Chem Soc Rev View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

REVIEW ARTICLE

Cite this: DOI: 10.1039/c4cs00160e

View Journal

Catalytic applications of layered double hydroxides: recent advances and perspectives Guoli Fan, Feng Li,* David G. Evans and Xue Duan This review surveys recent advances in the applications of layered double hydroxides (LDHs) in heterogeneous catalysis. By virtue of the flexible tunability and uniform distribution of metal cations in the brucite-like layers and the facile exchangeability of intercalated anions, LDHs—both as directly prepared or after thermal treatment and/or reduction—have found many applications as stable and recyclable heterogeneous catalysts or catalyst supports for a variety of reactions with high industrial and academic importance. A major challenge in this rapidly growing field is to simultaneously improve

Received 9th May 2014

the activity, selectivity and stability of these LDH-based materials by developing ways of tailoring the

DOI: 10.1039/c4cs00160e

electronic structure of the catalysts and supports. Therefore, this Review article is mainly focused on the most recent developments in smart design strategies for LDH materials and the potential catalytic

www.rsc.org/csr

applications of the resulting materials.

1. Introduction Layered double hydroxides (LDHs) or hydrotalcite-like compounds are a large family of two-dimensional (2D) anionic clay materials which can be represented by the general formula [M1x2+Mx3+(OH)2]x+[Ax/n]nmH2O.1–3 As shown in Fig. 1, LDHs are composed of brucite-like layers in which a fraction of the divalent metal cations (e.g., Mg2+, Fe2+, Co2+, Cu2+, Ni2+, or Zn2+) State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, P.O. Box 98, Beijing 100029, P. R. China. E-mail: [email protected]; Fax: +86-10-64425385; Tel: +86-10-64451226

Guoli Fan

Guoli Fan received his BS (2006) and PhD (2011) degrees from Beijing University of Chemical Technology, where he mainly studied synthesis of new catalysts based on layered double hydroxides, under the supervision of Prof. Feng Li. After obtaining his PhD in 2011, he continued working at Beijing University of Chemical Technology. His research interests focus on design and fabrication of catalytic materials and environmentally friendly catalysis.

This journal is © The Royal Society of Chemistry 2014

coordinated octahedrally by hydroxyl groups have been replaced isomorphously by the trivalent metal cations (e.g., Al3+, Cr3+, Ga3+, In3+, Mn3+ or Fe3+), giving positively charged layers.4 The value of x is equal to the molar ratio M2+/(M2+ + M3+) and is generally in the range 0.2–0.33; water and exchangeable inorganic or organic charge-compensating anions are present in the interlayer galleries. Each hydroxyl group in the LDH layers is oriented toward the interlayer region and may be hydrogen bonded to the interlayer anions and water molecules. LDH materials have a relatively weak interlayer bonding and as a consequence exhibit excellent expanding properties. The wide tunability of the types of metal cations, the M2+/M3+ molar ratios,

Feng Li studied at East China University of Science and Technology for both his undergraduate and graduate degrees. He was awarded a PhD degree from Beijing University of Chemical Technology (BUCT) in 1999. He was a visiting scholar at Universite´Blaise Pascal in 2001. He was appointed Professor of Chemistry at BUCT in 2005. His research interests are heterogeneous catalysis and materials chemistry. In 2004 he Feng Li was selected to participate in the ‘New Century Outstanding Talent’ scheme of the Ministry of Education. In 2013 he was supported by the National Science Foundation for Distinguished Young Scholars.

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

Fig. 1 The idealized structure of carbonate-intercalated LDHs with different M2+/M3+ molar ratios showing the metal hydroxide octahedra stacked along the crystallographic c-axis, as well as water and anions present in the interlayer region.

and the nature of interlayer compensating anions results in a huge variety of host–guest assemblies and nanoarchitectures with versatile physical and chemical properties, and the increasing interest in these materials is driven by this compositional flexibility. An additional advantage is that LDHs can be synthesized by well-established synthetic protocols—such as co-precipitation and homogeneous precipitation by urea hydrolysis—which are simple and amenable to scaling up for industrial production.1 Over the past two decades, the properties of LDHs have been tailored to produce materials designed to fulfil specific requirements for practical applications in a wide variety of fields, including as additives in polymers, adsorption materials, precursors for functional materials, and in pharmaceutics, photochemistry and electrochemistry.5 In particular, LDHs—both as directly prepared or after thermal treatment and/or reduction— have been widely employed in heterogeneous catalysis as actual catalysts, catalyst precursors and catalyst supports for various reactions including organic transformations, photodegradation of organic wastes, and the total decomposition of volatile organic compounds.6 In the past five years, there has been an

David G. Evans studied as both an undergraduate and a research student at Jesus College, Oxford, and obtained a DPhil under the supervision of Prof. D. M. P. Mingos FRS. After postdoctoral work at Bristol University with Prof. F. G. A. Stone FRS, he was appointed as a lecturer at Exeter University in 1985. He moved to Beijing University of Chemical Technology in 1996. His research interests focus on David G. Evans intercalation in layered solids. He was awarded an International Scientific and Technological Cooperation Award of the People’s Republic of China in 2005.

Chem. Soc. Rev.

Chem Soc Rev

extremely rapid growth in publications related to the catalytic applications of LDH-based materials. Recently, An and co-workers summarized advances in the design of nanocatalysts using LDHs as precursors.7 However, the rapid growth of the use of LDH materials in a range of fields related to heterogeneous catalysis warrants a broader assessment and critical discussion of recent developments. In this Review article, we summarize the most recent advances in the smart design and synthesis of LDHs and derivatives for use as catalysts and catalyst supports in the field of heterogeneous catalysis. Especially, a variety of types of LDH-based catalysts, which can be obtained by different synthetic pathways (Fig. 2), are emphatically reviewed in the following paragraphs. A major challenge in this rapidly growing field is to improve the functionalities of these LDH-based catalytic materials. Therefore, the advantages of LDHs and derivatives as catalysts and catalyst supports are highlighted and the structure–property correlations are addressed, which provides useful insights into ways to develop robust and promising catalysts and catalyst supports based on LDH materials. In addition, we also look at potential future developments in the use of LDH-derived materials in catalysis. It should be noted that throughout the paper, LDHs with different metal cations are usually abbreviated in the form M2+M3+–LDH (M2+ denotes divalent cations, M3+ denotes trivalent cations), if anions are not specified.

2. LDH-based catalysts 2.1.

Simple LDHs as catalysts in their own right

LDHs have two attractive features which impart catalytic properties to the materials. Firstly, the brucite-like layers have an abundance of basic sites allowing the materials to be used as heterogeneous solid base catalysts. Secondly, the two or more metal cations within the brucite-like layers are uniformly distributed at the atomic level without a segregation of ‘‘lakes’’ of separate cations, and where one of the cations is a catalytically

Xue Duan was elected as an Academician of the Chinese Academy of Sciences in 2007 and a Fellow of the Royal Society of Chemistry in 2009. He was awarded his BS degree from Jilin University in 1982 and his PhD degree from the Chinese Academy of Sciences in 1988. He was subsequently appointed to the staff of Beijing University of Chemical Technology (BUCT) and established the Applied Xue Duan Chemistry Research Institute in 1990. He was promoted to full professor in 1993. He is currently Executive Vice-Chair of the Academic Committee of the State Key Laboratory of Chemical Resource Engineering.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

Fig. 2

Review Article

Summary of the synthetic pathways for preparation of the LDH-based catalysts.

active transition metal this can lead to high catalytic activity and selectivity. 2.1.1 Solid base catalysts. By virtue of the abundance of hydroxyl groups, simple binary as-prepared materials such as MgAl–LDHs are potential heterogeneous solid base catalysts for a wide variety of organic transformations. Furthermore, after calcination at intermediate temperatures (450–600 1C), LDHs can be converted into well-dispersed mixed metal oxides (MMOs) with large surface area and numerous Lewis base sites.8–11 Through subsequent rehydration in the absence of CO2, MMOs can reform the layered structure with hydroxyl anions incorporated in the interlayer region producing activated LDHs with abundant Brønsted-type basic sites.12 The conditions under which the rehydration of the MMO is carried out control the number and the strength of OH sites. As a result, LDHs, MMOs or activated LDHs acting as heterogeneous solid base catalysts are currently attracting increasing attention from the viewpoint of environmental and economical concerns, because they can replace homogeneous base catalysts as environmentally benign and recyclable catalysts for several types of important organic reactions. These reactions involve C–C and CQC bond formation, such as Knoevenagel condensations,13,14 Michael additions,15 aldol and Claisen–Schmidt condensations,16–19 cyanoethylation of alcohols,20 as well as transesterification of

This journal is © The Royal Society of Chemistry 2014

triglycerides with methanol for synthesis of biodiesel.21–26 The catalytic performance of different materials depends strongly on structural features, such as the nature, strength and relative amounts of base sites. The catalytic performance of binary LDHs can be tailored by incorporation of a third cation. For example, Pavel et al.27 recently reported that modification of MgAl–LDH with Y3+ ions followed by calcination/rehydration gave improved activity and selectivity for styrene oxide by styrene epoxidation with hydrogen peroxide in acetonitrile as the solvent. Since its ionic radius is significantly larger than those of Mg2+ or Al3+, the Y3+ cations are present in the form of hydroxide species either in the interlayer galleries or on the LDH surface rather than being isomorphously substituted in the brucite-like layers, and the lower electronegativity of yttrium compared with that of aluminium resulted in an increase of the basicity, accounting for the observed enhanced catalytic activity. In addition to elemental composition, the surface structure and morphology of LDH-based solid base catalysts also have a significant effect on their catalytic behaviour. For example, Lei et al.28 reported a green route to the synthesis of fatty acid monoethanolamides using fatty acid methyl esters as the raw materials and activated MgAl–LDHs as solid base catalysts. They prepared two activated MgAl–LDHs—U-RLDH and S-RLDH—by

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

calcination/rehydration of LDH precursors prepared by the urea decomposition method and the separate nucleation and aging steps (SNAS) method5 developed in our laboratory, respectively. They found that S-RLDH showed a higher catalytic activity than U-RLDH in the amidation of methyl stearate, with 87.0% conversion at 120 1C. It was concluded that the smaller crystallite size and larger surface area of S-RLDH resulted in there being more basic sites on the catalyst surface, as well as the easier desorption of the final product. Although powdered LDHs can be used as catalysts on a laboratory scale, for large scale practical applications they need to be supported on a suitable substrate to facilitate masstransport. For example, Lei et al. fabricated structured MgAl– LDH films on a muscovite mica substrate through an epitaxial growth approach,29 as illustrated in Fig. 3. Atomic force microscopy (AFM) revealed that half-hexagonal shaped LDH platelets grew obliquely on the substrate with a high degree of dispersion. After being activated by a calcination–rehydration procedure, the rehydrated LDH (RLDH) platelets remained firmly immobilized on the muscovite substrate and retained the half-hexagonal platelet morphology (Fig. 3b). In the aldol condensation of acetone, a conversion of 23.0% was achieved over the RLDH/ muscovite acting as a structured solid base catalyst after 460 min. However, the powdered RLDH analogue synthesized by the same procedure exhibited an identical conversion after a longer reaction time of 560 min. In addition, it has been reported that small activated MgAl– LDH nanocrystallites supported on carbon nanofibers (CNFs) showed a catalytic activity almost 300 times higher than that of the bulk catalyst in the transesterification of glycerol with diethyl carbonate to obtain glycerol carbonate.30 In base-catalyzed reactions, the active sites are usually at the edges of LDH crystallites. Therefore, the significantly improved activity was attributed to the small size of the MgAl–LDH crystallites and the accessible pores of the CNFs, which gave rise to a large number of accessible active OH sites in catalysts. In an attempt to overcome the severe mass-transport limitations in biodiesel synthesis from viscous oils, Woodford et al.31 fabricated an alkali-free macroporous activated MgAl–LDH catalyst via physical templating around a sacrificial polystyrene

Fig. 3 (a) Proposed structure of LDH/muscovite based on the lattice match between MgAl–LDH and the muscovite phases, and (b) an AFM image of RLDH/muscovite. Reproduced with permission from ref. 29. Copyright 2012 American Chemical Society.

Chem. Soc. Rev.

Chem Soc Rev

bead template (ca. 350 nm in diameter). The resulting hierarchical macroporous–microporous solid base displayed exceptional activity in the transesterification of long-chain triglycerides and methanol with 24.7% conversion of trilaurin and 26.1% conversion of triolein after 24 h reaction time, due to a great improvement of the diffusion of bulky triglycerides and accessibility of active sites within the hierarchical macropore– micropore architecture. The MMOs with a rock-salt-like structure formed by calcination of multi-component LDHs, as mentioned above, possess abundant acid and basic sites associated with the presence of O2–Mn+ acid–base pairs. The nature, strength and relative amounts of these basic sites depend on the identity and molar ratio of the metal cations, the preparation method, the calcination temperature, and type of interlayer anions.32–39 For example, a LiAl–MMO was recently synthesized from a uniform LiAl–LDH precursor prepared by a separate nucleation and solvothermal aging steps method.40 The synthesis of the LDH precursor involved a very rapid mixing and nucleation process in a modified colloid mill5 followed by a separate ethanol-aided solvothermal aging process. The resulting MMO possessed higher surface basicity than the corresponding material prepared without the addition of ethanol, as a result of both the smaller crystallite size and the higher surface area, and exhibited higher catalytic activity for the Knoevenagel condensation of benzaldehyde (BA) and ethyl cyanoacetate with 83.8% conversion of BA after 12 h. Recently, Meyer et al.41 investigated the influence of the nature of the interlayer anions in MgAl–LDH precursors on the catalytic activity of the resulting MMO in the transesterification of glycerol trioctanoate with methanol. By comparing different intercalated anions, they concluded that MgAl–MMO derived from precursors with highly charged anions exhibited higher catalytic activity than those derived from precursors with long-chain anions. They attributed the higher activity to wider pores, enabling better accessibility for bulky triglyceride molecules. In addition, Ruiz and colleagues found that the solid base catalyst obtained by calcining delaminated MgAl–LDH containing interlayer hydroxyl ions provided excellent catalytic activity (100% conversion after 2 h) in the Meerwein– Ponndorf–Verley (MPV) reaction of benzaldehyde with 2-propanol, due to the high surface basicity of the Lewis type.42 Recently, Tichit and colleagues also prepared a series of new MMOs via alkaline-earth cations or La3+ containing anionic complexes intercalated MgAl–LDH precursor route. The formation of stable complexes was achieved using EDTA or citrate as chelating agent. They found that Sr- and Ba-containing MMOs exhibited basicity close to the ones of MgO and Mg(La)O mixed oxide possessing strong basicity, owing to the improved metal dispersion in precursors.43 2.1.2 Catalytic properties of LDHs with transition metal cations in the layers 2.1.2.1 Electrocatalysts. Replacing rare and expensive noble metal catalysts with inexpensive and abundant ones for various renewable energy-related chemical processes, as well as for the production of high value chemicals, is of considerable importance in making chemical processes more sustainable. LDH materials

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

containing transition metal cations, particularly Ni- and Co-containing LDHs, have been widely used in electrocatalytic reactions.44–49 For example, ZnCo–LDH has been used as an electrocatalyst and catalyst for water and alcohol oxidation, respectively.48 Compared with traditional monometallic Co(OH)2 and Co3O4 systems, ZnCo–LDH possessed a lower overpotential (by B100 mV) for the electrochemical oxidation of water, and the turnover frequency (TOF) of the LDH was 410 times that of Co(OH)2 and Co3O4 at the same applied potentials. Moreover, this material also exhibited good catalytic activity and selectivity for oxidation of alcohols at relatively low temperature. The authors claim that the relatively inactive Zn2+ cations can be regarded as a structural support which provides a synergistic effect in the catalytic system (possibly by assisting the interactions between the reactants and the catalyst). Recently, a core–shell structured CoNi–LDH was built from a Co–Ni hydroxide precursor in a topotactic process using iodine as an oxidizing agent.49 A low overpotential (B0.5 V, at 1.4 V vs. NHE) and high current density (B1 mA cm2) were achieved using the as-constructed CoNi–LDH as a catalyst for the oxygen evolution reaction (OER) without any sacrificial agent being required, which was attributed to both the presence of more Co active sites and the rapid movement of interlayer ions within the layered structure. In addition to the composition of LDH materials, the morphology of electroactive species can also play an important role in improving the electrocatalytic performance. For instance, Sun and colleagues successfully fabricated NiFe–LDH with a three-dimensional (3D) architecture.50 Compared with the conventional 2D planar architecture, an electrode based on the 3D porous architecture could facilitate electron transportation, promote electrolyte penetration and increase the electrochemically active surface area. As a result, the onset potential of the electrode in OER was significantly reduced to B1.46 V vs. RHE, yielding a high current density of B60 mA cm2 at small overpotentials; in addition the 3D architecture showed outstanding stability. 2.1.2.2 Photocatalysts. In view of the increasingly serious problems of environmental pollution and energy shortage, the search for suitable semiconductor photocatalysts for the decomposition of organic dyes and colourless pollutants, as well as water splitting by using solar energy, is one of the primary missions of materials science.51,52 It is well known that many layered compounds that contain Ti4+, Nb5+, or Ta5+ cations possess photocatalytic capability for water splitting, since they may provide spatially separate reduction and oxidation reaction sites.53,54 However, the biggest problem with these layered photocatalysts is the aggregation of photo-active species and thus the low energy-conversion efficiency. Fortunately, dispersing the active species in an inorganic LDH matrix can provide an effective solution to this problem. The appropriate incorporation of photo-active components (e.g., Zn, Ti, Fe or Cr) into LDHs can afford a variety of transition metal-containing LDHs which are semiconductor photocatalysts with a band gap normally in the range from 2.0 to 3.4 eV. There have been numerous investigations on the application of LDH-based

