Sana 1f
Sana 1f
Contents
Abstract. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2. Systems hosting single-metal anionic complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.1. Layered double hydroxides intercalated with halocomplexes. . . . . . . . . . . . . . . . 64
2.2. Layered double hydroxides intercalated with cyanocomplexes . . . . . . . . . . . . . . . 66
2.3. Layered double hydroxides intercalated with oxocomplexes . . . . . . . . . . . . . . . . 73
2.4. Layered double hydroxides intercalated with macrocyclic
ligand-containing complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3. Systems hosting oxometalates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.1. Layered double hydroxides intercalated with low-nuclearity oxometalates . . . . . . . . 87
3.2. Layered double hydroxides intercalated with medium-nuclearity oxometalates:
vanadates and molybdates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.2.1. Vanadates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.2.2. Thermal decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.2.3. Molybdates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.3. Layered double hydroxides intercalated with high-nuclearity oxometalates:
iso and hetero-polyoxometalates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4. Miscellaneous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
0010-8545/99/$ - see front matter © 1999 Elsevier Science S.A. All rights reserved.
PII: S 0 0 1 0 - 8 5 4 5 ( 9 8 ) 0 0 2 1 6 - 1
62 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Abstract
This paper reviews the synthesis, properties and applications of layered double hydroxides
(LDHs), also known as anionic clays or hydrotalcite-like materials, containing intercalated
anions constituted by metal complexes or oxometalates. After an introduction describing the
main features of these compounds, emphasis is put on the synthesis methods, characteriza-
tion and applications. © 1999 Elsevier Science S.A. All rights reserved.
Keywords: Anionic clays; Hydrotalcite-like materials; Layered double hydroxides; LDH; Oxometalates;
Anionic compounds of metal ions
1. Introduction
Layered double hydroxides (LDHs), also known as anionic clays, are a family of
compounds which are deserving much attention in recent years ([1–3] and refer-
ences therein). The structure of most of them corresponds to that of hydrotalcite,
a natural magnesium – aluminum hydroxycarbonate, discovered in Sweden around
1842, which occurs in nature in foliated and contorned plates and/or fibrous
masses. Its formula is Mg6Al2(OH)16CO3 · 4H2O, although due to the relationship
between its structure and that of brucite, Mg(OH)2, it is usually formulated as
[Mg0.75Al0.25(OH)2](CO3)0.125 · 0.5H2O. Brucite shows the well-known CdI2-type
structure, i.e. an hexagonal close-packing of hydroxyl ions, with all octahedral sites
every two interlayers occupied by Mg2 + ions. Partial Mg2 + /Al3 + substitution gives
rise to positively charged layers, thus leading to location of anions in the unoccu-
pied interlayers. In natural hydrotalcite these interlayer anions are carbonate, and
water molecules also exist in the interlayer space. Stacking of the layers can be
accomplished in two ways, leading to two polytypes with a rombohedral (3R
symmetry) or an hexagonal cell (2H symmetry); hydrotalcite corresponds to sym-
metry 3R, Fig. 1, while the 2H analogous is known as manasseite [2]. On the other
hand, the electric charge of the layers and the interlayer ions is just the opposite of
that found in silicate clays (cationic clays), and for all these reasons, these materials
are usually known as layered double hydroxides (LDH), anionic clays or hydrotal-
cite-like materials.
As with cationic clays, the interlayer anions are easily exchanged, and carbonate
has been exchanged for many different anions [4–9], including even hydroxyl
groups (meixnerite). The nature of the layer cations can be also changed, and,
although most of the studies reported in the literature refer to systems with
M2 + /M3 + cations in the layers, other are known with M2 + /M4 + [10,11] or the
rather well studied Li + /Al3 + system [12]. The value of the M3 + /M2 + ratio is
limited if pure materials are desired, and such a ratio, in addition to being
important, also determines the concentration of interlayer anions.
Also as cationic clays, hydrotalcites can be pillared with polynuclear anions
[1,13–17], although the thermal stability of anionic clays is markedly lower than
that of cationic clays. Its thermal decomposition leads to mixed oxides.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 63
Fig. 1. Idealized structure of a layered double hydroxide, with interlayer carbonate anions. Several
parameters are defined.
64 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
a change to a violet color, but the UV–vis bands expected for tetrahedral
[Co(NH3)4]2 + or octahedral [Co(NH3)6]2 + species did not appear; instead, all the
bands shifted to the high energy side of the spectrum. Gentle evacuation at room
temperature (r.t.) restored the original spectrum, indicating that the interaction
between ammonia and the Co(II) ions is rather weak. A similar experiment with
pyridine (a stronger field ligand than ammonia in the spectrochemical series) gave
rise to no color change, even after 1 week exposure at r.t. These results suggest
pyridine does not reach the complex, which should be located in the interlayer
space, and not simply adsorbed on the external surface of the particles.
UV–vis spectroscopy shows further evidence of reactivity in the interlayer space.
In the solid state, the UV – vis/DR (Diffuse Reflectance) spectrum of the LDH–
[NiCl4] compound prepared in ethanol shows, in addition to a doublet at 665 and
706 nm characteristic of [NiEt4]2[NiCl4] salt and of [NiCl4]2 − in solution, two bands
at 550 and 750 nm that have been ascribed by these authors [21] to the presence of
three-coordinated planar [NiCl3] − species [25], while the presence of hexacoordi-
nated Ni species, similar to those existing in NiCl2 [26], can be concluded from the
presence of a band at 450 nm.
In the interlayer space, chloride ligands in the coordination shell of Ni(II) ions
can be reversibly exchanged by water molecules, as concluded from UV–vis/DR
measurements. Outgassing the LDH – [NiCl4] compound under reduced pressure up
to 150–250°C leads to development of bands characteristic of [NiCl4]2 − and
[NiCl3] − , and the color of the sample changes from pale green to pale violet. These
changes can be related to stripping of ligand molecules (residual water or solvent
molecules) from hexacoordinated species NiCl4L2 (L=ligand). Exposure of the
outgassed solid to water vapor at r.t. results in a pale green solid, which spectrum
shows a broad band at 700 nm and a weak band at 390 nm, which may be ascribed
to [Ni(H2O)6]2 + ions. Reevacuation at 200°C for 2 h results in a spectrum similar
to that recorded before water adsorption, the pale-violet color being restored.
The green – blue colored Co-complex shows a somewhat different behavior. Its
UV–vis/DR spectrum shows the characteristic split band at 610 and 658 nm,
typical of tetrahedral Co(II) species, as in the precursor salt [27]; in addition, a
broad feature at 500 nm suggests the presence of Co(II) species with a higher
coordination number, probably [Co(H2O)6]2 + . Outgassing leads to a change in the
color of the sample to turquoise blue, removal of the band at 500 nm, and shift of
the split band to 570 and 608 nm, while a sharp band at 756 nm, with shoulders at
740 and 715 nm, develops. The anomalous position of the sharp band at 756 nm
suggests, from comparison with the spectra of halo-cobalt(II) complexes (halide=
Cl − , Br − , I − ) and other compounds, a decrease in the field strength around the
Co(II) cation in evacuated LDH – [CoCl4]. Exposure to water vapor removes the
756 nm band, suggesting formation of both tetrahedral and octahedral Co species,
coordinated by chloride and aquo ligands, contrary to the results with the Ni
analogous, where only octahedral species existed in the presence of water vapor.
Tetrachloronickelate(II) has been also introduced in the interlayer space of a
Li,Al–LDH by anionic exchange from the nitrate form [28]. Again the interlayer
space calculated (2.9 Å) from the spacing (7.7 Å) was too small for tetrahedral
66 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
species. However, from EXAFS measurements, the first neighbor Ni–Cl distances
were 2.10 Å, compatible with 4-fold coordination. In addition, no EPR signal was
recorded for the LDH – [NiCl4] compound, despite a strong signal was reported
both for octahedral (NiCl2) and tetrahedral ([Et4N]2[NiCl4] in nitromethane) spe-
cies, thus suggesting that a change in the geometry has occurred in the LDH
interlayer, forming square planar [NiCl4]2 − species which size, if the C4 axis is
perpendicular to the brucite-like layers, is compatible with the measured interlayer
space.
The LDH – [NiCl4] compound has been tested for chloride/bromide exchange in
butyl bromide in toluene suspension [29]. The exchange rate has been found to be
slightly lower than that for LDH in the chloride form. This rate sharply increases
with the reaction temperature; thus, only 18% exchange was achieved at 50°C after
150 min, while at 100°C an exchange degree of 84% was reached; the reaction is
inhibited in n-butanol and ethanol [30]. The exchanged bromide anion enters the
coordination shell of interlayer Ni(II) ions, as concluded from UV–vis/DR results.
This same compound, LDH[NiCl4], also catalyzes the halide exchange reaction
between benzyl chloride and butyl bromide [29,30], but only in DMF, and not in
toluene.
Anion-exchanged hydrotalcite-like clay-modified electrodes containing IrCl26 − in
the interlayer have been reported by Itaya et al. [31]. These were prepared by anion
exchange of the LDH in the chloride form (carbonate was not exchanged) by
shaking the electrodes (films of LDH –Cl on SnO2) with a 20 mM solution of the
iridium complex for ca. 1 h, leading to a solid with a basal spacing of 10.8 Å,
corresponding to an interlayer space of 6.03 Å. Steady voltammgrams recorded
after an initial potential of 0 V (vs. SCE) for 10 s indicate that all [IrCl6]2 − ions
incorporated into the film were reduced to [IrCl6]3 − at 0 V.
Fig. 2. X-ray powder diffractions and IR spectra of [Fe(CN)6]4 − – LDH. R-values correspond to molar
percentage of aluminum. Reprinted from S. Kikkawa, M. Koizumi, Ferrocyanide anion bearing Mg,Al
hydroxide, Mater. Res. Bull. 17 (1982) 191–198, © 1982, with permission from Elsevier Science.
Fig. 3. Calculated end-to-end distances in a hexacyanoferrate ion octahedron. Adapted from Ref. [34].
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 69
the porosity properties [50], with a value of 499 m2 g − 1 for a M2 + /M3 + ratio of
3.33.
The change in the capacity for carbon dioxide adsorption on a Mg,Al-
[Fe(CN)6]4 − LDH with the Mg/Al ratio (between 1 and 7) has been studied by
Mao et al. [51]. Maximum adsorption was observed for Mg:Al= 1.7, i.e. it depends
not only on the width of the interlayer space (ranging, although not steadily, from
10.64 to 10.96 Å), but also on the layer charge. The isosteric heat of adsorption was
calculated to be 43.3 kJ mol − 1, a value similar to that reported by Miyata and
Hirose [33], who found the adsorption capacity for CO2 of a Mg,Al–[Fe(CN)6]4 −
LDH to be 60% of that of zeolite 5A [52].
The interlayer space of LDHs provides a reaction medium for chemical reactions.
However, its utility is limited because of the interlayer space and the size of the
reagents. So, reactions have been studied on LDHs previously expanded with large
anions, such as hexacyanoferrates or terephthalate. Challier and Slade [53] have
reported the oxidative (due to Cu2 + ) polymerization of aniline between the layers
of a Cu,Al – [Fe(CN)6]4 − LDH.
Metalocyano-containing LDHs have been also used to prepare modified elec-
trodes, as the redox behavior of the interlayer anions makes them electrochemically
active. Itaya et al. [31] have reported on LDHs containing [Mo(CN)8]4 − or
[Fe(CN)6]3 − prepared by anionic exchange of a commercial Mg,Al–Cl precursor.
A film (thickness ca. 100 nm) of the chloride hydrotalcite was prepared on SnO2,
and exchange was achieved by shaking the electrodes in dilute (20 mM) solutions of
the metalocyano complexes. Interlayer spacings of 11.2 Å were measured, in
agreement with previous results described above. The peak currents for the
[Mo(CN)8]4 − /3 − couple incorporated in the interlayer are clearly observed at ca.
0.5 V (vs. SCE), which is almost in complete agreement with the half-wave potential
of this couple in the same solution. In addition, a steady voltammogram was
obtained after 10 – 20 cycles, without further decrease after 60 min cycling, indicat-
ing the anion is strongly held in the interlayer space of the LDH. Similar results
were obtained for [Fe(CN)6]3 − .
Hexacyanoferrate(III) was used by Shaw et al. [54] as an electroactive probe to
determine the availability of Zn2 + and Al3 + at the surface of a Zn,Al–Cl LDH
applied at the surface of a glassy carbon electrode. When [Fe(CN)6]3 − is reduced at
the electrode surface to [Fe(CN)6]4 − , this becomes anchored forming a multilayered
Prussian blue-like film, the nitrogen atom binding to Zn2 + and Al3 + ions from the
layers (as concluded from XPS measurements for the N(1s) energy level), so
demonstrating such an availability. Phenol was not electro-oxidized on this elec-
trode, while such an oxidation was observed on the Mg,Al analogue, where the
Prussian blue-like layer was not formed. From these results, and taking into
account that oxidation of phenol is pH-dependent, these authors were able to
assign effective pH values of \11.2 and 8.3, respectively, for the surfaces of
Mg,Al–Cl and Zn,Al – Cl.
Hexacyanoferrate(II) modified carbon paste electrodes have been also studied by
Labuda and Hudáková [55], who observe oxidation of ascorbic acid at a potential
0.4 V less positive than the oxidation potential at a bare carbon paste electrode, at
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 73
pH\ 4. According to these authors, this electrode is more sensitive and more
selective than carbon paste electrodes modified with organic salts and that hexa-
cyanoferrate(II) bound to poly(4-vinylpyridine), and permits good storage and
operational stability.
