You are on page 1of 12

Solid State Ionics 98 (1997) 73–84

The intercalation of carboxylic acids into layered double


hydroxides: a critical evaluation and review of the different
methods
Simon Carlino*
Department of Chemistry, University of Reading, Whiteknights, P.O. Box 224, Reading, RG6 2 AD, UK
Received 13 August 1996; accepted 9 December 1996

Abstract

This review paper details the literature which is currently available concerning the intercalation of aliphatic and aromatic
mono- and dicarboxylic acids into layered double hydroxides (LDHs) and their calcined oxides (LDOs). We present a critical
overview of the major synthetic routes and detail, where possible, the main mechanisms of intercalation. The structure of the
intercalates formed in the various methods is also detailed. A discussion of the main techniques of analysis which may be
applied to these materials is also presented along with an empirical method for accurately determining 13 C chemical shifts.

Keywords: Layered double hydroxide (LDH); Layered double oxide (LDO); Carboxylic acids

1. Introduction (Mg 6 Al 2 (OH) 16 CO 3 ? 4H 2 O), the [Mg 6 Al 2 (OH) 16 ]


layer has a charge of 1 2. Overall electrical neu-
Layered double hydroxides, LDHs or the so-called trality is maintained by the presence of anions, which
anionic clays, consist of layers of M II and M III are typically CO 322 or NO 32 [3], in the interlayer
cations which are octahedrally co-ordinated by six region between the metal hydroxide layers. The
oxygen anions, as hydroxides [1]. These layers octahedral layers are able to form infinite 2-D sheets
(henceforth referred to as M II –M III –(OH) x ) exist via edge sharing [4] and can stack via hydrogen
with a similar layered structure to that exhibited by bonding between the hydroxyl groups of adjacent
natural Mg(OH) 2 , also known as brucite. The substi- Mg–Al–(OH) x layers. A simplified view of the
tution of M III cations into the brucite-like hydroxide interlayer of an Mg–Al–CO 3 hydrotalcite-like LDH
layer imparts an overall positive charge on the is shown in Fig. 1. The overall structure of the LDHs
octahedral layer [2]. Thus, in the case of the are dependent upon the ratio of M II –M III –(OH) x
eponymous mineral of the group, hydrotalcite layers to interlayer anionic sheets present in the
structure.
The interlayer area between the adjacent M II –
III
M –(OH) x layers also contains water molecules
*Present address: 23 Yorke Gardens, Reigate, Surrey, RH2 which are able to move randomly about their C 2 axes
9HQ, UK. [5]. The presence of this interlayer water was

0167-2738 / 97 / $17.00  1997 Elsevier Science B.V. All rights reserved


PII S0167-2738( 96 )00619-4
74 S. Carlino / Solid State Ionics 98 (1997) 73 – 84

Table 1
Listing of the main literature on LDH–aliphatic mono- and
dicarboxylic acid intercalation reactions
Starting LDH Carboxylic acid Literature
M II –M III cations anion reference
Ca–Al citrate [47]
Ca–Al diglycolate [47]
Ca–Al maleate [47]
Ca–Al salicylate [47]
Ca–Al tartrate [47]
Fig. 1. A schematic view of the interlayer region of hydrotalcite
Li–Al citrate [47]
showing the carbonate, water molecules and hydrogen bonding Li–Al diglycolate [47]
between the Mg–Al–(OH) x layers. Li–Al laurate [15]
Li–Al maleate [47]
Li–Al myristate [15]
thought to be crucial to the preferential formation of Li–Al palmitate [15]
certain polytypes of LDHs over others in nature [6]. Li–Al salicylate [47]
Li–Al tartrate [47]
Although there have, as yet, been few studies cited Mg–Al acetate [29]
in the literature, there is current interest in layered Mg–Al adipate [29]
double hydroxides or their calcined oxides (hence- Mg–Al caprate [49]
Mg–Al citrate [47]
forth referred to as LDO or layered double oxides) Mg–Al diglycolate [47]
which are intercalated with organic acids. Interest in Mg–Al ethanoate [47]
these materials came about because it was thought Mg–Al glutarate [29]
Mg–Al laurate [15]
that they might form useful intermediates for the Mg–Al maleate [47]
preparation of LDHs which are ‘pillared’ with Mg–Al malonate [29]
macromolecules such as phthalocyanines, dyes or Mg–Al myristate [15]
Mg–Al oxalate [29]
polyoxometallates [7–10]. The scope of the current Mg–Al palmitate [15]
review article does not include details of the incorpo- Mg–Al propionate [29]
ration of photoactive anionic species (such as cinna- Mg–Al salicylate [47]
Mg–Al sebacate [17–19,47]
mate anions) into LDHs or their oxides or industrial Mg–Al succinate [29,47]
applications of alkylcarboxylate pillared LDHs. For Mg–Al tartrate [47]
details of these materials readers should consult Mg–Al n-valerate [29]
Ni–Al sebacate [18]
references ([11–14] respectively). Zn–Al acetate [32]
Zn–Al adipate [27]
Zn–Al citrate [47]
Zn–Al diglycolate [47]
2. Review of the current literature Zn–Al maleate [47]
Zn–Al malonate [27]
Zn–Al oxalate [27]
The more important LDH- / LDO-intercalated Zn–Al salicylate [47]
mono- and dicarboxylic aliphatic acids together with Zn–Al sebacate [27,47]
Zn–Al succinate [27,47]
their literature references are listed in Table 1. Zn–Al tartrate [47]
Similarly, the more important aromatic mono- and Zn–Cr acetate [32]
dicarboxylate intercalated LDHs or LDOs are given Zn–Cr citrate [47]
Zn–Cr diglycolate [47]
in Table 2. Zn–Cr ethanoate [47]
Zn–Cr maleate [47]
Zn–Cr salicylate [47]
2.1. Synthesis of carboxylic acid intercalated Zn–Cr sebacate [47]
LDHs Zn–Cr succinate [47]
Zn–Cr tartrate [47]

