Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Accepted Manuscript

Title: Recent Advances on Polyoxometalates Intercalated


Layered Double Hydroxides: From Synthetic Approaches to
Functional Material Applications

Author: Solomon Omwoma Wei Chen Ryo Tsunashima


Yu-Fei Song

PII: S0010-8545(13)00194-X
DOI: http://dx.doi.org/doi:10.1016/j.ccr.2013.08.039
Reference: CCR 111769

To appear in: Coordination Chemistry Reviews

Received date: 22-5-2013


Revised date: 28-8-2013
Accepted date: 29-8-2013

Please cite this article as: S. Omwoma, W. Chen, R. Tsunashima, Y.-F. Song,
Recent Advances on Polyoxometalates Intercalated Layered Double Hydroxides: From
Synthetic Approaches to Functional Material Applications, Coordination Chemistry
Reviews (2013), http://dx.doi.org/10.1016/j.ccr.2013.08.039

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Edited Aug 29

Recent Advances on Polyoxometalates Intercalated Layered Double

Hydroxides: From Synthetic Approaches to Functional Material

Applications

t
Solomon Omwoma,1 Wei Chen,1 Ryo Tsunashima,2 Yu-Fei Song1*

ip
1
State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical

cr
Technology, Beijing 100029, P. R. China
2
Graduate School of Science and Engineering, Yamaguchi University, Yamaguchi, 753 8512,
Japan.

us
Corresponding Author: Yu-Fei Song; Email: songyufei@hotmail.com or
songyf@mail.buct.edu.cn; Fax/Tel: +86-10-64431832.

an
M
d
p te
ce
Ac

1
Page 1 of 28
Edited Aug 29

Contents

1. Introduction..............................................................................................................................4
2. Synthesis of the POM/LDH nanocomposites...........................................................................4
2.1. Polyoxometalates (POMs) ...............................................................................................4
2.2. Layered double hydroxides (LDHs) ................................................................................5
2.3. The POM/LDH nanocomposites......................................................................................6

t
2.3.1. Ion exchange pathway...........................................................................................7

ip
2.3.2. Reconstitution synthetic pathway .........................................................................9
2.3.3. Co-precipitation synthetic pathway ....................................................................10

cr
2.3.4. Electrochemical reduction ..................................................................................10
2.3.5. Ultrasound treatment...........................................................................................11
2.3.6. The delamination technology..............................................................................11

us
3. Material Applications of the POM/LDH nanocomposites.....................................................14
3.1. Catalytic Oxidation reactions using the POM/LDH nanocomposites............................14
3.2. Dehydrogenation reactions using the POM/LDHs as catalysts .....................................18

an
3.3. Esterification reactions using the POM/LDH nanocomposites .....................................18
3.4. Oximation of aromatic aldehydes to aldoximes and ketoximes.....................................19
3.5. Photo-luminescent materials..........................................................................................20
M
3.6. Dyes removal using the POM/LDH nanocomposites as adsorbent aterials...................23
4. Conclusions............................................................................................................................23
Acknowledgements .............................................................................................................24
References ...........................................................................................................................24
d
p te
ce
Ac

2
Page 2 of 28
Edited Aug 29

Abstract: Polyoxometalates (POMs) exhibit attractive properties and great potential


to meet with contemporary society demands regarding environment, materials, energy,
health and information technologies etc. The development of POM-based advanced
functional materials is of significance to effectively utilize POMs in meeting various
challenges. In recent years, the intercalation of POM anions into layered double

t
ip
hydroxides (LDHs) has been a versatile and important approach for the development
of POM-based multifunctional materials. The current review summarizes the latest

cr
progress on the preparation of POM intercalated LDHs (denoted as POM/LDHs) and
their material application.

us
Keywords: Polyoxometalate, Layered double hydroxide, Functional materials

an
M
d
p te
ce
Ac

3
Page 3 of 28
Edited Aug 29

1. Introduction

Polyoxometalates (POMs) are a class of discrete anionic metal oxides of groups 5 and
6 and POMs exhibit attractive properties such as thermal and oxidative stability,
remarkable electronic and magnetic properties, and Brønsted acidity etc, which result
in intriguing applications ranging from medicine, catalysis to material science. We
have witnessed significant progress during past decades that have been
summarized in different books, reviews and thematic journal issues [1-4], including

t
books by Pope in 1984 and Müller in 2004 [1], a review by Pope and Müller in 1991

ip
[2], the thematic POM issues in Chemical Reviews and in Chemical Society Reviews
[3], and several contributions from Coordination Chemistry Reviews [4].

cr
POMs are most well-known as catalysts and have been widely applied in industry [5].
The development of highly efficient, easily recovered and reusable POM-based
heterogeneous catalysts has long been and will continue to be the focus for the

us
practical application of POMs. As a result, two strategies [6], namely “solidification”
and “immobilization” of catalytically active POMs, have been adopted in the
development of POM-based molecular heterogeneous catalysts. One way for the

an
immobilization of POMs is the intercalation of POMs into LDH layers, which results
in the formation of a number of interesting nanocomposites with unique properties.

The first scientific publication of the POM/LDH nanocomposites was reported by


M
Pinnavaia in 1988 [7]. It should be mentioned that an interesting POM/LDH
composite material was reported to be able to apply in exhaust gas and hydrocarbon
conversion in a US patent [8] in 1984. The POM/LDH nanocomposites are a class of
very important and versatile materials. Previously, Wang et al. (1997) [10], Pinnavaia
d

et al. (1998) [11], Rives & Ulibarri (1999) [12], Hu and Li (2004) [6], and Rives et al
(2010) [6] reviewed this type of materials from different perspective, and most of
te

them mainly focused on the introduction of the preparation methods, characterization


and some application in catalytic reactions. In this review, after brief introduction of
the classical synthetic approaches (ion exchange, reconstitution, and co-precipitation)
p

of the POM/LDH nanocomposites, we summarize some other synthetic pathways that


have been employed in engineering the POM/LDH composite materials in recent
ce

years. The important applications of the POM/LDH nanocomposites in material


science have been highlighted and discussed.

2. Synthesis of the POM/LDH nanocomposites


Ac

2.1. Polyoxometalates (POMs)

POMs are built from condensation of metal oxide polyhedra (MOx, M = WVI, MoVI,
VV, NbV, TaV etc., and x = 4 – 7) with each other through corner-, edge-, or rarely in a
face-sharing manner [5, 13]. The metal atoms are referred to as addenda atoms. The
atoms that can function as addenda are those that can change their coordination with
oxygen from 4 to 6 as the MOx polyhedra condense in solution upon acidification
[14]. Although oxygen is the ligand that mostly coordinates with the addenda atoms,
other atoms/groups such as sulphur, bromine, nitrosyl and alkoxy, have been
substituted in some of the POM clusters reported [15,16].

4
Page 4 of 28
Edited Aug 29

t
ip
cr
Fig 1. Classical POM structures in polyhedral representations.

When the POM framework exclusively contains addenda metals (from groups 5

us
and/or 6) and oxygen, the cluster is called isopolymetalate, and the Lindqvist type
anion [M6O19]2- is an archetypical example (Fig. 1). When the POM shows additional
elements besides addenda metals and oxygen, it is known as a heteropoly complex,
which is formed by condensation of MOx polyhedra around a central heteroatom as

an
the solution is acidified [14]. Many different elements can act as heteroatoms in the
heteropoly complex with various coordination numbers: 4-coordinate (tetrahedral) in
Keggin and Wells-Dawson structures (e. g., PO43-, SiO44-, AsO43-); 6-coordinate
(octahedral) in Anderson-Evans structure (e. g., Al(OH)63-, TeO66-); 12-coordinate
M
(Silverton) in [(UO12)Mo12O30]8−. See Figure 1 for some common POM structures in
polyhedra representations.
d

2.2. Layered double hydroxides (LDHs)


p te
ce
Ac

Fig 2. Schematic representation of the V10O28/LDHs structure (H2O molecules in the interlayer:
spacefill style) [12].

5
Page 5 of 28
Edited Aug 29

LDHs comprise an unusual class of layered materials with positively charged


nanosheets/layers and the balancing anions located in the interlayer region. LDHs
consist of a broad range of compositions of the type:
2+ 3+ x+ n- 2+ 3+
[M 1-xM x(OH)2] (A x/n)·zH2O; where M and M metal ions occupy octahedral
positions in the hydroxide layers, An- is the gallery anion, x = 0.17–0.33 and it is
defined as the M3+/(M2++M3+) ratio [17]. The interlayer anions are linked with
positively charged host layer by means of electrostatic force and hydrogen bonding
interactions through the water molecule of the interlayer or the hydroxyl group on the

t
sheets (Fig 2) [18]. In contrast to LDHs, most other layered materials have negatively

ip
charged sheets with charge-balancing cations in hydrated interlayer regions, such as
phylosilicate clays and zirconium phosphates etc [19].

cr
LDHs have been known for more than 150 years since the discovery of the mineral
hydrotalcite, [Mg6Al2(OH)16]CO3·4H2O [12]. Partial substitution of the Mg2+ ions by
the Al3+ ions in hydrotalcite compositions leads to the formation of positively charged

us
layers. Stacking of these layers creates charge imbalance that can be compensated by
the anions in the interlayer region such as carbonate etc. The layers can be stacked in
two ways, either with a rhombohedral (3R symmetry) or a hexagonal cell (2H

an
symmetry). Hydrotalcite corresponds to the 3R symmetry, while the 2H analogue is
referred to as manasseite [20]. The electric charge of the layers and the interlayer ions
is just the opposite of that found in silicate clays (cationic clays). Due to the above
structural characteristics, hydrotalcite compositions are sometimes referred to as
M
layered double hydroxides, anionic clays or hydrotalcite-like materials. LDHs have
found a wide range of applications. For instance, [Mg6Al2(OH)16]CO3·4H2O (denoted
as Mg3Al-CO3) has been used to improve the heat retention properties of low density
polyethylene [19, 21] that is a good fire retardant material [22]; Mg3Al-CO3 can be
d

used to stabilize poly(vinyl chloride) [23]. The thermal stability and UV


photo-stability of C.I. Pigment Red 48:2 can be largely improved after it is
te

intercalated into the Mg3Al-NO3 LDH precursor [24].