Qiu and Villemure [56] have found an enhanced reduction current when
[Fe(CN)6]4 − or [Mo(CN)8]4 − are exchanged in Ni,Al or Ni,Fe–LDH-modified
electrodes. These authors found that Ni2 + in these LDHs can be oxidized up to
4 + if sufficiently positive potentials are used, but they can be reduced only to
2.7+ ; reversible reduction to 2+ is attained only up to a maximum oxidation
value of 3.6+, the layered structure being stable, at least, as far as X-ray
diffraction diagram concerns. The enhancement in the cathodic current has been
attributed to electron transfer between the intercalated anions and the oxidized Ni
sites in the brucite-like layer, as no enhancement was observed when the oxidized
forms of the cyanocomplexes were intercalated. Similar results were obtained in
systems containing Mn or Co in the brucite-like layers, the voltammetric waves
being smaller, however, if the interlayer anions were carbonate or chloride, instead
of [Fe(CN)6]4 − , [Mo(CN)8]4 − or [Ru(CN)6]4 − . When the potentials of Ni, Co or
Mn LDH modified electrodes were scanned in mixtures of the iron and ruthenium
or molybdenum cyanocomplexes, electron transfer from [Fe(CN)6]4 − to electro-
chemically oxidized [Ru(CN)6]3 − or [Mo(CN)8]3 − , mediated by the LDH electroac-
tive metal centres, was observed, but such a transfer was not observed for the redox
inactive Zn,Al – Cl LDH.
intercalated in the Zn,Al – LDH. The increase in the number of terminal oxygen
atoms bonded to the MoVI ion, as concluded from XANES studies [60], follows
decoordination of the carboxylate ligands, this being the first step which induces
sufficient positive charge on the MoVI site to allow an oxygen atom transfer from
a contiguous water molecule without any change in the oxidation state of Mo.
Mg,Al– LDH has been also intercalated with peroxomolybdate(VI) anions con-
taining tartrate as ligand [64], [Mo2O2(O2)4(C4H2O6)]4 − . Intercalation was achieved
after swelling the LDH with terephthalate, a method widely used to intercalate
large anions [65], leading to samples with spacings of 12.4 and 14.3 Å, respectively,
for the terephthalate and peroxomolybdate-containing compounds. When sus-
pended in water, this compound releases the peroxo group at 80°C to yield
oxomolybdate, the process being reversible by addition of hydrogen peroxide;
oxygen is released when heated in the solid state at ca. 120°C.
Pinnavaia et al. [66] have studied the effect of intercalation in different hosts on
the properties of the excited-state of dioxorhenium(V) ions. From previous studies
with polypyridyl complexes, it was concluded that the excited-state properties are
preserved upon intercalation [67 – 70]. However, in the case of metal–oxo com-
pounds and, particularly, d2 trans-dioxo species [71], it was expected that its
location in the interlayer space of LDHs and layered silicate clays would probably
modify the orientation of the trans-ReO2+ core, thus affecting luminiscence proper-
ties. A Mg,Al – ReO2(CN)4 LDH was prepared by coprecipitation of the hydrotal-
cite in the presence of the complex anion, while intercalates in layered silicate clays
were prepared with [ReO2py4] + and [ReO2en2] + (py= pyridine; en=ethylendi-
amine) in hectorite and fluorhectorite. In all three cases, exchange corresponded to
ca. 15% of the exchange capacity of the clay.
Location of these complexes in the interlayer space of the clays gives rise only to
minor structural distortions, as revealed by IR and Raman (in the resonance
Raman effect mode) spectroscopies. Basal spacings (from XRD studies) are en-
larged, consistenly with incorporation of the complexes. The luminescent hectorite
intercalate is largely unperturbed, emission from the fluorhectorite intercalate is
significantly attenuated, and no luminiscence was observed from the ReO2 –LDH
intercalate. From basal spacing measurements and calculations of charge density of
the layers, these differences have been related by these authors [66] to the different
possibilities of orientation of the ReO2 complexes in the interlayer space, Fig. 5. So,
in the case of the hectorite intercalate, the complex should be located with its
O–Re–O axis perpendicular to the layers, and the oxygen atoms can then be
‘keyed’ in the hexagonal cavities of the layered silicate clay; for the fluorhectorite
intercalate, the higher charge density (27 vs. 80 Å2 per charge unit for hectorite)
precludes such an orientation, but data are consistent with the complex oriented
with its C4 axis parallel to the layers. Finally, for the LDH intercalate, the
d-spacing is consistent with the effective C3 axis of the pesudooctahedral
ReO2(CN)34 − complex perpendicular to the layers, in agreement with previous
studies which have demonstrated that the preferred orientation of intercalated
anions in LDHs either maximizes the hydrogen bonding interactions of the protons
of the brucite-like layers with the guest species, and/or minimizes the charge
76 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
separation distance between the positive layers and the gallery anions [33,72,73].
Contrary to the structure of hectorite, in LDHs there are no cavities in which the
ReO2+ core can key, and oxygen atoms of this core are probably hydrogen-bonded
directly to the hydroxide layer. Because proton-donating solvent effects efficiently
quench ReO2+ excited states by hydrogen bonding interactions [74], the non-radi-
active decay rates of electronically-excited ReO2(CN)34 − ions in LDH are expected
to be exceedingly fast.
One of systems most widely studied in recent years concerning LDHs, corre-
sponds to those where the interlayer anion is a coordination compound with
macrocyclic ligands; also, where the interlayer anion is the anionic ligand itself.
These nanocomposite materials prepared with intercalated metalloporphyrines and
Fig. 5. Proposed orientations of the trans-dioxorhenium(V) core in (a) hectorite, (b) fluorhectorite, and
(c) a layered double hydroxide. Adapted from Ref. [66].
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 77
Zn/Al ratio was increased. From the basal spacings and the size of the intercalated
anions [90], these authors conclude [88,89] that the anion (pTCPP and pTSPP)
should be located in a perpendicular orientation relatively to the layers, with the
four anionic groups in tight interactions with the hydroxylated sheets, in a fashion
similar to that above described for the Mg,Al–TSPP LDH [86] material. Such an
arrangement would also be consistent with the layer charge density, that corre-
sponds to 33 Å2 per unit charge for the sample with a Zn/Al ratio of 3; an
arrangement of the interlayer anion in horizontal position would correspond to 49
Å2 per unit charge, while in the vertical position it corresponds to 15 Å2 per unit
charge. On the contrary, although the basal spacing measured for the LDH–
oTCPP compound (18.5 Å) could be fitted also with a vertical disposition (this
molecule is ca. 3.2 Å smaller than the other two), such an arrangement would
decrease the hydrogen bonding interaction between the carboxylate groups and the
hydroxyl anions in the brucite-like layers. Consequently, a parallel disposition of
two molecules has been claimed in this case [89].
The photochemical properties of this sort of macromolecules can be modified
when located in the interlayer space of an LDH. Tagaya et al. [91] have studied the
intercalation of colored organic anions (pTSPP and pTCPP) in Mg,Al and Zn,Al
LDHs by the reconstruction method. As in the cases above described [86,88,89], the
basal space increase is consistent with a vertical orientation of the anions (spacings
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 79
up to 30.7 Å, depending on the nature of the porphyrin), although the guest/host
ratio varied from 48 to 100%. The absorption maxima of these colored organic
anions are solvent-dependent, although no clear correlation exists between the
wavelengths of the absorption maxima and the dielectric constant of the solvents
[91]. When intercalated, a shift towards the red was observed for both anions,
which has been explained on the basis of a close packing of anion molecules in the
interlayer space, an effect similar to that previously reported for methylene blue
adsorbed on a clay [92].
Incorporation of metal cations coordinated by macroligands in the interlayer
space of LDHs has also deserved very much attention. In the case of the
Zn,Al–pTCPP system, complexation of copper in the interlayer nanospace was
attained simply by suspending the organic-LDH solid in a solution of copper
nitrate [88], and metallation was confirmed by UV–vis and EXAFS spectra. In the
UV–vis region, metallation leads to a slight red-shifting of the Soret band recorded
at 400–450 nm for the free anion, while only two Q bands are recorded between
500 and 650 nm, and these results were observed for the Zn,Al–pTCPP–Cu(II)
system. From EXAFS spectra, radial distributions around the Cu(II) were similar
for Cu(NH3)4SO4 · H2O (with four equatorial nitrogen atoms and two oxygen atoms
from water molecules in trans geometry) and for Cu(II)–pTCPP, with four nitrogen
atoms in a planar environment around the copper cation at ca. 2.10 Å and Cu–Cu
interactions along the direction perpendicular to the chelating ligand. For the
Zn,Al–pTCPP – Cu(II) system, the radial distribution reveals a nearly regular
octahedral environment, probably due to completion of the copper coordination
sphere by two water molecules, thus further confirming the perpendicular arrange-
ment of the porphyrin pillars in the interlayer space, as well as the absence of
Cu–Cu interactions.
Intercalation of tetracarboxyphthalocyanine cobalt (II), TPC–C, in Zn,Al and
Mg,Al–LDHs was carried out by the reconstruction method by Tagaya et al. [91].
The guest/host ratio was 100% for the Zn,Al system, with a basal spacing of 24.8
Å, i.e. the plane of the guest anion perpendicular to the plane of the host layers.
However, for the Mg,Al system the basal spacing was 7.9 Å. Although these
authors claim this value being consistent with a small amount (not quantified) of
intercalated TPC – C, this basal spacing coincides with the value reported in the
literature for carbonate containing Mg,Al–LDHs [3], and so it is possible that no
intercalation of the macrocyclic was attained.
Complexes with macrocyclic ligands can be also incorporated in the interlayer
space of LDHs via anionic exchange. Dutta and Puri [39] have reported complete
ion exchange of nickel(II) phtalocyaninetetrasulfonate ion in the Al2Li–LDH in the
chloride form. Quite surprisingly, the basal spacing was 10.61 Å, a value quite close
to that reported for phosphate and sulphate-containing LDHs, indicating that the
phtalocyanines are parallel (‘flat’) to the aluminate layer and not arranged in a
stacked fashion, which would require a spacing close to 22 Å.
Cobalt(II) phtalocyanines, specifically Co(II) phtalocyanine-3,4%,4%%,4%%%-tetrasul-
fonate (hereafter CoPcTs) have been also introduced in the galleries of Mg,Al
LDHs by reconstruction of the hydrotalcite precursor calcined at 500°C and
80 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
exposed to excess aqueous Pc salt at 60°C [76] or hydrothermally at 100°C [93], the
solid displaying a basal spacing of 23.3–23.7 Å, Fig. 7. This is in agreement with
a tilted orientation with respect to the hydroxide layers, with the axis joining the
non-coordinating nitrogen atoms of the CoPcTs molecule oriented almost perpen-
dicular to the hydroxide layers. These Co(II)–PcTs-containing LDHs have been
tested as catalysts for autoxidation of 1-decanethiol [76] and 2,6-di-tert-butylphenol
[93,94] by Pinnavaia et al., who have concluded that the catalyst becomes extremely
stabilized in the gallery space of the LDH, if compared to its stability under
homogeneous catalysis conditions; deactivation of the LDH-supported catalysts
during 1-decanethiol oxidation was not observed even after five catalytic cycles for
a total of more than 770 turnovers, while the homogeneous catalyst was deactivated
after 25 turnovers; such a stabilization was even greater during 2,6-di-tert-butylphe-
nol oxidation (3200 vs. 25 turnovers).
Carrado et al. [95] have synthetized Mg,Al LDHs intercalated with CuPcTs by
hydrolysis of mixed aqueous salt solutions in the presence of NaOH, a method
previously proposed by Park et al. [96], who first reported the direct synthesis of
organic dyes into LDHs. Anionic exchange was achieved only when starting from
a freshly prepared slurry (‘wet’ anionic exchange) of the LDHs (in the nitrate form),
exchanged under carbon dioxide-free conditions, and an aqueous CuPcTs solution;
however, no satisfactory products were obtained when the exchange was performed
with dry LDH (also in the nitrate form) redispersed in water. Basal spacing for
Fig. 7. X-ray diffraction pattern (Cu–Ka ) of an oriented film sample of Mg,Al – [Co(PcTs)] LDH.
Reprinted from M. Chibwe, T.J. Pinnavaia, Stabilization of a cobalt(II) phtalocyanine oxidation catalyst
by intercalation in a layered double hydroxide host, J. Chem. Soc. Chem. Commun. (1993) 278 – 280, ©
1993, with permission from The Royal Society of Chemistry.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 81
both samples (prepared by direct synthesis and by anionic exchange) was 22.5–23.0
Å, but crystallinity was better for the sample prepared by wet anionic exchange.
This spacing is markedly larger than that obtained (14–16 Å) for hectorite
interlayered with cationic copper(II)-containing dyes, such as alcian blue. The
difference arises from the different layer charge density for these cationic and
anionic clays. The anionic clay prepared by these authors had a layer charge density
close to 34 Å2 charge − 1, a value only reached by micas among the cationic clays.
From elemental chemical analysis and molecular modelling, and taking into ac-
count the layer density charge, Carrado et al. have concluded [95] that the
phtalocyanine molecules are oriented in a tilted arrangement, in agreement with
similar previous results by Pinnavaia et al. [76,93]. On the contrary, in the case of
the hectorite clays, their lower layer charge density leads to a flat orientation of the
phtalocyanine anions.
In addition to the increased stability (probably because the immobilization
process inhibits the deactivating dimerization and self-oxidation reactions occurring
in the homogeneous catalyst), it should be noted the large reactivity of the
heterogeneous Co(PcTs) – LDH catalyst, despite the extremely low specific surface
area, 28 m2 g − 1, as determined from nitrogen adsorption, i.e. the nitrogen
molecules are merely adsorbed on the external surface of the crystallites, without
accessing the gallery space. So, 2,6-di-tert-butylphenol, with a much larger Van der
Waals radii, does not access either, and thus only the Co(II) ions held at crystallite
edge sites and external basal surface sites are able of participating in the oxidation
reaction. This conclusion is supported by the finding that a similar system, but with
a lower charge layer density (with Mg/Al ratio of 4, instead of 2), and thus with a
greater separation between the intercalated cobalt centers, shows a 5-fold increase
in activity.
This same cobalt phtalocyaninetetrasulphonate has been incorporated into a
Zn,Al LDH by coprecipitation at constant pH from a Zn and Al nitrates solution
and the Co complex. The solid isolated showed an interlamellar spacing of 23.0 Å,
in close agreement with data previously reported [76]. This compound has been
used for in situ studies of cyclohexene oxidation by combined EXAFS/XRD
techniques studies [97].
Similar compounds containing increasing amounts (2–90 mmol g − 1 of LDH
support) of Co(II) phtalocyanine tetracarboxylate or Co(II) phtalocyanine tetrasul-
fonate have been tested by Iliev et al. for 2-mercaptoethanol oxidation [98].
Incorporation of the complexes into the interlayer space was achieved in this case
by soaking the LDH (previously calcined at 450°C for 24 h in argon) with a
solution of the sodium salts of the phtalocyanines, at 60°C for 7 days in argon.
