Hu, Z. Huang, Z. Amghouz, M. Makkee, F. Xu, F. Kapteijn,
Mater. 2018, 28, 1703363. A. Dikhtiarenko, Y. Chen, X. Gu, X. Tang, Angew. Chem., Int. Ed.
[67] M. Gong, Y. Li, H. Wang, Y. Liang, J. Z. Wu, J. Zhou, J. Wang, 2014, 53, 3418.
T. Regier, F. Wei, H. Dai, J. Am. Chem. Soc. 2013, 135, 8452. [100] B. Yoon, H. Häkkinen, U. Landman, A. S. Wörz, J. M. Antonietti,
[68] C. Tang, H. S. Wang, H. F. Wang, Q. Zhang, G. L. Tian, J. Q. Nie, S. Abbet, K. Judai, U. Heiz, Science 2005, 307, 403.
F. Wei, Adv. Mater. 2015, 27, 4516. [101] J. M. Thomas, R. Raja, D. W. Lewis, Angew. Chem., Int. Ed. 2005,
[69] S. Chen, J. Duan, M. Jaroniec, S. Z. Qiao, Angew. Chem., Int. Ed. 44, 6456.
2013, 52, 13567. [102] J. Lu, C. Aydin, N. D. Browning, B. C. Gates, Angew. Chem., Int. Ed.
[70] X. Long, J. Li, S. Xiao, K. Yan, Z. Wang, H. Chen, S. Yang, Angew. 2012, 51, 5842.
Chem., Int. Ed. 2014, 53, 7584. [103] Y. Zhai, D. Pierre, R. Si, W. Deng, P. Ferrin, A. U. Nilekar, G. Peng,
[71] W. Ma, R. Ma, C. Wang, J. Liang, X. Liu, K. Zhou, T. Sasaki, ACS J. A. Herron, D. C. Bell, H. Saltsburg, M. Mavrikakis, M. Flytzani-
Nano 2015, 9, 1977. Stephanopoulos, Science 2010, 329, 1633.
[72] J. Zhao, J. Chen, S. Xu, M. Shao, Q. Zhang, F. Wei, J. Ma, M. Wei, [104] J. Wan, W. Chen, C. Jia, L. Zheng, J. Dong, X. Zheng, Y. Wang,
D. G. Evans, X. Duan, Adv. Funct. Mater. 2014, 24, 2938. W. Yan, C. Chen, Q. Peng, D. Wang, Y. Li, Adv. Mater. 2018, 30,
[73] R. Zhang, M. Shao, S. Xu, F. Ning, L. Zhou, M. Wei, Nano Energy 1705369.
2017, 33, 21. [105] H. Bielawa, O. Hinrichsen, A. Birkner, M. Muhler, Angew. Chem.,
[74] F. Ning, M. Shao, S. Xu, Y. Fu, R. Zhang, M. Wei, D. G. Evans, Int. Ed. 2001, 40, 1061.
X. Duan, Energy Environ. Sci. 2016, 9, 2633. [106] S. Kattel, W. Yu, X. Yang, B. Yan, Y. Huang, W. Wan, P. Liu,
[75] Y. Hou, M. R. Lohe, J. Zhang, S. Liu, X. Zhuang, X. Feng, Energy J. G. Chen, Angew. Chem., Int. Ed. 2016, 55, 7968.
Environ. Sci. 2016, 9, 478. [107] S. Kattel, P. Liu, J. G. Chen, J. Am. Chem. Soc. 2017, 139, 9739.
[76] Y. Dou, S. Zhang, T. Pan, S. Xu, A. Zhou, M. Pu, H. Yan, J. Han, [108] T. Komanoya, T. Kinemura, Y. Kita, K. Kamata, M. Hara, J. Am.
M. Wei, D. G. Evans, X. Duan, Adv. Funct. Mater. 2015, 25, 2243. Chem. Soc. 2017, 139, 11493.
[77] F. Mi, X. Chen, Y. Ma, S. Yin, F. Yuan, H. Zhang, Chem. Commun. [109] Y. Wang, D. Widmann, R. J. Behm, ACS Catal. 2017, 7, 2339.
2011, 47, 12804. [110] D. Baudouin, K. C. Szeto, P. Laurent, A. De Mallmann, B. Fenet,
[78] M. Shao, F. Ning, M. Wei, D. G. Evans, X. Duan, Adv. Funct. Mater. L. Veyre, U. Rodemerck, C. Coperet, C. Thieuleux, J. Am. Chem.
2014, 24, 580. Soc. 2012, 134, 20624.
[79] J. Ping, Y. Wang, Q. Lu, B. Chen, J. Chen, Y. Huang, Q. Ma, C. Tan, [111] L. Lin, W. Zhou, R. Gao, S. Yao, X. Zhang, W. Xu, S. Zheng,
J. Yang, X. Cao, Z. Wang, J. Wu, Y. Ying, H. Zhang, Adv. Mater. Z. Jiang, Q. Yu, Y. W. Li, C. Shi, X. D. Wen, D. Ma, Nature 2017,
2016, 28, 7640. 544, 80.