2.3. The POM/LDH nanocomposites


p

The POM/LDH nanocomposites have attracted wide interests as they have great
ce

advantages over other LDH compounds, especially as catalytically active and


microporous materials [25]. Both electrostatic interactions and hydrogen bonding
interactions can be found between the brucite-like layers of LDHs and the intercalated
POM anions. The general structure of the POM-LDH composition is illustrated in
Ac

Fig 2.

The first reported pillared microporous material of poly-acid type is Zn2Al-V10O28,


which shows an improved catalytic activity on photo-oxidation of isopropyl alcohol to
acetone in comparison to [V10O28]6- or Zn2Al-LDH [24]. The intercalation of the
Keggin type POM anions such as [H2W12O40]6- into LDHs shows that the Keggin
anions enter into the interlayer region with the C2 axis perpendicular to the inorganic
layers [25, 26], producing gallery height close to 10 Å and providing sufficient room
to allow the physical and chemical processes to occur at the interior active sites.

Although POM intercalated LDHs possess superior application capabilities than


POMs or LDHs themselves, the intercalation of POMs into LDHs faces several
challenges. First of all, during the anions exchange reactions between LDH precursor

6
Page 6 of 28
Edited Aug 29

and the POM anions, it is possible that the M2+/M3+ cations are leached out of the
LDHs, which occurs under the neutral to slightly acidic reaction conditions [27].
Secondly, the control over the final M2+:M3+ ratio in the POM pillared LDHs is
essential since this ratio determines the charge density of the LDH layers, and
therefore affects the basal spacing between the intercalated POMs [28]. To synthesize
the POM/LDH nanocomposites with different pore size distributions, it is necessary to
retain, as much as possible, the desired M2+:M3+ ratio from the initial step of synthesis
through to the ultimate product. Finally, some POM anions are hydrolytically unstable

t
at weakly acidic to basic pH [29] and the pillaring reactions are accompanied by the

ip
co-formation of an impurity phase, which is the M2+ rich salt of the POM that deposits
on the surface of the LDH crystallites [25]. Deposition of POM salt can block the
micropores of LDHs, resulting in low surface areas [25]. To date, several synthetic

cr
approaches can be used for the preparation of POM pillared LDHs. Since the first
three methods are most known and they have been discussed in other reviews, we
only briefly introduce them herein.

us
2.3.1. Ion exchange pathway

The decavanadate anion, [V10O28]6-, was among the very first POMs to be

an
successfully intercalated into LDHs by the ion exchange method [7,10]. Chemical
analyses and X-ray basal spacing (d001=11.9Å) were compatible with the indicated
formulae. Pillaring was achieved by ion exchange of NO3- in LDH interlayer with
[NH4]6[V10O28].6H2O at pH 4.5 and 25oC (Fig 3).[10] Chemical analysis indicated the
M
absence of NO3- and the presence of 0.17mol [V10O28]6- per LDH equivalent as
expected for complete exchange. The basal spacing (d001=11.9Å) corresponded to
gallery heights of 7.1 Å (three oxygen planes) and to a V10O286- orientation in which
d

one C2 symmetrical axis was perpendicular to the host layers as shown in Fig. 4.
Further verification of the intercalated [V10O28]6- was obtained from the 51V
te

MAS-NMR spectrum of the pillared Zn2Al-V10O28 composite.


p
ce
Ac

Fig 3. Schematic representation of POM/LDH nanocomposites through ion exchange synthetic


route (H2O, POM, CO32- and NO3- molecules in the interlayer in polyhedral representation, and
the anions are schematized) [10].

7
Page 7 of 28
Edited Aug 29

t
ip
cr
Fig 4. Structural orientation of Zn2Al-V10 O28.

It is difficult to obtain the POM/LDH nanocomposites in crystalline form, which is in

us
part because LDH hosts are basic and most POMs are acidic. Thus, the hydrolysis
reactions of the LDHs or POMs can result in the products that are poorly ordered with
amorphous or multi-crystalline phases [26]. However, the POM/LDH composites can

an
be prepared by ion exchange of a POM anion with a LDH precursor intercalated by an
organic anion [25]. Since the ion-exchange reactions are topotactic, any layer stacking
defects in the precursor could appear in the resulting pillared product. As such,
Pinnavaia et al. employed a swelling reagent to overcome this problem during the ion
M
exchange synthesis [26] (Fig 5).
d
p te
ce
Ac

Fig 5. The synthetic pathway for preparation of the POM/LDH nanocomposites using the ion
exchange pre-swelling agent method (H2O molecules, glycerol molecules, tosylate and POM
anions in the interlayer: spacefill style) [25].

The key to implementing the ion exchange method successfully is to utilize the
LDH-chloride or LDH-nitrate as precursor that has been prepared by co-precipitation
[25]. It is noted that the drying temperature could affect the orientation of the
intercalated POM anions. Generally, water molecules can be removed when the
POM/LDH nanocomposites are heated, which results in stronger electrostatic and

8
Page 8 of 28
Edited Aug 29

hydrogen bonding interactions between POMs and LDHs, and thereby a decreased
gallery height. Moreover, it can push POMs to adopt a thermodynamically stable
gallery orientation. At 100oC, the reorientation appears to be responsible for the
observed gallery contractions for the POM/LDH nanocomposites [25].

Pillaring of LDHs depends on the conditions of swelling of LDHs. In other words, it


can be affected by wetting the LDHs with deionized water for sufficiently long time
[26]. The nature of the anions initially present in LDHs can have a great influence on

t
the ion exchange reaction. As a result, the anions weakly held between the hydroxide

ip
layers are more appropriate for ion exchange as they can easily be replaced by the
pillaring anions. Wang et al. indicates that the order of preferred leaving anions is
NO3->Cl-> SO42-> CO32-. As indicated earlier, carbonate is the hardest one to

cr
exchange and can be avoided by using pre-degassed water under CO2 free
atmosphere. POM pillaring LDHs are also affected by the softness of the LDHs
powder and the slurring time of LDHs with water prior to the pillaring reactions [26].

us
The carbonate anions are difficult to replace in LDHs using the ion exchange method.
However, Pinnavaia et al. successfully replaced the CO32- anion by pre-treating the

an
LDHs carbonate with related polyol intercalates formed from glycerol or
triethyleneglycol [25]. This is particularly important as the [M2+1-xM3+x][CO3]x/2 are
attractive precursors for pillared derivatives owing, in part, to the exceptional ease of
synthesis of these compositions. Nevertheless, the relative ion-exchange inertness of
M
LDH carbonates has discouraged their use for pillaring reaction. Kooli et al.,
decomposed the carbonate anion by thermal treatment of the precursor at
temperatures between 300–600oC overnight and the resulting LDHs were suspended
in decarbonized water overnight at room temperature for reconstitution before
d

intercalation with POM anions [30]. This step seems to solve the problem of replacing
carbonate anions in LDH precursors before intercalation with other anions as
te

discussed below.

2.3.2. Reconstitution synthetic pathway


p

LDHs exhibit high affinity for CO32- anions, and ion exchange reactions are unsuitable
ce

very often using LDHs-CO3 as precursor in most cases [30]. However, the carbonate
anion can be removed by thermal decomposition and the oxides obtained can absorb
anions when suspended in solution to restore the hydrotalcite structure [30]. Some
calcined LDHs reconstruct to the original hydrotalcite structure (the memory effect)
Ac

on exposure to aqueous solution [31]. This memory effect is only possible for LDHs
heated below 600 oC and reconstituted in aqueous solution to reform the original LDH
structure [32]. Technically, since the carbonate is driven off as CO2 during the heating
step and no other anions are present during the reconstitution phase, the product
obtained from a calcined LDH such as [Mg6Al2(OH)16]CO3 is a synthetic form of
meixnerite, [Mg3Al(OH)8]OH [28]. The conversion of the meixnerite-like phase to a
second intermediate by the substitution of adipate anions for the inter-lamellar
hydroxyl groups greatly facilitates the desired POM anion exchange reaction [28].

This method yields LDH materials pillared with POMs possessing the Keggin,
Wells-Dawson and Finke structures, in which total BET surface areas of up to
136m2g-1 were obtained [25]. (BET is the Brunauer, Emmet and Teller theory). In
addition, as much as 50% of the total surface area was attributed to micropores.

9
Page 9 of 28
Edited Aug 29

However, other oxides do not reconstruct automatically and need hydrothermal


treatment between 80 and 250 oC [33] to reproduce the hydrotalcite structure. This
methodology is called reconstitution and has been employed to intercalate POM
anions such as [V10O28]6- or [Mo7O24]6- into LDH compositions [34]. The pH of the
solution before adding the pillaring species is important as is the time required to
calcine the precursor LDH. It was reported that new POM/LDH compositions can be
obtained by heating the LDH carbonate under nitrogen atmosphere at around 450oC
for 18 hours prior to addition of the pillaring species at pH 4.5 [34]. Jan et al. also

t
successfully intercalated α-[SiW9O37{Co(H2O)}3]10− into Mg3Al-LDH through the

ip
same methodology [35].