However, while the basal spacing was close to 22.7 Å in both series of samples (in
agreement with the values reported by the authors previously cited), in the case of
the carboxylate phtalocyanine, crystallization of its sodium salt, most likely on the
external surface of the LDH crystallites, was also observed. The ESR spectra of
these samples showed, for low Co and Cu(II) phtalocyanine concentrations (5 mmol
g − 1 of LDH support), a hyperfine splitting from 57Co (I =7/2) and from 63,65Cu
(I= 3/2), together with superhyperfine splitting from 14N (I=1) in the case of the
82 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
calculated [108]) to the clay layers. Finally, the basal spacing was 23.5 Å, and did
not change upon dehydration, for the LDH-hosted CoPcTs, a value similar to that
reported by these and other authors [76,93,95,98], Fig. 8.
While the UV – vis spectra of the hectorite and fluorhectorite systems do not
change sensibly upon exposure to O2 (suggesting that intercalated Co macrocycles
do not form adducts with O2, as observed in solution [109]), small changes in the
positions of the Soret and Q bands for the CoPcTs–LDH compound have been
related to changes in the electron density of the conjugated p orbitals of the
macrocycle ring caused by confinement in the interlamellar space of the LDH, by
interactions between neighboring CoPcTs and by changes in the inductive effect of
the sulfonate groups.
ESR results are consistent with the XRD results, for the CoTMPyP–fluorhector-
ite compound in the hydrated state, if the molecular plane is tilted 27° with respect
to the clay layers, and with one or two water molecules coordinating to the Co(II)
along the z-axis. Upon dehydration, water molecules are removed, and the hosted
84 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
molecules form a bi-layer, with the molecular plane oriented parallel to the clay
layers, a similar disposition as that concluded for CoTMPyP–hectorite, although in
this case only a monolayer of hosted molecules is formed, without coordinating
water molecules. The orientation of CoPcTs in the LDH gallery could not be
concluded from anisotropic ESR measurements, as no oriented films could be
obtained. From XRD measurements, similar conclusions to those by Carrado et al.
[95] were achieved, i.e. a perpendicular arrangement of the complex in the LDH
gallery; similar spacing values could be explained by a non-favorable trilayer
arrangement of the anions parallel to the LDH plane, but in this case the charge
layer density would not be matched, and the spacing would be expected to decrease
upon dehydration, but this was not observed. The ESR spectrum of CoPcTs–LDH
was similar to that of air-dried CoTMPyP–hectorite, so further confirming the
slightly tilted ‘upright’ orientation; however, it changes dramatically upon vacuum
dehydration, with a weak ESR signal at g: 2, probably because aggregation of the
hosted molecules upon water removal; this process is reversible upon rehydration
and water diffusion into the interlayers.
From these results, these authors conclude [105] on the suitability of the
fluorhectorite and LDH compounds, but not the hectorite-hosted one, for catalytic
reactions taking place on the Co(II) sites, accesible to reactant molecules for
electron-transfer reactions, behaving as biomimetic catalysts, as observed for 2,6-di-
tert-butylphenol oxidation [94].
These three host – guest systems have been also tested for reductive dechlorination
of carbon tetrachloride [110], and the results compared with those obtained for the
same Co compounds in homogeneous conditions, and by silica-supported
CoTMPyP. While under homogeneous conditions, both chloroform and
dichloromethane were formed, only the former was observed under heterogeneous
conditions, accounting for less than 30% of degradated CCl4, which is consistent
with previous studies of degradation by Co macrocycles [111–115]. Degradation of
CCl4 follows a first-order kinetics. Under homogeneous conditions, the initial rate
rapidly decreases, indicating deactivation of the catalyst, probably because aggrega-
tion in aqueous solution [116], but as aggregation is hindered in the clays gallery,
activity is maintained. The lack of formation of dimethylmethane has been at-
tributed [110] to a change in the reduction potential of the supported macrocycle
because of the charged layers. Initial rate constants for heterogeneous CCl4
degradation (as calculated along the first 30 min of reaction) decreases when using
the supports silica-gel\LDH \ fluorhectorite\hectorite. This decrease is in agree-
ment with accessibility of the reactant molecules to the Co(II) sites, as concluded
from XRD and ESR studies [94,105] for the layered materials. Overall, the
dechlorination reaction in these heterogeneous systems is very similar to enzyme-
catalyzed reactions, and the initial degradation rate (R0) can be fitted by the
Michaelis-Menten kinetic model:
R0 = nmax[CCl4]0/([CCl4]0 +Km) (6)
where nmax is the maximum reaction rate for a specified initial Co macrocycle
concentration and Km is the Michaelis constant. nmax values decrease for the
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 85
ometalates – intercalated LDHs and their applications have been published [84,122–
124]. In this section, classification has been made according to the nuclearity and
nature of the interlayer anion.
Fig. 9. Scheme of a layered double hydroxide photochemical assembly. Adapted from Ref. [119].
88 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Fig. 10. Effect of heating on interlayer spacings of Cu,Cr – X LDHs (X = Cl − , CrO24 − , Cr2O27 − ).
Reprinted from C. Depège, C. Forano, A. de Roy, J.P. Besse, [Cu – Cr] layered double hydroxides
pillared by CrO24 − and Cr2O24 − , Mol. Cryst. Liq. Cryst. 244 (1994) 161 – 166, © 1994, with permission
from Gordon & Breach Science Publishers SA.
in determining the interlayer space, due to the ability to adsorb water in the
interlayer space, as confirmed by thermogravimetric and elemental chemical analy-
ses. The spacings determined from X-ray diffraction were 8.95 Å and 8.42 Å,
respectively, for the dichromate and chromate compounds, these values being
markedly lower than those reported by Chibwe and Jones [32] for a Mg,Al–
Cr2O27 − LDH. These values, however, are close to that reported by Miyata and
Okada [72] for the chromate form, and so a dedimerization could happen. Both
phases undergo a slow evolution to contracted forms with basal spacings of 7.68 Å
(chromate) and 7.87 Å (dichromate) upon water removal, Fig. 10, and these
extremely low values cannot account for free oxoanions, but evidence an effective
pillaring on the hydroxylated sheets. The difference with the behavior observed with
other chromate-containing LDHs has been attributed [128] to the specific chemical
properties of copper, which can form a wide range of lamellar copper hydroxides
Cu2(OH)3A (A =Cl − , ClO4− , NO3− , MnO4− , etc.) [130] with short interlayer
distances. However, such pillaring is reversible, as evidenced by the ability to
exchange by chloride in the presence of an excess of KCl. A decrease in the basal
spacing to 7.10 and 7.30 Å is observed for the chromate and dichromate forms,
respectively, upon heating at 80°C [128], the decrease is not reversible, and the
anions cannot be further exchanged, indicating that this rather soft treatment has
been able to anchor the anions to the layers. Further heating at 200–300°C destroys
the lamellar structure, as confirmed by X-ray diffraction and XAS measurements
[129].
Similar spontaneous contraction of the layers has been also described for Zn,Al
and Zn,Cr – LDHs intercalated with CrO24 − and/or Cr2O27 − , but only after ageing
90 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Fig. 11. Two ways of grafting of ditetrahedral anions (Cr2O27 − , V2O27 − ) to layered double hydroxides.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 91
3.2.1. Vanadates
Among all metal-containing anions incorporated into the interlayer space of
hydrotalcites and other LDHs, oxovanadates represent the widest studied group.
Although mostly as decavanadate, some studies have been also devoted to interca-
lation of lower oligovanadates.
Oxometalate-pillared LDHs are in many cases prepared by a two-step anion
exchange method, through intermediate intercalation of a large organic anion, to
swell the brucite-like layers [65,134].
The polymerization degree of oxovanadates is pH-dependent, nuclearity increas-
ing as pH decreases. For a 0.1 M aqueous vanadate solution, these equilibria are as
follows:
V10O26(OH)2− 4, V10O27(OH)5 − , V10O628− , decavanadate,
pH =1 −3 l
lVO(OH)3, VO2(OH), V3O39 − , V4O412− , metavanadate,
pH = 4 −6 l
VO3(OH)2 − , HV2O37 − ,V2O47 − , pyrovanade, pH= 8−11l
lVO34 − , vanadate, pH \12.
In one of the pioneering works on LDH-intercalated vanadates, Twu and Dutta
[135] prepared Li,Al – LDH with different oxovanadate oligomers by ion exchange
of the Li,Al – Cl precursor with NH4VO3 aqueous solutions at different pH; it was
expected anionic exchange and intercalation of the oxovanadate more stable at each
pH. However, at pH larger than 13, selective Cl − /OH − exchange, but not
chloride/vanadate, was observed. When the pH was lowered to 8–11, complete
chloride/vanadate exchange was achieved, the basal spacing increasing from 7.8 to
10.8 Å. Although a compound with the same gallery height (6.0 Å, once the
thickness of the brucite-like layer, 4.8 Å, is considered) is obtained after exchange
at pH 5– 6, the Raman spectra of both samples were completely different. Finally,
at pH 3– 4, only partial exchange was achieved, as a vanadate/chloride competition
exists, because of the use of HCl to attain this low pH values. As different interlayer
anions could give rise to the same gallery height (ca. 6 Å), thus making impossible
to ascertain the actual nature of the interlayer anion, this was studied by Raman
spectroscopy, based on results by Griffith et al. [136,137] for oxovanadate species.
The most prominent Raman band of vanadate species, due to V–O stretching,
occurs in the 800 – 1000 cm − 1 range, shifting to higher wavenumbers as polymeriza-
tion and/or branching increases. From this, it was concluded that at pH 10 the
92 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Fig. 12. Two possible orientations of V2O27 − in layered double hydroxides. Adapted from Ref. [135].
interlayer species is V2O47 − (also the major component of the solution at this pH),
with the V – V axis parallel to the layers, Fig. 12(b), and also hosting interlayer
water molecules; at pH 5 – 6 the interlayer species is V4O412− , with a similar gallery
height, while at lower pH the predominant species still is V4O412− , with only a minor
amount of V10O628− , despite the decavanadate can be completely exchanged in other
hydrotalcites (see below).
On the other hand, the average charge per V atom changes from one oxovana-
date to another. Taking into account that the negative charge of the interlayer
anion should balance the positive charge of the Al-containing brucite-like layers,
Bhattacharyya et al. [138] have prepared, through a careful control of pH during
reaction and of the NaVO3/NaOH ratio in the starting solution (from 1:7 to 1:3),
Mg,Al and Mg – Zn,Al – LDHs containing different oxovanadate oligomers by a
single-step method. In all cases, an aqueous solution of the Mg and Al (or Mg, Zn,
and Al) cations was dropwise added to the NaVO3/NaOH solution at the desired
pH (10.8 for V2O47 − and 8.3 for V4O412− ), the gelatinous mixture being heated for
several hours at 80 – 90°C. The basal spacings determined by X-ray diffraction were
10.5 Å for the pyrovanadate, and 9.7 Å for the cyclotetravanadate.
The V2O47 − and HV2O37 − anions are constituted by two [VO4] tetrahedra sharing
a vertex, and in the interlayer space, two different arrangements are possible [139],
one with the V – V edge perpendicular to the brucite-like layers, with a calculated
basal spacing of 12.6 Å (from the size of the V2O47 − anion derived from single
crystal data for Mg2V2O7 [140]), and another where it is parallel to the layers, with
a calculated value of 9.8 Å, Fig. 12, a situation similar to that described by Chibwe
and Jones [32] for intercalated Cr2O27 − ; in addition, the presence of water molecules
can enlarge such a spacing. On the other hand, the smallest dimension of V4O412− is
expected to be the same as V2O47 − [138], and these authors conclude the anions
should be in all cases in a ‘parallel’ disposition.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 93
Similar studies have been carried out by Besse et al. [141] in a Cu,Cr–LDH,
analyzing also the role of swelling agents to favor the exchange and incorporation
of polyoxometalate species. At pH 10 –11, V2O47 − is selectively exchanged, whereas
as the pH is lowered, the major interlayer species are V2O412− (pH 6–7) and V10O628−
(pH 4–5). These authors conclude that the size of the precursor anion used to swell
the layers (terephtalate or dodecylsulphate), and hence the values of the interlayer
distances, must favor the intercalation of vanadates of similar hinderance.
Delmas et al. [142,143] have recently proposed an alternative route, by the
so-called ‘chimie douce’ (soft chemistry), to insert metavanadate in an LDH, Fig.
13. These authors claim the method overcomes the problem, usually found, that
chemical composition of LDH shows fluctuations due to the different pH at which
precipitation of M(OH)2 and M(OH)3 hydroxides occurs, thus leading to a chemical
composition determined by the intrinsic stability of the solid, rather than by the
composition of the starting solution. By high temperature synthesis methods these
authors obtained a layered NaNi1 − y Coy O2 solid, with a layer–layer distance of
5.18 Å; an oxidizing hydrolysis with NaClO and KOH leads to an expansion to
7.08 Å and insertion of potassium ions, followed by reduction in H2O2/NH4VO3
solution. The asymmetry of the X-ray diffraction lines usually indexed as due to
(101) and (111) planes, in some cases attributed to a turbostratic-like character
(parallel and equidistant layers disoriented with regard to one another along the
c-axis), has been attributed by these authors to local distortions within each layer,
probably related to a misfit of the oxygen in the layers (O–O distance 3.04 Å) and
in the metavanadate chains (2.91 Å). The interlayer spacing of the vanadium-con-
Fig. 13. Scheme showing the successive reaction steps involved in the preparation of a LDH by ‘chimie
douce’. Reprinted from K.S. Han, L. Guerlou-Demourgues, C. Delmas, A new metavanadate inserted
layered double hydroxide prepared by ‘chimie douce’, Solid State Ionics 84 (1996) 227 – 238, © 1996, with
permission from Elsevier Science.
94 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Fig. 14. Orientation of decavanadate anion, V10O628− , in the interlayer space of layered double
hydroxides, with its ‘main’ C2 axis parallel to the brucite-like layers.
taining material was 9.15 Å, and from IR data, the presence of polymeric,
metavanadate species, (VO3)nn − (consisting of [VO4] groups with a C26 symmetry)
was concluded, in agreement also with chemical analysis, which suggest one
negative charge per vanadium ion, while the presence of cyclic metavanadate
entities, such as V3O39 − or V4O412− , was excluded.