The five main experimental methods which may M II –M III refers to the starting LDH or LDO. For example: Mg–Al
is, in most cases, Mg–Al–CO 3 LDH or its calcined oxide.
be used in the preparation of carboxylic acid interca- Reference numbers refer to references cited at the end of this
lated LDHs or LDOs are summarised in Scheme 1. article.
S. Carlino / Solid State Ionics 98 (1997) 73 – 84 75

Table 2
List of the main literature on LDH–aromatic mono- and dicarboxylic acid intercalation reactions
Starting LDH Carboxylic acid Literature
M II –M III cations anion reference
Ca–Al anthranilate [47]
Ca–Al benzoate [47]
Ca–Al phthalate [47]
Ca–Al terephthalate [47]
Li–Al anthranilate [47]
Li–Al benzoate [47]
Li–Al phthalate [47]
Li–Al terephthalate [47]
Mg–Al anthranilate [47]
Mg–Al benzoate [47,51,56]
Mg–Al benzenehexacarboxylate [56]
Mg–Al benzenepentacarboxylate [56]
Mg–Al 1,2,4,5-benzenetetracarboxylate [56]
Mg–Al 1,3,5-benzenetricarboxylate [56]
Mg–Al 1-naphthoate [51]
Mg–Al phthalate [47]
Mg–Al terephthalate [17,47,51,56]
Zn–Al anthranilate [47]
Zn–Al benzoate [32,47]
Zn–Al phthalate [47]
Zn–Al terephthalate [32,47]
Zn–Cr anthranilate [47]
Zn–Cr benzoate [47]
Zn–Cr phthalate [47]
Zn–Cr terephthalate [32,47]
M II –M III refers to the starting LDH or LDO as in Table 1.
Reference numbers refer to references cited at the end of this article.

Brief details of these synthetic methods are presented acid into an LDH (1,12-dodecanedicarboxylic acid
below. pillared Mg–Al LDH) was made by Reichle [16]
who used what is termed a ‘coprecipitation reaction’.
This method requires the addition of an M II / M III salt
2.2. The direct ion exchange method
solution to a base solution (or vice versa) containing
the carboxylic acid as the anion. The solution is then
In the direct ion exchange method of intercalation,
crystallised using the same (hydrothermal) method
the incorporation of carboxylic acid anions into the
as employed in the preparation of the LDH host
LDH is achieved by shaking the LDH in a suitably
material [2,16].
concentrated solution of the desired carboxylic acid
Since that time a number of authors have used
or its salt. A modified version of this method was
modified versions of this technique. Probably the
used more recently by Borja et al. [15] who success-
best known example of the use of the coprecipitation
fully intercalated Mg–Al–Cl and Li–Al–Cl LDHs
technique is detailed in the paper by Drezdzon [17]
by shaking them with aliquots of the desired acid in
who termed the reaction a ‘digestion reaction’.
alcoholic solution.
Drezdzon made use of the coprecipitation reaction to
synthesise a number of organic acid intercalated
2.3. The coprecipitation method LDHs, however, he states that this method could not
be used effectively to prepare a sebacic acid (seba-
The earliest suggestion of intercalating an organic cate) intercalated LDH.
76 S. Carlino / Solid State Ionics 98 (1997) 73 – 84

Scheme 1.

2.4. The rehydration method preparing carboxylic acid exchanged LDHs. The
method uses the calcined LDH (referred to as LDO)
This reaction route was pioneered in the late 1980s as its precursor. This LDO is added to solution
[18] and is by far the most common method of containing the carboxylic acid or its salt. The mix-
S. Carlino / Solid State Ionics 98 (1997) 73 – 84 77

ture is then either simply left to stir or refluxed for an carboxylic acid or its salt. This method was used by
appropriate length of time. In order to obtain best Pinnavaia [21] who intercalated nonanoic and adipic
results the stirring and / or refluxing should be carried (hexanedioic) acids into a manasseite-like Mg–Al
out under a nitrogen atmosphere. LDH using a combination of rehydration to preswell
the LDO followed by glycerol intercalation of the
2.5. The thermal or ‘ melt’ reaction method desired acid salt.