2.3.3. Co-precipitation synthetic pathway

cr
In 1993, Pinnavaia et al. developed the approach, so called the co-precipitation
method, to synthesize the POM/LDH nanocomposites [26]. In this method, the M2+

us
and M3+ cations can react directly with the intercalated A- ions. For instance,
Pinnavaia et al. reacted Zn2+ and Al3+ ions with an acidic solution of α-[SiW11O39]8-,
resulting in directly the formation of a layered double hydroxide derivative with an

an
approximate composition of [Zn24Al8(OH)64][α-SiW11O39] with an exceptionally
well-ordered gallery height of 9.8 Å [26]. Much sharper X-ray reflections can be
achieved compared with the previous two methods and the compositions indicated
improved stacking order or scattering domain size along the stacking direction. The
main shortcoming of this method is the formation of a quasi-crystalline M2+/M3+ salt
M
of the POMs. However, this may be overcome by increasing the pH of the solution to
7.5 [26]. Note that the co-precipitation method is only suitable for acidic LDH
derivatives such as those with layer compositions of the type Zn1-xAlx(OH)2x+. The
d

efforts to prepare pillared forms of typically basic LDH derivatives, e.g., layer
compositions of the type Mg1-xAlx(OH)2x+, affords instead poorly ordered salts of the
te

POMs or base hydrolysis of the POMs [25].

2.3.4. Electrochemical reduction


p

POM anions show unmatched range of molecular structures that depend on the
ce

number of addenda metal centers and heteroatoms. The typical Keggin-type structure
contains twelve metal-oxygen octahedra, which form a shell surrounding a
tetrahedrally coordinated heteroatom. It is a bulky species with a diameter of c.a. 10Å.
For successful intercalation of these Keggin clusters into the LDHs, the POMs should
Ac

possess enough charge to compensate for excess positive charges of the hydroxide
layers on the LDHs [6]. In addition, the pH values should be not only suitable for
POMs to be resistant to hydrolytic decomposition, but also for the stability of LDHs.
Therefore, the minimum negative charge on a Keggin-type heteropoly anion
intercalated into LDHs should have charge density of at least -4. POMs without the
above characteristics are unfavourable for pillaring into the LDHs [36].

In terms of POMs that do not satisfy the above requirements, electrochemically


treated POMs can be used for intercalation into LDHs [36]. Electrochemical reduction
of POM anions leads to a facile reduction by several electrons per Keggin unit, giving
rise to heteropoly blues. The reduction renders the anions less acidic, making them
viable for pillaring into LDHs [36, 37]. However, it should be noted that this is only
suitable for a certain class of POMs, i.e. POMs of type I like the Keggin anions with a

10
Page 10 of 28
Edited Aug 29

single terminal oxygen atom on each addenda metal center that can be reversibly
reduced by several electrons without significant variations in the POM structures; on
the contrary, POMs of type II like β-octamolybdate with at least one addenda metal
center displaying two cis-related terminal O atoms cannot be reversibly reduced
because reductions promote skeletal transformations. Serwicka et al. managed to
intercalate [PMo12O40]3- anions into the [Mg0.67Al0.33(OH)2]0.33•nH2O using the
electrochemical reduction technique [36]. They reduced the [PMo12O40]3-
electrochemically to get [PMo12O40]7- before intercalation into LDHs. Recently, Xu et

t
al. [38] used the same technology to engineer the nanocomposite of

ip
[Zn2Al(OH)6][α-PW10Mo2O40]1/5 that can adsorb dyes (methylene blue) from waste
water.

cr
2.3.5. Ultrasound treatment

In 1987, Chatakondu et al. documented the use of ultrasound to significantly increase

us
the rates of intercalation of organic and organometallic compounds into various
layered inorganic oxide and sulfide host solids [39]. Consequently, Hu et al. in 1997,
successfully intercalated [SiW11O39Co(H2O)]6- into a Zn2Al-LDH by the use of

an
ultrasound technology to achieve [Zn2Al(OH)6][SiW11O39Co(H2O)]1/6•4H2O [25].
Kooli et al reported the successful intercalation of decavanadate-exchanged
Mg3Al-LDH using ultrasonic treatment of an aqueous suspension of the
carbonate-containing precursor [40]. The above results confirm that ultrasound can
M
stimulate inorganic and organic chemical reactions at or near room temperature [41].

Reactions with ultrasound treatment may be divided into two categories [42]: (i)
Those that can be accelerated by ultrasound and (ii) those which otherwise would not
d

take place in the absence of ultrasound. It is likely that POMs exchange with the LDH
precursors can be facilitated by the high dispersion of the agglomerated particles
te

following ultrasonic treatment. In addition, the ultrasonic treatment causes a rapid


movement of the particles in the aqueous environment and this effect results in a
continuous exposure of the particles to the anions, thereby improving the access of the
p

POM species to the inter-layer space.


ce

2.3.6. The delamination technology

The delamination of LDHs results in the positively charged thin platelets with a
thickness of a few atomic layers, which can be used as building units for engineering
Ac

new, designed organic/inorganic composite materials [44-46]. The delamination of


LDHs is more complicated than other cationic clays such as montorillonite and
laponite, which can be exfoliated into single clay sheets in aqueous suspension. In
contrast, LDH layers have high charge densities between the sheets, resulting in
strong interlayer electrostatic interactions. As such, extensive interlayer hydrogen
bonding networks lead to a tight stacking of the lamellae. The as-defined morphology,
therefore, prevents accessibility to the major part of the surface or exfoliation of the
sheets in water or in any other non-aqueous solvents.

11
Page 11 of 28
Edited Aug 29

Fig 6. The “top-down” and “bottom-up” strategies for LDH single layers [44].

t
ip
To synthesize LDH nanosize plates, two approaches are currently in use: “bottom-up”
and “top-down” (Fig 6) strategies. The “top-down” synthesis is the most widely
developed method. The method requires the modification of the inter-lamellar

cr
environment of LDHs and then the selection of an appropriate solvent system. For
instance, ion-exchange intercalation of anionic surfactants such as dodecyl sulfate into
LDHs, the aliphatic tails of dodecyl sulfate are elongated and present a high degree of

us
inter-digitation, which consequently enlarges the Brucite interlayer distance and
weakens the Brucite interlayer force. Delamination then occurs when it is dispersed in
a highly polar solvent, which solvates the hydrophobic tails of the intercalated anions.
For the bottom-up synthesis, aqueous co-precipitation system is introduced into an oil

an
phase with dodecyl sulfate as surfactant and n-butanol as co-surfactant. The reverse
micelles act as nano-reactors, in which LDH single layers can be formed due to
limited space and nutrients. Pioneering works on delamination of LDH nanosheets
(Fig 7) were reported by Forano et al, [47] Sasaki et al, [48] and Kannan [49] et al,
M
respectively.
d
p te
ce

Fig 7. Illustration of the possible delamination mechanism for LDHs in formamide [48].

Recently, Song et al developed new POM/LDH nanocomposites functional materials


Ac

by employing layer-by-layer (LbL) assembly technique. They managed to synthesize


two-colour luminescent ultra-thin films (UTFs) of (EuW10/LDH)m/
(BNMA@PVS/LDH)n (m = 3, 6, 9, and n = 0 - 21 on the basis of the hybrid
assemblies of EuW10 (Na9[EuW10O36]•H2O), PVS (Poly-vinyl sulfonate), BNMA
(Bis(N-methylacridinium), and exfoliated LDH monolayer step by step (Fig 8) [50].

12
Page 12 of 28
Edited Aug 29

t
ip
Fig 8. Two-colour luminescent ultra-thin films of (EuW10/LDH)m/(BNMA@PVS/LDH)n based on
the hybrid assemblies of EuW10, BNMA, and exfoliated LDH monolayers. BNMA =

cr
Bis(N-methylacridinium; PVS = Poly-(vinyl sulfonate) [50].

Layered rare-earth hydroxides (LRHs) are another type of LDHs, in which only one

us
type of cation (Ln3+) resides in the host layer. LRHs possess the general formula of
Ln8(OH)20(Am-)4/m.nH2O (Ln = rare-earth ions; A = intercalated anions). Pioneering
works were performed on layered yttrium hydroxyl nitrate by Byeon’s group [51].
They successfully exfoliated nitrate intercalated layered yttrium hydroxide

an
Y2(OH)5NO3·1.5H2O in toluene. The decrease of interlayer interactions by the
exchange between nitrate ions and oleate anionic surfactants resulted in a complete
exfoliation in non-aqueous solvents. The same group also reported successful
exfoliation/delamination of gadolinium hydroxide nanosheets [52]. Positively charged
M
layered gadolinium hydroxide nanosheets can be prepared by directly exfoliating
LGdH–NO3 in formamide that proceeds via incorporation of POM anions into the
interlayer space. In 2011, the successful intercalation of POM anions into layered
gadolinium hydroxides and layered europium hydroxides was reported [53], in which
d

the top-down strategy was applied. As a result, self-assembly of nanoparticle building


blocks into ordered superstructures was achieved [53]. It is worth noting that the most
te

striking feature is the sonication process of the layered rare earth composites in
aqueous media, resulting in the exfoliation of the positive nanosheets under different
pH environments and in the presence of POM anions, followed by self-assembly
p

giving rise to superstructures of LRH-POM compositions (Fig 9) [54]. Fig 10


presents the synthetic pathways for engineering new POM/LDH and POM/LRH
ce

nanocomposites.
Ac

Fig 9. Directional self-assembly of (LEuH/[H2W12 O42]10−) and LTbH/[H2W12O42]10−) hybrids with


different orientations of POM anions [54].