Pinnavaia et al. have shown [134] that it is possible to introduce polyoxovanadate
ions as pillars in LDHs containing Zn,Al or Zn,Cr or Ni,Al in the layers, in the
chloride form. At pH 5.5 – 10 a byproduct containing V4O412− was formed, but at pH
4.5 total exchange with V10O628− was observed, with solids possessing a basal
spacing of 11.9 Å, corresponding to a gallery height of 7.1 Å and a decavanadate
orientation in which the C2 axis is parallel to the host layers, Fig. 14. Retention of
the structure of the interlayer anions was confirmed by 51V MAS-NMR spec-
troscopy, while EXAFS studies at the Zn K-edge have shown [144] that no
structural distortion occurs for the brucite-like host lattices upon intercalation, in
agreement with electron microscopy studies which show that the exchange reactions
are topotactic. Exchange from the nitrate form has been reported by Woltermann
[145]. Photocatalytic oxidation of isopropyl alcohol to acetone was achieved on the
Zn,Al–V10O28 LDH in the presence of oxygen, this catalyst being more active than
the homogeneous catalyst, despite scattering of light by the host particles [134].
LDHs containing Zn and Al in the layers have been also prepared with interlayer
vanadates, by ionic exchange of chloride or carbonate precursors [146]. It has been
found that the Zn/Al ratio in the final Zn,Al–vanadate LDH decreases when the
pH is lowered, probably due to selective dissolution of Zn2 + . Vanadium K-edge
XAS data show that in samples prepared at low pH the major interlayer species is
V10O628− , while as the pH is increased, interlayer vanadates consist of tetrahedral
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 95
[VO4] units, and it has been even possible to estimate the fraction of V5 + ions
existing as decavanadate, tetravanadate or tetrahedra chains.
Drezdzon [65] has proposed an alternative method to prepare LDHs exchanged
with large oxometalates, by intermediate preparation of organic-exchanged materi-
als. It was expected that, since hydrotalcite-like materials have higher charge
density than cationic clays, they would be more difficult to swell and exchange. A
Mg,Al–terephthalate LDH was obtained from Mg and Al nitrates and terephtha-
late in basic (NaOH) medium, with a basal spacing of 14.4 Å (7.8 Å for the
carbonate form). After addition of this solid to an aqueous NH4VO3 solution at pH
4.5, the basal spacing of the layered material obtained was 11.9 Å (coincident with
the value reported by Pinnavaia et al. [134]). Acidification plays a double role: (i)
polymerization of the metalate, and (ii) protonation of the organic anions, so
making easy its removal from the interlayer space; however, it has been also
claimed [147] that the process is inhibited by the poor solubility of the organic acid
in water, and the resulting difficulty in its removal from the clay matrix.
A third method already mentioned to prepare other LDHs, different from
exchange and pre-swelling with organic anions, was used by Jones et al. [148], from
the known ability of calcined LDHs to recover the layered structure. An Mg–Al
LDH calcined at 450°C for 18 h, suspended in an aqueous solution of NaVO3,
recovers the layered structure, hosting decavanadate species, as concluded from a
basal spacing of 11.8 Å, after acidification with HCl at pH 4.5.
All these methods require the use of carbon dioxide-free conditions, in order to
avoid incorporation of carbonate anions in the interlayer region. For acidic
systems, such as the Zn,Cr – LDH system, Kooli and Jones have reported a direct
method for the synthesis of decavanadate-containing Zn,Cr–LDH [149], at a pH
where carbonate is not present in the solution. A solution containing Zn and Cr
chlorides was added to an aqueous solution of NaVO3 (pH 4.5), and the slurry
obtained aged overnight at 55°C. The basal spacing for (003) planes was 11.89 Å,
suggesting again an orientation of the decavanadate anion with the main C2 axis
parallel to the host layers. Expanded structures with interlayer V10O628− anions were
obtained for solutions with Zn/Cr ratios between 1 and 5, although the a parameter
(related to the average cation – cation distance in the brucite-like layers, and hence,
on the ionic radii of these cations and their nature and concentration) remains close
to 3.11 Å, whichever the starting Zn/Cr ratio, a result similar to that previously
reported by de Roy et al. [1]. However, decavanadate was incorporated even at pH
6.5 (if the decavanadate solution had been prepared at pH 4.5), but not at higher
pH, probably due to the preferred formation of other oxovanadate oligomers.
These authors also succeed to prepare by this direct method decavanadate interca-
lated Zn,Al LDHs, avoiding the use of ZnO.
Exchange of decavanadate by carbonate in the interlayer space of Mg–Al LDH
was easily achieved by ultrasonic treatment of a suspension of the carbonate form
in a vanadate solution at the pH required to polymerize VO3− , but without any
further pH control [150]. It is likely that decavanadate exchange is facilitated by the
high dispersion of the agglomerated particles following ultrasonic treatment, and
also by enhanced diffusion of decavanadate on temporary delaminated particles.
96 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
used. The same effect of the drying step was observed by Carrado et al. [95] for
CuPcTs exchange in a Mg,Al – LDH. At pH 8.5 no exchange was observed, but at
intermediate pH a biphasic material, with characteristic XRD peaks at 7.54 Å
(carbonate) and 11.7 Å (decavanadate), was obtained. With regards to the recon-
struction method, this was valid only if the sample was submitted to hydrothermal
treatment at autogenous pressure, and when the precursor had been calcined at
300°C, but for higher calcination temperatures, NiO was also formed. According to
Clause et al. [162], as the calcination temperature is increased, Al3 + ions migrate
from the mixed oxide phase formed upon carbonate removal, to the crystallite
surface, where from they are dissolved when the solid is suspended at pH 4.5; so,
the nickel in excess does not enter the reconstructed LDH structure, but remains as
an independent NiO phase. The intensities of the NiO XRD peaks decrease (and
those of the LDH material increase) as the temperature during hydrothermal
treatment is increased in the 80 – 150°C range. While at pH 4.5–5.5 reconstruction
leads to the decavanadate – LDH phase (together with the byproduct responsible for
the XRD peak at 10 Å), when the pH during reconstruction was increased the IR
and XRD data indicate formation of (VO3) chain-like polymeric metavanadate,
composed of [VO4] tetrahedra with C26 symmetry, and with the longitudinal chain
axis parallel to the host layers.
Similar studies were also carried out with the Mg,Cr and Ni,Cr systems [161,163],
but decavanadate-containing LDHs were obtained only when following the ex-
change method using pre-wet carbonate–LDH, and at pH lower than 6.5 for
Fig. 15. Diagram showing various routes followed for the synthesis of terephthalate and vanadate
materials. Reprinted from M.A. Ulibarri, F.M. Labajos, V. Rives, R. Trujillano, W. Kagunya, W.
Jones, Comparative study of the synthesis and properties of vanadate-exchanged layered double
hydroxides, Inorg. Chem. 33 (1994) 2592–2599, © 1994, with permission from The American Chemical
Society.
98 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Mg,Cr, but 5.5 for Ni,Cr. In other words, not only pH, but also the intrinsic nature
of the layer cations seems to play an important role on the ability to exchange
decavanadate for carbonate. With regards to vanadate-containing LDHs prepared
following the reconstruction method, differences are also observed depending on
the nature of the cations in the brucite-like layers. So, the Ni,Al and Mg,Cr systems
are easily reconstructed if the carbonate precursor had been calcined below 300°C,
while after calcination at 400°C only an amorphous material was obtained after
contacting the calcined product with the vanadate solution. However, the Ni,Cr
calcined precursor does not reconstruct at all, and the spinel (NiCr2O4) phase is
detected by XRD in the precursor calcined at 500°C. Probably, the additional
stability of the calcined product because of crystal field effects when containing
transition metal cations somewhat hinders recovering of the layered structure with
interlayer vanadates.
layers. This behavior has been ascribed to grafting of the intercalated anions to the
layer upon heating, Fig. 11 [176]. The grafting process has been followed also by
51
V MAS-NMR, even from the very first stages of exchange in a NiIICoIII –LDH
[177]; as soon as they are inserted in the interlayer to compensate the positive
charge in excess in the layers, the isolated diperoxovanadate ions (formed in the
NH4VO3/H2O2 medium) undergo a competition between polycondensation and
grafting: if the solid is maintained with the reducing solution for a long time, partial
grafting occurs, leading to dehydroxylation of the layers; however, if the solid is
removed early from the solution, polycondensation is favored, together with a low
extended grafting. Thermal treatment favors further grafting and defragmentation
of the polyoxovanadate, finally leading to collapsing of the layered structure.
Grafting (as concluded from an abnormal short layer–layer distance), even without
any thermal treatment, has been also observed for pyrovanadate species, V2O47 − ,
onto a Cu,Cr LDH [141]; in this case, however, two adjacent hydroxyl groups of
one OH layer are substituted by two oxygen atoms of the ditetrahedra, as proposed
also for a Cu,Cr – Cr2O7 LDH phase [128].
Isopolyvanadate has been also exchanged at pH 4.5 in Mg–Al LDHs in the
nitrate form. A detailed study by XRD and IR spectroscopy of the species formed
upon heating in air at increasing temperatures has been carried out by López-Sali-
nas and Ono [174]. The results were similar to those reported by Twu and Dutta
[139], despite the difference in the Mg/Al ratio, 2 or 3. The changes can be also
easily followed by IR spectroscopy: The decavanadate gives rise to a strong, sharp
absorption band at 957 cm − 1, together with weaker bands at 557, 598, 660, 740,
and 820 cm − 1, while the presence of metavanadates gives rise to bands at 840 and
920 cm − 1 (terminal VO stretching), and 550 and 680 cm − 1 (V–O stretching in
bridging V – O – V bonds) [178]. If decomposition is carried out under vacuum,
partial reduction of V5 + to V4 + species takes place, the highest concentration (6
V4 + /100 Vtotal) of reduced species being reached at 150°C [174]; such a reduction is
not reversed by oxygen treatment at 200°C during 1 h, probably because of the
hindered access of even small molecules into the interlayer space due to very close
adjacent isopolyanions.
The presence of decavanadate anions in the interlayer space of LDHs has also
important effects on the decomposition process of the material. First of all,
calcination of a Mg – Al LDH in the carbonate form usually leads to a weight loss
close to 35 – 50% of the initial weight, due to removal of water physically adsorbed
on the external surface of the crystallites (usually at low temperature), removal of
interlayer water molecules, and, finally, dehydration/dehydroxylation, due to con-
densation of hydroxyl groups from the brucite-like layers, and decarbonation, from
the interlayer carbonate anions [179,180]. If decavanadate, instead of carbonate, is
present in the interlayer, the second weight loss corresponds only to water removal
through condensation of layer hydroxyl groups, and the total weight loss is usually
lower than 20 – 25% of the initial sample weight. The nature of the solids formed
upon calcination also depends on the nature of the interlayer anion. This effect has
been particularly studied for Co,Cr [181] and Zn,Cr [182] LDHs by del Arco et al.
For Co,Cr – V10O628− and Zn,Cr – V10O628− LDHs prepared from a carbonate precur-
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 101
Fig. 16. Cr–K edges XANES spectra for Zn,Cr – CO3 (a) and Zn,Cr – V10O28 (b) LDHs calcined at 400
and 800°C. Inset: Cr K-XANES spectrum for crystalline K2CrO4. Reprinted from M. del Arco, V.
Rives, R. Trujillano, P. Malet, Thermal behaviour of Zn – Cr layered double hydroxides with hydrotal-
cite-like structures containing carbonate or decavanadate, J. Mater. Chem. 6 (1996) 1419 – 1428, © 1996,
with permission from The Royal Society of Chemistry.
sor by ion exchange at pH 4.5, the XANES at the vanadium K-edge features
(e.g. pre-edge peak intensity and position, main edge position, and post-edge
structure) are almost identical as for a crystalline decavanadate (n-
C6H13NH3)6(V10O28) · 2H2O, confirming the structure of the intercalated anion. It
has been also observed, from XAS studies [147], that the zinc shell is in accordance
with Lowenstein’s rule, which states that trivalent cations in aluminosilicates should
not be in adjacent metal sites [183]. On calcination, while the DTA profiles for these
carbonate LDHs showed the expected endothermic peaks due to water removal,
when recorded in air, an additional exothermic peak was recorded for the Zn,Cr–
LDH at 435°C (335°C for the Co,Cr–LDH), but was absent when recorded in
nitrogen, indicating an oxidation process, involving the Cr ions; according to Fuda
et al. [184], the presence of carbonate in the interlayer of hydrotalcites favors the
Cr3 + Cr6 + oxidation. The peak was absent in the DTA profiles of the vanadate-
containing LDHs. When the samples were calcined in air at increasing tempera-
tures, oxidation of Co2 + to Co3 + takes place at 400°C in the Co,Cr–CO3 LDH,
but not in the decavanadate analogue. In addition, a weak pre-edge peak character-
istic of chromate ions appears at the chromium K-edge XANES spectrum, Fig. 16.
Depolymerization of decavanadate species follows trends similar to those reported
by Twu and Dutta [139] for Mg,Al –decavanadate LDHs. Calcination at 650°C
leads to crystallization of CoIICoIIICrIIIO4 from the carbonate precursor, but of
102 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
CoIICr2O4 and CoII 2 V2O7 from the decavanadate precursor. In other words, the
presence of decavanadate and the absence of carbonate hinders oxidation of Co2 +
ions, thus modifying the nature of crystalline phases formed at high temperature.
For the Zn,Cr analogues, the same behavior is observed with respect to Cr
oxidation, and the crystalline phases formed were ZnO and ZnCr2O4 from the
carbonate precursor (through formation of chromate species at intermediate calci-
nation temperatures), and ZnV2O6 and Zn2V2O7 at 400 and 750°C, respectively,
from the vanadate precursor, together with ZnCr2O4. Formation of chromate and
Co3 + species in the case of the carbonate precursors, but not in those containing
interlayer decavanadate, has been also concluded from temperature-programmed
reduction studies [181,182], a technique that has been proved to be adequate to
determine redox processes in layered double hydroxides containing reducible
cations in the layers or in the interlayer anions [185].
3.2.3. Molybdates
While the chemistry of vanadates in the interlayer space of LDHs has been
throughout studied, that of molybdates has been restricted to the heptamolybdate,
Mo7O624− , and the papers published are rather scarce. The preswelling technique
with terephthalate was used by Drezdzon [65] to intercalate Mo7O624− in the
interlayer space of a Mg,Al carbonate precursor. The method was similar to that
above described for decavanadate, i.e. direct preparation of the Mg,Al–terephtha-
late precursor (from Mg2 + and Al3 + nitrates and terephthalic acid in NaOH
medium), and mixing of the slurry with a Na2MoO4 · 2H2O aqueous solution, and
further lowering of the pH to 4.4 – 4.7 with HNO3. The basal spacing was 12.2 Å,
corresponding to a heptamolybdate orientation in which the C2 axis is perpendicu-
lar to the brucite-like layers. Exchange reactions using other organic precursors did
not succeed (2,5-dihydroxy-1,4-benzendisulphonate; lauryl sulphate) or proceed
with difficulty (1,5-naphtalenedisulphonate), because of competition with complex
formation between the organic species and the metalate [186].