Recently an alternative reaction route by which


suitable organic acids may be intercalated into either 3. Mechanisms of intercalation and the
LDHs or LDOs has been developed [19]. This structure of the reaction products
reaction requires the intimate mixing of the organic
acid guest with either the LDH or LDO host prior to 3.1. The direct ion exchange mechanism
heating at a ramp-rate not greater than 108C min 21
up to a temperature which is approximately 108C This mechanism relies upon the direct exchange of
above the melting point of the particular mono- or the LDH’s interlayer anionic species with the desired
dicarboxylic acid used. The mixture is then held at acid species. Our own experiments indicate that
this temperature for ca. 8 h in order to attain an carboxylic acids do not readily undergo intercalation
equilibrium and then cooled to room temperature at a into the hydrotalcite-like Mg–Al–CO 3 LDH [6,19]
ramp-rate of 108C min 21 . The products are then owing to the charge balancing carbonate anion which
washed in hot ethanol and stored in a dessicator is tenaciously held in the LDH interlamellar region.
under an inert atmosphere. (NOTE: readers should These findings are confirmed by another author
note that no toxicological studies have been carried [22] who noted that the exchange of the interlayer
out on thermal reactions and the author accepts no carbonate anions was not greatly affected by the
responsibility, whatsoever, for the safety or effective- direct exposure of uncalcined LDHs to aqueous
ness of this reaction method.) solutions of common anionic species. The following
sequence of exchange for both mono- and divalent
2.6. Glycerol effected exchange anions in LDHs has been derived [23–27]:

CO 22 22
Apart from the methods mentioned above, the 3 4 SO 4 4
only other method listed in the literature for inter- OH . F . Cl . Br 2 . NO 2
2 2 2 2
3 .I
calating an organic acid into the interlayer of a
layered double hydroxide involves the use of gly- (where the divalent anions are more strongly held in
cerol or other glycols 1 as exchange media. One of the LDH interlayer than the monovalent anions;
the first reported uses of glycerol enhanced exchange carbonate is held most tenaciously of all).
in LDHs was given by Hansen and Taylor [20] who The above shows the derived exchange sequence
used this method to exchange the interlayer carbon- of mono- and divalent anions into LDHs (from Refs.
ate present in samples of natural pyroaurite (the [23–27]).
Mg–Fe–CO 3 LDH) and synthetic hydrotalcite for The direct ion exchange mechanism has more
SO 422 , NO 32 or Cl 2 anions. Glycerol exchange may recently been successfully modified by Dutta et al.
be carried out via two different methods. Firstly, the who reacted uncalcined LDHs with long-chain
LDH may be heated in glycerol vapour in order to monocarboxylic acids in alcoholic solution [15]. No
preswell its interlayer region. The preswelled LDH is formal mechanism was proposed for this reaction;
then reacted with an appropriate solution of the however, the authors noted the presence of several
carboxylic acid to be exchanged. different species (as evidenced by FTIR spectros-
The second method requires the LDH or LDO to copy) in the reaction products which, in turn, sug-
be stirred with a glycerolic solution of the desired gested that the reaction proceeded in a similar
manner to that of glycerol intercalation (vide supra)
1
According to Ref. [20], glycerol is better than other glycols (for – i.e. the alcohol molecules probably preswell the
example ethylene glycol) for preswelling treatments in LDHs. LDH interlayer regions.
78 S. Carlino / Solid State Ionics 98 (1997) 73 – 84