13
Page 13 of 28
Edited Aug 29

t
ip
cr
us
an
Fig 10. Summary of the synthetic pathways of engineering POM/LDHs and POM/LRH
nanocomposites
M
3. Material Applications of the POM/LDH nanocomposites

Table 1. Basal spacing of some POMs intercalated LDHs.


d
LDH POM type Basal spacing d001(Å) Ref
6-
Zn2Al, Zn2Cr, Ni3Al [V10O28] 11.9 [7]
te

6- 7-
Zn2Al [H2W12O40] , [SiV3W9O40] 14.6 [25]
Mg2Al [Mo7O24] 6-, [V10O24] 6- 12.2 [25]
p

6- 8-
Mg2Al [H2W12O40] , [SiW11O39] 14.8 [26]
8-
Zn3Al [SiW11O39] 14.6 [26]
ce

The POM/LDH nanocomposites show superior applicability in catalytic reactions


than POMs due to the large gallery heights exhibited by these materials (Table 1).
Upon intercalation, POM anions enter into the interlayer region of LDHs with their C2
Ac

axis perpendicular to the inorganic layers in most cases [55]. The POM intercalated
LDH nanocomposites possess superior application capabilities as compared with
parent materials. For instance, the intercalation of POM anions into LDH layers
results in an insoluble product in aqueous media and the POMs retains their catalytic
characteristics in the final product, converting homogeneous POM catalysts into
heterogeneous ones. The combined structural and chemical features of the
POM-pillared LDH materials give the final product its superior qualities.

3.1. Catalytic Oxidation reactions using the POM/LDH nanocomposites

14
Page 14 of 28
Edited Aug 29

Table 2 Oxidation reactions catalysed by POM-LDH intercalation compounds


Reaction Oxidant LDH POM Ref.
6-
Photo-Oxidation of O2 Zn2Al [V10O28] [7, 25]
iso-propanol to acetone
Alkene epoxidation H2 O2 Mg2Al or Zn2Al [Mo7O24]6-, [W7O24]6-, [H2W12O40]6- [56, 57]
Zn3Al or Ni2Al [SiW12O40]4-, [SiW11O39]8- [58]
5-
Benzaldehyde to benzonic H2 O2 Mg2Al [CoW12O40] [59]

t
acid

ip
o-xylene to o-tolualdehyde O2 LiAl2 [V2O7]4- [60]
1 2-
H2O2 to O2 H2 O2 Mg0.9 Al0.1 [MoO4] [61]

cr
Several oxidation reactions (Table 2), such as the photo-oxidation of iso-propanol to
acetone [7, 25], the shape selective epoxidation of alkenes [56-58], oxidation of
benzaldehyde to benzonic acid [59] and selective oxidation of o-xylene to

us
o-tolualdehyde [60] etc, can be catalyzed using the POM/LDH nanocomposites as
catalysts. According to Kwon et al., the intercalated Zn2Al-[V10O28] was substantially
more active than the vanadate cluster of [V10O28]6-, producing 6 moles of acetone per
mole of decavanadate under oxygen [7]. In addition, the Keggin POMs of

an
[BVW11O40]7-, [SiV3W9O40]7-, and [H2W12O40]6- intercalated Zn2Al-LDH showed the
same photo-activity for iso-propanol to acetone conversion in the presence of oxygen
[7].
M
Different degrees of gallery accessibility for 2-hexene and cyclohexene were observed
when POMs like [Mo7O24]6-, [W7O24]6- and [H2W12O40]6- were intercalated into LDHs
[56,57]. The molybdate-pillared LDH showed slower oxidation of cyclohexene due to
d

its smaller gallery height and lower accessibility of substrate to the inter-lamellar
catalytic sites. However, the decrease in the relative reactivity of cyclohexene is less
te

remarkable for the tungstate-pillared LDHs with a higher gallery space. The gallery
pores are believed to provide a unique chemical environment for these catalytic
reactions. This is particularly the case when POM anions alone are applied as
p

catalysts, resulting in the formation of a mixture of epoxide and diol products; while
ce

the POMs intercalated LDHs as catalysts afford the desired diol product [57]. It was
postulated that the surface acid-base properties of the LDHs promote ring opening
of the epoxide to the diol. Moreover, the intercalation of [SiW12O40]4- into LDHs
layers did not change its structure, and the intercalated compound of Zn3Al-[SiW12O40]
Ac

exhibited effective epoxidation of alkenes with O2 in the presence of isobutyraldehyde


[58]. The [SiW11O39]8- intercalated between the LDH layers are more stabilized than
free POM, and the resulting Zn3Al-[SiW11O39] exhibited remarkable shape selectivity
for the oxidation of cyclohexene in the presence of H2O2 [58].

Interestingly, Pinnavaia et al [57] provided experimental evidence that it is


unfavourable to dry Zn2Al-[Mo7O24] and Zn2Al-[W7O24] at 120oC before catalytic
reaction as this could decompose the catalyst. The preferred method is to air dry the
catalyst before performing the reaction at 75oC. Additionally, except for
Zn2A1-H2W12O40 catalyst, the resultant products under the original conditions do not
support the catalytic reactions. In another words, apart from Zn2A1-H2W12O40, small
amount of products observed was due to the decomposed products and/or the

15
Page 15 of 28
Edited Aug 29

accessible POM anions on the surface of the POM/LDH nanocomposites. The


impurity formed in the Zn2A1-H2W12O40 composite materials has a significant
contribution to the epoxidation reactions. This particular controversy is yet to be
solved bearing in mind that the calcined products also give mixed oxides of high
catalytic property [43].

Furthermore, a higher cis to trans ratio is observed for the oxidation of 2-hexene over
the Zn2Al-[Mo7O24]. Pinnavia et al. suggested that the improved cis to trans ratio

t
relative to homogeneous solution probably is due to a steric constraint imposed by the

ip
interstitial environment [62]. The gallery pores are believed to provide a unique
chemical environment for these catalytic reactions. This is particularly the case when
POM anions alone are applied as catalysts, resulting in the formation of a mixture of

cr
epoxide and diol products; while the POM intercalated LDHs as catalysts afford the
desired diol product. Note that the epoxidation reactions do not take place in the
presence of the LDH precursor alone nor in the absence of the POM catalysts. Similar

us
results have been observed by Carriazo et al.[63].

Another very important oxidation reaction by the POM/LDH nanocomposites is the

an
selective oxyfunctionalization of organic compounds by singlet molecular oxygen
(1O2) [61]. Molecular oxygen is a very important industrial product as it is used in
production of enylhydroperoxides, dioxetanes and endoperoxides [64]. Molecular
oxygen is preferred in these reactions due to its regio- and chemo-selectivity
M
properties as compared with other radicals or electrophilic agents such as RO• or
ROO• radicals or (in)organic peracids [65]. Van der Ven et al. developed a
molybdate-exchanged layered double hydroxide, which was a heterogeneous catalyst
for conversion of H2O2 into 1O2 [23, 61]. The use of such a heterogeneous catalyst
d

allowed working with catalytic amounts of molybdate without base dissolution and
provided more freedom in the choice of an appropriate solvent for the substrate
te

molecule [23, 61].

The highest product yield at complete H2O2 consumption was obtained with the same
p

weight of LDH catalyst containing less Mo. The LDH catalyst with high Mo loading
produced a high rate of 1O2 that caused depletion of the rather hydrophobic substrate
ce

1-methyl-1-cyclohexene around the surface occupied by the 1O2 generating centres.


Since 1O2 has a low lifetime in water-containing solutions, it may have been
deactivated by solvent quenching or by quenching on the solid material before it
could react with the substrate to form an oxygenated product SO2.
Ac

Alternatively, the 1O2–1O2 annihilation may have occurred. This phenomenon occurs
for the hypochlorite–hydrogen peroxide reaction, and arises when a large amount of
1
O2 is formed per unit of time and volume [23]. A similar situation may arise for the
LDH catalyst with the highest Mo loading and 1O2 generation rate. However, the
above explanations remain speculative as the researchers did not succeed in
detecting the 663.4 nm luminescence, resulting from the simultaneous decay of two
excited 1O2 molecules to the ground state, which then would be evidence enough for
these conclusions [64].

Technically though, the low Mo concentration in the catalyst (typically 0.2 mmol per
gram) is relatively appropriate for efficient oxygenation with the 1O2 generated from
H2O2 [61]. In addition, the change of the M2+:M3+ratio could affect the performance

16
Page 16 of 28
Edited Aug 29

of the catalysts with the best results being obtained from the LDH with a low Al3+
ratio [61]. The oxidizing species, like the monoperoxomolybdate, are not only
favoured at lower H2O2/Mo ratios, but also at lower pH. Hence it seems that with
Mg0.9Al0.1–LDH, which is the most basic support, the formation of such species is
sufficiently suppressed to avoid significant epoxidation.

The POM/LDH nanocomposites heterogeneous catalysts have also found a wide


application in green chemistry technology in recent years. For instance, the new

t
environmental legislations [66] require desulfurization of fuels and effluents from

ip
industries/factories. Catalytic oxidative desulfurization has been a promising
alternative technology to the current hydrodesulfurization process for reducing sulfur
in gasoline and diesel fuels [67]. In catalytic oxidative desulfurization process, the

cr
sulfides and thiophene derivatives can be converted into their corresponding
sulfoxides and sulfones, which are preferentially extracted due to their increased
polarity. The reactions are carried out in liquid phase and under mild conditions. The

us
hydrogen peroxide is mostly chosen as oxidizing agent owing to its high
effective-oxygen content (47%), cleanliness (it produces only water as by-product)
and acceptable safety in storage and operation.

an
M
d
p te

Fig 11. Oxidation of thioethers with H2O2 using WO4 ion intercalated into layered double
hydroxides as a catalyst [69].
ce

Although different types of catalysts have been used [68], the most attractive catalysts
are Mg2Al-MO4 (M = V, Mo and W etc) with ortho-tungstate, -molybdate and
-vanadate as guest species in the LDH layers. The POM/LDH composites were tested
Ac

in the oxidation of the organic sulfides and thiophene derivatives, which are the major
sulfur-containing compounds present in fuels [69]. The oxidation reaction was carried
out using dilute hydrogen peroxide (aqueous solution, 30%) at 40 oC in the presence
of MeCN as solvent. Under these conditions, the oxidation reaction of sulfides led to
the corresponding sulfoxides and sulfones as major products (Fig 11)

17
Page 17 of 28
Edited Aug 29

t
ip
cr
us
an
Fig 12. Possible reaction mechanisms for the tetrahydrothiophene oxidation using LDH-WO4
catalyst [70].