[80] X. Shu, Z. An, L. Wang, J. He, Chem. Commun. 2009, 5901. [112] S. M. Kim, P. M. Abdala, T. Margossian, D. Hosseini, L. Foppa,
[81] S. He, S. Zhang, J. Lu, Y. Zhao, J. Ma, M. Wei, D. G. Evans, A. Armutlulu, W. van Beek, A. Comas-Vives, C. Coperet, C. Muller,
X. Duan, Chem. Commun. 2011, 47, 10797. J. Am. Chem. Soc. 2017, 139, 1937.
[82] D. Wang, X. Chen, D. G. Evans, W. Yang, Nanoscale 2013, 5, 5312. [113] J. L. Rogers, M. C. Mangarella, A. D. D’Amico, J. R. Gallagher,
[83] G. Carja, E. F. Grosu, M. Mureseanu, D. Lutic, Catal. Sci. Technol. M. R. Dutzer, E. Stavitski, J. T. Miller, C. Sievers, ACS Catal. 2016,
2017, 7, 5402. 6, 5873.
[84] S. Xia, M. Shao, X. Zhou, G. Pan, Z. Ni, J. Mol. Catal. A: Chem. [114] S. He, C. Li, H. Chen, D. Su, B. Zhang, X. Cao, B. Wang, M. Wei,
2015, 406, 127. D. G. Evans, X. Duan, Chem. Mater. 2013, 25, 1040.
[85] Y. Zhao, X. Jia, G. Chen, L. Shang, G. I. Waterhouse, L. Z. Wu, [115] M. Behrens, F. Studt, I. Kasatkin, S. Kuhl, M. Havecker, F. Abild-
C. H. Tung, D. O’Hare, T. Zhang, J. Am. Chem. Soc. 2016, 138, Pedersen, S. Zander, F. Girgsdies, P. Kurr, B. L. Kniep, M. Tovar,
6517. R. W. Fischer, J. K. Norskov, R. Schlogl, Science 2012, 336, 893.
[86] L. Teruel, Y. Bouizi, P. Atienzar, V. Fornes, H. Garcia, Energy [116] S. Kuld, M. Thorhauge, H. Falsig, C. F. Elkjær, S. Helveg,
Environ. Sci. 2010, 3, 154. I. Chorkendorff, J. Sehested, Science 2016, 352, 969.
[87] Q. Feng, S. Zhao, Y. Wang, J. Dong, W. Chen, D. He, D. Wang, [117] S. Kattel, P. J. Ramírez, J. G. Chen, J. A. Rodriguez, P. Liu, Science
J. Yang, Y. Zhu, H. Zhu, L. Gu, Z. Li, Y. Liu, R. Yu, J. Li, Y. Li, J. Am. 2017, 357, 1296.
Chem. Soc. 2017, 139, 7294. [118] S. J. Tauster, S. C. Fung, R. L. Garten, J. Am. Chem. Soc. 1978, 9,
[88] Y. Liu, J. Zhao, Y. He, J. Feng, T. Wu, D. Li, J. Catal. 2017, 348, 170.
135. [119] S. J. Tauster, Acc. Chem. Res. 1987, 20, 121.
[89] W. Z. Li, L. Kovarik, D. Mei, J. Liu, Y. Wang, C. H. Peden, Nat. [120] S. J. Tauster, S. C. Fung, R. T. K. Baker, J. A. Horsley, Science 1981,
Commun. 2013, 4, 2481. 211, 1121.

Adv. Funct. Mater. 2018, 1802943 1802943  (19 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[121] X. Liu, M. H. Liu, Y. C. Luo, C. Y. Mou, S. D. Lin, H. Cheng, [137] W. Gao, Y. Zhao, H. Chen, H. Chen, Y. Li, S. He, Y. Zhang, M. Wei,
J. M. Chen, J. F. Lee, T. S. Lin, J. Am. Chem. Soc. 2012, 134, 10251. D. G. Evans, X. Duan, Green Chem. 2015, 17, 1525.
[122] A. Bruix, J. A. Rodriguez, P. J. Ramirez, S. D. Senanayake, J. Evans, [138] C. Li, Y. Dou, J. Liu, Y. Chen, S. He, M. Wei, D. G. Evans, X. Duan,
J. B. Park, D. Stacchiola, P. Liu, J. Hrbek, F. Illas, J. Am. Chem. Soc. Chem. Commun. 2013, 49, 9992.
2012, 134, 8968. [139] X. Ning, Z. An, J. He, J. Catal. 2016, 340, 236.