This journal is © The Royal Society of Chemistry 2014

Review Article

photocatalysts in the elimination of organic pollutants, and water reduction and water oxidation using UV or solar light as the irradiation source. (a) Pollutant elimination. The growing demand for the removal of organic pollutants in the effluent from chemical and pharmaceutical industries has stimulated the development of effective photocatalysts. Due to their excellent visible light absorbing abilities, ZnTi–LDH and ZnFe–LDH afford good visiblelight-induced photocatalytic performance in the degradation of organic pollutants.55,56 Recently, Zhao et al.57 prepared a series of MCr–X–LDHs (M = Cu, Ni, Zn; X = NO3, CO32) by the SNAS5 method. They demonstrated that the highly dispersed Cr3+ ions in the octahedral sites in the LDH matrix played a significant role in the photo-excitation of electrons. Notably, MCr–NO3–LDHs (M = Cu, Ni) showed 20-times higher activity than a standard Degussa P25 titania catalyst in the decomposition of dye and phenols under visible light irradiation, as well as excellent recycling ability. In addition, with the introduction of Co2+ into CuCr–LDH, Mohapatra et al. found that active superoxide anion radical species could be generated through the visiblelight-induced metal-to-metal charge transfer (MMCT) transitions of Cr(III)–O–Co(II) to Cr(IV)–O–Co(I) moieties,58 thus giving rise to the higher photocatalytic activity. For the purpose of obtaining immobilized and recyclable photocatalysts for practical water treatment application, CuCr–LDH films immobilized strongly on copper substrates were fabricated by an electrophoretic deposition method.59 Due to the large specific surface area and rich macro-/mesoporous structure, a structured CuCr–LDH film, with an optimized thickness of ca. 16.5 mm and high orientation along the c-axis, showed higher visible-light-driven photocatalytic activity for the decomposition of three representative organic contaminants (about 97% degradation percentage for 2,4,6-trichlorophenol and above 90% degradation percentage for sulforhodamine B (RB) and Congo red after 200 min irradiation) than the corresponding powder sample, as well as excellent recyclability and convenient manipulation. (b) Water reduction for H2 generation. The search for suitable semiconductor photocatalysts for water splitting using solar energy is one of the most important areas of contemporary physical science. In the case of a carbonate-intercalated ZnCr–LDH,60 the ligand-to-metal charge transfer (LMCT) transitions of CrO6 octahedra visible in the UV-vis spectra are thought to be the origin of its ability to photodegrade organic pollutants. More importantly, this material was subsequently found to be a visible-light-active photocatalyst for water splitting to generate hydrogen,61 because the carbonate radicals originating from the oxidation of intercalated carbonate anions by photogenerated holes were found to improve the light-induced electron–hole separation. Furthermore, replacing part of the Zn2+ with Ni2+ also resulted in MMCT transitions of Zn(II)/Ni(II)–O–Cr(III) to Zn(I)/Ni(I)–O–Cr(IV),62 accounting for the enhanced photocatalytic activity for hydrogen evolution. In addition, the high dispersion of TiO6 octahedra confined in the 2D matrix of a series of MTi–LDH (M = Ni, Zn, Mg) can suppress the carrier recombination under visible-light excitation,63 by taking advantage of the large number of surface Ti3+–O defects

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

serving as trapping sites to improve the electron–hole separation. The modified band gap of about 2.1 eV obtained by the formation of the M2+–O–Ti network greatly facilitated water splitting, accounting for the 1.23 eV thermodynamic overpotential and other kinetic overpotentials for both hydrogen and oxygen evolution. As a result, Ti-containing LDHs showed an extremely high photocatalytic H2-production rate of 31.4 mmol h1, which was 18 times higher than that of the control K2Ti4O9 sample. (c) Water oxidation for O2 evolution. In both thermodynamic and kinetic terms, water oxidation is the uphill step in solardriven water splitting. The formation of oxygen from water is challenging since it has to occur through several steps requiring four positive holes and the formation of O–O bonds.64 Recently, Ti- or Zn-containing LDH materials have been tested for their photocatalytic abilities in water oxidation.65,66 LDH-based photocatalysts for O2 evolution under visible irradiation, such as ZnCr– LDH, ZnTi–LDH and ZnCe–LDH, were first reported by Silva et al.65 Of these, ZnCr–LDH displayed two visible absorption bands centred at 410 and 570 nm, which are assigned to the LMCT transitions in the octahedral Cr sites, and possessed the highest apparent quantum yield of 60.9% for O2 generation. This is one of the highest values ever reported for a bifunctional chromophore–catalyst material without the involvement of a foreign photosensitizer. Recently, high-surface-area NiTi–LDH with two visible absorption bands in the blue and red wavelength regions was found to show a higher O2 generation rate than that of CuTi–LDH under visible excitation (49 vs. 31 mmol of O2) by using 200 mg of the photocatalyst and 1 mmol of AgNO3 as a sacrificial agent,66 indicative of the good water oxidation activity of Ti-embedded LDHs. In addition, Kim et al.67 demonstrated that the Co–O–Fe oxobridges connecting the different metal ions in CoFe–LDH helped to prevent the recombination of holes with electrons. In particular, an Fe-rich compound displayed the highest photocatalytic efficiency for water oxidation (about 45 mmol of O2 for 3 h) under visible light, which was attributed to its good crystallinity, enhanced light absorbance and low charge carrier recombination originating from the increase in the number of oxo-bridges. It is well known that the size, shape and morphology of semiconductors are key factors in determining the nature of their surface defects, which affects both the efficiency of charge separation and the photoconversion capability.68–70 In a very recent research, NiTi–LDH nanosheets with lateral dimensions in the range 30–60 nm were synthesized using a reverse microemulsion method,71 in order to increase the density of surface Ti3+ defects in Ti-based LDH materials. This material showed extraordinarily high photocatalytic activity for oxygen evolution from water (B2148 mmol g1 h1) using visible light, which is currently the most effective visible-light photocatalyst reported for O2 production. Moreover, the quantum yields reached 65.0% and 20.0% under monochromatic irradiation at 400 and 650 nm respectively. A proposed scheme that relates the structure and O2 evolution process is shown in Fig. 4. Such LDH

Chem. Soc. Rev.

Fig. 4 (A) Proposed structural model of energy states for NiTi–LDH and (B) schematic illustration of the O2 evolution process over NiTi–LDH nanosheets under visible-light irradiation. Reproduced from ref. 71. Copyright 2014 The Royal Society of Chemistry.

materials can be expected to act as a useful stepping stone for the exploration of new efficient photocatalysts for water oxidation. (d) CO2 reduction. Photocatalytic reduction of CO2 to CO, methanol and other organic species is an attractive way to meet the demands for green energy and environmental sustainability. In this regard, a study of the remarkably highly selective CO2 photoconversion to methanol (88 mol%) was reported using H2 and ZnCuGa–LDHs.72–74 The high catalytic selectivity was mainly attributed to the binding of CO2 at the Cu–O–Zn and Cu–O–Ga sites to form hydrogen carbonate species. In addition, MgIn– LDHs were found to exhibit better activity (B16.4 mmol of O2 and B2.4 mmol of CO after 10 h irradiation) than a simple metal hydroxide for the photocatalytic conversion of CO2 into CO in an aqueous system,75 by taking the advantage of high water tolerance of the surface base sites in MgIn–LDHs. 2.2

Intercalation catalysts

LDH materials have relatively weak interlayer bonding and as a consequence exhibit excellent expanding properties. During the past two decades, increasing interest has been devoted to the use of these layered inorganic solids as host materials in order to create host–guest supramolecular intercalation structures with desirable physical and chemical properties, usually via an ion-exchange route based on the electrostatic interactions between the intercalated anions and the positively charged layers. The versatile intercalation ability of LDHs greatly enlarges the family of available LDH materials. Specifically, the intercalation of catalytically active species (e.g., simple inorganic anions, complex anions, and biomolecules) in the interlayer galleries of LDHs is an effective immobilization approach, and has been proved to be capable of improving catalytic stability and recyclability when compared to the corresponding homogeneous catalyst. More importantly, the brucite-like layers may impose a restricted geometry on the catalytically active guests in the interlayer galleries leading to enhanced control of stereochemistry and the dispersion of the active species, and thus lead to increased reaction rates and modified product distributions.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

2.2.1 Polyoxometalate-intercalated catalysts. Polyoxometalates (POMs) are a subset of metal oxides with unique physical and chemical properties, which can be reliably modified through various techniques and have been used to develop sophisticated materials and devices.76 Since Pinnavaia first reported the synthesis of nanocomposites of bulky and stable POMs and LDHs,77 increasing attention has been paid to POM–intercalated LDHs (POM–LDHs).78–82 Their synthesis has been discussed in a recent review paper.83 POM–LDHs can be used widely as heterogeneous catalysts for various organic transformations including oxidation of alcohols,84 alkenes,85–87 thioethers88 and tetrahydrothiophene,89 and acid-catalyzed esterification,90 and have shown superior catalytic properties to pure POMs, owing to the large gallery height and unique interlamellar chemical environment, as well as the high dispersion of catalytically active species. Recently, a series of POM–intercalated LDHs were reported to exhibit much higher epoxide selectivity than the corresponding homogeneous Na–POM catalysts in the epoxidation of allylic alcohols with aqueous H2O2 as the oxidant without using an organic solvent.91 In particular, the combination of a sandwich-type [WZn3(ZnW9O34)2]12 (ZnWO) guest and a MgAl– LDH host endowed the resulting POM–LDH catalyst with 99% selectivity for the epoxide, 95% H2O2 efficiency and a TOF of 37 200 h1 without the need for any basic additives or pH control. The authors attributed the excellent catalytic performance to the beneficial effect of the basic LDH host in suppressing the acid-catalyzed hydrolysis of the epoxide. Although pristine ZnWO can be used as a catalyst for the oximation of aldehydes, its inherent acidity can cause subsequent acid-catalyzed Beckmann rearrangement and dehydration of the target oximation products leading to nitrile and amide by-products. In an attempt to solve this problem, Song and colleagues92,93 prepared a ZnWO-intercalated ZnAl–LDH with a Zn/Al molar ratio of 3 : 1 in the brucite-like layers, and found that the material showed significantly enhanced selectivity towards aldoximes and ketoximes in the oximation of aromatic aldehydes under organic solvent-free conditions with about 90% conversion of thiophene-2-carboxaldehyde and 85% selectivity to the oxime, showing that the surface basic sites of the ZnAl–LDH host (Fig. 5) do indeed suppress the acidity of the POM guest. The ZnWO–LDH catalyst could be reused at least ten times without significant loss of activity, indicative of its high chemical and structural stability.

Fig. 5 A sandwich POM-intercalated ZnAl–LDH active as a catalyst for highly selective oximation reactions. Reproduced from ref. 92. Copyright 2011 The Royal Society of Chemistry.

This journal is © The Royal Society of Chemistry 2014

Review Article

The intercalation of simple metal oxyanions and POMs in LDH hosts can also lead to photocatalytic capability. For example, when MoO42 and WO42 anions are intercalated in ZnY–LDHs, the MMCT transitions from the Zn–O–Y oxo-bridged linkages leads to an absorption band in the visible region.94 Moreover, the MoO42 intercalated LDH exhibited excellent photocatalytic activity with about 98% degradation percentage of the dye after 2 h irradiation, higher than that by the WO42intercalated analogue (only 80% degradation percentage), due to the contribution of unoccupied Mo 4d orbitals to the conduction band resulting in a smaller band gap. 2.2.2 Biomolecule-intercalated catalysts. In recent years, much attention has been paid to the intercalation of biomolecules (e.g., proteins,95,96 enzymes,97 and amino acids98–101) into the interlayer galleries of LDH nanomaterials, because the adjustable height of the 2D interlayer galleries in LDHs can provide great opportunities for a significant enhancement of their efficiency. He et al.102 presented a detailed investigation of the immobilization of L-proline in the interlayer galleries of LDHs using the calcination–rehydration method of synthesis, along with its application in a typical asymmetric aldol reaction between acetone and benzaldehyde. It was noted that the resulting L-proline-LDHs gave 94% enantioselectivity and 90% yield even when exposed to rigorous conditions (e.g. UV irradiation and thermal treatment). Moreover, the immobilization of L-proline in LDHs significantly inhibited its racemization, and thus enhanced its thermal and optical stability. Similarly, L-proline-LDHs also showed improved catalytic enantioselectivity compared with pure L-proline in the asymmetric Michael addition reaction between b-nitroalkene and acetone.103 Recently, Pitchumani et al.104 prepared sixteen MgAl–LDHs containing different amino acids (including leucine, isoleucine, phenylalanine, methionine, and tyrosine) intercalated in the interlayer galleries, and found that the intercalated amino acids exhibited higher activity in the chemoselective O-methylation of phenols and S-methylation of thiophenol with dimethyl carbonate (DMC), whereas the free amino acids showed negligible activity. The intercalation of amino acids thus clearly created an environment facilitating the reaction in a facile manner with high conversion and selectivity. Moreover, the as-obtained catalysts could be easily recycled and reused without decreasing the catalytic activity. Taking the leucine-intercalated LDH catalyst (Leu-LDH) as an example, a plausible reaction pathway was proposed. As shown in Fig. 6, the amine group of leucine was exposed and the elongated side chain created a hydrophobic pocket in Leu-LDH. The free amine could abstract a proton from phenol generating a phenolate anion, which in turn attacked DMC forming a six membered transition state, which further rearranged to give the anisole product. 2.2.3 Metal complex-intercalated catalysts. Many years ago, Pinnavaia and colleagues used LDHs as layered host systems for interlayer gallery immobilization of complex anions such as [Ru(4,7-diphenyl-1,10-phenanthrolinedisulfonate)]4 via an intercalation pathway.105 The intercalated complex anions are highly dispersed within the structure of LDHs owing to the confinement effect of the 2D LDH layers. In recent years, the intercalation of

Chem. Soc. Rev.

View Article Online

Review Article

Chem Soc Rev

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Fig. 7 Phospholipid bilayer in nature (A) and the schematic structure of the proposed biomimetic catalyst (B). Reproduced with permission from ref. 118. Copyright 2013 American Chemical Society.

Fig. 6 Plausible reaction mechanism for Leu–LDH catalyzed O-methylation of phenol with DMC. Reproduced with permission from ref. 104. Copyright 2013 Elsevier.

various metal complex anions (e.g., metal porphyrins,106 metal salens (salen-N,N 0 -ethylenebis(salicylimine)),107,108 and metal– Schiff base complexes109) into the interlayer space of LDHs has been a versatile and important approach for the immobilization of active metal complex species subsequently employed in a number of organic transformations. For instance, Bhattacharjee et al.110 reported that a chiral sulfonato-salen manganese(III)complex-intercalated ZnAl–LDH was a useful stereoselective epoxidation catalyst, and found that the catalyst gave 100% conversion, 91.7% selectivity, and 98.1% enantiomeric excess in the oxidation of (R)-limonene. Also, a Ce(III)-complex-intercalated ZnAl–LDH catalyst could efficiently catalyze the liquid phase oxidation of primary alcohols to their corresponding aldehydes using molecular oxygen at room temperature,111 and showed excellent yields and good recyclability. However, in same cases, the steric hindrance to the access of the substrates to the active sites of the intercalated species could occur,112 leading to the lowered activity. Homogenous catalysts suffer from the problems of low stability and selectivity, and difficulty in separating the catalytically active component from the product at the end of the reaction. Wei et al.113 synthesized a ZnAl–LDH cointercalated with trans-RhCl(CO)(TPPTS)2 and TPPTS (TPPTS = trisodium salt of 3,3 0 ,300 -phosphanetriyl benzenesulfonic acid), and found that the LDH-intercalated catalyst showed higher activity toward aldehyde formation in the hydroformylation of higher olefins with TOF values of 80–150 h1, compared with the corresponding water–oil biphasic catalytic system (TOFs = 20–25 h1) under similar reaction conditions, as well as better reusability because of the strong interactions between the LDH layers and the active rhodium complex species. Similarly, ZnAl–LDHs cointercalated with a Pd–TPPTS complex and PdCl42 were also found to be efficient green catalysts in the cycloisomerisation reaction of acetylenic carboxylic acids to the corresponding five-membered ring heterocycles with 100% atom efficiency and exclusive formation of the exo-isomer under heterogeneous conditions.114 The catalytic results were superior to those obtained using other heterogeneous systems for this reaction, and were comparable with those obtained under homogeneous conditions. Since no

Chem. Soc. Rev.

external base or external ligand was required, it was proposed that the basicity of the LDH support promotes a clean and selective reaction of acetylenic carboxylic acids. Recently, chiral sharpless titanium tartrate catalyst-intercalated MgAl–LDHs have been successfully prepared as catalysts for the sulfoxidation of the prochiral methyl phenyl sulfide.115–117 After intercalation in the LDH host, the catalysts showed significant increases in enantiomeric excess of up to B50%. Moreover, a significant further improvement in catalytic performance could be achieved by incorporating more highly dispersed active guest anions and/or increasing the interlayer spacing by in situ swelling with organic solvents. In these cases, the confinement effect of the LDH layers played a key factor in enhancing the catalytic properties of the intercalated active species. The formation of such swollen intercalated LDH structures opens up new possibilities for synthesising catalytic materials with enhanced performance. Inspired by the crucial role of the nature of hydrophobic regions in highly efficient enzymatic catalysis, He’s group have recently developed a heterogeneous Mn catalyst by encapsulating Mn(TPP)OAc (TPP = tetraphenylporphyrin dianion) in the biomimetic flexible 2D hydrophobic region of amphiphilic dodecyl sulfonate (DDS)-intercalated ZnAl–LDHs (Fig. 7),118 which has a similar bilayer structure to that of naturally occurring phospholipids. In the DDS-intercalated ZnAl–LDHs, the aliphatic tails of DDS were elongated and adopted a bilayer arrangement in order to maximize guest–guest dispersion interactions, which consequently enlarged the interlayer gallery height and weakened the interlayer interactions, thus facilitating the accommodation of Mn(TPP)OAc. The resulting catalyst exhibited superior catalytic activity and comparable selectivity to its homogeneous counterpart in the epoxidation of a variety of alkenes including cyclohexene, heptylene, phenylethylene, 3-methyl-3-buten-1-ol, and even functional alkenes such as ethyl cinnamate and chalcone. 2.3

Nanocomposite catalysts

Although confinement in the interlayer galleries of LDHs has been reported to be a successful way of enhancing the selectivity toward the desired products for some catalysts, this approach is not always effective owing to the complicated geometrical and chemical microenvironment in the interlayer galleries, as well as the inaccessibility of the inner surfaces of the host layers to the reagents. In this regard, the fabrication of heterostructured LDH hybrid nanocomposites from two functional components is another approach for the design and synthesis of efficient heterogeneous catalysts, which may also benefit from synergistic effects between the two components.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

2.3.1 Nanocatalysts based on LDH nanosheets. Another advantage of LDH (and other layered materials) is that under appropriate conditions they may be delaminated to give single highly anisotropic LDH nanosheets. Due to their high specific surface area and large number of surface functional groups, these nanosheets can be used as supports for immobilization of some macromolecules and biomolecules, including iron porphyrins and amino acids.119,120 Catalytically active anions can also be immobilized on the exfoliated LDH nanosheets to form novel nanocatalysts. However, constraining catalytic centers on the surface of solids usually reduces the reaction rate, due to the emergence of liquid/solid interfaces and diffusion limitations. To solve the above problem, He and co-workers successfully immobilized a vanadium(V) precursor coordinated to amino acid ligands on host ZnAl–LDH nanosheets formed by delamination of the amino acid-intercalated ZnAl–LDHs (Fig. 8).121–123 Their studies indicated that anchoring amino acid ligands on the delaminated ZnAl–LDH nanosheets led to an excellent catalytic performance with a product yield of 93% and enantiomeric excess values of 96% for trans isomers and 63% for cis isomers after 520 min in the asymmetric epoxidation of allylic alcohols. In contrast, the heterogeneous catalysts based on pure amino acid-intercalated only gave a product yield of 83% after a longer reaction time of 1050 min. The impressive enhancement in catalytic performance was explained as follows: (i) the delaminated LDH nanosheets provide a stable and rigid environment around the chiral centre, which significantly improves the enantiomeric selectivity by restricting or directing the access trajectory of reactant molecules; (ii) the easy access of the substrate to the catalytic centers located at the interface between the delaminated nanosheets and the solution led to the great increase in the reaction rate. Interestingly, such colloidal pseudo-homogeneous catalysts obtained by delamination in water could be directly recovered by simple liquid–liquid separation, and therefore reused with negligible loss of catalytic activity and enantioselectivity. Furthermore, the chiral induction

Fig. 8 The a-amino acids used as ligands attached to the brucitelike layers. (K: Zn, K: Al, K: N, K: C, K: O, K: H; blue: Zn–O octahedron, pink: Al–O octahedron). L-Glutamic acid (1), L-alanine (2), and L-serine (3) in pristine (homogeneous) state are labeled a; their anions (green) in intercalated (heterogeneous) or delaminated (colloidal) states are labeled as b and c, respectively. Reproduced with permission from ref. 121. Copyright 2011 Wiley-VCH.