Tetrahydrothiophene can be oxidised to sulfolane over V-, Mo- and W-containing


LDHs [70]. Sulfolane is a valuable industrial solvent with good thermal and
M
hydrolytic stability, high boiling point and high density, and it is widely used in
extraction and extractive distillation processes for the recovery of high-purity
aromatic hydrocarbons from streams refinery process [71]. It is also used in the
d

washing process to remove hydrogen sulfide, carbonyl sulfide and carbon dioxide
from synthetic gas or natural gas [71]. Industrial preparation of sulfolane is based on
te

the 1,4-addition reaction of SO2 to butadiene [71]. It can also be produced from
sulfoxidation reaction of tetrahydrothiophene by hydrogen peroxide [72]. Under these
sulfoxidation reactions, Maciuca and co-workers have successfully used WO42-
p

intercalated LDH as catalyst to achieve better yields of sulfolane (Fig 12) [70].
ce

3.2. Dehydrogenation reactions using the POM/LDHs as catalysts

Dehydrogenation reactions have been employed to produce important industrial


materials such as styrene [73]. The use of POM/LDH composites as catalysts in
Ac

dehydrogenation reactions has been reported [74]. For example, vanadium containing
catalysts such as “vanadia–alumina” give better yields of styrene than Fe-K-Cr
oxide-based catalyst in these dehydrogenation reactions [74]. The vanadium
containing catalysts help to solve the problem of catalyst deactivation that is
experienced by other catalysts [75] while the use of CO2 as a cofeed gas solves the
thermodynamic challenges [76]. The best result was achieved when vanadium was
intercalated into a Mg3Al-LDH and further calcined at 450oC for 5 hours [77]. The
resulting mixed oxides showed relatively higher catalytic activity than those of the
POM-LDH precursors [78].

3.3. Esterification reactions using the POM/LDH nanocomposites

Esterification of acetic acid with n-butanol to yield n-butyl acetate is an important

18
Page 18 of 28
Edited Aug 29

industrial process that finds application in the manufacture of artificial perfume,


leather, photographic films, lacquer, plastic and safety glass. Das et al [79] achieved a
higher selective esterification of acetic acid using n-butanol with a 15 wt. %
molybdophosphoric acid intercalated into Zn2Al-LDH as a heterogeneous catalyst.
The optimum acid:alcohol ratio for the reaction was 1:10. Increasing the alcohol
amount hinders the reaction because excess amounts of alcohol block the active sites
on the catalyst surface or prevent any nucleophilic attack by shielding the protonated
alcohol [80,81]. The reaction of acetic acid with n-butanol proceeded as a first order

t
reaction with respect to acetic acid and zero order with respect to n-butanol [82]. The

ip
esterification reaction was a Brønsted acid catalysed reaction. At first, there was
formation of carbonium ion due to the chemisorptions of n-butanol on the catalyst
surface following the Eley–Rideal mechanism. The carbonium ion formed, was stable

cr
and attacked the nucleophilic centre of acetic acid, leading to the formation of an
unstable intermediate which in the final step released a proton, thus, regenerating the
catalyst and forming ester. The role of an acid catalyst was to facilitate the formation

us
of the carbocation, and to help remove OH− from n-butanol [83].

3.4. Oximation of aromatic aldehydes to aldoximes and ketoximes

an
Industrial production of aldoximes and ketoximes, one type of important chemical
intermediates in synthetic organic chemistry and chemical industry [84], is performed
by the oximation of aldehydes and ketones with a mineral salt of hydroxylamine
M
(NH2OH·HC1 or NH2OH·H2SO4) [85]. However, the high cost of hydroxylamine and
the production of large amounts of unwanted inorganic salts is the main industrial
challenge in this process. Alternatively, several studies have focused on the use of
H2O2 as oxidant in the oximation reactions of aldehydes to aldoximes and ketoximes
d

[86]. However, the aldoximes formed undergo acid catalysed Beckmann


rearrangement to the amide or an acid-catalysed dehydration to yield nitriles, resulting
te

in decreased catalytic selectivity.


p
ce
Ac

Fig 13. Sandwich polyoxometalate intercalated layer double hydroxide (Zn3Al–Zn5WO) for
highly selective oximation reaction [89].

Sandwich type POM cluster of Na12[WZn3(H2O)2(ZnW9O34)2]·46H2O [Zn5WO] is


known to be an efficient catalyst for the oximation reactions with H2O2 as oxidant
[87]. However, the inherent acidity of the catalyst can lead to further transformation
of the oximation product. Therefore, it is necessary to develop catalyst to improve the
selectivity. Li et al. reported [88] that self-assembled sandwich POMs of
K11[WMn3(H2O)2(ZnW9O69O34)2]·27H2O (Zn2Mn3WO) and
K11[WFe3(H2O)2(ZnW9O69O34)2]·44H2O (Zn2Fe3WO) intercalated into LDHs hosts

19
Page 19 of 28
Edited Aug 29

are active catalysts for heterogeneous epoxidation reactions, and the presence of basic
hydroxyl groups on the layers of the LDHs may prevent subsequent acid-catalyzed
hydrolysis of the target epoxide product. Inspired by these results, Song et al prepared
POM/LDHs of Zn3Al-Zn5WO, Zn3Al-Zn2Mn3WO and Zn3Al-Zn2Fe3WO and tested
them for the oximation reaction. A large enhancement of selectivity, in the oximation
of aromatic aldehydes by applying the above POM/LDH composites as catalysts, has
been demonstrated through effectively suppressing the formation of by-products
(nitriles and amides) under mild and organic solvent-free conditions. Furthermore, the

t
catalysts can be easily separated, recycled and reused at least ten times without any

ip
decrease of activities (Fig 13). [89] It is worth noting, however, that the synergistic
interactions between LDHs and intercalated sandwich POM anions play key roles for
the high selectivity. Zn3Al–Zn5WO showed a better catalytic activity in the oximation

cr
of different aromatic aldehydes than Zn3Al–Zn2Mn3WO and Zn3Al–Zn2Fe3WO. This
may be due to the catalytic activities of the POMs themselves, since Na–Zn5WO
showed the best results among the three POM clusters [89]. By intercalation into

us
LDHs, the catalytic activities of the POM/LDHs remain close to or slightly lower than
those of the corresponding POMs but with the above indicated advantages [90].

an
3.5. Photo-luminescent materials

Lanthanide-containing polyoxometalates, especially the europium decatungstate


(EuW10 = Na9[EuW10O36]•H2O) [91-93], show excellent luminescence performance
M
when compared with other luminescent materials [94,95]. The photo-luminescence
properties of lanthanide-containing compounds show an attractive feature of line-like
emission, which results in the high colour purity of the emitted light [95]. The emitted
colour depends largely on the lanthanide ions and it is independent of the environment
d

of a given lanthanide ion. It has been observed that the intercalation of lanthanide (III)
complexes into a layered inorganic host exhibits beneficial effect on both the stability
te

and the luminescence performance. Similar observation has been found for lanthanide
complexes doped in sol-gel matrices [95].
p

Investigation of luminescent properties of [EuW10O36]9-, [Eu(BW11O39)(H2O)3]6-, and


[Eu(PW11O39)2]11- intercalated into a Zn–Al LDH host was performed, in which the
ce

host-guest interactions showed a significant influence on the nature of the resulting


intercalated species, although the emission spectra for the pillared LDHs were not
very different from that of K5H[Eu(BW11O39)(H2O)3] [96]. In 2011, Song et al [50]
reported the layer-by-layer (LbL) assembly of delaminated Mg2Al-LDH and
Ac

Na9[EuW10O36]•32H2O, resulting in the formation of a novel EuW10/LDH


nanocomposite-based ultra-thin films (Fig 14). The assembly process was monitored
by UV/Vis absorption and fluorescence spectroscopy measurements to get a stepwise
and regular growth of the UTFs upon increasing the deposited cycles.

20
Page 20 of 28
Edited Aug 29

t
ip
cr
us
an
Fig 14. (a) Schematic representation of the assembly process of the nanocomposite films of
(EuW10/LDH)n. (b) characterization of (EuW10/LDH) (n = 3-18) multilayer UTFs on quartz
substrates with UV/Vis absorption, the absorbance values at 265 nm as a function of the number
M
of deposition cycles, plot of the thickness of the UTFs as a function of n, and AFM images of the
above UTFs with n = 3, 9, 12 [50].