3.2. The coprecipitation reaction mechanism It is known that brucite, Mg(OH) 2 , is able to
expand in the direction of the c-axis (i.e. perpen-
The coprecipitation reaction relies upon the in situ dicular to the lamellae) from 4.75 A˚ at 208C to 4.81
formation of the LDH M II –M III –(OH) x octahedral ˚A at 2028C when heated [38]. The difference be-
layers around the ‘pillaring’ species in solution. The tween brucite and the Mg–Al–CO 3 LDH, aside from
LDH layers are thought to form in a similar manner aluminium substitution, is that the latter contains
to those of the parent M II –M III –X LDH. tenaciously held interlayer carbonate anions together
The method of coprecipitation has, more recently, with some mobile interlayer water molecules. The
been used by Slade [28] in the synthesis of a thermal decomposition of LDHs has been well
terephthalate pillared copper-LDH with the general documented [39–42]. It is known that the interlayer
formula Cu 4 M 2III (OH) 12 [ 2 OOC–C 6 H 4 –COO 2 ]? space in hydrotalcite-like materials still provides a
nH 2 O. This material gave a basal spacing of 15 A ˚ reactive environment upon the gentle heating associ-
and a gallery height of ca. 10.2 A. ˚ ated with conventional (solution) ion exchange re-
actions [43]. Calcination studies on the LDH indi-
cated that the oxide material produced via a slow
3.3. The rehydration reaction mechanism
heating process showed less of a deviation from the
original structure of the octahedral layers than that
The most commonly used method of intercalating
produced by flash calcination [6,42]. The same study
organic acids into LDHs relies upon the so-called
also found (by using in situ variable temperature
‘memory effect’ of the calcined LDH [29–31]. It 27
Al MAS NMR) that the LDH starts to loose this
was found [12,18] that the LDO, when placed in an
interlayer carbonate at a lower temperature than
aqueous solution containing 0.05 M sodium carbon-
traditionally expected. It was, therefore, decided to
ate, was able to rehydrate and revert back to a
try and exploit these two features and slowly heat the
structure which was similar to that of the parent
Mg–Al–CO 3 LDH in order to try to expand the
LDH. Using this principle, calcined LDHs have been
weak bonding in the c-direction of the LDH layer
intercalated by rehydration in solutions of carboxylic
structure [43].
acids [18]. Following the intercalation of sebacic
acid by this method other groups have studied the
3.5. Mechanism of intercalation using glycerol as
uptake of both mono- and dicarboxylic acids by
an exchange medium
calcined LDHs in aqueous solution [29,32].
Several mechanisms have been postulated for the
3.4. Thermal or ‘ melt’ reaction mechanism intercalation of organic species following preswel-
ling treatment with glycerol or other glycols [20].
The process of thermal intercalation has been Probably the most plausible mechanism of those
known for some time. For example, the use of suggested so far involves penetration of the glycerol
solid-state synthesis techniques for the preparation of molecules into the metal hydroxide interlayers result
a wide variety of oxide-type materials is well in a decrease in the hydrogen bonding of the
documented [33]. The selective solid-state intercala- interlayer carbonate anions with the layer hydroxyl
tion of cis- and trans-isomers of organic acids groups. This decrease in hydrogen bonding enables
(maleic, methylmaleic; fumaric and methylfumaric the CO 223 anions to be readily exchanged with the
acids) into montmorillonites was studied by Kato anions outside the interlayer region (i.e. sebacate
[34]. The selectivity of the cis-form of the acids etc.).
(maleic and methylmaleic acids) was thought to A second mechanism relies on the formation of
derive from the ability of the cis-isomers to form negatively charged glycerolate anions (CH 2 OH–
chelate complexes with Al(III) of the montmorillo- CHOH–CH 2 O 2) which react with the interlayer
nite layers. The same authors also studied the carbonate of the LDH and are then exchanged out by
sorption of acrylamide [35], n-alkylamines [36] and the pillaring species in solution outside the interlayer
2,29-bipyridine [37] onto montmorillonites. region.
S. Carlino / Solid State Ionics 98 (1997) 73 – 84 79

Other mechanisms suggest a decrease in the


positive layer charge. Certainly, in the case of
phosphonic acids [44] heated with the Mg–Al–CO 3
LDH there is some evidence to suggest that the
octahedral layers of the LDH may not be as stable as
expected. This may also apply to heating with
glycerol.
In reality, a combination of these postulated Fig. 2. A schematic view of the simple intercalation of a
mechanisms is most probably responsible for the carboxylate into an LDH – typically via direct ion exchange. (See
anion exchange phenomenon with glycerol. It is main text for details of the parameters ‘x’ and ‘y’.)
known that the temperature of glycerol treatment is
important [20]. For example, treatment of Mg–Al– these materials are, almost without exception, X-ray
CO 3 LDH with liquid glycerol gives a d 003 of 9.7 A˚ amorphous microcrystalline solids and, therefore, a
(an increase in interlayer spacing of ca. 2 A). ˚ variety of phases (including regions of unintercalated
Corresponding treatment with glycerol vapour gives host LDH) may co-exist in the reaction products.
an increase in interlayer spacing of 6.7 A˚ (d 003 of Typically, the PXRD pattern of a carboxylate inter-
˚
14.4 A). These correspond to one or two layers of calated LDH will have as its three major reflections:
glycerol intercalate respectively. (i) the large basal spacing (d 003 ); (ii) the half-height
harmonic (d 006 ) (where d 006 ¯ ]12 d 003 ) and (iii) the
d 009 reflection (where d 009 ¯ ]13 d 003 ). The remaining
4. Methods of analysis of intercalation reaction reflections can be indexed as either lesser reflections
products of the intercalated phase or from areas of uninterca-
lated LDH.
4.1. Powder X-ray diffraction studies The results of the studies by Lagaly et al. [46–48]
show that monocarboxylate intercalated LDHs have
4.1.1. Interpretation of PXRDs: assumptions and a more complex structure than that of their corre-
models sponding dicarboxylate intercalates. This is probably
Overall, the process of intercalation of carboxylic because the dicarboxylates with two terminal carbox-
acids into LDHs or LDOs may be summarised by yl functional groups, have a limited number of
Fig. 2. The unintercalated host LDH has (in most orientations between contiguous LDH layers. Where-
cases) a basal spacing of ca. 7.6 A ˚ and a corre- as, in the case of the monocarboxylates, bonding
2 ˚
sponding interlayer distance of 2.9 A. Upon interca- through the carboxyl group to an M II –M III –(OH) x
lation each of these parameters (in the direction of layer only fixes one end of the molecule, allowing
the c-axis) increases to some new value designated y the terminal methyl group to exhibit a hydrophobic
and x respectively in Fig. 2. interaction with the protons in the LDH Mg–Al–
It may be assumed from the observed PXRD data (OH) x layers [19,49].
that the reflection with the greatest d-spacing value A total of three situations may exist for LDH or
˚
(in angstrom) is the basal spacing of the intercalated LDO carboxylate intercalates. (i) they may form ‘flat
material. In reality this reflection may or may not stacks’ in the LDH / LDO interlayer region; (ii) they
correspond to the reflection with the greatest observ- may form monolayers of carboxylate anions in the
able percentage intensity in the powder pattern as LDH / LDO interlayer and (iii) they may form a
bilayer of carboxylate anions in the interlayer. The
exact nature of these structures also depends upon
2
The interlayer height of either the host LDH or intercalation the functionality of the carboxylate; i.e. if it is a
reaction products is simply found by subtracting a value of 4.77 A ˚
˚ is the
mono- or dicarboxylate. It is also known that aro-
from the observed basal spacing (d L ). Where 4.77 A
thickness of one layer of brucite (natural Mg(OH) 2 ) [45]. This is a matic anions such as the terephthalate salt ( 2 OOC–
good approximation of the thickness of a single Mg–Al–(OH) x C 6 H 4 –COO 2), will not form flat stacks, but will
layer. form pendant groups from the layers which abut in
80 S. Carlino / Solid State Ionics 98 (1997) 73 – 84