This journal is © The Royal Society of Chemistry 2014

Review Article

of amino acids in the direct aldol reaction was significantly improved by their attachment to the rigid LDH nanosheets, regardless of whether the amino acids acted as chiral ligands for Zn(II) centers or as the asymmetric catalysts in their own right. Theoretical calculations revealed that the nanosheet not only exerted steric effects promoting the chiral induction, but also assisted the formation of transition states via hydrogen bonds. Recently, the direct combination of MgAl–LDH nanosheets and an ionic Mn(II)–salen complex to afford an ordered nanosized sandwich-structured composite was achieved via a modified flocculation method using oppositely charged compounds.124 In this way, the interlayer gallery of the LDH was completely opened up, resulting in no steric hindrance for the immobilization of large guests. In the N-oxidation of 4-picoline using H2O2 as an oxidant, the material exhibited superior thermal stability and 2–10 fold activity over both the corresponding LDH material prepared using a traditional ion exchange method, and the homogeneous analogue. It is believed that the weak basicity of the nanosheets helped to cleave the peroxide bond in H2O2 through the formation of hydrogen bonds between LDH and H2O2, while the nanoscale thickness of the materials significantly decreased the distance needed for diffusion of the substrate to the active site. By taking advantage of the electrostatic forces or/and hydrogen bonding between oppositely charged particles, layer-by-layer (LBL) assembly of LDH nanosheets with negativity charged 2D sheets can be employed to synthesize nanocomposites with excellent catalytic performances. For example, Gunjakar et al.125 reported the assembly of exfoliated positively-charged ZnCr–LDH nanosheets with negatively charged layered titanate nanosheets to form heterolayered nanohybrids (ZCT) by the LBL approach. As shown in Fig. 9, the cross-sectional high resolution transmission electron microscopy (HRTEM) lattice image of the sample exhibited parallel-aligned dark lines with two different spacings of B0.5 nm and B0.7 nm, which can be attributed to the ZnCr–LDH layers and titanate layers, respectively. The effective electrostatic coupling between cationic and anionic inorganic nanosheets resulted in a strong visible light absorption ability and a remarkably depressed photoluminescence signal; the layer-by-layer stacking of the two layered materials also resulted in an enhanced porosity, which could be controlled by changing the ratio of layered titanate to ZnCr–LDH. The resulting ZCT heterolayered nanohybrids were found to be extremely active

Fig. 9 Cross-sectional HRTEM image of the ZCT nanohybrid (a) and its enlarged view and structural model (b). Reproduced with permission from ref. 125. Copyright 2011 American Chemical Society.

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

Chem Soc Rev

catalysts for O2 generation with a rate of B1.18 mmol h1 g1 under visible light, which is higher than that for the pristine ZnCr–LDH (B0.67 mmol h1 g1). The superior photocatalytic activity for O2 production was rationalized by considering the ordering of the heterolayered nanosheets. In addition, the chemical stability of the LDH was greatly improved due to the protection afforded by the layered titanate. This work verified the efficacy of hybridization between two types of nanosheets in constructing performance-enhanced nanocatalysts, and this approach should be capable of extension to hybrids with a range of other layered materials. The exfoliated positively-charged LDH nanosheets have the ultimate in 2D anisotropy which facilitates their use as building blocks for assembly into functional multilayer nanostructures with active species by employing the LBL technique.126,127 Wei’s group assembled cobalt phthalocyanine/LDH ultrathin films (UTFs) using an electrostatic LBL technique.126 They found that the surface of the UTFs was continuous and uniform with a long range stacking order in the direction normal to the substrate. A modified electrode based on this material exhibited a low detection limit of 3.2  107 mol L1, fast response of B2 s and excellent long-term stability for the determination of dopamine. In addition, the LDH nanosheet can be used as a 2D matrix to immobilize metal nanoparticles (NPs). For example, Zhao et al.127 successfully fabricated LDH nanosheets/Au NPs UTFs using the LBL technique. The LDH nanosheets provided a confined and stable microenvironment for the immobilization of Au NPs with a long range stacking order, and the uniform dispersion of positive charge in the LDH nanosheets resulted in the high dispersion of Au NPs. As a result, the multilayers possessed higher electrocatalytic activity and selectively for the oxidation of glucose than that of previously reported Au NP-based glucose sensors. Also, Lee et al.128 immobilized Pt NPs on exfoliated MgAl–LDH nanosheets through an electrostatic self-assembly processes. The resulting catalyst exhibited higher catalytic activity for the reduction of p-nitrophenol into p-aminophenol by NaBH4 with a TOF value of 0.75 min1, in comparison with the commercial Pt/C catalyst (TOF = 0.56 min1). 2.3.2 Nanocatalysts based on multifunctional nanocomposites. Co-assembly of nanocomposites consisting of various LDHs and other functional materials is an efficient strategy for developing multi-component or bifunctional heterogeneous catalysts. These functional components usually are semiconductors, metals and carbon-based materials.

the degradation percentage of phenol increased significantly from 27% over the pristine ZnLDH to 85% over the hybrid TiO2/ ZnLDH after 240 min UV illumination. In addition to the most widely used semiconductor TiO2,129–131 CeO2 has also been assembled with MgAl–LDH to form efficient CeO2/MgAl–LDH photocatalysts for the degradation of phenol and chlorinated phenols under UV irradiation.132 Photocatalysis by these bifunctional basic/semiconducting photocatalysts with a band gap energy of B3.2 eV was shown to be most likely triggered either by UV-induced excitation of electrons from the valence band (VB) to the conduction band (CB) of CeO2 or by the formation of a charge-transfer complex at the defect sites near to the CeO2/LDH interfaces. CeO2 has also been incorporated into ZnTi–LDH to form an efficient hybrid CeO2/Zn2TiO4 photocatalyst via spontaneous formation and organization of CeO2 NPs on a ZnTi–LDH matrix using a method based on LDH calcination plus hydration in an aqueous solution of Ce(SO4)2, followed by a final calcination process.133,134 Compared with CeO2/MgAl–LDH, the resulting CeO2/Zn2TiO4 photocatalyst exhibited a much high efficiency for phenol removal (about 90%). It was argued that the insertion of Zn2+ and Ti4+ within the LDH network resulted in a good dispersion of the metal cations within the layers which acted as charge separation centers enhancing the degradation efficiency. In a recent report, Shao et al.135 prepared novel hierarchical nanoarrays based on a ZnO nanowire core and a CoNi–LDH platelet shell on fluorine-doped tin oxide (FTO) electrodes. The resulting ZnO@CoNi–LDH core–shell nanoarray showed a high photocurrent density of 1.5 mA cm2 at +0.5 V under illumination, about three times larger than that of pristine ZnO nanowires, as well as excellent stability in photoelectrochemical water splitting. This was attributed to the successful photogenerated electron–hole separation originating from the intimate contact at the interface between the ZnO core and the electrocatalytically active Co-containing LDH shell. In addition, Ag/AgBr nanomaterials with plasmonic photocatalytic performance were combined with a nitrate-intercalated CoNi–LDH to form bifunctional visible-light-induced nanocomposite photocatalysts for degradation of anionic dyes.136 In this nanocomposite structure, ‘‘hot’’ electrons from the Ag surface plasmon resonance induced by incident visible light were transferred to the CB of AgBr under visible light, and trapped by the dioxygen molecule to form O2 or other reactive oxygen species. Moreover, the intimate attachment of AgBr to the surface of the LDH further favoured the formation of  OH radicals from OH groups of LDH induced by holes in the VB of AgBr.

(a) Semiconductor/LDH nanocomposites. By hybridization with other semiconductors in composites, the photocatalytic activity of LDHs is expected to be enhanced because of the resulting increase in the lifetime of electrons and holes. Recently, a mesoporous zinc-based LDH (ZnLDH)-supported TiO2 was constructed by reconstruction of ZnLDH in an aqueous solution of TiOSO4.129 Notably, the high surface area of TiO2/ZnLDH favoured light harvesting, while the complex heterojunctions between the two components facilitated the transfer of photogenerated electron– hole pairs during the course of photocatalysis. As a result,

(b) Carbon-based nanocomposites. Carbon-based materials, such as carbon nanotubes (CNTs), carbon fibres and graphene, have been widely used as catalyst supports in a variety of heterogeneous reactions, due to their fascinating physicochemical properties including their high aspect ratio, high specific surface areas, versatile and abundant surface functional groups after surface modification, high mechanical strength, excellent thermal stability, and good electronic conductivity. Among carbon-based materials, CNTs are commonly considered as ideal candidates for assembling functional nanocomposites

Chem. Soc. Rev.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

used in photoelectrochemical or electrochemical applications. Incorporating CNTs and LDHs into nanocomposites has attracted the interest of many researchers. In LDH/CNT nanocomposites, LDHs can retain their intrinsic structure, while the CNT support enhances the dispersion of LDHs and the heat and mass transfer during the reaction, and favours electron transfer throughout the composites. Therefore, combining CNTs with LDHs can result in enhanced performance to meet the requirements of advanced catalytic applications. For example, multi-walled CNTs have been explored as a functional component to enhance the electrocatalytic activity of NiAl– LDH.137 By means of a simple in situ coprecipitation method, the NiAl–LDH nanocrystallites were highly dispersed and anchored on the surface of a modified CNTs matrix by virtue of the strong interfacial electrostatic interactions between them. An electrode modified by the NiAl–LDH/CNTs nanocomposite exhibited eight times higher electrocatalytic activity for glucose electrooxidation than those modified by either pristine NiAl–LDH or CNTs. The enhanced electrocatalytic activity was attributed to the fact that CNTs efficiently promoted the charge transport between the active Ni centers and the electrode, and the CNTs also constructed a porous network-like structure which enhanced the diffusion of the reactant. Recently, a NiFe–LDH/CNTs hybrid was assembled by direct nucleation and growth of ultrathin NiFe–LDH nanoplates on the functional groups of mildly oxidized multi-walled CNTs.138 Fig. 10 shows the interconnected electrically conducting networks of the NiFe–LDH/CNTs nanocomposite, where the NiFe–LDH nanosheets were dispersed on the surface of CNTs. The resulting hybrid NiFe–LDH/CNTs exhibited higher electrocatalytic activity for oxygen evolution in alkaline media, and higher stability, than either commercial precious metal Ir catalysts or the NiFe–LDH nanosheets themselves. Moreover, the TOF value of the NiFe–LDH/CNTs catalyst was about 3 times higher

Fig. 10 Ultrathin NiFe–LDH nanoplates grown on CNTs. (a) Scanning electron microscopy (SEM) image of NiFe–LDH nanoplates grown on a network of multi-walled CNTs. (b) and (c) Transmission electron microscopy (TEM) images of a NiFe–LDH/CNTs hybrid. (d) X-ray diffraction (XRD) patterns of NiFe–LDH/CNTs and a control b-Ni(OH)2/CNTs sample. (e) Schematic showing the hybrid architecture and LDH crystal structure. Reproduced with permission from ref. 138. Copyright 2013 American Chemical Society.

This journal is © The Royal Society of Chemistry 2014

Review Article

than that of mixed nickel and iron oxide electrocatalysts previously reported and also comparable to that of the most active perovskite-based catalyst.139 The authors attributed the high activity of the nanocomposite catalyst to the strong interactions between NiFe–LDH and CNTs facilitating electron transport. As a zero band gap semiconductor carbon-based material, graphene with a special 2D sheet-like structure possesses remarkable electrical, thermal and mechanical properties.140,141 In particular, its robust and flexible structure with high carrier mobility affords an excellent electrical conductivity.140 Therefore, graphene has been intensively investigated as a substrate to grow and anchor functional nanomaterials. Recently, Hwang et al. reported the immobilization of ZnCr–LDH nanoplates on the oppositely-charged graphene nanosheets.142 The strong electronic coupling between the two components in the hybrid ZnCr–LDH/ graphene nanocomposite gave rise to significantly improved visible light absorption. Consequently, the nanohybrid exhibited excellent photocatalytic activity for O2 generation under visiblelight irradiation with a rate of B1.20 mmol h1 g1, much higher than that over pristine ZnCr–LDH (B0.67 mmol h1 g1). More recently, a nanocomposite of NiTi–LDH/reduced graphene oxide (RGO) nanosheets was synthesized by an in situ growth approach.143 The hybrid composite possessed excellent photocatalytic activity for water oxidation under visible light irradiation, showing an O2 generation rate of 1.968 mmol g1 h1 and a very high quantum efficiency of 61.2% at 500 nm, owing to the efficient electron transfer from NiTi–LDH to the RGO matrix. In addition, Xu et al.144 also reported the fabrication of NiAl–LDH/ graphene nanocomposites using a conventional coprecipitation method under low-temperature conditions and subsequent reduction of the supporting graphene oxide. Notably, the NiAl–LDH/graphene modified electrode exhibited an oxidation current of 2.06 mA for dopamine electrooxidation, which was 2.8 times higher than that over the pristine NiAl–LDH modified electrode. The above results confirm the benefits arising from hybridization between photoactive or electroactive LDH materials and electron-accepting graphene nanosheets which lead to highly efficient catalysts and electrocatalysts with improved activity. (c) Magnetic nanocomposites. Magnetically separable nanocomposite catalysts have attracted tremendous interest because of the minimization of the consumption of additional materials, energy and time required for their separation from a reaction mixture, which can give significant economic and environmental benefits.145,146 Particular interest has focused on magnetic nanocatalysts with the core@shell structure, which combine the advantages of both a magnetic core and an active shell and the possible synergistic effects between them.147 In recent years, the synthesis of core–shell type magnetic LDH composite has been developed by several research groups. For example, Li et al.148 assembled core–shell Fe3O4@MgAl–LDH using a LBL method with delaminated LDH nanosheets and a magnetic matrix. After loading W7O246 anions, the magnetic nanocomposite showed 93.6% degradation percentage of hexachlorocyclohexane after 12 h under visible light irradiation,

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

Fig. 11 SEM (a) and (c) and TEM (b) and (d) images of Fe3O4@LDH-1 (a) and (b) and Fe3O4@LDH-2 (c) and (d). Reproduced from ref. 150. Copyright 2013 The Royal Society of Chemistry.

and the activity remained unchanged after six cycles or reuse. Another example involves a hierarchical core–shell Fe3O4@MgAl– LDH with a honeycomb-like structure, which was prepared by the selective deposition and formation of MgAl–LDH onto the surface of Fe3O4 submicrospheres (220 or 450 nm in diameter).149 Interestingly, the loading of Au NPs on the surface of resulting nanocomposite afforded a material giving 99% yield of acetophenone in 1-phenylethanol oxidation. The same research group also synthesized two hierarchical core–shell Fe3O4@CuMgAl–LDH magnetic nanocatalysts,150 including LDH particles horizontally coated onto the core surface (Fe3O4@LDH-1) and hexagonal plate-like LDH particles vertically grown on the core surface (Fe3O4@LDH-2) (Fig. 11). The two catalysts both showed good magnetic susceptibility and significantly higher catalytic activity in phenol hydroxylation than the pure CuMgAl–LDH. The improved performance was explained by a Fenton-like oxidation process, which was enhanced by the HO–Cu2+–O–Fe2+–O– linkages formed at the shell–core interface. It is worth noting that the horizontal growth of CuMgAl–LDH particles afforded a more effective nanocatalyst, which can be attributed to the larger synergistic effect between the thin LDH shell and the Fe3O4 core. In addition, the as-prepared nanocatalysts showed excellent ease of separation, and high stability and efficiency of reuse. 2.4

Complex and mixed metal oxide catalysts

LDHs can be used as precursors to generate a variety of MMO materials by controlled thermal decomposition. The transformation of the LDH precursors into MMOs is topotactic owing to the ordering and uniform distribution of metal cations in the layers. MMOs have large specific surface areas (100–300 m2 g1), basic properties, a homogeneous and thermally stable dispersion of the metal ion components, synergistic effects between

Chem. Soc. Rev.