Song et al. [97] also used the above technique to develop well-ordered two-colour
d

luminescent UTFs of (EuW10/LDH)m/(BNMA@PVS/LDH)n (m = 3, 6, 9, and n = 0 –


21) based on the hybrid assemblies of EuW10, BNMA, and exfoliated Mg2Al-LDH
te

monolayers (Fig 15). The fluorescence of the (EuW10/LDH)9/(BNMA@PVS/LDH)n


composite materials remained almost of the same intensity as that of (EuW10/LDH)9
UTFs with the increase in n from 3 to 21. No clear red- or blue shift of the absorption
p

and fluorescence spectra could be observed for the as-prepared UTFs with different
numbers of (EuW10/LDH) and (BNMA@PVS/LDH) bilayers [98]. As a result, the
ce

(EuW10/LDH)9/(BNMA@PVS/LDH)n UTFs maintained the spectral properties of


individual chromophores. Interestingly, the red to green luminescence can be
achieved by tuning the numbers of the (BNMA@PVS/LDH) bilayers that were
assembled onto (EuW10/LDH)9. The polarized luminescence property of the two-color
Ac

(EuW10/LDH)9/(BNMA@PVS/LDH)n UTFs were exploited to characterize the


microenvironment of the guest molecules in the restricted space of the LDHs. Typical
polarized photoluminescence spectra of the (EuW10/LDH)9/(BNMA@PVS/LDH)n (n
= 6 or 9) UTFs showed that the anisotropy values for the resulting UTFs ranged from
1.0 to 3.5, whereas the values of r (anisotropy) for the pristine EuW10 and BNMA
powder and solution samples were almost 0. These results suggested that the films
had long-range stacking order in the normal direction of the substrate, and the
polarization scrambling by means of Förster transfer was minimal in the UTFs. The
assembly of guest molecules into the LDHs was able to generate two-dimensional
uniform orientation of the guest anions within the rigid LDH matrix, which was
favourable for the polarized photoluminescence. It was suggested [97] that the
pronounced orientation effect of guest anions in the two colour UTFs might be caused
by non-covalent interactions between EuW10 and hydroxyl groups on the LDHs, and

21
Page 21 of 28
Edited Aug 29

the ordered alignment of the conjugated polymers.

t
ip
cr
us
an
Fig 15. The assembly process for preparation of the (EuW10 /LDH)m/(BNMA@PVS/LDH)n UTFs.
(BNMA = Bis(N-methylacridinium); PVS = Poly-(vinyl sulfonate)) nitrate; UTFs = Ultrathin
films; EuW10 = [EuW10O36]9-) [97].
M
The use of individual hydroxocation layers as building blocks with polyoxoanion
controlled crystallization enables the engineering of a variety of POM/layered rare
earth hydroxides (LRH) composites [53]. The combination of the luminescent
d

properties of layered rare earth hydroxides (LRHs) and POMs yields unique optical
properties related to the hierarchical structure. Hence, the use of nanosheets produced
te

from LRHs as building blocks for the construction of the defined inorganic/inorganic
hybrid superstructures yields fascinating materials such as the rose-like hierarchical
superstructure that exhibited enhanced photoluminescence of LRH-POM hybrids [53].
p

Interestingly, Beyon et al [53] reported that the 5D0-7F2 emission intensity of


LEuH-Mo7O24 excited at 394 nm was comparable to that of commercial Y2O3:Eu red
ce

phosphor excited at 260 nm. The increased emission intensity arose from the fact that
the incorporation of POM anions (Mo7O246-) reduced the number of hydroxyl groups
onto the LEuH layers, and consequently, the number of hydroxyls coordinated to the
Eu3+ ions is responsible for the luminescence quench. Furthermore, the electric dipole
Ac

5
D0 - 7F2 transitions were sensitive to the chemical bonds in the coordination sphere of
Eu3+ [98]. As such, the increased intensity of this transition was attributed to a
significant change of Eu3+ coordination polyhedra by the incorporation of the
Mo7O246- groups into the LEuH layer.
Replacing the Mo7O246- anions with other POM anions such as heptatungstate,
[W7O24]6− and paratungstate [H2W12O42]10− revealed similar reaction mechanisms
[99]. However, in the case of the W7O246- anion, the assembly occurred at pH 4, while
for [H2W12O42]10−, it occurred at pH 7 due to the ability of the above anions to
interconvert at different pH values [100, 101]. Since heating acidified solutions of
[H2W12O42]10− accelerates its conversion to [W7O24]6−, the assembly reaction should
be carried out at room temperature to avoid such process.

22
Page 22 of 28
Edited Aug 29

3.6. Dyes removal using the POM/LDH nanocomposites as adsorbent materials

Contamination of industrial waste water with dyes from textiles, paper, printing,
plastics, and leather and so on, is a major concern due to their toxicity and
carcinogenicity that affects both aquatic life and human life by extension [102].
Various physicochemical and biological methods are used for dye removal, such as
adsorption, precipitation, coagulation and filtration. Adsorption is a more favourable
procedure due to economic feasibility, simplicity of design, recycle of adsorbent and

t
nonexistence of harmful residues [103]. Commercial activated carbon is an effective

ip
adsorbent for colour dyes removal, but the high cost of activated carbon restrict its
widespread use. In their search for low-cost and effective adsorbents, Bi et al. [38]
engineered the POM/LDH nanocomposite of [Zn2Al(OH)6][α-PW10Mo2O40]1/5,

cr
which was used to adsorb methylene blue dye from waste water.

Technically, the dye was removed by the adsorption process (Fig 16). The

us
Lagergren's first order kinetic model and the Ho's pseudo second order model [104]
were used to experimentally fit the data of methylene blue removal with
[Zn2Al(OH)6][α-PW10Mo2O40]1/5 in order to predict the sorption process mechanisms.

an
The reaction order fits pseudo second order kinetic model and is comparable to the
sorption mechanisms of methylene blue by dodecylsulfate- and dodecyl benzene
sulfonate-intercalated hydrotalcite [105]. In addition, the adsorption process obeyed
the Langmuir model better than Freundlich model suggesting that the adsorption
M
process took place at monolayer sites.
d
p te
ce

Fig 16. The proposed adsorption mechanism for methylene blue with
Ac

[Zn2Al(OH)6][α-PW10 Mo2O40]1/5 [38].

4. Conclusions

Engineering of the POM/LDH nanocomposites has been realized through ionic


exchange, co-precipitation, reconstitution, electrochemical reduction, ultra sonication
and adsorption mechanisms. During anions exchange reactions between LDH
precursors and the POM anions, it is likely to leach the M2+/M3+cations out of the
LDHs layers. The leaching process occurs under the neutral to slightly acidic reaction
conditions, which are typically employed during anion exchange reactions. In addition,
the control over the final M2+:M3+ ratio in the POM pillared product is essential since
this ratio determines the charge density of the LDH layers, and therefore influences
the basal spacing between the POM pillars. To synthesize a series of POM/LDH
nanocomposites with different pore size distributions, it is necessary to retain, as
23
Page 23 of 28
Edited Aug 29

much as possible, the desired M2+:M3+ ratio from the initial step of synthesis through
to the ultimate product.

Some POM anions are hydrolytically unstable at weakly acidic to basic pH and the
pillaring reactions are accompanied by the co-formation of an impurity phase, which
is the M2+ rich salt of the POM that deposits on the surface of the LDH crystallites.
The formation of this salt phase, however, may be greatly reduced by performing the
anion exchange reaction at elevated temperature (80oC) and with minimal aging time

t
after complete exchange has occurred.

ip
The POM/LDH nanocomposites have been widely applied in catalytic reactions such
as epoxidation of alkenes, molecular oxygen generation, esterification of acetic acid,

cr
oxidation of thioether and thiophene etc. These composite materials have found a
wide range of application in enhancing the photo-luminescent properties of lanthanide
anions and promoting the adsorption of harmful materials like dyes from waste water.

us
Recently, delamination of LDH nanosheets before use in the development of the
POM/LDH nanocomposites has been employed to engineer novel materials in this

an
field. This new technology may solve the above stated problems as well as retain the
original composition of POMs apart from opening a new field of application as
photo-luminescent materials based on the ultra-thin film. We believe that the
ultra-thin films based on the POM/LDHs or POM/LRH composites can be employed
M
to develop novel composite materials with diverse applications.

Acknowledgements
d

This research was supported by the National Basic Research Program of China
(2014CB932104), National Science Foundation of China, the Beijing Nova Program
te

(2009B12), the Fundamental Research Funds for the Central Universities (ZZ1227).
p

References
ce

1. M. T. Pope, Heteropoly and Isopolyoxometalates, Springer Verlag, Tokyo, 1984; A. Müller


and S. Roy, in: C. N. R. Rao, A. Müller, A. K. Cheetham (Eds), The Chemistry of
Nanomaterials: Synthesis, Properties and Applications, Wiley-VCH, Weinheim, 2004, pp.
452-475.
Ac

2. M. T. Pope, A. Müller, Angew. Chem., Int. Ed. Engl. 30 (1991) 34;


3. C. L. Hill, Chem. Rev. 98 (1998) 1; L. Cronin, A. Müller, Chem. Soc. Rev. 41 (2012) 7333.
4. E. Coronado, C. Giménez-Saiz, C. J. Gómez-Garcían, Coord. Chem. Rev. 249 (2005) 1776;
T. Akutagawa, D. Endo, S.-I. Noro, L. Cronin, T. Nakamura, Coord. Chem. Rev. 251
(2007) 2547; P. Putaj, F. Lefebvre, Coord. Chem. Rev. 255 (2011) 1642; N. Mizuno, K.
Kamata, Coord. Chem. Rev. 255 (2011) 2358: D.-Y. Du, L.-K. Yan, Z.-M. Su, S.-L. Li,
Y.-Q. Lan, E.-B. Wang, Coord. Chem. Rev. 257 (2013) 702.
5. A. Proust, B. Matt, R. Villanneau, G. Guillemot, P. Gouzerh, G. Izzet, Chem. Soc. Rev. 41
(2012) 7605; D-L. Long, L. Cronin. Dalton Trans. 41 (2012) 9815; Y-F. Song, R.
Tsunashima, Chem. Soc. Rev. 41 (2012) 7384; N. Mizuno, K. Kamata, K. Yamaguchi, Top.
Catal. 53 (2010) 876.
6. C. Hu, D. Li, in: F. Wypych (Ed), Clay Surfaces: Fundamentals and Applications, Elsevier
Inc. San Diego, 2004, pp. 374 - 402; V. Rives, D. Carriazo, C. Martin, in: A. Gill, S. A.
Korili, R. Trujillano and M. A. Vicentè (Eds), Pillared clays and related catalysis, Springer,
New York, 2010, pp. 319 - 422; C. Hu, D. Li, Y. Guo, E. Wang, Chin. Sci. Bull. 46 (2001)

24
Page 24 of 28
Edited Aug 29

1061; H. Nijis, M. de Bock, E. F. Vansant, J. Porous Mater. 6 (1999) 101; N. Mizuno, K.