the interlayer region (see, for example, Ref. [44] on


phenylphosphonic acid).

4.1.2. Observed PXRD patterns


Typical PXRD patterns for the coprecipitation,
rehydration and thermal methods of intercalation for
an Mg–Al LDH system are shown in Fig. 3(a)–(c)
respectively. Although it is correct to talk of the d 003
(basal spacing) a typical PXRD pattern obtained
from any of the different synthetic methods will
typically show a broad spread of reflections which
make up this main reflection. This indicates that, in
reality, there are several intercalating species at
slightly varying orientations or environments. The
observed maximum d-spacings may also vary slight-
ly from method to method owing to the degree of
hydration of each sample.
An example of a PXRD pattern observed for the
coprecipitation reaction is shown in Fig. 3(a). This
pattern of the Mg–Al–[Sebacate] intercalate shows a
large d 003 of 17.6 A ˚ which corresponds to an
˚ This may be compared
interlayer spacing of 12.8 A.
with a typical pattern for Mg–Al–[Sebacate] ob-
tained by the rehydration method (Fig. 3(b)). This
PXRD pattern has a d 003 of 16.3 A ˚ which corre-
sponds to an interlayer spacing of 11.5 A. ˚ The
˚
half-height harmonic (d 006 ) of 8.1 A is also present
˚ (i.e. ]1 (d 003 )). The
along with the d 009 value of 5.4 A 3
difference between the observed d-spacings of these
two methods is probably accounted for by the
structure of the LDO which may have areas which
are ill-defined or varying degrees of hydration (see
also Section 3.4). The coprecipitated product is also
known to contain no carbonate in the interlayer
region [18] which may explain the difference in
spacing.
A typical example of a PXRD for the Mg–Al– Fig. 3. Powder X-ray diffraction patterns for Mg–Al–[Sebacate]
[Sebacate] system obtained using the thermal or obtained by (a) rehydration method, (b) coprecipitation method
‘melt’ reaction is shown in Fig. 3(c). The maximum and (c) the thermal or ‘melt’ method. All patterns are indexed
using JCPDS Card No. 22-700 (Synthetic Hydrotalcite) and Refs.
spacing obtained for this material compares well [18,19]. Note that pattern (c) contains reflections from a phase
with those of the two previous methods indicating corresponding to a region of unintercalated host LDH (marked as
that the products are the same (i.e. a d 003 of 19.3 A). ˚ ‘*’).
The main differences between this and other methods
are (i) that there is, in this case, no reflection
corresponding to the half-height harmonic (d 006 ) and materials which are truly X-ray amorphous. There
that (ii) there are also some areas of unintercalated was thought to be an interaction between the carbon-
LDH as evidenced by the d 003 of 7.7 A ˚ of the host ate in the LDH interlayer and the carboxylate anions.
LDH. The thermal reaction method, therefore, gives This may explain the relatively large observed basal
S. Carlino / Solid State Ionics 98 (1997) 73 – 84 81

carboxylate anions are, in fact, parallel to the layers


of the LDH (or LDO) as ‘flat stacks’ (see Fig. 4).
In the case of the coprecipitation and rehydration
methods (vide infra) the reaction products are always
monolayers (Fig. 5(a) and (b)) or bilayers (Fig. 6) of
carboxylate anions.