Chem Soc Rev

the elements, and the possibility of structure reconstruction under mild conditions, all of which are very important attributes for catalysts. Furthermore, the structures and properties of MMOs can be tuned by changing the composition of the LDH precursors, resulting in a huge family of new materials.151,152 The resulting multi-metal MMOs, especially those containing transition metals, can be used as heterogeneous catalysts for photodegradation of organic pollutants, reduction of NOx and SOx emissions, oxidation/total combustion of (volatile) organic compounds, and chemical transformations for the synthesis of fine chemicals, and usually exhibit higher activities and longer lifetimes than catalysts prepared by traditional methods. 2.4.1 Photodegradation of organic pollutants. In recent years, environmentally friendly complex metal oxide photocatalysts with long life and excellent catalytic activity derived from LDHs have attracted increasing attention. For example, Li and colleagues reported a novel synthesis of high-surfacearea ZnAl2O4 complex metal oxides with well-developed interconnected mesopore networks.153 The synthesis procedure involved the formation of a two-phase composite (spinel and zinc oxide) by thermal conversion of single-source ZnAl–LDH precursors at 500 1C or above, followed by selective leaching of the ZnO phase from the resulting calcined products. Compared with materials prepared by the conventional solid-state method, the as-prepared spinels showed better photocatalytic activity for degradation of phenol in aqueous solution under UV irradiation due to their high surface areas of 253 m2 g1 and small pore size. Furthermore, to facilitate practical applications, a ZnFe2O4 film supported on a sulfonated silicon substrate was also fabricated using the above method,154 and exhibited greatly enhanced photocatalytic activity in the degradation of rhodamine B compared with pure ZnFe2O4 powder prepared by the conventional solid-state method. Such structured ZnFe2O4-based film photocatalysts may have potential applications because of their high thermal stability, strong resistance to acid and alkali, and ease of separation. Recently, Zhao et al.155 demonstrated that ZnAl2O4 particles were uniformly dispersed in a ZnO matrix after controlled thermal conversion of ZnAl–LDH. The heterojunctions formed, and the band gap coupling between them, improved the spatial separation of photogenerated carriers under UV irradiation leading to much enhanced photocatalytic performance in the degradation of methyl orange in comparison with ZnO or ZnO/ ZnAl2O4 catalysts prepared by coprecipitation or physical mixing. In addition, a new strategy for constructing efficient and recyclable MMO-type photocatalysts with hierarchical biomimetic structures was reported by Wei and colleagues via a biotemplate-assisted LDH precursor route.156 They used a legume as a sacrificial biotemplate to prepare a biomorphic hierarchical ZnAl–MMO with a meso/macroporous framework (Fig. 12). The as-fabricated hierarchical structured MMO material was applied for UV-lightdegradation of two dyes, RB and a azobenzene-containing polymer poly{1-4[4-(3-carboxy-4-hydroxyphenylazo)benzene sulfonamido]1,2-ethanediyl sodium salt} (PAZO). The catalytic degradation by the MMO powder was 8.1% for PAZO and 74% for RB in water after 2 h. In contrast, the fraction of the decomposed dye

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

Fig. 12 Schematic representation of the fabrication of a biotemplated LDH film and MMO framework from a legume. A uniform Al2O3 coating was deposited on the surface of the biotemplate with an atomic layer deposition (ALD) process; the film of ZnAl–LDH was prepared by an in situ growth technique; finally, the ZnAl–MMO framework obtained by calcination of the LDH precursor. Reproduced with permission from ref. 156. Copyright 2009 American Chemical Society.

was 29% for PAZO and 81% for RB by the biomorphic MMO framework. The enhanced photocatalytic activity was attributed to the rather high specific surface area as well as the presence of hierarchical porosity. As in the case of semiconductor photocatalysts, the high exposure of active facets and a high degree of dispersion also play important roles in improving the catalytic performance. For example, hexagonal ZnO nanoplatelets with a high percentage of exposed (0001) facets and high dispersion were grafted on a hierarchical flower-like matrix,157 based on an in situ topotactic transformation of a ZnAl–LDH precursor grown on amorphous alumina (Fig. 13). The resulting ZnO-based nanostructure possessed a lower band gap of 2.90 eV than the control samples including ZnAl–MMO powder (3.0 eV), ZnO nanoplates (3.04 eV) and ZnO nanorods, and exhibited much

Fig. 13 SEM images of ZnAl–MMO: (a) an overall view of the ZnAl–MMO microspheres; (b) a single ZnAl-microsphere; (c) the nanoflakes of a ZnAl– MMO microsphere; (d) a portion of a ZnAl–MMO nanoflake with hexagonal ZnO nanoplatelets embedded on its surface (inset: a cross-section of the ZnAl–MMO nanoflake). Reproduced from ref. 157. Copyright 2011 The Royal Society of Chemistry.

This journal is © The Royal Society of Chemistry 2014

Review Article

higher photocatalytic performance for visible-light-degradation of RhB compared with the above control samples, mainly because of the presence of more defects at the highly active exposed (0001) facets. On the other hand, semiconductor photocatalysts formed by coupling two kinds of metal oxides together were found to be able to improve the separation efficiency of electron–hole pairs through the proper modulation of electronic structure. Recently, Fan et al.158 reported the synthesis of visible-lightinduced heterostructured ZnAlIn–MMO nanocomposite photocatalysts from ZnAlIn–LDH precursors. In the composites, well dispersed amorphous In2O3 domains were intimately attached to ZnO nanocrystallites, thus forming the heterostructure of ZnAlIn–MMO. Under visible light irradiation, the fraction of the decomposed methylene blue could reach 73% by the ZnAlIn– MMO nanocomposite with an estimated band gap of B2.50 eV. In constrast, the catalytic degradation by pure In2O3 was only 48%. This was attributed to the efficient separation and transportation of photogenerated charge carriers originating from the heterostructure, as well as the large specific surface area. Besides MMOs containing Zn, In and Fe, other photoresponsive MMOs can be obtained when active components (e.g., Ni, Ti) are incorporated into LDH precursors. Ti-containing inorganic solids have attracted much attention owing to their multifunctionality, especially in energy conversion and catalytic reactions. One example is a Ti-containing NiO–NiTiO3 nanocomposite obtained by the thermal conversion of a NiTi–LDH precursor at 900 1C and above.159,160 The Ti-based nanocomposite exhibited good photocatalytic activities for the degradation of methylene blue under both UV- and visible-light irradiation, which was attributed to the synergistic effect between the mixed wide band gap of semiconductor NPs in the UV region and the photoabsorption of NiTiO3 in the visible-light region. In general, the studies addressed above provide very useful design and synthesis strategies for efficient and versatile MMO materials by the rational combination of the elements in the LDH layers and the structural tunability of LDH precursors, leading to new and effective photocatalysts for the elimination of hazardous pollutants. 2.4.2 Elimination of NOx and SOx emissions. The increasing use of fossil fuels such as coal, natural gas and petroleum for power generation and transportation has greatly increased NOx (N2O, NO and NO2) and SOx (SO2 and SO3) emissions, leading to serious environmental problems. Thus, the elimination of NOx and SOx is becoming a more and more urgent mission for chemists and related researchers. In recent years, many researchers have explored different types of transition metal-containing MMOs derived from LDH precursors for the reduction of NOx emission. In particular, LDH-derived catalysts containing different transition metals (e.g., Co, Mn, Ni, Cu) that can either directly decompose NOx or effectively reduce NOx in the presence of excess O2 are widely sought.161–166 The NOx storage/reduction MMOs derived from Co– and Ca–LDHs,167,168 which are used in engines that alternately operate under lean-burn and rich-burn conditions, may be good candidates. The increased surface redox species,

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

alkalinity and reducibility of the catalysts are believed to be associated with their improved catalytic activity. Recently, alkali metal (Li, Na, K, Rb, Cs) promoted CoMnAl– MMOs (with a molar ratio of alkali metal/Co = 0.037), derived from CoMnAl–LDHs with the molar ratio Co : Mn : Al = 4 : 1 : 1, were found to show an increased catalytic activity for N2O decomposition in inert gas and simulated waste gas from HNO3 production, which decreased in the order Cs 4 Rb 4 K 4 Na = Co4MnAlOx 4 Li, i.e., the catalytic activity increased with the increasing ionic radius of the alkali metal cation for the same surface coverage.169,170 The promotional effect of alkali metals was related to their ionization potential, the charge donation from alkali metal cations to the oxide form of surface oxygen and further to cobalt and manganese, and a decrease in the binding energies of the catalyst components. This work strongly indicated that transition metal-containing MMOs for effective NOx transformation can be rationally synthesized by the combination of transition metals in the LDHs precursor and other promoters, and are potential important catalysts for environmental applications. Calcined transition metal-containing LDHs have been investigated as potential materials for the reduction of SOx emission from the fluid catalytic cracking (FCC) process in oil refineries.171 Corma et al. found that CuMgAl–MMO could catalyze both the oxidation of SO2 to SO42 in the FCC regenerator, and the reduction of sulfates to H2S in the reducing atmosphere of the cracking zone. Notably, CuMgAl–MMO172 and CoMgAl–MMO,173 subsequently activated by heating under H2, could simultaneously remove SOx and NOx. Moreover, CuAl– MMOs and MgAl–MMOs were found to be able to capture a large amount of SOx.174,175 Introduction of different metal species (e.g. Ce, Fe, Cu, Mn) to LDHs, either by impregnation or coprecipitation, has been proven to be the best way to generate the mixed oxides/spinel solid solutions with both basic and redox properties for the SOx emissions.174–178 Polato et al.179 demonstrated that the SOx storage capacity of transition metal doped MgAl–MMOs with a transition metal/total metal molar ratio of around 0.11 decreased gradually in the order Cu(II) 4 Co(II) 4 Fe(III) 4 Cr(III). Polato et al.180 also prepared Mn, Mg, Al-spinels from LDHs, and found that the SO2 uptake capacity of Mn, Mg, Al-spinels could reach 2890 mmol of SO2 per g only after 10 min time on stream (TOS), much higher than that of Mg, Al-spinels (about 400 mmol of SO2 per g after 4 h TOS). In addition, the incorporation of CeO2 in Mn, Mg, Al-spines could not only increase the amount of SO2 uptake but also favour the catalyst regeneration efficiency.181 2.4.3 Oxidation of volatile organic compounds. Emission of volatile organic compounds (VOCs) from many industrial processes and transport vehicles represents a serious environmental problem. Compared with the elimination of VOCs by thermal combustion, the catalytic total combustion of the pollutants to carbon dioxide and water is more effective and economical in terms of energy and cost. In general, noble metals exhibit excellent catalytic oxidation properties. However, the high cost and low stability of noble metal catalysts greatly limits their large-scale practical applications. Therefore, in terms of

Chem. Soc. Rev.

Chem Soc Rev

economic and environmental sustainability, numerous efforts have been made to design efficient and recoverable MMOs for the total combustion of some commonly encountered VOCs, such as methane and ethanol, by employing transition metal (Cu, Mn, Co, and Ni)-containing LDHs as catalyst precursors.182–191 Recently, Li et al.192 found that Pd-modified CoAl–MMO derived from a CoAl–LDH with the molar Co/Al ratio of 3.0 was much more active in toluene elimination than the catalyst prepared by a traditional thermal combustion method, owing to the combined effects of the high surface area, small mean crystallite size of the nano-sized support and highly dispersed PdO particles, as well as the presence of a large number of oxygen vacancies on the support surface serving as the site of oxygen activation. Moreover, the strong synergistic effects between Co3O4 and PdO also contributed to the high catalytic activity. In another example of use of LDH-derived ˘ et al.193 synthesized a MMOs for elimination of VOCs, Urda range of rare-earth metal-promoted MgAl–MMOs derived from LDH precursors, and found that the most active and highly stable catalyst was a Ce-containing material containing about 4.5 at% surface Ce with respect to total cations, which gave total conversion in methane oxidation at 700 1C. The correlation observed between the intrinsic catalyst activity and H2 consumption in temperature-programmed reduction experiments suggested the involvement of an oxidation–reduction process based on the involvement of the Ce(IV)/Ce(III) redox couple in the catalytic reaction. 2.5

Metal catalysts

In heterogeneous catalysis, synthesis of catalyst NPs with controlled size and uniform distribution is a major challenge. In both academia and industry, the most widely used preparation approaches for supported metal catalysts are deposition–precipitation and incipient wetness impregnation. In most cases, however, it is difficult to obtain well-dispersed metal NPs with uniform size and good thermal stability, especially at high loadings. This results from the inhomogeneous distribution of active precursors on the supports as well as the weak interaction between them. In addition, the agglomeration of metal NPs during the catalytic reaction—which can further decrease the dispersion of metal species—is always a problem. Therefore, the design and synthesis of heterogeneous metal catalysts with high dispersion and stability is very necessary. The layered matrix of LDHs offers multiple advantages as a host for NPs, such as ease of manipulation of the brucite-like layer composition by incorporating different active metal anions, and the good dispersion of the metal cations within the layers. Upon calcination and subsequent reduction, such LDHs usually give rise to metal particles well-dispersed in a mixed oxide matrix. This structural transformation endows LDH materials with extraordinary capability as a precursor to metal catalysts for use in a variety of reactions. Compared with the traditional methods for the synthesis of supported metal catalysts, the use of LDH precursors has three advantages: (i) as noted above, the uniform distribution of active metal cations at the atomic level without a segregation of ‘‘lakes’’ of separate cations within the

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

brucite-like layers, as verified by Grey et al.2,194 and Cadars et al.195 by means of multinuclear nuclear magnetic resonance spectroscopy for MgAl–LDH, ensures high homogeneity in a LDH precursor, which enhances the catalytic activity of transition metal containing LDHs as catalysts in their own right; however, it also facilitates the formation of highly dispersed and stable metallic species upon calcination and subsequent reduction; (ii) the strong metal–metal oxide support interactions between the metal oxide and metal NPs—formed concomitantly during calcination of the LDH precursor—can prevent sintering/ aggregation of the metal NPs during the reaction; (iii) the particle size and specific morphology of metal NPs can be controlled in situ by the topotactic nature of the transformation of LDHs into MMOs, and/or the effect of confinement of a catalyst precursor in the LDH interlayer galleries.196 The above characteristics of LDHs open opportunities for developing powerful new metal-based catalyst systems. 2.5.1 Integrated monometallic nanocatalysts (a) H2 generation and hydrogenation. Hydrogen represents an important alternative energy feedstock for both environmental and economic reasons. Over the past decade, due to their high activity and low cost, numerous transition metal catalysts such as copper, nickel, cobalt on a metal oxide matrix have been prepared by calcination and reduction of LDH precursors, and employed in a variety of heterogeneous reactions for the purpose of H2 production; these include the partial oxidation of methane,197 steam reforming of methane,198,199 steam reforming of methanol and ethanol,200–205 and the water–gas shift reaction.206,207 Moreover, the use of these LDH precursors allows the concentration of each metal component in the mixed oxides to be controlled, while a more homogeneous active phase distribution can lead to better catalytic activity. ´ et al.208 investigated the steam reforming Recently, Montane of ethanol (SRE) over Ni-based catalysts derived from NiFe– LDHs. It was found that the best catalyst, with a Ni/Fe ratio of 1 and calcined at 500 1C, rendered high and stable hydrogen selectivity of up to ca. 60%, low methane content, and consisted of a mixture of Ni(Fe)Ox solid solution and NiFe2O4 with very small crystallites. This was because lower Ni/Fe ratios strongly promoted ethanol dehydrogenation and acetaldehyde decarbonylation followed by steam reforming likely due to the promotion effect of Fe/Fe2+ on the reducibility of nickel species and the formation of small nickel particles. However, crystallization and growth of NiFe2O4 present in catalysts at high calcination temperature (800 1C) led to lower activity and fast deactivation, due to active Ni0 sintering and higher carbon deposition. Moreover, the promoter effect of various metals (La, Ce, Ca, etc.) has also been assessed in some cases.209,210 For example, Assaf et al.209 demonstrated that the intercalation of Ce- and La-containing complexes into NiMgAl–LDH precursors by anion-exchange resulted in better H2 production per mol of ethanol in the feed at 550 1C on the resultant Ni-based catalysts in ESR. Especially, on the Ce-modified Ni catalyst, the highest ethanol and acetaldehyde conversion rates were achieved, possibly due to a favoring of water adsorption on the weakly interacting clusters of CeO2 on the surface. More importantly,

This journal is © The Royal Society of Chemistry 2014

Review Article

the catalysts did not favor the formation of significant carbon deposits throughout the time on stream. The development of highly selective, efficient and economical catalysts for H2 generation from hydrogen storage materials (e.g., NH3BH3, N2H4H2O) has also been one of the most active research areas. Transition metals such as metallic Co, Cu and Ni are very efficient catalysts for NH3BH3 dehydrogenation. Recently, Wei’s group reported a series of Cu-based catalysts derived from CuCoAl–LDH precursors for the dehydrogenation of NH3BH3, and found that the catalyst with a Cu/Co molar ratio of 1/1 yielded a hydrolysis completion time of less than 4.0 min with a rate of 1000 mL min1 gcat1 under ambient conditions,211 which is comparable to that reported for the best noble metal catalysts (e.g., Ru, Pt). It was shown that the synergistic effects between highly dispersed metallic Cu and Co3O4 spinel species played a key role in the significantly enhanced activity of these Cu-based catalysts. In addition, the synergistic effect between the small highly dispersed Ni NPs and the strong basic sites on the surface could be formed in noble-metal-free Ni-based catalysts derived from NiAl–LDH precursors,212 giving 93% selectivity to H2 for the decomposition of N2H4H2O at ambient temperature. Recently, the hydrogen-free synthesis of 1,2-propanediol (1,2-PDO) from glycerol was performed efficiently over Cu-based catalysts derived from CuMgAl–LDHs.213 It was found that the dehydration of glycerol occurred mainly on the basic sites of the catalyst, and the acetol intermediate formed was further hydrogenated to 1,2-PDO over metallic copper NPs using active hydrogen atoms from the dehydrogenation of ethanol as the hydrogen source. The activity of catalysts depended mainly on its basicity, while the presence of metallic copper NPs was indispensable for dehydrogenation of the ethanol. The maximum conversion of glycerol reached 95.1%, with a selectivity to 1,2-PDO of higher than 90%. Recently, Li and co-workers employed a facile single-source precursor route to achieve green, well dispersed supported Cr-free Cu-based catalysts from CuMgAl–LDH precursors and then used these for the vapour-phase hydrogenation of dimethyl-1,4cyclohexane dicarboxylate (DMCD) to 1,4-cyclohexane dimethanol (CHDM), a commercially important diol.214 Their studies indicated that the catalyst could achieve a long-lasting (up to 200 hours) 100% conversion of DMCD with 99.8% selectivity to CHDM. They ascribed the unprecedented catalytic hydrogenation performance to the synergistic effect between the active Cu0 sites highly dispersed on the surface and the surface Lewis base sites from neighbouring MgO (Fig. 14). It was demonstrated that Cu0 species were effective catalytic centers for the hydrogenation

Fig. 14 Schematic representation of the hydrogenation mechanism over Cu-based catalysts derived from CuMgAl–LDH precursors. Reproduced from ref. 214. Copyright 2013 The Royal Society of Chemistry.

Chem. Soc. Rev.

View Article Online

Review Article

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

of DMCD. The quantity of dissociative hydrogen on the active sites directly determined the rate of the reaction. Additionally, surface Lewis basic (LB) sites coming from MgO, especially LB sites in intimate contact with metallic Cu0 sites, could easily interact with the p*-acceptor orbital of the CQO group by forming a zwitterionic, tetrahedral intermediate with nucleophilic reactivity at the oxygen atom, thus improving the reactivity of the ester group in the DMCD molecule. (b) Growth of carbon materials. The use of the regularly arranged transition metal cations in the brucite-like layers of LDHs as precursors for the in situ formation of highly dispersed metal NPs (Fe, Co, and Ni) through calcination and reduction, has been proposed as a valid approach for the growth of carbon nanomaterials from catalytic chemical vapour deposition (CCVD). The chemical composition, size, shape, and crystallographic orientation of the NPs can be easily tuned through controlled calcination and reduction of LDHs and selection of the metal cation combination, and has a significant effect on the microstructures and morphologies of the carbon materials formed. Meanwhile, the interaction between the active metal and the support is responsible for the quantity and quality of carbon materials. Li and colleagues first synthesized multi-walled CNTs with a uniform diameter ranging from 20 to 30 nm by CCVD of acetylene at 700 1C over highly dispersed Co NPs derived from a CoAl–LDH precursor.215 Such an LDH precursor route for a Co-based catalyst enables scalable preparation of CNTs at low cost. Furthermore, three types of carbon materials (nanotubes, caterpillar-like fibers and interwoven spheres) could be grown on a silicon-supported Co-based catalyst derived from a CoAl– LDH precursor.216 The morphologies of the carbon materials could be easily tuned by simply adjusting the growth duration. A growth mechanism was proposed, based on an overgrowth of the initially formed tubular nanostructures. Zhang et al.217 investigated the effect of the composition of NiMgAl–LDHs on the formation of CNTs from CCVD of acetylene. The yield of carbon increased with increasing Mg content, because the spinel phases (Ni1xMgxAl2O4) formed by calcination of LDHs inhibited the growth of active Ni particles, thus leading to good metal dispersion and small catalyst particle sizes. Moreover, larger amounts of Mg also induced the formation of coiled CNTs with a helical structure and an outer diameter of ca. 20 nm. The same workers also synthesized uniform twist-shaped single-helical CNFs with a coil pitch of B80 nm, and herringbone-type double-helical carbon fibers with a coil diameter of B1–2 mm over NiCuAl–LDH and NiCuMgAl– LDH, respectively.218 The growth of fibrous carbons with different morphology was correlated with the diffusion of carbon species on different crystal planes of the two different active metals (Ni and Cu), which controlled the deposition of graphitic layers. After loading with Pt, an electrode supported on the herringbonetype double microcoiled carbon fibers showed much larger Pt active surface area than that supported on commercial carbon black. Furthermore, compared with the commercial material, the new electrocatalyst exhibited a four-fold enhancement of activity

Chem. Soc. Rev.