The so-called reconstruction method, from a carbonate precursor calcined at
450°C and an acidified (pH 4.5) solution of ammonium heptamolybdate, led also to
intercalation of heptamolybdate species [148,187] with the same spacing as that
obtained by Drezdzon. However, Misra and Perrotta have reported [126] the
preparation of a Mg,Al – Mo7O24 LDH with a basal spacing of 9.6 Å from a
carbonate precursor calcined at 500°C, but without acidification. This spacing is
markedly lower than that reported by Drezdzon for samples prepared by the
terphthalate intermediate [65], and by Chibwe and Jones [148] for samples prepared
by reconstruction. The difference has been attributed by these authors [126] to the
different Mg/Al ratio in their sample (1.88) and in Drezdzon’s sample (2.0).
However, the difference is very small and, moreover, the value by Misra and
Perrotta is almost coincident with that reported by Chibwe and Jones (1.82), which
led an interlayer space of 12.0 Å. So, the difference should be more probably due
to: (i) a different orientation of the interlayer anion, (ii) the presence of a different
oxomolybdate anion (due to partial depolymerization), or (iii) some sort of
grafting, as reported in the case of vanadate [141] and chromate [127,128,131], as
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 103
remaining of organic molecules in the interlayer space should be ruled out since
these were not used in the method followed by Chibwe and Jones [148]; unfortu-
nately, these authors [126] provide no other experimental data, in addition to XRD,
to support their conclusions. The same value, 12.0 Å, has been recently reported by
Hibino and Tsunashima [188] for samples prepared in an ethanol–water solution
by anionic exchange, to avoid partial dissolution of the brucite-like layers, due to
the acidic medium provided by the molybdate solution. A value of 12.2 Å has been
also reported [189] by Twu and Dutta in materials (Mg/Al = 2.0) prepared follow-
ing the Drezdzon’s method; although the layer structure is destroyed at 300°C, the
Mo7O624− moiety is stable up to 400°C, then forming MgAl2O4, MgMo2O7 and
MgMoO4, as concluded from XRD and Raman spectroscopy studies. However,
intercalation of MoO24 − in a Li,Al LDH failed, due to hydrolysis of the molybdate
anion, even at r.t.
Levin et al. [190,191] have reported synthesis of a layered ammonium zinc
molybdate using as a precursor a Zn,Al–LDH calcined at 500°C and Mo7O624−
(although this depolymerizes along the reaction) at r.t. These authors find a similar
reactivity with LDHs containing, in addition to Zn and Al, Cu2 + , Co2 + or Ni2 + ,
but not with LDHs containing Ni2 + or Mg2 + only as the divalent cation. From
27
Al MAS-NMR studies, it was concluded that a high fraction of tetrahedral Al3 +
ions are required in the calcined precursor to yield the layered molybdate.
with charge balance reasons: as the diameter of the Keggin unit is close to 9.8 Å,
an area of 83 Å2 is required to accomodate a Keggin unit; as the charge density in
the LDH used was 16.6 Å2 (although this value has been corrected to 25 Å2 by
Clearfield et al. [122]), those Keggin units with a formal charge lower than − 5 will
be unable to enter in the interlayer space to compensate the positive charge of the
layers.
In any case, the products obtained, Zn,Al–a-[H2W12O40]6 − and Zn,Al–a-
[SiV3W9O40]7 − , are crystallographically well-ordered phases, with a basal spacing
of 14.5 Å, corresponding to a gallery height of 9.8 Å, in agreement with crystallo-
graphic data for Keggin units, and up to six diffraction harmonics were recorded in
the XRD diagram. The extremely large swelling of the layers is accompanied by a
substantial increase in the specific surface area, from 26 m2 g − 1 for the nitrate
precursor, to 63 and 155 m2 g − 1, respectively, for the solids intercalated with
a-[H2W12O40]6 − and a-[SiV3W9O40]7 − , and then the term ‘pillared’ seems to be
more adequately used than in other cases where, despite intercalation of large
anions, such an increase is not observed. Further evidence for the retention of the
Keggin structure inside the layers was attained by IR and 29Si and 51V MAS-NMR
spectroscopies. With respect to the orientation of the anion in the interlayer, these
authors [194] conclude that the C2 axis of the oxygen framework should be
perpendicular to the brucite-like layers, Fig. 17, as in this orientation the number of
hydrogen bonds to layer OH groups is maximized. This orientation has been also
proposed by Liu et al. [195] for [PW11O39Cr(H2O)]4 − , [PW11TiO40]5 − and
[PW11VO40]4 − intercalated in Zn,Al – LDHs prepared by ion exchange.
In order to intercalate Keggin-type anions in rather basic LDHs, Dimotakis and
Pinnavaia have proposed [196,197] an alternative method, consisting in preparation
of the Mg3Al – OH LDH (meixnerite) by reconstruction of a calcined carbonate
precursor; meixnerite is then exchanged with p-toluensulfonate or adipate in the
presence of glycerol as a swelling agent, yielding a well-crystallized phase with
extremely well ordered organic anions, which gallery height is very close (14.4 Å for
the adipate) to that of the Keggin derivative, from which anion exchange led to
microporous, single crystalline phases with d= 14.8 Å, after exchanging with
[H2W12O40]6 − or [SiW11O39]8 − .
Fig. 17. A Keggin unit, [XM12O40]n − , formed by 12-corner-sharing octahedra, in two different
orientations.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 105
Fig. 18. XRD patterns (Cu–Ka ) of oriented film samples of Mg,Al – LDH metatungstate reaction
products obtained from (A) meixnerite, (B) LDH glycerolate, and (C) LDH triethyleneglycolate
precursors. All diffraction patterns were recorded at r.t. For the LDH – POM intercalated derived from
the triethyleneglycolate precursors, the XRDs were taken at r.t. after 1 h preheating under nitrogen at
(D) 100, (E) 200, and (F) 250°C. Reprinted from S.K. Yun, V.R.L. Constantino, T.J. Pinnavaia, New
polyol route to keggin ion-pillared layered double hydroxides, Microporous Mater. 4 (1995) 21 – 29, ©
1995, with permission from Elsevier Science.
aqueous solution without any organic swelling agent. Direct anion exchange in
aqueous solution was also used by Guo et al. [201] to prepare Zn2Al–LDH
intercalated with [PVn W12 − n O40](3 + n) − (n=1–4). Exchange is easier if the LDH
precursor is thoroughly wet [122,154], either by preparing wet solids or by soaking
the dried product for 3 – 4 h (a similar role of the wetness state of the precursor has
been also shown by Carrado et al. [95] and Kooli and Jones [149] to incorporate
phtalocyanines or decavanadate in the interlayer region, respectively). Also the
exchange is easier with soft powders than with glassy particles. Moreover, if the
solid is slurried enough, pillaring seems to be almost independent on the charge of
the Keggin unit: for a wet Ni2Al– NO3 precursor soaked for at least 3 h, total
exchange was obtained with all Keggin ions with net charge ranging from − 3 to
− 7 (a 90% exchange was already attained after 5–10 min). This apparent indepen-
dence on the net charge of the Keggin unit is rather surprising, as the area required
to accomodate a Keggin unit is 83 Å2, and the area per layer charge unit is 25 Å2
(16.6 Å2 according to Kwon and Pinnavaia [194]); then the charge per Keggin unit
should be, at least, −4 (or − 5, according to Kwon and Pinnavaia). The experi-
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 107
mental finding that even Keggin units with a net charge of − 3 are also exchanged,
can be only explained assuming an alteration of the layers during the exchange
process, leading to an increase in the layer charge density.
On the other hand, the interlayer spacing required to host the Keggin unit is close
to 14.2 Å, but the interlayer spacing measured for the pillared Ni,Al–LDHs was
only ca. 12 Å. For the Mg,Al – LDHs the spacing was 14.7 Å (the corresponding
harmonics being also recorded), and an additional broad reflection, similar to that
reported by Narita et al. [152], was also recorded at 11–13 Å. These authors
suggest a partial dissolution of the divalent anion in the acidic medium (pH 4–5)
during exchange, together with partial removal of hydroxyl groups from the layers,
thus creating vacancies where the Keggin ions could fit, leading to interlayer
spacings lower than expected, and thus explaining intercalation of the [PW12O40]3 −
Keggin unit. Computer graphics models support such relationship between ‘pene-
tration’ of the Keggin units into the brucite-like layers, and the spacing.
To overcome the difficulty in intercalation of Keggin units with low negative
charge, Serwicka et al. [202] have subjected the [PMo12O40]3 − anion to electrochem-
ical reduction prior to exposure to the LDH; it is known [192] that these anions
undergo facile reduction to give so-called heteropoly-blues, and that reduction
renders the anions less acidic, so helping also to overcome decomposition by the
reaction with the rather basic Mg,Al –NO3 LDH. In this case, reduction led to a
transfer of four electrons to the Keggin unit, and the authors find that the
unreduced heteropolyanion actually reacts with the LDH, XRD show diffraction
maxima at 11.1 Å, too small to correspond to the intercalated Keggin unit,
probably corresponding to decomposed species; the atomic Mo:P ratio was 4.5.
However, for the reduced heteropolyanion, the spacing was 14.8 Å, with corre-
sponding harmonics, although an intense reflection was also recorded at 10.8 Å,
ascribed to formation of a non-layered byproduct.
Direct exchange reaction in aqueous solution has been used by Hua et
al. [203,204] to intercalate peroxoheteropolyanions with the Keggin structure,
such as [SiW11(TiO2)O39]6 − , [SiW9(TiO2)3O37]10 − , [PW11(TiO2)O39]5 − and
[PW9(TiO2)3O37]9 − , in Zn2Al – LDHs in the nitrate form, obtaining solids with
basal spacings of 14.7 Å.
Weber et al. [205] have studied by TEM the partial exchange of [SiV3W9O40]7 −
in a Mg2Al – LDH. These authors find that the average crystallite size of the
exchanged product was larger than that of the original LDH, suggesting exchange
proceeds via dissolution and topotactic reprecipitation of the exchanged LDH, as
also proposed by Pinnavaia et al. when following formation of triethylenglycol
intermediates [198]. On the other hand, the results obtained on the local chemical
composition (Mg/Al and W/Al ratios) indicate that in partially intercalated solids,
the resulting structures consist of stacks of completely substituted layers superposed
on unchanged layers.
The synthesis of many other LDHs containing intercalated Keggin-type anions
has been described in the literature. Hu et al. [206,207] have reported the intercala-
tion in Zn2Al – LDH of different POMs, such as [PVW11O40]4 − ,
[XW11O39Z(H2O)]n − (X = Si, B; Z =Co, Ni, Cu, Al), [Ln(XW11O39)2]n − (Ln= La,
108 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
Contrary to the results with the Dawson’s anion, intercalation of the Finke anion
led to layered materials with a basal spacing of 17.79 0.3 Å, whichever (adipate or
meixnerite) the precursor used, suggesting that both intercalated products have the
C2 axis of the heteropolyanion perpendicular to the LDH layers. The microporous
structure is maintaned up to 200°C when heated in N2. Upon heating, the gallery
height decreases by 2 – 3 Å, by removal of water molecules, resulting in stronger
electrostatic and hydrogen bonding interactions between the oxygen ions of the
heteropolyanion and the LDH hydroxyl groups, leading to reorientation of the
former. Retention of the Dawson and Finke structures in the intercalated state was
verified by FTIR spectra, which showed the expected bands due to P–O linkages
close to 1100 and 1030 cm − l, and W –O linkages close to 960, 930, and 750 cm − l,
the precise positions depending on the particular POM [211,212].
As in most of the cases previously described [153,154,205], the XRD diagrams of
these materials showed, in addition to the lines due to the layered material, a broad
peak close to 11 Å. Although different explanations have been previously proposed
for the nature and origin of the material responsible for this reflection, these
authors conclude that salt formation from cations depleted from the layers and
non-gallery POM remains the favored explanation, and the XRD features of this
impurity were the same as of the solid obtained by grinding a carbonate–LDH or
even Mg(OH)2 and a POM in the solid state. This report [210] represents the most
widely study on this byproduct, detected upon reaction of POMs with LDHs.
Larger POMs have been introduced in the gallery space of Mg,Al and Zn,Al
LDHs by Evans et al. [147] by ion exchange and direct synthesis; solids have been
prepared with gallery heights ranging from 7.1 to 16 Å including species such as
[Nbx W6 − x O19](x + 2) − (x = 2 – 4), [V2W4O19]4 − , [Ti2W10PO40]7 − , and
l4 −
[NaP5W30O110] . The largest yields to highly crystalline solids were obtained
following the exchange method, with precursors containing chloride or nitrate.
These POMs, in addition, are stable in a wide pH range, and so can be incorpo-
rated into the interlayer space of strongly basic Mg,Al–LDH. Probably, one of the
most interesting products prepared by these authors [147] is that containing the
Preyssler anion, [NaP5W30O110]14 − , one of the largest known anions, and a rather
abnormal chemical assembly with a C5 axis. Its structure consists of a cyclic
arrangement of five [PW6O22] units, each formally derived from the Keggin-type ion
[PW12O40]3 − by removal of two sets of corner-shared [WO6] groups. The basal
spacing for the Zn,Al – Preyssler derivative obtained from the nitrate precursor was
21 Å, in agreement with the ion orienting its shortest dimension parallel to the host
layers. When ion exchange was performed from benzenecarboxylate-containing
precursors (YC6H4COO − , Y =COO − , OH, CH3), low (ca. 3.5–4.5) pH values
were required to protonate the carboxylate group, to induce removal of the free
organic acid. Direct synthesis led to incorporation of the Preyssler ion in a
Zn,Al–LDH, with a spacing of 21 Å. Retaining of the POM structure was checked
by EXAFS [147]. As expected, a by-product responsible for a reflection at 7–11°
(2u, Cu– Ka ) and which nature has been discussed above is formed in most of the
cases.
Studies on the thermal decomposition of Zn2Al–LDHs intercalated with differ-
ent Keggin ions have shown [213] that after dehydration (200°C), dehydroxylation
is completed at 350°C, leading to amorphous solids; the layer structure is destroyed
between 200 and 250°C, depending on the precise nature of the interlayer Keggin
ion. The layered material can be rehydrated by immersion of the calcined solid in
water, but the pillared, layered structure is not reconstructed by simply rehydration.