4.1.4. Monolayer structures


Using the terminology of Lagaly [46–48] the
following equation may be used to derive the
interlayer spacing and physical constants such as
slant angle, a, of the carboxylic chains from the
normal:
Fig. 4. Schematic view of the proposed structure of the LDH–
[carboxylate] reaction products with a ‘flat stack’ of carboxylate
d L 5 d 0 1 1.27n sin a 1 d 1 (1)
anions in the LDH interlayer region. where d L 5observed basal spacing in the PXRD
experiment; d 0 5distance between the terminal ion-
ised carboxyl group and the centre of the Mg–Al–
spacing of this material compared to those obtained (OH) x layer; d 1 5distance between the terminal
for the coprecipitation and rehydration methods (vide methyl group and the centre of the LDH Mg–Al–
infra). (OH) x layer; a 5the slant angle of the carboxylate
carbon chains from the normal of the LDH / LDO
4.1.3. Low loading of the LDH or LDO with M II –M III –(OH) x layers.
carboxylic acids The relationship between the number of carbon
When the intercalation reaction takes place with a atoms in the aliphatic carbon chain of the mono-
low percentage of intercalate present in the reaction carboxylic acid, n, and the slant angle, a, is ac-
mixture the product may not contain a phase corre- counted for by the 1.27 and 2.54 constant terms in
sponding to a monolayer or bilayer (vide supra) of Eq. (1) and Eq. (2). This relationship is accepted and
intercalated acid. In the case of the thermal reaction
mechanism (q.v.), when the reaction takes place with
a low initial acid:LDH (LDO) concentration the
observed PXRD patterns suggest that the mono-

Fig. 5. Schematic view of the proposed structure of the LDH– Fig. 6. Schematic view of the proposed structure of the LDH–
[carboxylate] reaction products with mololayer of carboxylate: (a) [carboxylate] reaction products with a full bilayer of carboxylate
hydrated phase with a monolayer of water and (b) dehydrated in the LDH interlayer region. (See main text for details of the
phase. (See main text for details of the parameters d L and d 1 .) parameters d L , d 0 , d 2 , a and n.)
82 S. Carlino / Solid State Ionics 98 (1997) 73 – 84

documented for other systems including zirconium 4.2.2. Interpretation of 13 C MAS NMR spectra
phosphates [50]. The 13 C MAS NMR spectra of the hydrotalcite-
Monocarboxylic acids are able to form two differ- like Mg–Al–CO 3 LDH host material is detailed in
ent types of monolayer. Firstly, they may be hy- the literature [19]. This material shows a single
drated monolayers; i.e. with a minimum of one layer resonance at ca. 1170.6 ppm which has been
of water molecules between the hydrophobic termi- assigned to the interlayer charge balancing carbonate
nal methyl group of the carboxylic acid and the anions of the LDH [6,19]. The 13 C MAS NMR
adjacent LDH M II –M III –(OH) x layer (Fig. 5(a)). spectra of the reaction products will, in most cases,
Secondly, they may be dehydrated monolayers; i.e. contain a signal corresponding to this interlayer
without any water molecules between the terminal carbonate. The exact intensity of this resonance will
methyl group and the M II –M III –(OH) x layer (Fig. depend upon the method of intercalation used and
5(b)). prevalent reaction conditions.
It is possible to assign observed resonances in the
13
4.1.5. Bilayer structures C NMR spectra to the carbons in a carboxylic acid
Bilayer structures may be accurately described by skeleton using an empirical ‘additive-shift’ method
the following equation [49]: of chemical shift calculation [52–54]. The Lin-
deman–Adams Equation used for these calculations
d L 5 d 0 1 2.54n sin a 1 d 2 (2) was of the form [55]:

where d L , d 0 and a are the same as for Eq. (1). The d(i) 5 2 2.5 1 na 1 nb 1 ng 1 nd 1 ne 1 n . . . (3)
value of d 2 corresponds to the distance between the
two facing terminal methyl groups in the bilayer where d(i) is the calculated chemical shift (in ppm
structure. This is always taken as 3 A.˚ from TMS) of the ith carbon atom in the acid chain.
The value of 22.5 is the value for the chemical shift
4.2. Solid-state MAS NMR studies on the reaction of methane. The a, b, g, d and e terms refer to the
products chemical shift parameters of the carbon atoms one,
two, three, four and five bonds removed from the ith
4.2.1. 27 Al MAS NMR spectroscopic studies carbon in question. The values used for these car-
Both one- and two-dimensional 27 Al MAS NMR bons (from reference [55]) are: a, 19.1 ppm; b,
spectroscopic studies on the common hydrotalcite- 19.4 ppm; g, 22.5 ppm; d, 10.3 ppm and e, 10.1
like Mg–Al–CO 3 LDH are detailed in the literature ppm. The term n is simply the number of a, b, g, d
[42]. In this LDH all of the aluminium is present as or e carbon atoms joined to the ith carbon. This
octahedrally co-ordinated species; i.e. as Al(OH) 6 equation does not utilise the so-called ‘corrective
with possibly some Al(O) 6 [6,42]. In the case of the term’ of Grant and Paul [53] for the various steric
monocarboxylate anion intercalated reaction products hindrances associated with branched paraffins.
the thermal intercalation reaction show a single Overall, this method permits the accurate calcula-
resonance at (typically) 18.8 ppm from TMS. This tion of chemical shifts and has been found to be
indicated that the LDH layers were unchanged from diagnostic for both C 2 and C 3 methylene carbons of
their original (octahedral) structure. the dicarboxylic acid carbon skeleton. In particular, it
In contrast, the coprecipitation reaction product enables a distinction to be made between dissociated
showed a complex 1-D 27 Al MAS NMR spectrum and undissociated (i.e. intercalated and ‘free’ acid)
with resonances which could be assigned to both carboxylate groups. Assignment of other chemical
octahedrally and tetrahedrally co-ordinated alumin- shifts (i.e. as C 4 or C 5 ) was found to be tentative.
ium species [51]. The rehydration method showed From our results on mono- and dicarboxylic acids
tetrahedral, octahedral aluminium resonances for it was found that the adjacent carbon to a dissociated
terephthalate, benzoate and naphthoate. Some evi- (i.e. intercalated) carboxyl group showed a chemical
dence was also found for 5-fold co-ordinated alu- shift of ca. 138 ppm in the 13 C spectrum [6,19,49].
minium species [51]. A value of ca. 28 to 29 ppm was obtained for C 3
S. Carlino / Solid State Ionics 98 (1997) 73 – 84 83