Chem Soc Rev

and excellent anti-poisoning ability in the electrocatalytic oxidation of methanol, which was attributed to the combined beneficial effects of the novel microstructure and the special composition of helical carbon fibers. In addition, new submicrometer-scale flat carbon fibers (SFCFs) were successfully synthesized over a NiAl– LDH compound by CCVD of acetylene in the absence of H2.219 During CCVD, the active crystal facets of the NiAl2O4 spinel phase formed favoured deposition of carbon atoms which grew into SFCFs. After supporting Pt on the SFCFs, the resulting Pt/SFCFs electrocatalyst exhibited much higher activity for methanol oxidation than a commercial Pt/carbon black catalyst, owing to the combined beneficial effects of the microstructure of the carbon support, the improved electrical conductivity, and the enhanced dispersion of Pt particles. When a small amount of active metal component is introduced into the LDH layers, metal NPs formed after calcination and reduction can also catalyze the growth of CNTs with fewer walls and smaller diameter, such as single-walled CNTs and double-walled CNTs. Zhao et al.220 demonstrated that the Fe NPs derived from MgAlFe–LDH with a Mg/Al/Fe molar ratio of 2 : 0.9 : 0.1 could catalyze the growth of single-walled CNTs, because the incorporation of Mg into the LDH layers stabilized the active metal NPs. Furthermore, Wei and colleagues221,222 reported the synthesis of single-walled CNT-array double helices by well dispersed and stable active Fe NPs with a small diameter of 1–14 nm, an extremely high density (1014 to 1016 particles per m2) and good thermal stability at 900 1C, which were obtained from calcined MoO42-intercalated MgAlFe– LDHs based on a ‘‘pinning effect’’ of Mo around the Fe NPs (Fig. 15). In comparison, a carbonate-type CoMgAlFe–LDH was demonstrated to be effective for the growth of multi-walled CNT double helices.223 Furthermore, the same research group fabricated nanocomposites of single/double-walled CNTs and MMO flakes using MgAlFe–LDH precursors with different Fe contents,224 where single/double-walled CNTs were interlinked between 2D MMO flakes. When incorporated in a polyimide matrix, the resulting nanocomposite possessed much higher tensile strength and Young’s modulus than the corresponding composite containing pure CNTs. In addition to Fe-containing LDH precursors, Co- and Ni-containing LDHs have been employed for the growth of single-walled CNTs in a fluidized bed reactor.225,226 In particular, high selectivity for metallic single-walled CNTs was achieved over the CoMgAl–LDH catalyst. It is not only transition metal cations located in the layers of LDHs that can give rise to metal particles after calcination and reduction—noble metal complex-intercalated LDHs can also

Fig. 15 Schematic illustration to show the formation of high density Fe NPs by the incorporation of MoO42 in MgAlFe–LDH. Reproduced with permission from ref. 221. Copyright 2010 American Chemical Society.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

afford metal particles with a high degree of dispersion on mixed oxides when they are subjected to similar post-synthesis treatment for the growth of carbon materials.227,228 For example, recently, Li and colleagues demonstrated that noble metal Pd nanocatalysts on MgAl–MMOs derived from PdCl42intercalated MgAl–LDH precursors efficiently catalyzed the growth of uniform bamboo-like CNTs with few defects by CCVD of methane.228 They found that the number of structural defects and the diameter of the CNTs were reduced with increasing metal dispersion. These results suggest that the synthesis of other metal oxide supported noble metal catalysts from LDH precursors will make it possible to prepare nanocomposites of CNTs and noble metals, which should be promising catalysts with many applications in view of the availability of complex geometric shapes of noble metal particles with specific facets exposed and their high degree of dispersion, as well as the high mechanical strength of CNTs as supports. 2.5.2 Supported metal nanocatalysts (a) Alumina as a support. Fabricating supported metal NPs with simultaneously enhanced activity and stability is of vital importance in heterogeneous catalysis and remains a challenging goal. Because of its good stability, high specific surface area, and excellent physical strength, alumina has been extensively used as a catalyst support in a variety of reactions. Recently, Li et al. reported the in situ growth of NiMgAl–LDH microcrystallites on a microspherical g-alumina support through surface activation of the support followed by a homogeneous coprecipitation process.229 They found that as compared with the analogous material prepared by conventional impregnation, Ni/MgO–Al2O3 obtained from NiMgAl–LDH/g-Al2O3 showed higher metal dispersion and stability, due to the formation of well-developed 2D platelet-like LDHs grown perpendicularly on the support surface, as well as the presence of lattice orientation and confinement effect of LDH platelets. With the introduction of Mg into the LDH layers, the ordered arrangement of larger NiMgAl–LDH platelets stabilized active nickel clusters, and the amorphous phases, including MgO and Al2O3, formed around the Ni particles impeded their agglomeration. Consequently, the highly dispersed small Ni NPs efficiently catalyzed the growth of uniform CNTs. In a similar way, Feng and colleagues synthesized Ni and Pd monometallic catalysts supported on spherical g-alumina through in situ growth of active NiAl–LDH and PdCl42/MgAl– LDH precursors.230,231 Because the introduction of the LDHs lowered the acidity of the support and strengthened the metal– support interaction, compared with materials prepared by conventional impregnation methods the resulting Ni-based and Pd-based catalysts exhibited not only higher activity but also better selectivity and stability in the vapour phase hydrodechlorination of chlorobenzene and the selective hydrogenation of acetylene, respectively. In addition, amorphous spherical alumina was utilized to synthesize a surface defect-promoted Ni nanocatalyst with a high dispersion and high particle density embedded on a hierarchical Al2O3 matrix.232 The high dispersion of active Ni species in the LDH precursor induced a uniform size distribution of Ni particles

This journal is © The Royal Society of Chemistry 2014

Review Article

Fig. 16 HRTEM images of Ni/Al2O3: (a) an overall image, (b) one nanoflake with Ni NPs embedded onto its surface (inset, a single-crystalline Ni NP), (c) a cross-sectional image of one nanoflake with Ni NPs embedded on both sides of the substrate (inset, a cross-section of a single-crystalline Ni NP), and (d) scanning transmission electron microscopy (STEM) image of the Ni NPs dispersed on the nanoflakes (inset, a portion of the nanoflake and the size distribution based on more than 200 particles counted). Reproduced with permission from ref. 232. Copyright 2013 American Chemical Society.

(5.0  0.9 nm) and extremely high particle density (B2.4  1016 particles per m2) (Fig. 16), far exceeding the densities of metal NPs reported previously.221 The as-fabricated supported Ni nanocatalyst exhibited excellent low temperature activity for CO2 methanation, owing to an abundance of clusters on surface vacancies which served as active sites. Moreover, the intrinsic topotactic nature of the phase transformation from the LDH matrix to supported Ni NPs ensured a strong anchoring effect, which gave rise to extremely high reaction stability, with no sintering and/or aggregation of active species being observed during long-term use up to 260 hours. (b) Carbon materials as supports. In the field of heterogeneous catalysis, carbon materials have been extensively investigated as catalyst supports to disperse and stabilize metal NPs.233,234 For example, Li and colleagues235 successfully prepared highly-dispersed, carbon-supported nickel catalysts with Ni loadings from 6.8 to 20.2 wt% via in situ self-reduction of NiAl–LDH/carbon nanocomposite precursors with interwoven hybrid structure, which involved the crystallization of Ni/Al– LDH with concomitant carbonization of glucose under mild hydrothermal conditions. Notably, the carbon component in precursors serving as the reducing agent could reduce Ni2+ species to Ni0 in situ upon heating under an inert atmosphere. A Ni nanocatalyst with a Ni loading amount of 18.7 wt% displayed excellent catalytic performance in the liquid phase hydrodechlorination of chlorobenzene, giving a very high

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

conversion of 99.3%. Li and co-workers further prepared a series of highly-dispersed, ZnO-modified supported nickel nanocatalysts (Ni–ZnO/C) via in situ self-reduction of hybrid NiZnAl–LDH/carbon (NiZnAl–LDH/C) nanocomposite precursors.236 They found that ZnO addition significantly modified the catalytic properties in the hydrogenation of citral, giving an improved selectivity for citronellol; the maximum yield of citronellol achieved was B92%. The promotional effect of ZnO addition was mainly attributed to the presence of a ZnO–metal interaction, which was proposed to be responsible for enhanced adsorption of the CQO bond in the citral molecule on the catalyst surface and thus activation of the CQO bond. Since such as-synthesized supported Ni catalysts have the advantages of low cost, excellent chemical stability, and facile separation and recovery using a magnetic field, they may have many potential applications in heterogeneous catalysis. In addition, Li and co-workers have made great efforts to improve the dispersion and activity of metal catalysts by combining the unique properties of one-dimensional CNTs with LDHs.237,238 First, they established a novel approach to synthesize highly dispersed Ni NPs by reduction of NiAl–LDH supported on nitric acid modified multi-walled CNTs in the presence of L-cysteine as a bridging linker.237 Due to the support effect of CNTs and the strong metal–support interaction, a high dispersion of Ni NPs having a small modal diameter of about 5 nm and narrow size distribution in the range 3–7 nm was obtained. The resulting Ni nanocatalysts exhibited excellent catalytic performance in liquid phase hydrogenation, with 97.1% conversion of o-chloronitrobenzene and 98.3% selectivity toward o-chloroaniline in 300 min, a much better performance than that of CNT-supported Ni catalysts prepared by incipient wetness impregnation. Secondly, to obtain dispersion-enhanced Ni-based nanocatalysts, poly(acrylic acid)-functionalized CNTs were also utilized to support NiAl–LDH (Fig. 17),238 and the resulting catalyst exhibited superior catalytic performance in the liquid phase selective hydrogenation of o-chloronitrobenzene to

Fig. 17 Schematic illustration of the synthesis of highly dispersed Ni NPs over poly(acrylic acid) – functionlized CNTs. Reproduced from ref. 238. Copyright 2013 The Royal Society of Chemistry.

Chem. Soc. Rev.

Chem Soc Rev

Fig. 18 SEM (a) and TEM (b) images of Ni/G. Inset in (b) shows a HRTEM image of a single crystalline Ni nanoparticle. Reproduced from ref. 239. Copyright 2014 The Royal Society of Chemistry.

o-chloroaniline when compared with a catalyst prepared by the conventional impregnation method, with the highest yield being 98.1% in 150 min. The superior performance of the new catalyst can be attributed to the higher degree of dispersion of the metal particles, as well as the presence of both the electron-rich support and electron-deficient metal species. In a very recent work, 2D graphene has also been utilized as a catalyst support to synthesize a supported Ni nanocatalyst (Ni/G) via self-reduction of a hybrid NiAl–LDH/graphene precursor.239 It was found that NiAl–LDH nanoplatelets were homogeneously dispersed on both sides of an exfoliated, structurally flexible graphene sheet in the hybrid precursor. Furthermore, the graphene component, serving as a reducing agent, reduced Ni2+ species in situ to Ni0 upon heating under an inert atmosphere, thus facilitating the formation of highly dispersed Ni NPs (Fig. 18). This work provides a simple approach to fabricate a large number of graphene-supported, metal-based heterogeneous catalysts with many potential applications. 2.5.3 Intermetallic nanocatalysts. In recent years, intermetallic compounds (IMCs) have attracted extensive research interest due to their unique geometric/electronic effects, specific compositions and fascinating catalytic properties.240–242 The wide versatility of LDH compositions and architectures makes a large number of element combinations accessible, which renders LDHs as a platform precursor for many IMC systems. Some IMCs, with or without noble metal being present in the LDH layers or in the interlayer space—such as Pt–Cu,243,244 Pt–Ga,245,246 Pt–In,247 Pd–Cu,248 Pd–Ga,249 and Pd–Zn250—have been formed from LDH precursors and shown to exhibit superior catalytic activity to the conventional monometallic catalysts. Chemoselective hydrogenation is of great importance in the chemical industry, and IMCs have attracted extensive interest as efficient catalysts. Recently, several Ni–In IMCs with a tunable particle size and Ni/In ratio were prepared successfully by a reduction process carried out during the topotactic transformation of NiIn–LDH precursors.251 The Ni–In IMCs exhibited excellent catalytic activity and selectivity toward the hydrogenation of various a,b-unsaturated aldehydes (e.g., furfural, 1-phenyltanol, crotonaldehyde, and 2-hexenal). Ni K-edge X-ray-absorption finestructure spectroscopy (XAFS) characterization and periodic density functional theory (DFT) calculations revealed that electron transfer from the In atom to the Ni atom, as well as active-site

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

isolation—which favoured the nucleophilic addition process of a CQO group instead of the electrophilic addition of CQC—could account for the significantly enhanced hydrogenation selectivity compared with that of conventional Ni–In IMCs. By taking advantage of the tunable composition and the topotactic character of the thermal transformation of the LDH precursors, supported Ni–Ga IMCs (Ni3Ga, Ni5Ga3, and NiGa) were also synthesized and exhibited significantly improved catalytic activity and selectivity in the hydrogenation of phenylacetylene to styrene,252 compared with most previously reported Ni catalysts. Subsequently, Feng and colleagues developed an effective synthetic approach for highly dispersed bimetallic Pd–Ga catalysts, which involved in situ growth of MgGaAl–LDH on spherical alumina, followed by the introduction of PdCl42 on the surface of the positively charged LDH layer which afforded an opportunity to realize the uniform dispersion of Pd species.253 Upon thermal reduction of the supported LDH precursor, highly stable and dispersed Pd–Ga IMCs were formed on the support surface. Owing to the high dispersion and synergistic effects, Pd–Ga/ MgO–Al2O3 catalysts possessed comparable activity in the partial hydrogenation of acetylene compared with the monometallic Pd/MgO–Al2O3. In particular, high selectivities (81.7–87.2%) were achieved by Pd–Ga/MgO–Al2O3 with different Pd/Ga ratios. In contrast, the selectivity was only 54.7% by the monometallic Pd/MgO–Al2O3. More significantly, the surface confinement effect of the catalysts suppressed the migration and aggregation of Pd–Ga IMCs, leading to excellent catalytic stability. Recently, Wei and colleagues254 developed a bifunctional NiFe-alloy/MgO catalyst, containing both active sites and solid base sites, via a NiFeMg–LDH precursor route. The material exhibited 100% conversion of hydrazine hydrate and up to 99% selectivity toward H2 generation at room temperature, comparable to the most effective noble metal catalysts (e.g., Rh, Pt). The high catalytic activity and selectivity were attributed to the Ni–Fe electronic interactions, and the strong surface basic sites in the NiFe-alloy/MgO catalyst, respectively. In addition, a platinummodified Ni catalyst derived from a NiAl–LDH precursor was found to exhibit 100% conversion and 98% H2 selectivity in the decomposition of hydrazine hydrate at room temperature;255 the TOF was increased sevenfold compared with a Pt-free catalyst. It was shown that the formation of a Pt–Ni alloy significantly weakens the interaction between adspecies produced (including H2 and NHx) and surface Ni atoms, accounting for the greatly enhanced reaction rate, as well as increased H2 selectivity.

3. LDH-based catalyst supports In addition to their use as catalysts in their own right, and as catalyst precursors, both LDHs and the MMOs formed by their calcination have received much attention as supports for catalysts. In particular—as for supported metal catalysts—in addition to the characteristics of the active metal species including the metal loading, the particle size and the metal surface morphology, the importance of the nature of the support, the metal–support interaction and metal–support synergy (metal–base bifunctional catalysis)

This journal is © The Royal Society of Chemistry 2014

Review Article

have been highlighted in interpreting the reactivity of LDH- and MMO-supported catalysts. 3.1

Supported Au nanocatalysts

Both LDH platelets themselves and MMOs have proved to be effective supports for Au NPs. There are many examples of the catalytic performance of Au NPs being enhanced after being supported in this way, including the aerobic homocoupling of phenylboronic acid with Au NPs/MgAl–MMOs,256 oxidation reactions with Au NPs/LDH,257–259 as well as deoxygenation reactions with Au NPs/LDH.260–263 In these cases, the presence of a synergistic effect between Au NPs and the basic sites of the support surface has been proposed. Control of the dispersion and size of Au NPs, as well as the metal–support interaction, is of vital importance in enhancing the catalytic performance of nanocatalysts.264 Zhang et al.265 found that Au NPs with a narrow size distribution of 2–3 nm were preferentially deposited on the lateral {101% 0} facets of large hexagonal LDH crystals, because of the high density of dangling bonds on these facets resulting in relatively active chemical characteristics, as well as strong interactions between the lateral {101% 0} crystal facets and the Au NPs. These crystalface-selective supported Au NPs were shown to be highly efficient catalysts for the epoxidation of styrene to styrene oxide using tert-butyl hydroperoxide as an oxidant. Recently, highly dispersed Au NPs supported on ZnAl–LDH and ZnCeAl–LDH and the corresponding MMOs were found to induce a surface plasmon resonance absorption in the visible light region at around 550 nm, enhancing photocatalytic water splitting under visible light illumination.266 The catalytic activity for H2 generation was dependent on both the size of the Au NPs and the composition of the LDHs. Under solar illumination, the highest H2 production activity was achieved in the case of Ce-containing LDHs, mainly due to the formation of a larger population of positively charged Au+ and Au3+ species. It has often been found that the nature of the support is a critical parameter affecting the catalytic performance of metal nanocatalysts. Xiang et al.267 recently prepared gold nanocrystals supported on CeO2-containing MMOs derived from LDH precursors by a reduction–deposition approach, and found that the incorporation of cerium into the support significantly enhanced the selectivity towards CQC bond formation (with a hydrocinnamaldehyde yield of ca. 83%) in cinnamaldehyde hydrogenation compared with the supported catalyst with no cerium (with a hydrocinnamaldehyde yield of ca. 42%) with a high conversion of above 91%. It was shown that increasing the amount of Ce3+ in the catalysts led to more oxygen vacancies. The surface electron density of Au was reduced due to the presence of these oxygen vacancies. Consequently, the CQC bond adsorption on Au was promoted by the lower electron density which led to a decrease in the repulsive interactions. Metal–support synergy also has remarkable effects on the activity and selectivity obtained with active metals.268,269 Liu et al.270 combined Au NPs with MgCr–LDH as a support and synthesized very highly efficient heterogeneous Au NPs/LDH catalysts for soluble-base-free green aerobic alcohol oxidation.