[123] H. Tang, J. Wei, F. Liu, B. Qiao, X. Pan, L. Li, J. Liu, J. Wang, [140] H. Chen, S. He, X. Cao, S. Zhang, M. Xu, M. Pu, D. Su, M. Wei,
T. Zhang, J. Am. Chem. Soc. 2016, 138, 56. D. G. Evans, X. Duan, Chem. Mater. 2016, 28, 4751.
[124] H. Tang, F. Liu, J. Wei, B. Qiao, K. Zhao, Y. Su, C. Jin, L. Li, J. J. Liu, [141] J. Liu, S. Xu, W. Bing, F. Wang, C. Li, M. Wei, D. G. Evans, X. Duan,
J. Wang, Angew. Chem., Int. Ed. 2016, 128, 10764. ChemCatChem 2015, 7, 846.
[125] H. Tang, Y. Su, B. Zhang, A. F. Lee, M. A. Isaacs, K. Wilson, L. Li, [142] C. Li, Y. Chen, S. Zhang, S. Xu, J. Zhou, F. Wang, M. Wei,
Y. Ren, J. Huang, M. Haruta, Sci. Adv. 2017, 3, e1700231. D. G. Evans, X. Duan, Chem. Mater. 2013, 25, 3888.
[126] C. T. Campbell, Nat. Chem. 2012, 4, 597. [143] J. Zhou, Y. Yang, C. Li, S. Zhang, Y. Chen, S. Shi, M. Wei, J. Mater.
[127] M. Xu, S. He, H. Chen, G. Cui, L. Zheng, B. Wang, M. Wei, ACS Chem. A 2016, 4, 12825.
Catal. 2017, 7, 7600. [144] S. Carenco, Y. Hu, I. Florea, O. Ersen, C. Boissière, N. Mézailles,
[128] H. Chen, S. He, M. Xu, M. Wei, D. G. Evans, X. Duan, ACS Catal. C. Sanchez, Chem. Mater. 2012, 24, 4134.
2017, 7, 2735. [145] B. M. Barry, E. G. Gillan, Chem. Mater. 2009, 21, 4454.
[129] Q. Hu, L. Yang, G. Fan, F. Li, J. Catal. 2016, 340, 184. [146] S. T. Oyama, T. Gott, K. Asakura, S. Takakusagi, K. Miyazaki,
[130] C. Li, Y. Chen, S. Zhang, J. Zhou, F. Wang, S. He, M. Wei, Y. Koike, K. K. Bando, J. Catal. 2009, 268, 209.
D. G. Evans, X. Duan, ChemCatChem 2014, 6, 824. [147] Y. Chen, C. Li, J. Zhou, S. Zhang, D. Rao, S. He, M. Wei,
[131] F. Studt, I. Sharafutdinov, F. Abild-Pedersen, C. F. Elkjaer, D. G. Evans, X. Duan, ACS Catal. 2015, 5, 5756.
J. S. Hummelshoj, S. Dahl, I. Chorkendorff, J. K. Norskov, Nat. [148] F. Zhang, X. Zhao, C. Feng, B. Li, T. Chen, W. Lu, X. Lei, S. Xu, ACS
Chem. 2014, 6, 320. Catal. 2011, 1, 232.
[132] A. García-Trenco, E. R. White, A. Regoutz, D. J. Payne, [149] J. H. Lee, H. Kim, Y. S. Lee, D. Y. Jung, ChemCatChem 2014, 6, 113.
M. S. P. Shaffer, C. K. Williams, ACS Catal. 2017, 7, 1186. [150] Y. Tan, X. Y. Liu, L. Zhang, A. Wang, L. Li, X. Pan, S. Miao,
[133] Y. He, L. Liang, Y. Liu, J. Feng, C. Ma, D. Li, J. Catal. 2014, 309, M. Haruta, H. Wei, H. Wang, F. Wang, X. Wang, T. Zhang, Angew.
166. Chem., Int. Ed. 2017, 56, 2709.
[134] E. Nikolla, J. Schwank, S. Linic, J. Catal. 2009, 263, 220. [151] K. Mori, T. Taga, H. Yamashita, ACS Catal. 2017, 7, 3147.
[135] H. Wang, W. Zhou, J. X. Liu, R. Si, G. Sun, M. Q. Zhong, H. Y. Su, [152] L. Wang, Y. Zhu, J. Q. Wang, F. Liu, J. Huang, X. Meng,
H. B. Zhao, J. A. Rodriguez, S. J. Pennycook, J. C. Idrobo, W. X. Li, J. M. Basset, Y. Han, F. S. Xiao, Nat. Commun. 2015, 6, 6957.
Y. Kou, D. Ma, J. Am. Chem. Soc. 2013, 135, 4149. [153] J. Wang, L. Zhao, H. Shi, J. He, Angew. Chem., Int. Ed. 2011, 50,
[136] Y. Xiang, R. Barbosa, X. Li, N. Kruse, ACS Catal. 2015, 5, 2929. 9171.

Adv. Funct. Mater. 2018, 1802943 1802943  (20 of 20) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like