methylene carbons in the dissociated acid chain. therefore, best applied to non-carbonate LDHs and
Similarly, it was found that the carbon of the ionised has been found to be effective if carried out in
carboxyl group itself appeared at a chemical shift of alcoholic solution. The thermal reaction method
ca. 182 ppm. produces PXRD amorphous materials which include
some areas of non-intercalated LDH or LDO host
material. This route leads to a variety of products
5. FTIR spectroscopic studies which are dependant upon the initial ratio of acid to
LDH / LDO host material.
Our own results suggest that FTIR studies on the Glycerol effected exchange has been found to be
reaction products are inconclusive. This was thought effective in the exchange of carboxylic acids into
to be because of the fact that, in each case, several LDHs. The method depends upon the exact heating
environments of carboxylates were noted which conditions employed and is thought to depend upon a
added to the complexity of the FTIR spectra. For combination of several mechanisms including the
example, we noted that there may be a difference in formation of glycerolate anions.
spectra obtained for mono- and dicarboxylates re- Powder X-ray diffraction is still the principal
lated to the orientation (i.e. the ‘fixed’ end) of the technique for discerning the structure of the interca-
molecule to the LDH layers which resulted in the lates. The technique of 13 C MAS NMR was also
absence of characteristic group vibrations (i.e. found to be effective at determining the presence of,
[nasym (CO)]. This anomaly may possibly be ex- for example, the interlayer carbonate anions of the
plained by the change in the structure of the LDH LDH. Finally, the use of the ‘additive-shift’ method
metal hydroxide layers resulting in a change in the of 13 C chemical shift calculation was found to be a
FTIR extinction coefficient. Details of the FTIR useful and diagnostic tool for ionised and unreacted
analyses of carboxylate intercalated LDHs are pre- intercalating carboxylate groups.
sented in the paper by El Malki et al. [32].

Acknowledgments
6. Conclusions
The author would like to thank Dr W. Jones of the
This review shows that there are five main types
Department of Chemistry, University of Cambridge
of reaction routes which may be used to intercalate
for his permission to reproduce the PXRD pattern of
both mono- and dicarboxylic acids into layered
the Mg–Al–[Sebacate] LDH prepared via the rehy-
double hydroxides or their oxides. From the com-
dration method shown in Fig. 3(b).
parison of the various methods (principally by
PXRD and NMR studies) it was found that the
coprecipitation and rehydration methods were the
most effective. These two methods gave, on the References
whole, single phase reaction products as indicated by
PXRD. It has also been demonstrated that both 1- [1] A.F. Wells, Structural Inorganic Chemistry, 5th ed., Oxford
and 2-D 27 Al solid-state MAS NMR techniques have University Press, Oxford, 1984, p. 259.
[2] W.T. Reichle, Solid State Ionics 22 (1986) 135.
their role to play in structure elucidation of the
[3] L. Pesic, S. Salipurovic, V. Markovic, D. Vucelic, W.
intercalates. Such studies indicated that there were Kagunya, W. Jones, J. Mater. Chem. 2 (1992) 1069.
many different environments or orientations taken up [4] H.W.F. Taylor, Mineral. Mag. 39 (1973) 377.
by the acid even in environments which PXRD [5] G. Marcelin, N.J. Stockhausen, J.F.M. Post, A. Schutz, J.
studies suggested had only contained a single phase Phys. Chem. 93 (1989) 4646.
[6] S. Carlino, Ph.D. Thesis, University of Reading, 1994.
of reaction product. The direct ion exchange method
[7] E. Narita, P. Kaviratna and T.J. Pinnavaia, Chem. Lett.
was found to be the least effective owing to the (1991) 805.
carbonate anion which is present in the interlamellar [8] T. Kwon, T.J. Pinnavaia, Chem. Mater. 1 (1989) 381.
region of most synthesised LDHs. This method is, [9] H. Hayashi, M.J. Hudson, J. Mater. Chem. 5 (1995) 781.
84 S. Carlino / Solid State Ionics 98 (1997) 73 – 84