Yamaguchi, K. Kamata, Catal. Surv. Asia 15 (2011) 68.
7. T. Kwon, G. A. Tsigdinos, T. J. Pinnavaia, J. Am. Chem. Soc. 110 (1988) 3654; A.
Udomvech, P. Kongrat, C. Pakawatchai, H. Phetmung, Inorg. Chem. Commun. 17 (2012)
132.
8. G. M. Woltermann, US. Pat., 4 454 244 (1984).
9. P. Gouzerh, M. Che, L’Actualité Chimique 298 (2006) 9.
10. L. Xu, W. Hu, E. B. Wang, J. Nat. Gas Chem. 6(2) (1997) 155.
11. E. A. Gardner, S. K. Yun, T. Kwon, T. J. Pinnavaia, Appl. Clay Sci. 13 (1998) 479.

t
12. V. Rives, M. A. Ulibarri, Coord. Chem. Rev. 181 (1999) 61.

ip
13. P. Yin, G. Li, T. Liu, Chem. Soc. Rev. 41 (2012) 7368.
14. C. W. L. Baker, D. C. Glick. Chem. Rev. 98 (1998) 3; P. Gouzerh, M. Che, Ĺactualitè
Chimique, 298(2006) 1.

cr
15. R. J. Errington, R. L. Wingad, W. Clegg, M. R. J. Elsegood, Angew. Chem. Int. Ed. 39(21)
(2000) 3884.
16. P. Gouzerh, Y. Jeannin, A. Proust, F. Robert, S. G. Roh, Molecular Engineering 3 (1993)
79.

us
17. X. Duan, D. G. Evans, Structure and Bonding, Springer-Verlag Berlin Heidelberg, 2006, pp.
1-223; Q. Tao, H. He, R. L. Frost, P. Yuan, J. Zhu, Appl. Surf. Sci. 255 (2009) 4334.
18. M. Q. Zhao, Q. Zhang, J.-Q. Huang, F. Wei, Adv. Funct. Mater. 22 (2012) 675; S. J. Mills,
A. G. Christy, J. M. Génin, T. Kameda, F. Colombo, Mineral. Mag. 76(5) (2012) 1289.

an
19. D. G. Evans, X. Duan, Chem. Commun. (2006) 485; J. Sun, Y. Li, X. Liu, Q. Yang, J. Liu,
X. Sun, D. G. Evans, X. Duan, Chem. Commun. 48 (2012) 3379; C. Yu, J. He, Chem.
Commun. 48 (2012) 4933; Z. An, J. He, Chem. Commun. 47 (2011), 8653.
20. F. Trifiro, A. Vaccari, in: J. L. Atwood, J. E. D. Davies, D. D. MacNicol, F. Vogtle, J. M.
M
Lehn, G. Alberti and T. Bein, (Eds.). Comprehensive Supramolecular Chemistry,
Pergamon, Oxford, 1996, 251.
21. M. Hancock, Plasticulture 79 (1988) 4.
22. G. Camino, A. Maffezzoli, M. Braglia, M. De Lazzaro, M. Zammarano, Polym. Degrad.
Stab. 74 (2001) 457.
d

23. S. Miyata, M. Kuroda, US Pat., 4 299 759 (1981); L. Van der Ven, M. L. M. Van Gemert,
L. F. Batenburg, J. J. Keen, L. H. Gielgens, T. P. M. Koster, H. R. Fischer, Appl. Clay Sci.
te

17 (2000) 25.
24. S. Guo, D. Li, W. Zhang, M. Pu, D. G. Evans, X. Duan, J. Solid State Chem.177 (2004)
4597; T. Kwon, G. A. Tsigdinos and T. J. Pinnavaia, J. Am. Chem. Soc., 1988, 110, 3654
25. E. D. Dimotakis, T. J. Pinnavaia, Inorg. Chem. 29 (1990) 2393; T. Kwon, T. J. Pinnavaia, J.
p

Mol. Catal. 74 (1992) 23; S. K. Yun, V. R. L. Constantino, T. J. Pinnavaia, Microporous


Materials 4 (1995) 21; J. Evans, M. Pillinger, J. Zhang, J. Chem. Soc. Dalton Trans. 2963
ce

(1996); S. K. Yun, T. J. Pinnavaia, Inorg. Chem. 35 (1996) 6853.


26. J. Wang, Y. Tian, R. Wang, A. Clearfield, Chem. Mater. 4 (1992) 1276; E. Narita, P. D.
Kaviratna, T. J. Pinnavaia, Chem. Commun. 1 (1993) 60; Z. Wang, T. J. Pinnavaia, Chem.
Mater. 10 (1998) 1820.
Ac

27. J. Wang, Y. Tian, R. C. Wang, J. L. Colon, A. Clearfield, Mater. Res. Soc. Symp. Proc. 233
(1991) 63.
28. M. R. Weir, R. A. Kydd, Microporous Mesoporous Mater. 20 (1988) 339.
29. G. B. McGarvey, J. B. Moffat, Mol Catal. (1991) 137.
30. F. Kooli, C. Dèpege, A. Ennaqadi, A. de Roy, J. P. Besse, Clays. Clay Miner. 45(1) (1997)
92.
31. S. Miyata, Clays Clay Miner. 28(1) (1980) 50; T. Sato, H. Fujita, T. Endo, M. Shimanda, A.
Tsunashima, React. Solids 5(2-3) (1988) 219.
32. W. T. Reichle, Chem. Tech. 58 (1986).
33. F. Kooli, M. A. Ulibarri, V. Rives, Mater. Sci. Forum 152-153 (1994) 375; T. Sato, M.
Tezuka, T. Endo, M. Shimada, J. Chem. Technol. Biotechnol. 39(4) (2007) 275.
34. K. Chibwe, W. Jones, Chem. Mater. 1 (1989) 489.
35. S. K. Jana, Y. Kubota, T. Tatsumi, J. Catal. 255 (2008) 40.
36. E. Serwicka, P. Nowak, K. Bahranowski, W. Jones, F. Kooli, J. Mater. Chem. 7 (1997)
1937.

25
Page 25 of 28
Edited Aug 29

37. A. G. Tsigdinos, Top. Curr. Chem. 76 (1978) 1.


38. B. Bi, L. Xu, B. Xu, X. Liu, Appl. Clay Sci. 54 (2011) 242.
39. K. Chatakondu, M. L. H. Green, M. E. Thompson, S. J. Suslick, J. Chem. Soc., Chem.
Commun. 900 (1987).
40. F. Kooli, W. Jones, J. Mater. Sci. Lett. 16 (1997) 27.
41. R. Roy, D. K. Agrawal, V. Srikanth. Mater. Res. 2412 (1991).
42. T. J. Mason, J. P. Lorimer, Sonochemistry: Theory, Application and Uses of Ultrasound in
Chemistry, Ellis Horwood Chichester, 1978.
43. Y. Guo, D. Li, C. Hu, E. Wang, Microporous Mesoporous Mater. 56 (2002) 153.

t
44. Q. Wang, D. O’Hare, Chem. Rev. 112 (2012) 4124.

ip
45. S. O’Leary, D. O’Hare, G. Seeley, Chem. Commun. 1506 (2002); B. G. Li, Y. Hu, R.
Zhang, Z. Y. Chen, W. C. Fan, Mater. Res. Bull. 38 (2003) 1567.
46. J. H. Choy, J. Phys. Chem. Solids 65 (2004) 373; Q. Wu, A. Olafsen, Ø. B. Vistad, J. Roots,

cr
P. Norby, J. Mater Chem. 15 (2005) 4695.
47. M. Adachi-Pagano, C. Forano, J. P. Besse, Chem. Commun. 1 (2000) 91; F. Leroux, M.
Adchi-Pagano, M. Intissar, S. Chauviere, C. Forano, J. P. Besse, J. Mater. Chem. 11 (2001)
105.

us
48. L. Li, R. Ma, Y. Ebina, N. Iyi, T. Sasaki, Chem. Mater. 17 (2005) 4386; Z. Liu, R. Ma, M.
Osada, N. Iyi, Y. Ebina, K. Takada, T. Sasaki, J. Am. Chem. Soc. 128 (2006) 4872; L. Li,
R. Ma, Y. Ebina, K. Fukuda, T. Sasaki, J. Am. Chem. Soc. 129 (2007) 8000.
49. N. Iyi, Y. Ebina, T. Sasaki, Langmuir 24 (2008) 5591.

an
50. J. Xu, S. Zhao, Z. Han, X. Wang, Y.-F. Song, Chem. Eur. J. 17 (2011) 10365.
51. C-H. Joh, B-I. Lee, J-H. You, J-K. Kang, S-H Byeon, Chem. Mater. 23(17) (2011) 4083.
52. K-H. Lee, B-I. Lee, J-H. You, S-H. Byeon, Chem. Commun. 46 (2010) 1461.
53. B-I. Lee, S-H. Byeon, Chem. Commun. 47 (2011) 409.
M
54. H. Jeong, B-I. Lee, S-H. Byeon, Dalton Trans. 41 (2012) 14055.
55. J. Wang, Y. Tian, R. Wang, A. Clearfield, Chem. Mater. 4 (1992) 1276; E. Narita, P. D.
Kaviratna, T. J. Pinnavaia, Chem. Commun. 1 (1993) 60; Z. Wang, T. J. Pinnavaia, Chem.
Mater. 10 (1998) 1820.
56. T. Tatsumi, K. Yamamoto, H. Tajima, H. O. Tominaga, Chem. Lett. (1992) , 815.
d

57. E. Gardner, T. J. Pinnavaia, Appl. Catal., A 167 (1998) 65.


58. Y. Watanabe, K. Yamamoto, T. Tatsumi, J. Mol. Catal. A: Chem. 145 (1999) 281.
te

59. X. Wei, Y. Fu, L. Xu, F. Li, B. Bi, X. Liu, J. Solid State Chem. 181 (2008) 1292.
60. J. Twu, P. K. Dutta, J. Phys. Chem. 93 (1989) 7863.
61. F. M. P. R. Van Laar, D. E. de Vos, F. Pierard, A. K. de Mesmaeker, L. Fiermans, P. A.
Jacobs, J. Catal. 197 (2001) 139.
p

62. T. J. Pinnavaia, M. Chibwe, V. R. L. Constantino, S. K. Yun, Appl. Clay Sci. 10 (1995)


117.
ce

63. D. Carriazo, S. Lima, C. Martın, M. Pillinger, A. A. Valente, V. Rives, J. Phys. Chem.


Solids 68 (2007) 1872.
64. A. A. Frimer, “Singlet Oxygen” Vols. I–IV. CRC Press, Boca Raton, FL, 1985.
65. R. A. Sheldon, J. K. Kochi, Metal Catalyzed Oxidations of Organic Compounds, Academic
Ac

Press, San Diego, 1981.