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

In the MgCr–LDH supported Au NPs, the Au–support synergy increased with decreasing Au NP size and increasing Cr content.271 The strong Au–support synergy was related to a Cr3+ 2 Cr6+ redox cycle at the Au/support interface, where O2 activation took place accompanied by electron transfer from the LDH to Au NPs. The interfacial Cr6+ species were reduced by surface Au–H hydride and negatively charged Au species. Moreover, increasing the Cr content and decreasing the size of Au NPs enhanced the Au–LDH interactions and, accordingly, increased the rate of O2 and O–H bond cleavage. In another example of the use of LDHs as a support, MAl–LDH (M = Mg, Ni, Co) supported Au nanocluster (Au NCs) catalysts were synthesized using water-soluble glutathione-capped Au nanoclusters as a precursor.272 The ultrafine Au NCs (ca. B1.5 nm in diameter) were well dispersed on the surface of the LDH with a loading of Au below 0.23 wt% by virtue of the strong interactions between Au NCs and MAl–LDH. The resulting Au NCs/MgAl–LDH exhibited much higher catalytic performance with a TOF of 46 500 h1 in the oxidation of 1-phenylethanol in toluene, compared with a catalyst prepared by conventional deposition–precipitation with a TOF value of 2040 h1. The catalyst could be applied in the oxidation of a wide range of alcohols without any basic additive being required, and reused without loss of activity or selectivity. It was found that the presence of transition metal cations in the LDH support further enhanced the alcohol oxidation activities of catalysts such as Au NCs/NiAl–LDH and Au NCs/CoAl–LDH. In particular, Au NCs/NiAl–LDH exhibited the highest activity with a TOF value of 46 500 h1 for the aerobic oxidation of 1-phenylethanol under solvent-free conditions, which was correlated with it having the strongest Au–support synergy among the catalysts studied. 3.2

Supported Pd nanocatalysts

Recently, Lu and colleagues employed a facile process involving an ion-exchange reaction and a subsequent reduction reaction and in situ growth of Pd NPs to synthesize LDH-supported Pd nanocatalysts (Fig. 19).273 The morphology of Pd NPs anchored on the surface of MgAl–LDH flakes could be selectively controlled by manipulating the reduction kinetics of the reaction through the synergistic effects between the oxidative etching of Pd(0) and the controlled release of PdCl42 in the PdCl42–LDH system. In the Suzuki cross-coupling reaction, LDH-supported Pd nanocubes with exposed {100} facets showed much higher catalytic activity compared to the other two supported Pd NP catalysts, namely truncated octahedra with exposed {100} and {111} facets and triangular nanoplates with exposed {111} facets, because the Pd{100}-oriented surface had a lower surface atom density, a higher degree of unsaturation of the crystal plane

Fig. 19 Schematic representation of the synthesis of Pd NP/MgAl–LDH. Reproduced from ref. 273. Copyright 2013 The Royal Society of Chemistry.

Chem. Soc. Rev.

Chem Soc Rev

atoms and a stronger interaction with the LDH support than Pd{111} facets. There was no significant loss of catalytic activity when the LDH-supported Pd nanocube catalyst was reused in four continuous cycles, indicative of its good reusability and high stability. Due to their unique nano/microstructure, specific meso/ macroporous networks, and excellent accessibility for active sites, LDH-based 2D arrays or hierarchical architectures have been developed for a variety of novel applications.274–280 For example, finely dispersed and uniform Pd NPs were anchored to well-ordered vertically aligned CoAl–LDH nanowalls (LDHNWs) by using an in situ spontaneous deposition route.281 The integrated LDH-NWs, which provided a confined and stable microenvironment for the in situ formation of Pd NPs, played the roles of both a hierarchical support and a reductant. Based on the effective exposure of the Pd active sites and the elaborate network architecture, in comparison with a commercial Pd/C catalyst the Pd/LDH-NW heterogeneous material afforded a significantly improved catalytic activity with a forward oxidation current density and a reverse oxidation current of 2.01 and 2.20 mA cm2, respectively, as well as robust durability, in ethanol electrooxidation. The enhancement in electrocatalytic properties was attributed to the synergistic effect between the metal and the support, by means of which the LDH support stabilized the Pd NPs via the formation of a Pd–HO bond, accompanied by an electron transfer from LDH to Pd NPs. 3.3

Supported Pt nanocatalysts

Recently, Xiang and co-workers synthesized MgAl–LDH supported Pt catalysts by a mild and environmentally benign solution chemistry method using sucrose as a reducing agent.282 The reduction of the metal precursor, the deposition of Pt NCs and the crystallization of the support were achieved in a one-step green process with no involvement of any toxic reagents. The size of the Pt NCs could be finely tuned by varying the ratio of surfactant to metal precursor ions. The supported Pt catalysts showed high activity (TOF 1757 h1) for the hydrogenation of cinnamaldehyde in neat water, with high selectivity for cinnamyl alcohol (85%). In this catalyst system, the hydroxyl groups in the MgAl–LDH inhibited the adsorption of the CQC bond by electrostatically repelling the phenyl ring of cinnamaldehyde, without the need for a foreign promoter (e.g. an alkali). Also, the hydrophilicity of LDH had a beneficial effect on the orientation of the hydrophilic CQO moiety on the active sites on the surface of the catalyst. Particularly noteworthy is that the hydrogenation products could be easily separated from water by simple extraction. This study offers a new green strategy for catalyst preparation, reaction and product separation. In addition, Xu et al.283 simultaneously immobilized a photosensitizer (rose bengal) and Pt NPs on MgAl–LDH for photocatalytic H2 evolution, and found that self-quenching of the photosensitizer was suppressed and that the compact arrangement of the photosensitizer and highly dispersed Pt NPs facilitated the electron transfer. As a result, the supported system showed a 13-fold enhancement in total turnover number (TON) when compared with the non-immobilized catalyst.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Chem Soc Rev

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

3.4

Other supported catalysts

LDHs can also be used as supports to immobilize functional organic molecules, metal complexes and other catalytically active species.284–286 For example, a structured catalyst was fabricated by immobilizing cobalt phthalocyanine tetrasulfonate (CoPcS) on a MMO film derived from a MgAl–LDH film grown in situ on an alumina substrate.287 The structured catalyst was composed of MMO nanoflakes whose ab-plane was perpendicular to the substrate, with excellent mechanical strength and high adhesion to the substrate, and exhibited high conversion (85.7%) for the oxidation of mercaptan to disulphide, markedly higher than that of the corresponding powder catalyst (38.7%), as a result of the high exposure of active species, as well as the synergistic effects between the oxidation centers (CoPcS) and the moderate basic sites. Moreover, these eco-friendly structured catalysts showed excellent stability and recyclability for mercaptan sweetening. By taking advantage of the ability of LDHs and LDH-derived materials to act as solid base catalysts or supports, one-pot reactions involving two or more steps in a single reaction vessel for the direct synthesis of structurally complex organic substances, without need the intermediate isolation and purification, have been successfully carried out.288–293 For example, a-alkylation of various nitriles with carbonyl compounds was performed using LDH-supported Pd NPs as a multifunctional catalyst. The alkylated nitriles were formed through an aldol reaction at base sites on the surface of the LDH followed by hydrogenation by molecular hydrogen on the Pd NPs.292 In a recent publication, a series of direct one-step cascade reactions were performed over an acid– base bifunctional core–shell structured MgAl–MMO@Al-containing mesoporous silica (MgAl–MMO/Al-MS) hexagonal noncatalyst.293 MgAl–LDH was chosen as both the base source and as a hard template for the acidic mesoporous silica coating leading to core–shell structured MgAl–MMO@Al-MS hexagonal nanoplates (Fig. 20). In this system, the Al-MS shell was not only capable of protecting the MgAl–LDO core, but also offered a high surface area for the functional acid catalytic sites. Due to the successful

Fig. 20 (a) Synthesis procedure for acid–base bifunctional core–shell structured MgAl–MMO@Al-MS nanoreactors. (b) One-pot cascade reaction sequence conducted in the core–shell structured MgAl–MMO@Al-MS nanoreactor. Reproduced from ref. 293. Copyright 2014 The Royal Society of Chemistry.

This journal is © The Royal Society of Chemistry 2014

Review Article

coexistence and site isolation of the acid and base active sites, and the confinement and enrichment effects of the reaction species provided by the mesoporous structure, this integrated nanocatalyst served as an acid–base bifunctional nanoreactor for the efficient catalysis of one-pot deprotection–Knoevenagel cascade reaction sequences. In the future, such multifunctional heterogeneous catalysts involving solid base LDHs and MMOs will surely act as pivotal tools in the development of economically and environmentally friendly chemical processes.

4. Conclusions LDHs represent one of the most technologically promising materials as a consequence of their low cost, relative ease of preparation, and the large number of composition/preparation variables that may be adopted. LDH-based catalytic materials have been rationally designed and synthesized by well-developed methodologies by taking advantage of the flexible tunability of the metal cations in the layers, and the exchangeability of the intercalated anions in interlayer galleries. In this review, we have attempted to summarize the methods for design and synthesis of LDH materials for specific catalytic applications. For instance, the incorporation of transition metals in LDH precursors allows a wide variety of highly dispersed metal and metal oxide catalysts to be obtained by appropriate post-synthesis treatment—such as calcination and reduction—which exhibit superior catalytic properties to the corresponding materials prepared by the conventional methods such as incipient wetness impregnation. The superiority of the LDH-derived catalysts results from the uniform dispersion of the different metal cations in the layers of the LDH precursor which leads to more efficient dispersion and less agglomeration of the active sites than is generally possible when using conventional methods of preparation. In addition the assembly or integration of LDHs with distinct micro/mesoscopic or hierarchical structures affords multifunctional nanocomposites with improved catalytic performance. At present, even though new and exciting applications for LDH-based materials in the field of heterogeneous catalysis are springing up at an increasing rate, it remains a challenge to achieve a precise tailoring of catalyst structures on multi-scale levels, accurate identification of the active sites and detailed insights into the mechanistic understanding. Therefore, in the future, more detailed and thorough studies should focus on the following areas: (i) The design and synthesis of low-cost and high-quality LDH-based catalysts with careful compositional, structural, and morphological control; important targets are improving costefficiency by replacing rare and noble metals with cheaper and more abundant ones, making processes greener and more environmentally benign, enhancing the ease of recovery and recycling, and developing scalable production protocols so that LDHs can be employed in practice on a large scale. (ii) The application of advanced physicochemical techniques for in situ monitoring of the topotactic transformation of LDH precursors into MMO with concomitant formation of metal catalysts,

Chem. Soc. Rev.

View Article Online

Review Article

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

in order to understand the lattice orientation/lattice confinement effect on the catalytically active species, as well as the interaction between the metal and the support. (iii) The development of new characterization tools and methods to understand the nature of active sites, the synergy between two or more active components and the resulting catalytic mechanisms. We believe that such studies will ensure that the application of LDH-based materials in heterogeneous catalysis will continue to expand rapidly for the foreseeable future.

Acknowledgements We gratefully acknowledge the financial support from the 973 Program (2011CBA00506), the National Natural Science Foundation of China and the Program for Changjiang Scholars and Innovative Research Teams in Universities (IRT1205). We also thank all the colleagues and students who have contributed to our work in this area.

Notes and references ` and A. Vaccari, Catal. Today, 1991, 1 F. Cavani, F. Trifiro 11, 173. 2 P. J. Sideris, U. G. Nielsen, Z. Gan and C. P. Grey, Science, 2008, 321, 113. 3 Q. Wang and D. O’Hare, Chem. Rev., 2012, 112, 4124. 4 D. G. Evans and R. C. T. Slade, Struct. Bonding, 2006, 119, 1. 5 D. G. Evans and X. Duan, Chem. Commun., 2006, 485. 6 F. Li and X. Duan, Struct. Bonding, 2006, 119, 193. 7 S. He, Z. An, M. Wei, D. G. Evans and X. Duan, Chem. Commun., 2013, 49, 5912. 8 D. Tichit and B. Coq, CATTECH, 2003, 7, 206. 9 D. Tichit, D. Lutic, B. Coq, R. Durand and R. Teissier, J. Catal., 2003, 219, 167. ´rardin, R. Durand and B. Coq, Top. Catal., 10 D. Tichit, C. Ge 2006, 39, 89. 11 Z. P. Xu, J. Zhang, M. O. Adebajo, H. Zhang and C. H. Zhou, Appl. Clay Sci., 2011, 53, 139. 12 J. C. A. A. Roelofs, D. J. Lensveld, A. J. van Dillen and K. P. de Jong, J. Catal., 2001, 203, 184. 13 B. M. Choudary, M. L. Kantam, V. Neeraja, K. K. Rao, F. Figueras and L. Delmotte, Green Chem., 2001, 3, 257. ´rez, B. M. de Souza, E. C. Lima, 14 C. O. Veloso, C. N. Pe A. G. Dias, J. L. F. Monteiro and C. A. Henriques, Microporous Mesoporous Mater., 2008, 107, 23. 15 B. M. Choudary, M. L. Kantam, C. R. V. Reddy, K. K. Rao and F. Figueras, J. Mol. Catal. A: Chem., 1999, 146, 279. ´ erveny´, J. ˇ ´, P. Kac ˇer, F. Kovanda, M. Kuzma 16 J. C Splı´chalova ˇ erveny´, J. Mol. Catal. A: Chem., 2008, 285, 150. and L. C 17 S. K. Sharma, P. A. Parikh and R. V. Jasra, J. Mol. Catal. A: Chem., 2007, 278, 135. ´, S. Dhir, G. Colet and J. Pe ´rez-Ramı´rez, Appl. 18 S. Abello Catal., A, 2007, 325, 121.

Chem. Soc. Rev.

Chem Soc Rev

´, F. Medina, D. Tichit, J. Pe ´rez-Ramı´rez, J. E. Sueiras, 19 S. Abello P. Salagre and Y. Cesteros, Appl. Catal., B, 2007, 70, 577. 20 J. S. Valente, H. Pfeiffer, E. Lima, J. Prince and J. Flores, J. Catal., 2011, 279, 196. 21 D. G. Cantrell, L. J. Gillie, A. F. Lee and K. Wilson, Appl. Catal., A, 2005, 287, 183. 22 Y. Xi and R. J. Davis, J. Catal., 2008, 254, 190. 23 Y. Xi and R. J. Davis, J. Catal., 2009, 268, 307. 24 L. Gao, G. Teng, J. Lv and G. Xiao, Energy Fuels, 2010, 24, 646. ´lez, J. Me ´rida25 M. J. Campos-Molina, J. Santamarı´a-Gonza Robles, R. Moreno-Tost, M. C. G. Albuquerque, S. Bruque´mez, E. Rodrı´guez-Castello ´n, A. Jime ´nez-Lo ´pez and Ga P. Maireles-Torres, Energy Fuels, 2010, 24, 979. 26 W. Xie, Y. Liu and H. Chun, Catal. Lett., 2012, 142, 352. 27 O. D. Pavel, B. Cojocaru, E. Angelescu and V. I. Parvulescu, Appl. Catal., A, 2011, 403, 83. 28 X. Lei, W. Lu, Q. Peng, H. Li, T. Chen, S. Xu and F. Zhang, Appl. Catal., A, 2011, 399, 87. ¨, X. Guo and F. Zhang, Ind. Eng. Chem. Res., 29 X. Lei, Z. Lu 2012, 51, 1275. ´lvarez, A. M. Frey, J. H. Bitter, A. M. Segarra, K. P. 30 M. G. A de Jong and F. Medina, Appl. Catal., B, 2013, 134–135, 231. 31 J. J. Woodford, J. P. Dacquin, K. Wilson and A. F. Lee, Energy Environ. Sci., 2012, 5, 6145. 32 P. Guerrero-Urbaneja, C. Garcı´a-Sancho, R. Moreno-Tost, ´rida-Robles, J. Santamarı´a-Gonza ´lez, A. Jime ´nezJ. Me ´pez and P. Maireles-Torres, Appl. Catal., A, 2014, Lo 470, 199. ´nez-Sanchidria ´n and J. R. Ruiz, Appl. Catal., A, 2014, 33 C. Jime 469, 367. ´ndez, 34 A. Navajas, I. Campo, G. Arzamendi, W. Y. Herna L. F. Bobadilla, M. A. Centeno, J. A. Odriozola and L. M. Gandı´a, Appl. Catal., B, 2010, 100, 299. 35 P. Liu, M. Derchi and E. J. M. Hensen, Appl. Catal., A, 2013, 467, 124. 36 P. Liu, M. Derchi and E. J. M. Hensen, Appl. Catal., B, 2014, 144, 135. 37 S. H. Wang, Y. B. Wang, Y. M. Dai and J. M. Jehng, Appl. Catal., A, 2012, 439–440, 135. 38 I. Cota, E. Ramı´rez, F. Medina, J. E. Sueiras, G. Layrac and D. Tichit, Appl. Catal., A, 2010, 382, 272. 39 Q. Ma, T. Zhao, D. Wang, W. Niu, P. Lv and N. Tsubaki, Appl. Catal., A, 2013, 464–465, 142. 40 J. Wang, Z. Lei, H. Qin, L. Zhang and F. Li, Ind. Eng. Chem. Res., 2011, 50, 7120. 41 O. Meyer, F. Roessner, R. A. Rakoczy and R. W. Fischer, ChemCatChem, 2010, 2, 314. ´nez-Sanchidria ´n and J. R. Ruiz, 42 J. M. Hidalgo, C. Jime Appl. Catal., A, 2014, 470, 311. 43 R. Chebout, D. Tichit, G. Layrac, A. Barama, B. Coq, I. Cota, E. R. Rangel and F. Medina, Solid State Sci., 2010, 12, 1013. 44 Z. Yin, J. Wu and Z. Yang, Biosens. Bioelectron., 2011, 26, 1970. 45 Z. Zhu, L. Qu, Y. Guo, Y. Zeng, W. Sun and X. Huang, Sens. Actuators, B, 2010, 151, 146.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