Crystallization of anhydrous mixed oxides (e.g. ZnWO4) is observed after dehy-
droxylation, leading to an exothermic peak in the DSC trace. Similar results have
been reported by Guo et al. [214].
The structure of salts of Keggin ions with bulk cations (e.g. Cs3[PWl2O40]) is
stable even up to 700°C. The results by Kwon and Pinnavaia [213] suggest that the
mixed oxide formed upon decomposition of the brucite-like layers reacts easily with
the hosted Keggin ion. This low thermal stability in some sort of way limits the
application of these materials to catalytic processes taking place at rather low or
even r.t., such as photocatalytic oxidation of isopropanol to acetone [213].
Although LDHs are mostly basic solids, incorporation of POMs in the interlayer
space not only increases the gallergy height and the thermal stability, but also may
provide electron acceptor sites and acid sites. Thereof, these LDH–POM com-
pounds are interesting because of their acid–base properties. In these LDH–POM
systems, basic sites are located on the layers, while acid sites are on the interlayer
anions [215]. As a consequence, the relative acid–base strength could be changed by
different anionic exchange ratios, or even thermal treatments to yield materials with
tailored acid – base properties. Putyera et al. [216] have prepared a series of
Mg,Al–LDHs intercalated with molybdate and tungstate, in order to assess the
relationship existing between the composition and the acid–base properties of the
pillared materials and of their derivatives obtained upon calcination. As expected,
bearing in mind the close relationship existing between the nuclearity of the POM
and the pH, (WO24 − is stable at pH \ 8, W12O12 42
−
at pH 7.8, and W12O639− at pH
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 111
5.7, while for the oxomolybdate species, MoO24 − is stable at pH\7, but Mo7O624−
at lower pH values), a change in pH during synthesis gives rise to different acid
sites/basic sites ratio in the intercalated derivatives. Upon calcination, formation of
high nuclearity oxometalates (Mo7O624− or W12O10 −
41 ) again produces changes in the
acid–base properties.
Polyoxometalates are well known as oxidation catalysts, and so it is expected that
POM-catalyzed oxidation processes on POM hosted in the interlayer region of an
LDH can be constrained in a shape-selective environment. Tatsumi et al. [217,218]
have reported the epoxidation of alkenes (e.g. 2-hexene, cyclohexene and b-methyl-
styrene) with H2O2, catalyzed by LDHs intercalated with POM derivatives of Mo
and W. When the catalytic activity results are compared with those for the
unhosted POM, an steric effect is observed, and so epoxidation of large alkenes is
less favored on the LDH – POM catalyst, than when smaller alkenes are used. On
the other hand, the steric hinderance for the alkenes to access the interlayer region
seems to be less important for LDH – W12O41 than for LDH–Mo7O24, in agreement
with a larger basal spacing in the former than in the latter (12.2 vs. 9.9 Å,
respectively). These results suggest a possible shape selectivity control on changing
the size of the hosted species. In addition, hydrolysis of epoxides to yield di-ol
species is slowed down if compared with that observed for the unhosted catalysts,
probably because of the basic properties of the brucite-like layers. However,
Gardner and Pinnavaia have recently pointed out [219] that the co-product
generally formed (characterized by a broad diffraction maximum close to 10–11 Å)
when preparing LDH – POM intercalates can have important catalytic conse-
quences, and be even catalytically more important than the LDH phase.
The catalytic oxidation of benzaldehyde to benzoic acid using H2O2 in a biphasic
liquid–solid system, has been studied by Hu et al. [220] on Zn,Al–LDHs interca-
lated with [SiW11O39]8 − and [SiW11O39Z(H2O)]6 − (Z=Co2 + , Ni2 + , Cu2 + ). The
largest catalytic activity has been observed for the cobalt-derivative, and this
finding has been tentatively related to the easy change in the oxidation state of Co,
suggesting that Co2 + becomes oxidized to Co3 + by H2O2, and Co3 + oxidizes
benzaldehyde to benzoic acid.
Guo et al. [221] have studied the O2-oxidation of cyclohexene on LDHs interca-
lated with [XW11O39Z(H2O)]n − (X = P, Si; Z= Mn2 + , Fe3 + , Co2 + , Ni2 + , Cu2 + ).
The LDH – POM system is more active than the LDH in its nitrate form, and also
more active than the alkaline salts of the POM. An effect of the transition metal
cation existing in the POM moiety (Z) has been observed. It was concluded that the
cations in the layers or in the POM group behave as active sites for oxygen transfer
in cyclohexene oxidation, and so the catalytic properties could be hopefully tuned
by precise changes in the nature and concentration of transition metal cations in
both types of sites (layers and POM units).
Zheng et al. [222] have reported alkylation of iso-butane with butene on
Zn,Al–LDHs intercalated with [SiW12O42]8 − , and of Ni,Al–LDH intercalated with
[PW12O42]7 − , with good results with respect to activity and selectivity. For this
reaction, alkylation may proceed on basic (LDH) or acid (POM) sites, i.e. the
POM-pillared LDH can behave as a bifunctional acid–base catalyst. On calcination
112 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
at 300°C, Ni2Al– PW12O42 shows a much higher butene conversion than the
uncalcined material, as well as a higher content of C12 and C16 products, probably
due to a lower steric hinderance upon removal of interlayer water molecules.
Clearfield et al. [38] have used LDH–POM materials for catalytic conversion of
isopropanol to acetone or propene. This is a test reaction widely used for catalyst
characterization, as it proceeds to propene on acid sites, while to acetone on basic
sites. According to these authors, LDHs such as Mg,Al–CO3 behave as basic
catalysts, leading to acetone [215]. However, insertion of heteropolyacid anions
alters the selectivity drastically towards propene, showing that the acid character of
the POM predominates. Minor changes in acid/base catalytic properties have been
also correlated to the precise nature of the cations (Co2 + , Fe3 + , Ni2 + , Mg2 + ,
Al3 + ,...) in the brucite-like layers. Studies by Kagunya and Jones [165] on the aldol
condensation of acetaldehyde on Mg,Al[SiW12O40] have correlated the catalytic
activity with the surface area available, while selectivity seems to be related to the
number and strength of the basic sites, responsible for the activity in this reaction.
Ethanolysis of propene oxide to yield glycol ether (a reaction that can proceed
catalytically both on acid and basic sites) has been studied by Jones et al. [223] as
a way to assess the nature of active sites in Mg,Al–LDHs intercalated with POMs
such as [PW12O40]3 − and [SiW12O40]4 − . While for the LDHs lacking interlayer
POMs these authors report the hydroxyl groups as being the active sites, in the
LDH–POM systems, up to three sites are present: (i) oxide anions directly linked
to metal atoms (strong basicity), (ii) oxide ions bonded to atoms adjacent to metal
centers (medium basicity), and (iii) surface hydroxyl groups (weak basicity).
The acid – base functionality of the Mg,Al–LDH–[H2W12O640− ] association,
which synthesis has been described above, has been examined by Pinnavaia et al.
[198] using 2-methyl-3-butyn-2-ol (MBOH) as a reactive probe. When the catalyst
was obtained via a triethylenglycolate intermediate, a high reactivity for the
base-catalyzed disproportionation was observed, whereas when obtained via the
glycolate or meixnerite it was rather inactive. The difference has been attributed to
the different porosity in the samples obtained when using alternative intermediates.
Keita et al. [224,225] have prepared oxometalate-clay-modified electrodes con-
taining metatungstate. Glassy carbon electrodes have been modified with LDHs
containing Zn and Al in the brucite-like layers, and the interlayer anions have been
substituted for [H2W12O40]6 − under mild acid conditions. Clay films were prepared
by dropping a colloidal solution of the LDH onto the glassy carbon surface;
incorporation of the POM was accomplished readily by cycling the electrode in the
solution containing the oxometalate in the potential domain of the first redox
system of [H2W12O40]6 − to monitor the progress of the incorporation. Migration of
the highly-charged POM in the interlayer region of the clay results in the final
LDH–POM system. The best results were obtained with a Zn,Al–LDH precursor
containing the relatively large (if compared to chloride or nitrate) terephthalate
dianion at pH 5, acidic enough to ensure protonation of the organic anion (thus
favoring its removal from the interlayer region), but basic enough to avoid
dissolution of the layers.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 113
4. Miscellaneous
5. Conclusions
From the results in this review, it is obvious the interest that layered double
hydroxides have deserved in recent years. Its structure, similar to that of layered
silicates, but with a change in the sign of the electric charges of the layers and the
interlayer ions, makes them true companions in systematizing the study of these
solids. On the other hand, as the layer cations and the interlayer anions can be
almost chosen from any one in the Periodic Table, the opportunities for synthesis
chemistry are enormous. This obviously constitutes a challenge for chemists. In
addition, the promising role that these materials, as obtained or after adequate
thermal treatments, can play as catalysts, sensors, electrodes, etc., makes them
worthwhile to be studied in a systematic way to modulate and to improve their
properties. Probably, in the next few years we will witness a lot of new work on
these compounds, expanding the nature of intercalated metal-containing anions.
Acknowledgements
The authors would like to thank the collaboration of their co-workers in the
Universities of Córdoba and Salamanca (Spain), as well as of Dr W. Jones
(University of Cambridge, UK), Dr P. Malet (Universidad de Sevilla, Spain), and
Dr F. Kooli (currently at NIRIM, Tsukuba, Japan). Finantial support by Junta de
Castilla y León (Consejerı́a de Educación y Cultura, grant SA45/96), Junta de
Andalucı́a (grant FQM-214) and Ministerio de Educacion y Cultura (grant PB96-
1307-C03) is acknowledged.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 115
References
[1] A. de Roy, C. Forano, K. El Malki, J.-P. Besse, in: M.L. Occelli, H.E. Robson (Eds.), Synthesis
of Microporous Materials, vol. 2, Expanded Clays and Other Microporous Systems, Van Nos-
trand Reinhold, New York, 1992, pp. 108 – 169.
[2] F. Trifirò, A. Vaccari, in: J.L. Atwood, J.E.D. Davies, D.D. MacNicol, F. Vögtle, J.-M. Lehn, G.
Alberti, T. Bein (Eds.), Comprehensive Supramolecular Chemistry, vol. 7, Solid-State Supramolec-
ular Chemistry: Two- and Three-Dimensional Inorganic Networks, Pergamon, Oxford, 1996, pp.
251 – 291.
[3] F. Cavani, F. Trifirò, A. Vaccari, Catal. Today 11 (1991) 173.
[4] M. Zikmund, K. Hrnciarová, Chem. Listy 91 (1997) 169.
[5] W.T. Reichle, Solid State Ionics 22 (1986) 135.
[6] M. Chibwe, J.B. Valim, W. Jones, in: C.A.C. Sequeira, M.J. Hudson (Eds.), Multifunctional
Mesoporous Solids, Kluwer, Amsterdam, 1993, pp. 191 – 206.
[7] K.A. Carrado, A. Kostapapas, S.L. Suib, Solid State Ionics 26 (1988) 77.
[8] M. Meyn, K. Beneke, G. Lagaly, Inorg. Chem. 29 (1990) 5201.
[9] W. Jones, M. Chibwe, in: I.V. Mitchell (Ed.), Pillared Layered Structures: Current Trends and
Applications, Elsevier, London, 1990, pp. 67 – 77.
[10] R.M. Taylor, Clay Miner. 19 (1984) 591.
[11] S. Velu, V. Ramaswamy, A. Ramani, B.M. Chanda, S. Sivasanker, J. Chem. Soc. Chem.
Commun. (1997) 2107.
[12] C.J. Serna, J.L. Rendón, J.E. Iglesias, Clays Clay Miner. 10 (1982) 180.
[13] M.A. Drezdzon, ACS Symp. Ser. 437 (1990) 140.
[14] A. Clearfield, M. Kuchenmeister, J. Wang, K. Wade, in: P.A. Jacobs, N.I. Jaeger, L. Kubelková,
B. Wichterlová (Eds.), Zeolite Chemistry and Catalysis, Elsevier, Amsterdam, Stud. Surface Sci.
Catal., vol. 69, 1991, pp. 485–497.
[15] A. Clearfield, in: C.A.C. Sequeira, M.J. Hudson (Eds.), Multifunctional Mesoporous Solids,
Kluwer, Amsterdam, 1993, pp. 159–178.
[16] S. Yamanaka, Mater. Sci. Forum 152–153 (1994) 69.
[17] K. Ohtsuka, Chem. Mater. 9 (1997) 2039.
[18] R. Szostak, C. Ingram, in: H.K. Beyer, H.G. Karge, I. Kiricsi, J.B. Nagy (Eds.), Catalysis by
Microporous Materials, Elsevier, Amsterdam, Stud. Surface Sci. Catal., vol. 94, 1995, pp. 13 – 38.
[19] T.J. Pinnavaia, M. Chibwe, V.R.L. Constantino, S.K. Yun, Appl. Clay Sci. 10 (1995) 117.
[20] A. Corma, Chem. Rev. 97 (1997) 2373.
[21] E. López-Salinas, Y. Ono, Microporous Mater. 1 (1993) 33.
[22] A.B.P. Lever, E. Montovani, B.S. Ramaswany, Can. J. Chem. 49 (1971) 1957.
[23] L.E. Alzamora, J.R.H. Ross, E.C. Kruissink, L.L. van Reijden, J. Chem. Soc. Faraday Trans. I
77 (1981) 665.
[24] J.R. Weisner, R.C. Srivastava, C.H.L. Kennard, M. di Vaira, E.C. Lingafelter, Acta Crystallogr.
23 (1967) 565.
[25] J. Brynestad, G.P. Smith, J. Am. Chem. Soc. 92 (1970) 3198.
[26] A.B.P. Lever, Inorganic Electronic Spectroscopy, 2nd edn, Elsevier, Amsterdam, 1984, p. 507.
[27] F.A. Cotton, D.M.L. Goodgame, M. Goodgame, J. Am. Chem. Soc. 83 (1961) 4690.
[28] K. Okada, F. Matsushita, S. Hayashi, Clay Min. 32 (1997) 299.
[29] E. López-Salinas, N. Tomita, T. Matsui, E. Suzuki, Y. Ono, J. Mol. Catal. 81 (1993) 397.
[30] E. López-Salinas, Y. Ono, E. Suzuki, Mater. Res. Symp. Proc. 368 (1995) 363.
[31] K. Itaya, H.-C. Chang, I. Uchida, Inorg. Chem. 26 (1987) 624.