66. H. Mei, B. W. Mei, T. F. Yen, Fuel 82 (2003) 405.
67. F. M. Collins, A. R. Lucy, C. J. Sharp, Mol. Catal. A 117 (1997) 397; W. Gore, S. Bonde, G.
E., Dolbear, E. R. Skov, US Patent, 20.020.035.306, 2002; F. Figueras, J. Palomeque, S.
Loridant, C. Feche, N. Essayem, G. Gelbard, J. Catal. 226 (2004) 25.
68. J. Palomeque, J. M. Clacens, F. Figueras, J. Catal. 211 (2002) 103; Y. Shiraishi, T. Naito, T.
Hirai, Ind. Eng. Chem. Res. 42 (2003) 6034; J. Palomeque, F. Figueras, G. Gelbard, Appl.
Catal. A 300 (2006) 100; P. Moreau, M. Ziolek, Catal. Today 90 (2004) 145.
69. A. Maciuca, C. Ciocan, E. Dumitriu, F. Fajula, V. Hulea, Catal. Today 138 (2008) 33.
70. A. Maciuca, E. Dumitriu, F. Fajula, V. Hulea, Appl. Catal. Gen. A 338 (2008) 1.
71. H. G. Franck, J. W. Stadelhofer, Industrial Aromatic Chemistry, Springer,
Berlin/Heidelberg/New York, 1988; K. Weissermel, H-J. Arpe, Industrial Organic
Chemistry, 3rd Ed., VCH, Weinheim, 1997; S. Lee, H. Kim, J. Chem. Eng. Data 43 (1998)
358.
72. N. N. Trukhan, A. Y. Derevyankin, A. N. Shmakov, E. A. Paukshtis, O. A. Kholdeeva, V.

26
Page 26 of 28
Edited Aug 29

N. Romannikov, Micropor. Mesopor. Mater. 44-45 (2001) 603; J. Palomeque, J. M. Clacens


and F. Figueras, J. Catal. 211 (2002) 103.
73. E. H. Lee, Catal. Rev. Eng. Sci. 8 (1973) 285.
74. W. T. Reichle, Chem. Tech., 1986, 58; W. Organowski, J. Hanuza, L. Kepinski, Appl.
Catal. A, 171, (1998), 145; Y. Sakurai, T. Suzaki, N. Ikenaga, T. Suzuki, Appl. Catal. A,
192, (2000), 281; Y. Sakurai, T. Suzaki, K. Nakagawa, N. Ikenaga, H. Aota, T. Suzuki, J.
Catal. 209 (2002) 16; V. P. Vislovskiy, J. S. Chang, M. S. Park, E. S. Park, Catal. Commun.
3 (2002) 227
75. M. Sugino, H. Shimada, T. Turuda, H. Miura, N. Ikenaga, T. Suzuki, Appl. Catal. A 121

t
(1995) 125; M. Nimura, M. Saito, Catal. Lett. 58 (1999) 59.

ip
76. E. A. Mamedov, V. C. Corberan, Appl. Catal. A 127 (1995) 1.
77. G. Carja, R. Nakamura, T. Aida, H. Niyama, J. Catal. 218 (2003) 104.
78. M. L. Labajos, V. Rives, P. Malet, M. A. Centeno, M. A. Ulibarri, Inorg. Chem. 35 (1996)

cr
1154.
79. J. Das, K. M. Parida, J. Mol. Catal. A: Chem. 264 (2007) 248.
80. S. R. Kirumakki, N. Nagaraju, S. Narayanan, Appl. Catal. A: Gen. 273 (2004) 1.
81. B. R. Jeremy, A. Pandurangan, J. Mol. Catal., A: Chem. 237 (2005) 146.

us
82. W. D. Bossaert, D. E. De Vos, W. M. Van Rhijn, J. Bullen, P. J. Grobert, P. A. Jacobs, J.
Catal. 182 (1999) 156.
83. A. K. Chakraborty, A. Basak, V. J. Grover, J. Org. Chem. 64 (1999) 8014.
84. G. Roman, E. Comanita, B. Comanita, Tetrahedron 58 (2002) 1617; X. Xu, T. Henninger,

an
D. Abbanat, K. Bush, B. Foleno, J. Hilliard, M. Macielag, Bioorg. Med. Chem. Lett. 15
(2005) 883.
85. G. Goto, K. Kawakita, T. Okutani, T. Miki, Chem. Pharm. Bull. 34 (1986) 3202; N. Jain, A.
Kumar, S. M. S. Chauhan, Tetrahedron Lett. 46 (2005) 2599; Y. X. Liu, B. L. Cai, Y. H. Li,
M
H. B. Song, R. Q. Huang, Q. M. Wang, J. Agric. Food Chem. 55 (2007) 3011; J. Mokhtari,
M. R. Naimi-Jamal, H. Hamzeali, M. G. Dekamin, G. Kaupp, ChemSusChem. 2 (2009)
248.
86. T. Sooknoi, V. Chitranuwatkul, J. Mol. Catal. A: Chem. 236 (2005) 220; F. Song, Y. M.
Liu, H. H. Wu, M. Y. He, P. Wu, T. Tatsumi, J. Catal. 237 (2006) 359; S. Mukherjee, M.
d

Nandi, K. Sarkar, A. Bhaumik, J. Mol. Catal. A: Chem. 301 (2009) 114; Z. H. Li, R. Z.
Chen, W. H. Xing, W. Q. Jin, N. P. Xu, Ind. Eng. Chem. Res. 49 (2010) 6309.
te

87. D. Sloboda-Rozner, R. Neumann, Green Chem. 8 (2006) 679.


88. P. Liu, H. Wang, Z. Feng, P. Ying, C. Li, J. Catal. 256 (2008) 345.
89. S. Zhao, J. Xu, M. Wei, Y-F. Song, Green Chem. 13 (2011) 384.
90. S. Zhao, L. Liu, Y-F. Song, Dalton Trans. 41 (2012) 9855.
p

91. R. D. Peacock, T. J. R. Weakley, J. Chem. Soc. A (1971) 1836.


92. J. Iball, J. N. Low, T. J. R. Weakley, J. Chem. Soc., Dalton Trans. (1974) 2021.
ce

93. M. Sugeta, T. Yamase, Bull. Chem. Soc. Jpn. 66 (1993) 444.


94. T. Yamase in: K. A. Gschneidner Jr., J. C. Bünzli, and V. Pescharsky (Eds), Handbook on
the Physics and Chemistry of Rare Earths, Elsevier Amsterdam, 2009, pp. 243.
95. K. Binnemans, Chem. Rev. 109 (2009) 4283.
Ac

96. F. L. Sousa, M. Pillinger, R. A. Sa Ferreira, C. M. Granadeiro, A. M. V. Cavaleiro, J.


Rocha, L. D. Carlos, T. Trindade, H. I. S. Nogueira, Eur. J. Inorg. Chem. (2006) 726.
97. Z. Han, Y. Guo, R. Tsunashima, Y-F. Song, Eur. J. Inorg. Chem. (2013) 1475.
98. J. A. Capobianco, P. P. Proulx, M. Bettinelli, F. Negrisolo, Phys. Rev. B: Condens. Matter
42 (1990) 5936.
99. S. Shigeo, M. William, Phosphor Handbook, CRC Press, Washington, DC, 1998.
100. J. J. Hastings and O. W. J. Howarth, J. Chem. Soc. Dalton Trans. (1992) 209.
101. C. Hu, B. Yue, T. Yamase, Appl. Catal. A 194/195 (2000) 99.
102. M. Rafatullah, O. Sulaimana, R. Hashima, A. Ahmad, J. Hazard. Mater. 177 (2010) 70.
103. R. Marangoni, A. Mikowski, F. Wypych, J. Colloid. Interf. Sci. 351 (2010) 384.
104. Y. S. Ho, G. McKay, Process Biochem. 34 (1999) 451.
105. M. Bouraada, M. Lafjah, M. S. Ouali, L. C. de Menorval, J. Hazard. Mater. 153 (2008) 911.

27
Page 27 of 28
Edited Aug 29

Highlights:

1. The POMs and LDHs are two types of important functional materials.

2. The POMs/LDHs nanocomposites have experienced tremendous development with some


new synthetic methods and applications.

t
3. This review summarizes the latest advances on the POMs/LDHs nanocomposites

ip
106.

cr
us
an
M
d
p te
ce
Ac

28
Page 28 of 28

You might also like