46 Y. Wang, D. Zhang, W. Peng, L. Liu and M. Li, Electrochim. Acta, 2011, 56, 5754. 47 J. Ju, J. Bai, X. Bo and L. Guo, Electrochim. Acta, 2012, 78, 569. 48 X. Zou, A. Goswami and T. Asefa, J. Am. Chem. Soc., 2013, 135, 17242. 49 Y. Zhang, B. Cui, C. Zhao, H. Lin and J. Li, Phys. Chem. Chem. Phys., 2013, 15, 7363. 50 Z. Lu, W. Xu, W. Zhu, Q. Yang, X. Lei, J. Liu, Y. Li, X. Sun and X. Duan, Chem. Commun., 2014, 50, 6479. 51 W. J. Youngblood, S. H. A. Lee, K. Maeda and T. E. Mallouk, Acc. Chem. Res., 2009, 42, 1966. 52 K. Maeda, K. Teramura, D. Lu, T. Takata, N. Saito, Y. Inoue and K. Domen, Nature, 2006, 440, 295. 53 X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891. ´rez, T. Marino, R. Molinari and H. Garcı´a, 54 C. G. Silva, R. Jua J. Am. Chem. Soc., 2011, 133, 595. 55 M. Shao, J. Han, M. Wei, D. G. Evans and X. Duan, Chem. Eng. J., 2011, 168, 519. 56 K. M. Parida and L. Mohapatra, Chem. Eng. J., 2012, 179, 131. 57 Y. Zhao, S. Zhang, B. Li, H. Yan, S. He, L. Tian, W. Shi, J. Ma, M. Wei, D. G. Evans and X. Duan, Chem. – Eur. J., 2011, 17, 13175. 58 K. Parida, L. Mohapatra and N. Baliarsingh, J. Phys. Chem. C, 2012, 116, 22417. 59 L. Tian, Y. Zhao, S. He, M. Wei and X. Duan, Chem. Eng. J., 2012, 184, 261. 60 K. Parida and L. Mohapatra, Dalton Trans., 2012, 41, 1173. 61 K. Parida, L. Mohapatra and N. Baliarsingh, J. Phys. Chem. C, 2012, 116, 22417. 62 N. Baliarsingh, L. Mohapatra and K. Parida, J. Mater. Chem. A, 2013, 1, 4236. 63 Y. Zhao, P. Chen, B. Zhang, D. Su, S. Zhang, L. Tian, J. Lu, Z. Li, X. Cao, B. Wang, M. Wei, D. G. Evans and X. Duan, Chem. – Eur. J., 2012, 18, 11949. 64 R. Nakamura, A. Okamoto, H. Osawa, H. Irie and K. Hashimoto, J. Am. Chem. Soc., 2007, 129, 9596. ´s and H. Garcı´a, J. Am. Chem. 65 C. G. Silva, Y. Bouizi, V. Forne Soc., 2009, 131, 13833. 66 Y. Lee, J. H. Choi, H. J. Jeon, K. M. Choi, J. W. Lee and J. K. Kang, Energy Environ. Sci., 2011, 4, 914. 67 S. J. Kim, Y. Lee, D. K. Lee, J. W. Lee and J. K. Kang, J. Mater. Chem. A, 2014, 2, 4136. 68 J. Pan, G. Liu, G. Q. Lu and H.-M. Cheng, Angew. Chem., Int. Ed., 2011, 50, 2133. 69 N. Wu, J. Wang, D. N. Tafen, H. Wang, J.-G. Zheng, J. P. Lewis, X. Liu, S. S. Leonard and A. Manivannan, J. Am. Chem. Soc., 2010, 132, 6679. 70 Y. Zhao, C. Eley, J. Hu, J. S. Foord, L. Ye, H. He and S. C. E. Tsang, Angew. Chem., Int. Ed., 2012, 51, 3846. 71 Y. Zhao, B. Li, Q. Wang, W. Gao, C. J. Wang, M. Wei, D. G. Evans, X. Duan and D. O’Hare, Chem. Sci., 2014, 5, 951. 72 N. Ahmed, Y. Shibata, T. Taniguchi and Y. Izumi, J. Catal., 2011, 279, 123.

This journal is © The Royal Society of Chemistry 2014

Review Article

73 N. Ahmed, M. Morikawa and Y. Izumi, Catal. Today, 2012, 185, 263. 74 M. Morikawa, N. Ahmed, Y. Yoshida and Y. Izumi, Appl. Catal., B, 2014, 144, 561. 75 K. Teramura, S. Iguchi, Y. Mizuno, T. Shishido and T. Tanaka, Angew. Chem., 2012, 124, 8132. 76 Y. F. Song and R. Tsunashima, Chem. Soc. Rev., 2012, 41, 7384. 77 T. Kwon, G. A. Tsigdions and T. J. Pinnavaia, J. Am. Chem. Soc., 1988, 110, 3653. 78 B. M. Choudary, B. Bharathi, C. V. Reddy, M. L. Kantam and K. V. Raghavan, Chem. Commun., 2001, 1736. 79 A. Severeyns, D. E. De Vos and P. A. Jacobs, Green Chem., 2002, 4, 380. 80 E. A. Gardner, S. K. Yun, T. Kwon and T. J. Pinnavaia, Appl. Clay Sci., 1998, 13, 479. 81 B. F. Sels, D. E. De Vos and P. A. Jacobs, J. Am. Chem. Soc., 2007, 129, 6916. 82 V. Rives and M. A. Ulibarri, Coord. Chem. Rev., 1999, 181, 61. 83 S. Omwoma, W. Chen, R. Tsunashima and Y.-F. Song, Coord. Chem. Rev., 2014, 258–259, 58. 84 D. Carriazo, S. Lima, C. Martı´n, M. Pillinger, A. A. Valente and V. Rives, J. Phys. Chem. Solids, 2007, 68, 1872. 85 A. Udomvech, P. Kongrat, C. Pakawatchai and H. Phetmung, Inorg. Chem. Commun., 2012, 17, 132. 86 Y. Liu, K. Murata, T. Hanaoka, M. Inaba and K. Sakanishi, J. Catal., 2007, 248, 277. 87 P. Liu, H. Wang, Z. Feng, P. Ying and C. Li, J. Catal., 2008, 256, 345. 88 A. L. Maciuca, C. E. Ciocan, E. Dumitriu, F. Fajula and V. Hulea, Catal. Today, 2008, 138, 33. 89 A. L. Maciuca, E. Dumitriu, F. Fajula and V. Hulea, Appl. Catal., A, 2008, 338, 1. 90 J. Das and K. M. Parida, J. Mol. Catal. A: Chem., 2007, 264, 248. 91 P. Liu, C. Wang and C. Li, J. Catal., 2009, 262, 159. 92 S. Zhao, J. Xu, M. Wei and Y.-F. Song, Green Chem., 2011, 13, 384. 93 S. Zhao, L. Liu and Y.-F. Song, Dalton Trans., 2012, 41, 9855. 94 L. Mohapatra, K. Parida and M. Satpathy, J. Phys. Chem. C, 2012, 116, 13063. 95 Z. An, S. Lu, J. He and Y. Wang, Langmuir, 2009, 25, 10704. 96 H. Nakayama, N. Wada and M. Tsuhako, Int. J. Pharm., 2004, 269, 469. 97 Z. An, J. He, S. Lu and L. Yang, AIChE J., 2010, 56, 2677. 98 S. Aisawa, H. Kudo, T. Hoshi, S. Takahashi, H. Hirahara, Y. Umetsu and E. Narita, J. Solid State Chem., 2004, 177, 3987. 99 M. X. Reinholdt and R. J. Kirkpatrick, Chem. Mater., 2006, 18, 2567. 100 S.-P. Li, Colloids Surf., A, 2006, 290, 56. 101 A. Dhakshinamoorthy, A. Sharmila and K. Pitchumani, Chem. – Eur. J., 2010, 16, 1128. 102 Z. An, W. Zhang, H. Shi and J. He, J. Catal., 2006, 241, 319. 103 S. Vijaikumar, A. Dhakshinamoorthy and K. Pitchumani, Appl. Catal., A, 2008, 340, 25.

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

104 T. Subramanian, A. Dhakshinamoorthy and K. Pitchumani, Tetrahedron Lett., 2013, 54, 7167. 105 E. P. Giannelis, D. G. Nocera and T. J. Pinnavaia, Inorg. Chem., 1987, 26, 203. 106 K. Lang, P. Bezdicˇka, J. L. Bourdelande, J. Hernando, ´fun ˇ kova ´, F. Kovanda, P. Kuba ´t, J. Mosinger I. Jirka, E. Ka ´ and D. M. Wagnerova, Chem. Mater., 2007, 19, 3822. 107 S. Bhattacharjee, T. J. Dines and J. A. Anderson, J. Phys. Chem. C, 2008, 112, 14124. 108 S. Bhattacharjee, K.-E. Jeong, S.-Y. Jeong and W.-S. Ahn, New J. Chem., 2010, 34, 156. 109 K. M. Parida, M. Sahoo and S. Singha, J. Catal., 2010, 276, 161. 110 S. Bhattacharjee, T. J. Dines and J. A. Anderson, J. Catal., 2004, 225, 398. 111 S. Singha, M. Sahoo and K. M. Parida, Dalton Trans., 2011, 40, 11838. 112 M. Halma, K. A. D. de Freitas Castro, C. Taviot-Gueho, ´vot, C. Forano, F. Wypych and S. Nakagaki, J. Catal., V. Pre 2008, 257, 233. 113 M. Wei, X. Zhang, D. G. Evans, X. Duan, X. Li and H. Chen, AIChE J., 2007, 53, 2916. ˆrvulescu, C. M. 114 F. Neat-u, L. Protes- escu, M. Florea, V. I. Pa Teodorescu, N. Apostol, P. Y. Toullecc and V. Michelet, Green Chem., 2010, 12, 2145. 115 H. Shi, C. Yu and J. He, J. Phys. Chem. C, 2010, 114, 17819. 116 H. Shi, C. Yu and J. He, J. Catal., 2010, 271, 79. 117 H. Shi and J. He, J. Catal., 2011, 279, 155. 118 Y. Liu, Z. An, L. Zhao, H. Liu and J. He, Ind. Eng. Chem. Res., 2013, 52, 17821. 119 F. Wypych, G. A. Bubniak, M. Halma and S. Nakagaki, J. Colloid Interface Sci., 2003, 264, 203. 120 T. Hibino, Chem. Mater., 2004, 16, 5482. 121 J. Wang, L. Zhao, H. Shi and J. He, Angew. Chem., Int. Ed., 2011, 123, 9337. 122 L.-W. Zhao, H.-M. Shi, J.-Z. Wang and J. He, Chem. – Eur. J., 2012, 18, 9911. 123 L.-W. Zhao, H.-M. Shi, J.-Z. Wang and J. He, Chem. – Eur. J., 2012, 18, 15323. 124 L. Dai, J. Zhang, X. Wang and Y. Chen, RSC Adv., 2013, 3, 19885. 125 J. L. Gunjakar, T. W. Kim, H. N. Kim, I. Y. Kim and S. J. Hwang, J. Am. Chem. Soc., 2011, 133, 14998. 126 J. Han, X. Xu, X. Rao, M. Wei, D. G. Evans and X. Duan, J. Mater. Chem., 2011, 21, 2126. 127 J. Zhao, X. Kong, W. Shi, M. Shao, J. Han, M. Wei, D. G. Evans and X. Duan, J. Mater. Chem., 2011, 21, 13926. 128 J. H. Lee, H. Kim, Y. S. Lee and D.-Y. Jung, ChemCatChem, 2014, 6, 113. 129 G. Carja, A. Nakajima, S. Dranca, C. Dranca and K. Okada, J. Phys. Chem. C, 2010, 114, 14722. 130 R. Lu, X. Xu, J. Chang, Y. Zhu, S. Xu and F. Zhang, Appl. Catal., B, 2012, 111, 389. 131 E. M. Seftel, M. Mertens and P. Cool, Appl. Catal., B, 2013, 134–135, 274. 132 J. S. Valente, F. Tzompantzi and J. Prince, Appl. Catal., B, 2011, 102, 276.

Chem. Soc. Rev.

Chem Soc Rev

133 H. Fan, J. Zhu, J. Sun, S. Zhang and S. Ai, Chem. – Eur. J., 2013, 19, 2523. 134 S. J. Xia, F. X. Liu, Z. M. Ni, W. Shi, J. L. Xue and P. P. Qian, Appl. Catal., B, 2014, 144, 570. 135 M. Shao, F. Ning, M. Wei, D. G. Evans and X. Duan, Adv. Funct. Mater., 2014, 24, 580. 136 E. M. Seftel, M. C. Puscasu, M. Mertens, P. Cool and G. Carja, Appl. Catal., B, 2014, 150–151, 157. 137 H. Wang, X. Xiang and F. Li, J. Mater. Chem., 2010, 20, 3944. 138 M. Gong, Y. Li, H. Wang, Y. Liang, J. Z. Wu, J. Zhou, J. Wang, T. Regier, F. Wei and H. Dai, J. Am. Chem. Soc., 2013, 135, 8452. 139 J. Suntivich, K. J. May, H. A. Gasteiger, J. B. Goodenough and Y. Shao-Horn, Science, 2011, 334, 1383. 140 A. K. Geim, Science, 2009, 324, 1530. 141 J. A. Matthew, C. T. Vincent and B. K. Richard, Chem. Rev., 2010, 110, 132. 142 J. L. Gunjakar, I. Y. Kim, J. M. Lee, N.-S. Lee and S.-J. Hwang, Energy Environ. Sci., 2013, 6, 1008. 143 B. Li, Y. Zhao, S. Zhang, W. Gao and M. Wei, ACS Appl. Mater. Interfaces, 2013, 5, 10233. 144 M. Li, J. E. Zhu, L. Zhang, X. Chen, H. Zhang, F. Zhang, S. Xu and D. G. Evans, Nanoscale, 2011, 3, 4240. ¨nemann and W. R. Thiel, Angew. Chem., 145 S. Shylesh, V. Schu Int. Ed., 2010, 49, 3428. 146 V. Polshettiwar, R. Luque, A. Fihri, H. Zhu, M. Bouhrara and J.-M. Basset, Chem. Rev., 2011, 111, 3036. ¨th, Angew. Chem., 147 A.-H. Lu, E. L. Salabas and F. Schu Int. Ed., 2007, 46, 1222. 148 L. Li, Y. Feng, Y. Li, W. Zhao and J. Shi, Angew. Chem., Int. Ed., 2009, 48, 5888. 149 F. Mi, X. Chen, Y. Ma, S. Yin, F. Yuan and H. Zhang, Chem. Commun., 2011, 47, 12804. 150 H. Zhang, G. Zhang, X. Bi and X. Chen, J. Mater. Chem. A, 2013, 1, 5934. 151 L. Zou, X. Xiang, J. Fan and F. Li, Chem. Mater., 2007, 19, 6518. 152 X. Xiang, H. I. Hima, H. Wang and F. Li, Chem. Mater., 2008, 20, 1173. 153 L. Zou, F. Li, X. Xiang, D. G. Evans and X. Duan, Chem. Mater., 2006, 18, 5852. 154 C. Nan, G. Fan, J. Fan and F. Li, Mater. Lett., 2013, 106, 5. 155 X. Zhao, L. Wang, X. Xu, X. Lei, S. Xu and F. Zhang, AIChE J., 2012, 58, 573. 156 Y. Zhao, M. Wei, J. Lu, Z. L. Wang and X. Duan, ACS Nano, 2009, 3, 4009. 157 S. He, S. Zhang, J. Lu, Y. Zhao, J. Ma, M. Wei, D. G. Evans and X. Duan, Chem. Commun., 2011, 47, 10797. 158 G. Fan, W. Sun, H. Wang and F. Li, Chem. Eng. J., 2011, 174, 467. 159 X. Shu, J. He and D. Chen, J. Phys. Chem. C, 2008, 112, 4151. 160 X. Shu, J. He and D. Chen, Ind. Eng. Chem. Res., 2008, 47, 4750. 161 J. J. Yu, J. Cheng, C. Y. Ma, H. L. Wang, L. D. Li, Z. P. Hao and Z. P. Xu, J. Colloid Interface Sci., 2009, 333, 423.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

´, K. Jira ´tova ´, F. Kovanda, M. Vala ´ˇskova ´, J. Balaba ´nova ´ 162 L. Obalova ´, J. Mol. Catal. A: Chem., 2006, 248, 210. and K. Pacultova ´, K. Pacultova ´, J. Balaba ´nova ´, K. Jira ´tova ´, 163 L. Obalova ´ ˇ ´ Z. Bastl, M. Valaskova, Z. Lacny´ and F. Kovanda, Catal. Today, 2007, 119, 233. ´, K. Kara ´skova ´, K. Jira ´tova ´ and F. Kovanda, Appl. 164 L. Obalova Catal., B, 2009, 90, 132. 165 P. A. Kumar, M. P. Reddy, L. K. Ju, B. Hyun-Sook and H. H. Phil, J. Mol. Catal. A: Chem., 2008, 291, 66. ¨hler, 166 C. He, M. Paulus, W. Chu, J. Find, J. A. Nickl and K. Ko Catal. Today, 2008, 131, 305. 167 J. J. Yu, X. P. Wang, L. D. Li, Z. P. Hao, Z. P. Xu and G. Q. Lu, Adv. Funct. Mater., 2007, 17, 3598. 168 J. J. Yu, X. P. Wang, Y. X. Tao and Z. P. Hao, Ind. Eng. Chem. Res., 2007, 46, 5794. ´skova ´, L. Obalova ´, K. Jira ´tova ´ and F. Kovanda, 169 K. Kara Chem. Eng. J., 2010, 160, 480. ´, K. Kara ´skova ´, A. Wach, P. Kustrowski, 170 L. Obalova ´-Kutla ´kova ´, S. Michalik and K. Jira ´tova ´, Appl. K. Mamulova Catal., A, 2013, 462–463, 227. 171 A. Corma, A. E. Palomares and F. Rey, Appl. Catal., B, 1994, 4, 29. ´rquez, J. Catal., 172 A. Corma, A. E. Palomares, F. Rey and F. Ma 1997, 170, 140. ´pez-Nieto, F. J. La ´zaro, A. Lo ´pez 173 A. E. Palomares, J. M. Lo and A. Corma, Appl. Catal., B, 1999, 20, 257. 174 G. Centi and S. Perathoner, Appl. Catal., B, 2007, 70, 172. 175 T. Kameda, A. Kodama, Y. Fubasami, S. Kumagai and T. Yoshioka, Chemosphere, 2012, 88, 250. 176 H.-T. Kang, C.-Y. Zhang, K. Lv and S.-L. Yuan, Ceram. Int., 2014, 40, 5357. ´pez-Salinas, J. S. Valente and R. Montiel, ´, E. Lo 177 M. Cantu Environ. Sci. Technol., 2005, 39, 9715. 178 J. S. Valente and R. Quintana-Solorzano, Energy Environ. Sci., 2011, 4, 4096. 179 C. M. S. Polato, C. A. Henriques, A. C. C. Rodrigues and J. L. F. Monteiro, Catal. Today, 2008, 133–135, 534. 180 C. M. S. Polato, A. C. C. Rodrigues, J. L. F. Monteiro and C. A. Henriques, Ind. Eng. Chem. Res., 2010, 49, 1252. 181 H. B. Pereira, C. M. S. Polato, J. L. F. Monteiro and C. A. Henriques, Catal. Today, 2010, 149, 309. 182 C. Gennequin, S. Siffert, R. Cousin and A. Aboukaı¨s, Top. Catal., 2009, 52, 482. 183 E. Genty, R. Cousin, S. Capelle, C. Gennequin and S. Siffert, Eur. J. Inorg. Chem., 2012, 2802. ´squez, A. Echavarrı´a, A. Faro, F. R. 184 L. A. Palacio, J. Vela Ribeiro and M. F. Ribeiro, J. Hazard. Mater., 2010, 177, 407. 185 D. A. Aguilera, A. Perez, R. Molina and S. Moreno, Appl. Catal., B, 2011, 104, 144. ˇ, D. Tichit, I. Sa ˇndulescu, 186 S. Tanasoi, N. Tanchoux, A. Urda F. Fajula and I.-C. Marcu, Appl. Catal., A, 2009, 363, 135. 187 Z. Jiang, J. Yu, J. Cheng, T. Xiao, M. O. Jones, Z. Hao and P. P. Edwards, Fuel Process. Technol., 2010, 91, 97. 188 L. Chmielarz, Z. Piwowarska, M. Rutkowska, M. Wojciechowska, B. Dudek, S. Witkowski and M. Michalik, Catal. Commun., 2012, 17, 118.