[32] K. Chibwe, W. Jones, J. Chem. Soc. Chem. Commun. (1989) 926.
[33] S. Miyata, T. Hirose, Clays Clay Min. 26 (1978) 441.
[34] S. Kikkawa, M. Koizumi, Mater. Res. Bull. 17 (1982) 191.
[35] L.H. Jones, Inorg. Chem. 2 (1963) 777.
[36] F.M. Labajos, V. Rives, M.A. Ulibarri, Spectrosc. Lett. 24 (1991) 499.
[37] P.S. Braterman, C. Tan, J. Zhao, Mater. Res. Bull. 29 (1994) 1217.
116 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
[38] J.D. Wang, G. Serrette, Y. Tian, A. Clearfield, Appl. Clay Sci. 10 (1995) 103.
[39] P.K. Dutta, M. Puri, J. Phys. Chem. 93 (1989) 376.
[40] S. Idemura, E. Suzuki, Y. Ono, Clays Clay Min. 37 (1989) 553.
[41] H.G. Drickamer, S.C. Fung, G.K. Lewis Jr., Adv. High Pressure Res. 3 (1969) 1.
[42] M.J. Holgado, V. Rives, M.S. San Román, P. Malet, Solid State Ionics 92 (1996) 273.
[43] H.C.B. Hansen, C.B. Koch, Clays Clay Min. 42 (1994) 170.
[44] J.A. Olabe, H.O. Zerga, Inorg. Chem. 22 (1983) 4156.
[45] I. Crespo, C. Barriga, V. Rives, M.A. Ulibarri, Solid State Ionics 101 – 103 (1997) 729.
[46] E. Suzuki, S. Idemura, Y. Ono, Clays Clay Min. 37 (1989) 173.
[47] F.A.P. Cavalcanti, A. Schutz, P. Biloen, in: B. Delmon, P. Grange, P.A. Jacobs, G. Poncelet
(Eds.), Preparation of Catalysts IV, Elsevier, Amsterdam, 1987, pp. 165 – 174.
[48] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, J. Rouquerol, T. Sieminiewska, Pure Appl.
Chem. 57 (1985) 603.
[49] J.M. Fernández, M.A. Ulibarri, F.M. Labajos, V. Rives, J. Mater. Chem. 8 (1998) 2507.
[50] H. Nijs, P. Cool, E.F. Vansant, Interface Sci. 5 (1997) 83.
[51] G. Mao, M. Tsuji, Y. Tamaura, Clays Clay Min. 41 (1993) 731.
[52] D.W. Breck, W.G. Eversole, R.M. Milton, T.B. Reed, T.L. Thomas, J. Am. Chem. Soc. 78 (1956)
5963.
[53] T. Challier, R.C.T. Slade, J. Mater. Chem. 4 (1994) 367.
[54] B.R. Shaw, Y. Deng, F.E. Strillacci, K.A. Carrado, M.G. Fessehaie, J. Electrochem. Soc. 137
(1990) 3136.
[55] J. Labuda, M. Hudáková, Electroanalysis 9 (1997) 239.
[56] (a) J. Qiu, G. Villemure, J. Electroanal. Chem. 395 (1995) 159. (b) J. Qiu, G. Villemure, J.
Electroanal. Chem. 428 (1997) 165.
[57] A. Cervilla, A. Corma, V. Fornés, E. Llopis, P. Palanca, F. Rey, A. Ribera, J. Am. Chem. Soc.
116 (1994) 1595.
[58] A. Cervilla, E. Llopis, A. Ribera, A. Corma, V. Fornés, F. Rey, J. Chem. Soc. Dalton Trans.
(1994) 2953.
[59] A. Corma, V. Fornés, F. Rey, A. Cervilla, E. Llopis, A. Ribera, J. Catal. 152 (1995) 237.
[60] A. Corma, F. Rey, J.M. Thomas, G. Sankar, G.N. Greaves, A. Cervilla, E. Llopis, A. Ribera, J.
Chem. Soc. Chem. Commun. (1996) 1613.
[61] P. Palanca, T. Picher, V. Sanz, P. Gómez-Romero, E. Llopis, A. Doménech, A. Cervilla, J. Chem.
Soc. Chem. Commun. (1990) 531.
[62] V. Sanz, T. Picher, P. Palanca, E. Llopis, J.A. Ramı́rez, D. Beltrán, A. Cervilla, Inorg. Chem. 30
(1990) 3113.
[63] E. Llopis, A. Doménech, J.A. Ramı́rez, A. Cervilla, P. Palanca, T. Picher, V. Sanz, Inorg. Chim.
Acta 189 (1991) 29.
[64] H. Kominami, S. Kurimoto, M. Kubota, R. Shiozaki, Y. Kera, J. Ceram. Soc. Jpn. 105 (1997) 707
(in Japanese).
[65] M.A. Drezdzon, Inorg. Chem. 27 (1988) 4628.
[66] M.D. Newsham, E.P. Giannelis, T.J. Pinnavaia, D.G. Nocera, J. Am. Chem. Soc. 110 (1988) 3885.
[67] E.P. Giannelis, D.G. Nocera, T.J. Pinnavaia, Inorg. Chem. 26 (1987) 203.
[68] P.K. Ghosh, A.J. Bard, J. Phys. Chem. 88 (1984) 5519.
[69] R.A. Della Guardia, J.K. Thomas, J. Phys. Chem. 87 (1983) 990.
[70] H. Nijs, J.J. Fripiat, H. Van Damme, J. Phys. Chem. 87 (1983) 1279.
[71] J.R. Winkler, H.B. Gray, J. Am. Chem. Soc. 105 (1983) 1373.
[72] S. Miyata, A. Okada, Clays Clay Min. 25 (1977) 14.
[73] (a) S. Miyata, Clays Clay Min. 23 (1975) 369. (b) S. Miyata, Clays Clay Min. 31 (1983) 305.
[74] J.R. Winkler, H.B. Gray, Inorg. Chem. 24 (1985) 346.
[75] L. Barloy, J.P. Lallier, P. Battioni, D. Mansuy, Y. Piffard, M. Tournous, J.B. Valim, W. Jones,
New J. Chem. 16 (1992) 71.
[76] M.E. Pérez–Bernal, R. Ruano-Casero, T.J. Pinnavaia, Catal. Lett. 11 (1991) 55.
[77] S. Fukuzumi, S. Mochizuki, T. Tanaka, Isr. J. Chem. 23 (1987 – 1988) 29.
[78] L. Gaillon, F. Bedioui, J. Devinck, P. Battioni, J. Electroanal. Chem. 347 (1993) 435.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 117
[79] H. Van Damme, M. Crespin, F. Obrecht, M.I. Cruz, J.J. Fripiat, J. Colloid Interface Sci. 66 (1978)
43.
[80] H. Kameyama, H. Suzuki, A. Amano, Chem. Lett. (1988) 1117.
[81] C. Mousty, S. Therias, C. Forano, J.P. Besse, J. Electroanal. Chem. 374 (1994) 63.
[82] S. Therias, C. Mousty, Appl. Clay Sci. 10 (1995) 147.
[83] J. Martinsen, J.L. Stanton, R.L. Greene, J. Tanaka, B.M. Hoffman, J.A. Ibers, J. Am. Chem. Soc.
107 (1985) 6915.
[84] T.J. Pinnavaia, Adv. Chem. Ser. 245 (1995) 283.
[85] F. Bedioui, Coord. Chem. Rev. 144 (1995) 39.
[86] I.Y. Park, K. Kuroda, C. Kato, Chem. Lett. (1989) 2057.
[87] K. Sakoda, K. Kominami, M. Iwamoto, Jpn. J. Appl. Phys. 27 (1988) L1304.
[88] S. Bonnet, L. Bigey, C. Forano, A. de Roy, J.P. Besse, P. Maillard, M. Momenteau, in: M.L.
Occelli, H. Kessler (Eds.), Synthesis of Porous Materials: Zeolites, Clays and Nanostructures,
Marcel Dekker, New York, 1997, pp. 627 – 640.
[89] S. Bonnet, C. Forano, A. de Roy, J.P. Besse, P. Maillard, M. Momenteau, Chem. Mater. 8 (1996)
1962.
[90] A. Stone, E.B. Fleisher, J. Am. Chem. Soc. 90 (1968) 2735.
[91] H. Tagaya, A. Ogata, T. Kuwahara, S. Ogata, M. Karasu, J. Kadokawa, K. Chiba, Microporous
Mater. 7 (1996) 151.
[92] R.A. Schoonheydt, L. Heughebaert, Clay Miner. 27 (1992) 91.
[93] M. Chibwe, T.J. Pinnavaia, J. Chem. Soc. Chem. Commun. (1993) 278.
[94] M. Chibwe, L. Ukrainczyk, S.A. Boyd, T.J. Pinnavaia, J. Mol. Catal. A: Chem. 113 (1996) 249.
[95] K.A. Carrado, J.E. Forman, R.E. Botto, R.E. Winans, Chem. Mater. 5 (1993) 472.
[96] I.Y. Park, K. Kuroda, C. Kato, J. Chem. Soc. Dalton Trans. (1990) 3071.
[97] I.J. Shannon, T. Maschmeyer, G. Sankar, J.M. Thomas, R.D. Oldroyd, M. Sheehy, D. Madill,
A.M. Waller, R.T. Townsend, Catal. Lett. 44 (1997) 23.
[98] V.I. Iliev, A.I. Ileva, L.D. Dimitrov, Appl. Catal. A: Gen. 126 (1995) 333.
[99] V. Iliev, J. Mol. Catal. 85 (1993) L269.
[100] J.M. Assour, W.K. Kahn, J. Am. Chem. Soc. 87 (1965) 207.
[101] J.F. Boas, P.E. Fielding, A.G. McKay, Austr. J. Chem. 27 (1974) 7.
[102] V. Iliev, A. Andreev, D. Wohrle, G. Schulz-Ekloff, J. Mol. Catal. 66 (1991) L5.
[103] A. Skorobogaty, T.D. Smith, J. Mol. Catal. 16 (1982) 131.
[104] J. Zwart, H.C. van der Weide, N. Broeker, C. Rummens, G.C.A. Schuit, A.L. German, J. Mol.
Catal. 3 (1977–1978) 151.
[105] L. Ukrainczyk, M. Chibwe, T.J. Pinnavaia, S.A. Boyd, J. Phys. Chem. 98 (1994) 2668.
[106] R. Allmann, Chimia 24 (1970) 99.
[107] M.B. McBride, Clays Clay Miner. 27 (1979) 97.
[108] E.P. Giannelis, Chem. Mater. 2 (1990) 627.
[109] J. Subramanian, in: K.M. Smith (Ed.), Porphyrins and Metalloporphyrins, Elsevier, Amsterdam,
1975, p. 568.
[110] L. Ukrainczyk, M. Chibwe, T.J. Pinnavaia, S.A. Boyd, Environ. Sci. Technol. 29 (1995) 439.
[111] U.E. Krone, R.K. Thauer, H.P.C. Hogenkamp, Biochemistry 28 (1989) 4908.
[112] U.E. Krone, K. Laufer, R.K. Thauer, H.P.C. Hogekamp, Biochemistry 28 (1989) 10061.
[113] C.J. Gantzer, L.P. Wackett, Environ. Sci. Technol. 25 (1991) 715.
[114] N. Assaf-Amid, K.F. Hayes, T.M. Vogel, Environ. Sci. Technol. 28 (1994) 246.
[115] G.M. Klecka, S.J. Gonsior, Chemosphere 3 (1984) 391.
[116] Y.C. Yang, J.R. Ward, R.P. Seiders, Inorg. Chem. 24 (1985) 1765.
[117] M. Zikmund, K. Putyera, K. Hrnciarova, Chem. Papers 50 (1996) 262.
[118] L. Gaillon, F. Bedioui, J. Devynck, P. Battioni, J. Electroanal. Chem. 347 (1993) 435.
[119] D.S. Robins, P.K. Dutta, Langmuir 12 (1996) 402.
[120] R.F. Pasternack, L. Francesconi, D. Raff, E. Spiro, Inorg. Chem. 12 (1973) 2606.
[121] K. Kalyanasundaram, M. Naumann-Spallart, J. Phys. Chem. 86 (1982) 5163.
[122] J. Wang, Y. Tian, R.-C. Wang, J.L. Colón, A. Clearfield, in: R.L. Bedard, T. Bein, M.E. Davis,
J. Garces, V.A. Maroni, G.D. Stucky (Eds.), Synthesis/Characterization and Novel Applications
of Molecular Sieve Materials, Materials Research Society, Pittsburgh, 1991, pp. 63 – 80.
118 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
[123] T. Tatsumi, K. Yamamoto, Trans. Mater. Res. Soc. Jpn. 15 (1994) 141.
[124] C.W. Hu, Q.L. He, E.B. Wang, Progr. Nat. Sci. 6 (1996) 524.
[125] S. Miyata, T. Kumura, H. Hattori, K. Tanabe, Nippon Kagaku Zasshi 92 (1971) 514 (in
Japanese).
[126] C. Misra, A.J. Perrotta, Clays Clay Min. 40 (1992) 145.
[127] K. El Malki, A. de Roy, J.P. Besse, Eur. J. Solid State Inorg. Chem. 26 (1989) 339.
[128] C. Depège, C. Forano, A. de Roy, J.P. Besse, Mol. Cryst. Liq. Cryst. 244 (1994) 161.
[129] L. Bigey, C. Depège, A. de Roy, J.P. Besse, J. Phys. IV France, C2, 7 (1997) 949.
[130] S. Yamanaka, T. Sako, K. Seti, M. Hattori, Solid State Ionics 53 – 56 (1992) 527.
[131] C. Forano, A. de Roy, C. Depège, M. Khaldi, F.Z. El Metoui, J.P. Besse, in: M.L. Occelli, H.
Kessler (Eds.), Synthesis of Porous Materials: Zeolites, Clays and Nanostructures, Marcel Dekker,
New York, 1997, pp. 607–625.
[132] E. Suzuki, Y. Ono, Bull. Chem. Soc. Jpn. 61 (1988) 1008.
[133] H. Shimada, K. Saito, Nippon Kagaku Kaishi (1997) 335 (in Japanese).
[134] T. Kwon, G.A. Tsigdinos, T.J. Pinnavaia, J. Am. Chem. Soc. 110 (1988) 3653.
[135] J. Twu, P.K. Dutta, J. Phys. Chem. 93 (1989) 7863.