This journal is © The Royal Society of Chemistry 2014

Review Article

´tova ´, J. Ludvı´kova ´ and H. Raabova ´, Appl. 189 F. Kovanda, K. Jira Catal., A, 2013, 464–465, 181. 190 S. Tanasoi, G. Mitran, N. Tanchoux, T. Cacciaguerra, ˘ndulescu, D. Tichit and I.-C. Marcu, Appl. F. Fajula, I. Sa Catal., A, 2011, 395, 78. ´tova ´, Z. Mikulova ´, J. Klempa, T. Grygar, Z. Bastl and 191 K. Jira F. Kovanda, Appl. Catal., A, 2009, 361, 106. 192 P. Li, C. He, J. Cheng, C. Y. Ma, B. J. Dou and Z. P. Hao, Appl. Catal., B, 2011, 101, 570. ˘, I. Popescu, T. Cacciaguerra, N. Tanchoux, 193 A. Urda D. Tichit and I.-C. Marcu, Appl. Catal., A, 2013, 464–465, 20. 194 P. J. Sideris, F. Blanc, Z. Gan and C. P. Grey, Chem. Mater., 2012, 24, 2449. ´rardin, M. Deschamps, 195 S. Cadars, G. Layrac, C. Ge J. R. Yates, D. Tichit and D. Massiot, Chem. Mater., 2011, 23, 2821. 196 X. Zhao, F. Zhang, S. Xu, D. G. Evans and X. Duan, Chem. Mater., 2010, 22, 3933. 197 K. Takehira, T. Shishido, P. Wang, T. Kosaka and K. Takaki, J. Catal., 2004, 221, 43. 198 K. Takehira, T. Kawabata, T. Shishido, K. Murakami, T. Ohi, D. Shoro, M. Honda and K. Takaki, J. Catal., 2005, 231, 92. ´pez, V. La Parola, M. A. Pen ˜a and J. L. G. Fierro, 199 R. Guil-Lo Catal. Today, 2006, 116, 289. 200 Y. Tang, Y. Liu, P. Zhu, Q. S. Xue, L. Chen and Y. Lu, AIChE J., 2009, 55, 1217. ´ndez, J. Walmsley, 201 L. He, H. Berntsen, E. Ochoa-Ferna E. Blekkan and D. Chen, Top. Catal., 2009, 52, 206. 202 G. Busca, U. Costantino, T. Montanari, G. Ramis, C. Resini and M. Sisani, Int. J. Hydrogen Energy, 2010, 35, 5356. 203 C. Resini, T. Montanari, L. Barattini, G. Ramis, G. Busca, S. Presto, P. Riani, R. Marazza, M. Sisani, F. Marmottini and U. Costantino, Appl. Catal., A, 2009, 355, 83. 204 M. Li, X. Wang, S. Li, S. Wang and X. Ma, Int. J. Hydrogen Energy, 2010, 35, 6699. 205 G. Wu, C. Zhang, S. Li, Z. Huang, S. Yan, S. Wang, X. Ma and J. Gong, Energy Environ. Sci., 2012, 5, 8942. 206 S. Nishimura, T. Shishido, K. Ebitani, K. Teramura and T. Tanaka, Appl. Catal., A, 2010, 387, 185. 207 S. Nishimura, T. Shishido, J. Ohyama, K. Teramura, A. Takagaki, T. Tanaka and K. Ebitani, Catal. Sci. Technol., 2012, 2, 1685. ´, E. Bolshak and D. Montane ´, Appl. Catal., A, 2013, 208 S. Abello 450, 261. ´dio, J. D. A. Bellido and E. M. Assaf, Appl. Catal., 209 A. F. Lucre A, 2010, 388, 77. 210 A. J. Vizcaı´no, M. Lindo, A. Carrero and J. A. Calles, Int. J. Hydrogen Energy, 2012, 37, 1985. 211 C. Li, J. Zhou, W. Gao, J. Zhao, J. Liu, Y. Zhao, M. Wei, D. G. Evans and X. Duan, J. Mater. Chem. A, 2013, 1, 5370. 212 L. He, Y. Huang, A. Wang, X. Wang, X. Chen, J. J. Delgado and T. Zhang, Angew. Chem., Int. Ed., 2012, 51, 6191. 213 S. Xia, L. Zheng, L. Wang, P. Chen and Z. Hou, RSC Adv., 2013, 3, 16569. 214 S. Zhang, G. Fan and F. Li, Green Chem., 2013, 15, 2389.

Chem. Soc. Rev.

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Review Article

215 F. Li, Q. Tan, D. G. Evans and X. Duan, Catal. Lett., 2005, 99, 151. 216 H. I. Hima, X. Xiang, L. Zhang and F. Li, J. Mater. Chem., 2008, 18, 1245. 217 L. Zhang, F. Li, X. Xiang, M. Wei and D. G. Evans, Chem. Eng. J., 2009, 155, 474. 218 L. Zhang and F. Li, Electrochim. Acta, 2010, 55, 6695. 219 L. Zhang, C. Zhang, X. Xiang and F. Li, Chem. Eng. Technol., 2010, 33, 44. 220 Y. Zhao, Q. Jiao, C. Li and J. Liang, Carbon, 2007, 45, 2159. 221 M. Q. Zhao, Q. Zhang, W. Zhang, J. Q. Huang, Y. Zhang, D. S. Su and F. Wei, J. Am. Chem. Soc., 2010, 132, 14739. 222 M. Q. Zhao, J. Q. Huang, Q. Zhang, J. Q. Nie and F. Wei, Carbon, 2011, 49, 2148. 223 M. Q. Zhao, Q. Zhang, G. L. Tian, J. Q. Huang and F. Wei, ACS Nano, 2012, 6, 4520. 224 M. Q. Zhao, Q. Zhang, X. L. Jia, J. Q. Huang, Y. H. Zhang and F. Wei, Adv. Funct. Mater., 2010, 20, 677. 225 M. Q. Zhao, Q. Zhang, J. Q. Huang, J. Q. Nie and F. Wei, Carbon, 2010, 48, 3260. 226 M. Q. Zhao, Q. Zhang, J. Q. Huang and F. Wei, Adv. Funct. Mater., 2012, 22, 675. 227 Q. Chen, J. Wang and F. Li, Ind. Eng. Chem. Res., 2011, 50, 9034. 228 T. Sun, G. Fan and F. Li, Ind. Eng. Chem. Res., 2013, 52, 5538. 229 M. Li, G. Fan, H. Qin and F. Li, Ind. Eng. Chem. Res., 2012, 51, 11892. 230 J. T. Feng, Y. J. Lin, D. G. Evans, X. Duan and D. Q. Li, J. Catal., 2009, 266, 351. 231 X. Y. Ma, Y. Y. Chai, D. G. Evans, D. Q. Li and J. T. Feng, J. Phys. Chem. C, 2011, 115, 8693. 232 S. He, C. Li, H. Chen, D. Su, B. Zhang, X. Cao, B. Wang, M. Wei, D. G. Evans and X. Duan, Chem. Mater., 2013, 25, 1040. 233 S. H. Joo, S. J. Choi, I. Oh, J. Kwak, Z. Liu, O. Terasaki and R. Ryoo, Nature, 2001, 412, 169. 234 J. Z. Zhang, Y. T. Tsai, K. L. Sangkaewwattana and J. G. Goodwin Jr., J. Catal., 2011, 280, 89. 235 J. Wang, G. Fan and F. Li, RSC Adv., 2012, 2, 9976. 236 L. Yang, Z. Jiang, G. Fan and F. Li, Catal. Sci. Technol., 2014, 4, 1123. 237 J. Wang, G. Fan, H. Wang and F. Li, Ind. Eng. Chem. Res., 2011, 50, 13717. 238 J. Wang, G. Fan and F. Li, Catal. Sci. Technol., 2013, 3, 982. 239 R. Xie, G. Fan, Q. Ma, L. Yang and F. Li, J. Mater. Chem. A, 2014, 2, 7880. ¨ster, K. Kovnir, M. Friedrich, D. Teschner, 240 M. Armbru G. Wowsnick, M. Hahne, P. Gille, L. Szentmiklosi, M. Feuerbacher, M. Heggen, F. Girgsdies, D. Dosenthal, ¨gl and Yu. Grin, Nat. Mater., 2012, 11, 690. R. Schlo 241 S. E. Collins, J. J. Delgado, C. Mira, J. J. Calvino, S. Bernal, ´s and A. L. Bonivardi, D. L. Chiavassa, M. A. Baltana J. Catal., 2012, 292, 90. 242 H. Abe, F. Matsumoto, L. R. Alden, S. C. Warren, H. D. ˇa and F. J. Disalvo, J. Am. Chem. Soc., 2008, 130, 5452. Abrun

Chem. Soc. Rev.

Chem Soc Rev

´bal, S. Contreras, N. Barrabe ´s, J. LIorca, D. Tichit 243 A. Aristiza and F. Medina, Appl. Catal., B, 2011, 110, 58. ´s, A. Frare, K. Fo ¨ttinger, A. Urakawa, J. Llorca, 244 N. Barrabe G. Rupprechter and D. Tichit, Appl. Clay Sci., 2012, 69, 1. 245 P. Sun, G. Siddiqi, M. Chi and A. T. Bell, J. Catal., 2010, 274, 192. 246 G. Siddiqi, P. Sun, V. Galvita and A. T. Bell, J. Catal., 2010, 274, 200. 247 P. Sun, G. Siddiqi, W. C. Vining, M. Chi and A. T. Bell, J. Catal., 2011, 282, 165. 248 S. Xia, Z. Yuan, L. Wang, P. Chen and Z. Hou, Appl. Catal., A, 2011, 403, 173. ¨ster, M. Behrens, D. Rosenthal, 249 A. Ota, M. Armbru M. Friedrich, I. Kasatkin, F. Girgsdies, W. Zhang, ¨l, J. Phys. Chem. C, 2011, 115, 1368. R. Wagner and R. Schlo 250 A. Ota, E. L. Kunkes, I. Kasatkin, E. Groppo, D. Ferri, B. Poceiro, R. M. N. Yerga and M. Behrens, J. Catal., 2012, 293, 27. 251 C. Li, Y. Chen, S. Zhang, S. Xu, J. Zhou, F. Wang, M. Wei, D. G. Evans and X. Duan, Chem. Mater., 2013, 25, 3888. 252 C. Li, Y. Chen, S. Zhang, J. Zhou, F. Wang, S. He, M. Wei, D. G. Evans and X. Duan, ChemCatChem, 2014, 6, 824. 253 Y. He, L. Liang, Y. Liu, J. Feng, C. Ma and D. Li, J. Catal., 2014, 309, 166. 254 W. Gao, C. Li, H. Chen, M. Wu, S. He, M. Wei, D. G. Evans and X. Duan, Green Chem., 2014, 16, 1560. 255 L. He, Y. Huang, A. Wang, Y. Liu, X. Liu, X. Chen, J. J. Delgado, X. Wang and T. Zhang, J. Catal., 2013, 298, 1. 256 L. Wang, H. Wang, W. Zhang, J. Zhang, J. P. Lewis, X. Meng and F. S. Xiao, J. Catal., 2013, 298, 186. 257 T. Mitsudome, A. Noujima, T. Mizugaki, K. Jitsukawa and K. Kaneda, Green Chem., 2009, 11, 793. 258 T. Mitsudome, A. Noujima, T. Mizugaki, K. Jitsukawa and K. Kaneda, Adv. Synth. Catal., 2009, 351, 1890. 259 S. Nishimura, A. Takagaki and K. Ebitani, Green Chem., 2013, 15, 2026. 260 T. Mitsudome, A. Noujima, Y. Mikami, T. Mizugaki, K. Jitsukawa and K. Kaneda, Angew. Chem., Int. Ed., 2010, 49, 5545. 261 A. Noujima, T. Mitsudome, T. Mizugaki, K. Jitsukawa and K. Kaneda, Angew. Chem., Int. Ed., 2011, 50, 2986. 262 W. Fang, Q. Zhang, J. Chen, W. Deng and Y. Wang, Chem. Commun., 2010, 46, 1547. 263 W. Fang, J. Chen, Q. Zhang, W. Deng and Y. Wang, Chem. – Eur. J., 2011, 17, 1247. 264 J. Huang, T. Akita, J. Faye, T. Fujitani, T. Takei and M. Haruta, Angew. Chem., Int. Ed., 2009, 48, 7862. 265 F. Zhang, X. Zhao, C. Feng, B. Li, T. Chen, W. Lu, X. Lei and S. Xu, ACS Catal., 2011, 1, 232. 266 G. Carja, M. Birsanu, K. Okada and H. Garcia, J. Mater. Chem. A, 2013, 1, 9092. 267 Z. Tian, X. Xiang, L. Xie and F. Li, Ind. Eng. Chem. Res., 2013, 52, 288. 268 B. Q. Xu, X. Zhang and H. Shi, Angew. Chem., Int. Ed., 2005, 44, 7132. 269 J. Lenz, B. C. Campo, M. Alvarez and M. A. Volpe, J. Catal., 2009, 267, 50.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 08 July 2014. Downloaded by University of Windsor on 17/07/2014 10:58:40.

Chem Soc Rev

270 P. Liu, Y. Guan, R. A. van Santen, C. Li and E. J. M. Hensen, Chem. Commun., 2011, 47, 11540. 271 P. Liu, V. Degirmenci and E. J. M. Hensen, J. Catal., 2014, 313, 80. 272 L. Li, L. Dou and H. Zhang, Nanoscale, 2014, 6, 3753. 273 Q. Zhang, J. Xu, D. Yan, S. Li, J. Lu, X. Cao and B. Wang, Catal. Sci. Technol., 2013, 3, 2016. 274 K. Mandel, A. D. Tuhtan, F. Hutter, C. Gellermann, H. Steinmetz and G. Sextl, J. Mater. Chem. A, 2013, 1, 1840. 275 B. M. Choudary, S. Madhi, N. S. Chowdari, M. L. Kantam and B. Sreedhar, J. Am. Chem. Soc., 2002, 124, 14127. 276 J.-T. Feng, X.-Y. Ma, D. G. Evans and D.-Q. Li, Ind. Eng. Chem. Res., 2011, 50, 1947. 277 Z. L. Hsieh, M. C. Lin and J. Y. Uan, J. Mater. Chem., 2011, 21, 1880. 278 J. Liu, Y. Li, X. Huang, G. Li and Z. Li, Adv. Funct. Mater., 2008, 18, 1448. 279 M. Shao, F. Ning, J. Zhao, M. Wei, D. G. Evans and X. Duan, J. Am. Chem. Soc., 2012, 134, 1071. 280 J. Yuan, C. Yu, X. Fan, Z. Ling, J. Qiu and Y. Gogotsi, J. Mater. Chem. A, 2013, 1, 1963. 281 J. Zhao, M. Shao, D. Yan, S. Zhang, Z. Lu, Z. Li, X. Cao, B. Wang, M. Wei, D. G. Evans and X. Duan, J. Mater. Chem. A, 2013, 1, 5840.

This journal is © The Royal Society of Chemistry 2014

Review Article

282 X. Xiang, W. He, L. Xie and F. Li, Catal. Sci. Technol., 2013, 3, 2819. 283 J. Hong, Y. Wang, J. Pan, Z. Zhong and R. Xu, Nanoscale, 2011, 3, 4655. 284 K. A. D. F. Castro, A. Bail, P. B. Groszewicz, G. S. Machado, W. H. Schreiner, F. Wypych and S. Nakagaki, Appl. Catal., A, 2010, 386, 51. 285 Z. Y. Sun, L. Jin, Y. F. Zhao, S. He, S. D. Li, M. Wei and L. R. Wang, Ind. Eng. Chem. Res., 2014, 53, 4595. 286 M. L. Zhang, Q. F. Yao, W. J. Guan, C. Lu and J.-M. Lin, J. Phys. Chem. C, 2014, 118, 10441. 287 Z. Sun, L. Jin, S. He, Y. Zhao, M. Wei, D. G. Evans and X. Duan, Green Chem., 2012, 14, 1909. 288 M. Ohara, A. Takagaki, S. Nishimura and K. Ebitani, Appl. Catal., A, 2010, 383, 149. 289 K. Motokura, D. Nishimura, K. Mori, T. Mizugaki, K. Ebitani and K. Kaneda, J. Am. Chem. Soc., 2004, 126, 5662. 290 K. Motokura, N. Fujita, K. Mori, T. Mizugaki, K. Ebitani and K. Kaneda, J. Am. Chem. Soc., 2005, 127, 9674. 291 K. Motokura, N. Fujita, K. Mori, T. Mizugaki, K. Ebitani, K. Jitukawa and K. Kaneda, Chem. – Eur. J., 2006, 12, 8228. 292 K. Motokura, N. Fujita, K. Mori, T. Mizugaki, K. Ebitani and K. Kaneda, Tetrahedron Lett., 2005, 46, 5507. 293 P. Li, Y. Yu, P. P. Huang, H. Liu, C. Y. Cao and W. G. Song, J. Mater. Chem. A, 2014, 2, 339.

Chem. Soc. Rev.

Catalytic applications of layered double hydroxides: recent advances and perspectives.

This review surveys recent advances in the applications of layered double hydroxides (LDHs) in heterogeneous catalysis. By virtue of the flexible tuna...
6MB Sizes 1 Downloads 5 Views