[136] W.P. Griffith, T.D. Wolkins, J. Chem. Soc. A (1966) 1087.
[137] W.P. Griffith, P.J.B. Lesniak, J. Chem. Soc. A (1969) 1066.
[138] A. Bhattacharyya, D.B. Hall, T.J. Barnes, Appl. Clay Sci. 10 (1995) 57.
[139] J. Twu, P.K. Dutta, J. Catal. 124 (1990) 503.
[140] R. Gopal, C. Calvo, Acta Crystallogr. Sect. B 30 (1974) 2491.
[141] C. Depège, L. Bigey, C. Forano, A. de Roy, J.P. Besse, J. Solid State Chem. 126 (1996) 314.
[142] K.S. Han, L. Guerlou-Demourgues, C. Delmas, Solid State Ionics 84 (1996) 227.
[143] K.S. Han, L. Guerlou-Demourgues, C. Delmas, Solid State Ionics 98 (1996) 85.
[144] M. Doeuff, T. Kwon, T.J. Pinnavaia, Synthetic Metals 34 (1989) 609.
[145] G.M. Woltermann, US patent 4,454,244, Ashland Oil Co., June 12, 1984.
[146] C. Barriga, W. Jones, P. Malet, V. Rives, M.A. Ulibarri, Inorg. Chem. 37 (1998) 1812.
[147] J. Evans, M. Pillinger, J. Zhang, J. Chem. Soc. Dalton Trans. (1996) 2963.
[148] K. Chibwe, W. Jones, Chem. Mater. 1 (1989) 489.
[149] F. Kooli, W. Jones, Inorg. Chem. 34 (1995) 6237.
[150] F. Kooli, W. Jones, V. Rives, M.A. Ulibarri, J. Mater. Sci. Lett. 16 (1997) 27.
[151] F. Kooli, M.J. Holgado, V. Rives, S. San Roman, M.A. Ulibarri, Mater. Res. Bull. 32 (1997) 977.
[152] E. Narita, P. Kaviratna, T.J. Pinnavaia, Chem. Lett. (1991) 805.
[153] E. Narita, P.D. Kaviratna, T.J. Pinnavaia, J. Chem. Soc. Chem. Commun. (1993) 60.
[154] J. Wang, Y. Tian, R.-C. Wang, A. Clearfield, Chem. Mater. 4 (1992) 1276.
[155] F. Kooli, V. Rives, M.A. Ulibarri, W. Jones, Mater. Res. Soc. Symp. Proc. 371 (1995) 143.
[156] G. Mascolo, O. Marino, Miner. Mag. 43 (1980) 619.
[157] M.A. Ulibarri, F.M. Labajos, V. Rives, W. Kagunya, W. Jones, R. Trujillano, Mol. Cryst. Liq.
Cryst. 244 (1994) 167.
[158] M.A. Ulibarri, F.M. Labajos, V. Rives, W. Kagunya, W. Jones, R. Trujillano, Inorg. Chem. 33
(1994) 2592.
[159] I.C. Chisem, W. Jones, J. Mater. Chem. 4 (1994) 1737.
[160] F. Kooli, V. Rives, M.A. Ulibarri, Inorg. Chem. 34 (1995) 5114.
[161] F. Kooli, V. Rives, M.A. Ulibarri, Mater. Sci. Forum. 152 – 153 (1994) 375.
[162] O. Clause, B. Rebours, E. Merlen, F. Trifiro, A. Vaccari, J. Catal. 133 (1992) 231.
[163] F. Kooli, V. Rives, M.A. Ulibarri, Inorg. Chem. 34 (1995) 5122.
[164] J. Guo, Q.Z. Jiao, G. Xiong, H.J. Lu, D.Z. Jiang, E.Z. Min, Chin. Chem. Lett. 7 (1996) 531.
[165] W. Kagunya, W. Jones, Appl. Clay Sci. 10 (1995) 95.
[166] V. Rives, F.M. Labajos, M.A. Ulibarri, P. Malet, Inorg. Chem. 32 (1993) 5000.
[167] P. Malet, J.A. Odriozola, F.M. Labajos, V. Rives, M.A. Ulibarri, Nucl. Instr. Methods Phys. Res.
B 97 (1995) 16.
[168] F. Kooli, I. Crespo, C. Barriga, M.A. Ulibarri, V. Rives, J. Mater. Chem. 6 (1996) 1199.
[169] A. Corma, J.M. López-Nieto, N. Paredes, M. Pérez, Appl. Catal. 97 (1993) 159.
[170] A. Corma, J.M. López-Nieto, N. Paredes, Appl. Catal. 104 (1993) 161.
V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120 119
[171] M.A. Chaar, D. Patel, M.C. Kung, H.H. Kung, J. Catal. 105 (1987) 483.
[172] F. Roozeboom, J. Medema, P.J. Gellings, Z. Phys. Chem. (Wiesbaden) 111 (1978) 215.
[173] R.L. Coustumer, B. Taouk, M. le Meur, E. Payen, M. Guelton, J. Grimblot, J. Phys. Chem. 92
(1988) 1230.
[174] E. López Salinas, Y. Ono, Bull. Chem. Soc. Jpn. 65 (1992) 2465.
[175] F. Kooli, C. Martin, V. Rives, Langmuir 13 (1997) 2303.
[176] C. Delmas, Y. Borthomieu, J. Solid State Chem. 104 (1993) 345.
[177] M. Ménétrier, K.S. Han, L. Guerlou-Demourgues, C. Delmas, Inorg. Chem. 36 (1997) 2441.
[178] S. Onodera, Y. Ikegami, Inorg. Chem. 19 (1980) 615.
[179] M. del Arco, C. Martin, I. Martin, V. Rives, R. Trujillano, Spectrochim. Acta Part A 49 (1993)
1575.
[180] L. Pesic, S. Salipurovic, V. Markovic, D. Vucelic, W. Kagunya, W. Jones, J. Mater. Chem. 2
(1992) 1069.
[181] M. del Arco, M.V.G. Galiano, V. Rives, R. Trujillano, P. Malet, Inorg. Chem. 35 (1996) 6362.
[182] M. del Arco, V. Rives, R. Trujillano, P. Malet, J. Mater. Chem. 6 (1996) 1419.
[183] L. Lowenstein, Am. Miner. 39 (1954) 92.
[184] K. Fuda, K. Suda, T. Matsunaga, Chem. Lett. (1993) 1479.
[185] V. Rives, M.A. Ulibarri, A. Montero, Appl. Clay Sci. 10 (1995) 83.
[186] B.I. Intorre, A.E. Martell, J. Am. Chem. Soc. 82 (1960) 358.
[187] F. Rey, V. Fornés, J.M. Rojo, J. Chem. Soc. Faraday Trans. 88 (1992) 2233.
[188] T. Hibino, A. Tsunashima, Chem. Mater. 9 (1997) 2082.
[189] J. Twu, P.K. Dutta, Chem. Mater. 4 (1992) 398.
[190] D. Levin, S.L. Soled, J.Y. Ying, Chem. Mater. 8 (1996) 836.
[191] D. Levin, S.L. Soled, J.Y. Ying, ACS Symp. Ser. 622 (1996) 237.
[192] M.P. Pope, Heteropoly and Isopoly Oxometalates, Springer-Verlag, New York, 1983.
[193] C. Preyssler, Bull. Soc. Chim. Fr. 1 (1970) 30.
[194] T. Kwon, T.J. Pinnavaia, Chem. Mater. 4 (1989) 381.
[195] Y.Y. Liu, C.W. Hu, Z.P. Wang, J.Y. Zhang, E.B. Wang, Sci. China Ser. B Chem. 39 (1996) 86 (in
Chinese).
[196] E.D. Dimotakis, T.J. Pinnavaia Jr., in: E.W. Corcoran Jr., M.J. Ledoux (Eds.), Synthesis and
Properties of New Catalysts: Utilization of Novel Materials Components and Synthetic Tech-
niques, Materials Research Society, Pittsburgh, 1990, pp. 77 – 80.
[197] E.D. Dimotakis, T.J. Pinnavaia, Inorg. Chem. 29 (1990) 2393.
[198] S.K. Yun, V.R.L. Constantino, T.J. Pinnavaia, Microporous Mater. 4 (1995) 21.
[199] H.C.B. Hansen, R.M. Taylor, Clay Miner. 26 (1991) 311.
[200] M.R. Weir, J. Moore, R.A. Kydd, Chem. Mater. 9 (1997) 1686.
[201] J. Guo, T. Sun, J.P. Shen, D.Z. Jiang, E.Z. Min, Chem. J. Chin. Univ. 16 (1995) 512 (in Chinese).
[202] E. Serwicka, P. Nowak, K. Bahranowski, W. Jones, F. Kooli, J. Mater. Chem. 7 (1997) 1937.
[203] R.N. Hua, Q.J. Shan, J.A. Gong, L.Y. Qu, Chem. J. Chin. Univ. 17 (1996) 1500 (in Chinese).
[204] R.N. Hua, Q.J. Shan, B.Z. Zhao, Y.H. Wang, B.L. Li, L.Y. Qu, Acta Chim. Sin. 55 (1997) 773
(in Chinese).
[205] R.S. Weber, P. Gallezot, F. Lefebvre, S.L. Suib, Microporous Mater. 1 (1993) 223.
[206] C.-W. Hu, Q.-L. He, Y.-H. Zhang, Y.-Y. Liu, Y.-F. Zhang, T.-D. Tang, J.-Y. Zhang, E.-B. Wang,
J. Chem. Soc. Chem. Commun. (1996) 121.
[207] C.W. Hu, Y.Y. Liu, Z.P. Wang, E.B. Wang, Acta Chim. Sin. 55 (1997) 49 (in Chinese).
[208] C.W. Hu, X. Zhang, Q.L. He, E.B. Wang, S.W. Wang, Q.L. Guo, Transit. Metal Chem. 22 (1997)
197.
[209] E. López Salinas, P. Salas Castillo, Y. Ono, Mater. Res. Soc. Symp. Proc. 371 (1995) 163.
[210] S.K. Yun, T.J. Pinnavaia, Inorg. Chem. 35 (1996) 6853.
[211] R.G. Finke, M.V. Droege, P.J. Domaille, Inorg. Chem. 26 (1987) 3886.
[212] C. Rocchiccioli-Deltcheff, R. Thouvenot, R. Franck, Spectrochim. Acta Part A 32 (1976) 587.
[213] T. Kwon, T.J. Pinnavaia, J. Mol. Catal. 74 (1992) 23.
[214] J. Guo, Q.Z. Jiao, J.P. Shen, H.J. Lu, D. Liu, D.Z. Jiang, E.Z. Min, Acta Chim. Sin. 54 (1996)
357 (in Chinese).
120 V. Ri6es, M. Angeles Ulibarri / Coordination Chemistry Re6iews 181 (1999) 61–120
[215] M.A. Drezdzon, in: R. Terry, K. Baker, L.L. Murrel (Eds.), Novel Materials in Heterogeneous
Catalysis, American Chemical Society, Washington, DC, 1990, pp. 141 – 148.
[216] K. Putyera, J. Jagiello, T.S. Bandosz, J.A. Schwarz, J. Chem. Soc. Faraday Trans. 92 (1996) 1243.
[217] T. Tatsumi, K. Yamamoto, H. Tajima, H. Tominaga, Chem. Lett. (1992) 815.
[218] T. Tatsumi, H. Tajima, K. Yamamoto, H. Tominaga, in: L. Guczi, F. Solymosi, P. Teteny (Eds.),
New Frontiers in Catalysis, Elsevier, Amsterdam, 1993, pp. 1703 – 1706.
[219] E. Gardner, T.J. Pinnavaia, Appl. Catal. A: Gen. 167 (1998) 65.
[220] C. Hu, Q. He, Y. Zhang, E. Wang, T. Okuhara, M. Misono, Catal. Today 30 (1996) 141.
[221] J. Guo, Q.Z. Jiao, J.P. Shen, D.Z. Jiang, G.H. Yang, E.Z. Min, Catal. Lett. 40 (1996) 43.
[222] X. Zheng, W. Yue, H. Heming, J. Dazben, in: H. Hattori, M. Misono, Y. Ono (Eds.), Acid-Base
Catalysis II, Elsevier, Amsterdam, Stud. Surface Sci. Catal. 90 (1994) 279.
[223] W. Kagunya, Z. Hassan, W. Jones, Inorg. Chem. 35 (1996) 5970.
[224] B. Keita, A. Belhouari, L. Nadjo, J. Electroanal. Chem. 314 (1991) 345.
[225] B. Keita, A. Belhouari, L. Nadjo, J. Electroanal. Chem. 355 (1993) 235.
[226] T. Sato, T. Wakabayashi, M. Shimada, I&EC Prod. Res. Develop. 25 (1986) 89.
[227] S.W. Rhee, M.J. Kang, H. Kim, C.H. Moon, Environ. Tech. 18 (1997) 231.
[228] L. Châtelet, J.Y. Bottero, J. Yvon, A. Bouchelaghem, Colloids Surf. A. Physicochem. Eng. Asp.
111 (1996) 167.
[229] W. Stumm, J.J. Morgan, Aquatic Chemistry — An Introduction Emphasizing Chemical Equilibria
in Natural Waters, 2nd ed., Wiley, New York, 1981, Ch. 5.
[230] M.J. Kang, S.W. Rhee, H. Moon, V. Neck, T. Fanghanel, Radiochim. Acta 75 (1996) 169.
[231] T. Yamagishi, Y. Oyanagi, E. Narita, Nippon Kagaku Kaishi, (1993) 329 (in Japanese).
[232] T.J. Pinnavaia, M. Rameswaran, E.D. Dimotakis, E.P. Giannelis, E.G. Rightor, Faraday Discuss.
Chem. Soc. 87 (1989) 227.
[233] C.R. Eady, P.F. Jackson, B.F.G. Johnson, J. Lewis, M.C. Malatesta, M. McPartlin, J.H. Nelson,
J. Chem. Soc. Dalton Trans. (1980) 383.
[234] E.P. Giannelis, E.G. Rightor, T.J. Pinnavaia, J. Am. Chem. Soc. 110 (1988) 3880.
[235] F. Bergaya, H. Van Damme, J. Chem. Soc. Faraday Trans. II 79 (1983) 505.
[236] T. Sato, H. Okuyama, T. Endo, M. Shimada, React. Solids 8 (1990) 63.
[237] T. Sato, K. Masaki, T. Yoshiaki, A. Okuwaki, J. Chem. Tech. Biotechnol. 58 (1993) 315.