[10] L. Barloy, J.P. Lallier, P. Battioni, D. Mansuy, Y. Piffard, M. [35] C. Kato, M. Ogawa, K. Kuroda, Chem. Lett. (1989) 1659.
Tournoux, J.B. Valim, W. Jones, New J. Chem. 16 (1992) 71. [36] C. Kato, M. Ogawa, T. Hashizume, K. Kuroda, Inorg. Chem.
[11] J. Valim, B.M. Kariuki, J. King, W. Jones, Mol. Cryst. Liq. 30 (1991) 584.
Cryst. 211 (1992) 271. [37] C. Kato, K. Kuroda, M. Ogawa, K. Kato, Clay Sci. 8 (1990)
[12] M. Chibwe, J.B. Valim, W. Jones, in: Multifunctional 31.
Mesoporous Inorganic Solids, NATO-ASI Series C, Vol. 400, [38] S.A.T. Redfern, B.J. Wood, Am. Mineral. 77 (1993) 1129.
eds. M.J. Hudson and C.A.C. Sequira, Van Nostrand [39] G.W. Brindley, S. Kikkawa, Clay. Clay Miner. 28 (1980) 87.
Reinhold, Dordrecht, 1993. ´ J.M. Rojo, J. Chem. Soc., Faraday Trans.
[40] F. Rey, V. Fornes,
[13] E.P. Giannelis, D.G. Nocera, T.J. Pinnavaia, Inorg. Chem. 26 88 (1992) 2233.
(1987) 203. [41] W.T. Reichle, S.Y. Kang, D.S. Everhardt, J. Catal. 101
[14] D.S. Robins, P.K. Dutta, Langmuir 12 (1996) 402. (1986) 352.
[15] M. Borja, P.K. Dutta, J. Phys. Chem. 96 (1992) 5434. [42] M.J. Hudson, S. Carlino, D.C. Apperley, J. Mater. Chem. 5
[16] W.T. Reichle, J. Catal. 94 (1985) 547. (1995) 323.
[17] M.A. Drezdzon, Inorg. Chem. 27 (1988) 4628. [43] A. Vaccari, CEA-PLS Newsl. 3 (1992) 8.
[18] K. Chibwe, W. Jones, J. Chem. Soc., Chem. Commun. [44] S. Carlino, M.J. Hudson, S.W. Husain, J.A. Knowles, Solid
(1989) 926. State Ionics 84 (1996) 117.
[19] S. Carlino, M.J. Hudson, J. Mater. Chem. 4 (1994) 99. [45] J.R. Smyth, D.L. Bish, Crystal Structures and Cation Sites of
[20] H.C.B. Hansen, R.M. Taylor, Clay Miner. 26 (1991) 311. the Rock-forming Minerals, Allen and Unwin, London,
[21] E.D. Dimotakis, T.J. Pinnavaia, Inorg. Chem. 29 (1990) 1988, p. 69.
2393. [46] M. Meyn, K. Beneke, G. Lagaly, Inorg. Chem. 32 (1993)
[22] D.L. Bish, Bull. Mineral. 103 (1980) 170. 1209.
[23] S. Miyata, Clays Clay Miner. 33 (1975) 369. [47] M. Meyn, K. Beneke, G. Lagaly, Inorg. Chem. 29 (1990)
[24] S. Miyata, Clays Clay Miner. 28 (1980) 50. 5201.
[25] S. Miyata, Clays Clay Miner. 31 (1983) 305. [48] H. Kopka, K. Beneke, G. Lagaly, J. Colloid Interfac. Sci.
[26] S. Miyata, A. Okada, Clays Clay Miner. 25 (1977) 14. 123 (1988) 427.
[27] S. Miyata, T. Kumura, Chem. Lett. (1975) 843. [49] S. Carlino, M.J. Hudson, J. Mater. Chem. 5 (1995) 1433.
[28] R.C.T. Slade, CEA-PLS Newsl. 4 (1993) 8. [50] D. O’Hare, in: Inorganic Materials, eds. D.W. Bruce and D.
[29] E. Narita, T. Yamagishi, in: Proc. 9th Int. Clay Conf., eds. O’Hare, John Wiley and Sons, London / New York, 1992.
V.C. Farmer and Y. Tardy, Strasbourg, 1989, p. 145. [51] M. Vucelic and W. Jones, in: Multifunctional Mesoporous
[30] F.M. Labajos, V. Rives, M.A. Ulibarri, in: Multifunctional Inorganic Solids, NATO-ASI Series C, Vol. 400, eds. M.J.
Mesoporous Inorganic Solids, NATO-ASI Series C, Vol. 400, Hudson and C.A.C. Sequira, Van Nostrand Reinhold, Dor-
eds. M.J. Hudson and C.A.C. Sequira, Van Nostrand drecht, 1993.
Reinhold, Dordrecht, 1993. [52] L.P. Lindeman, J.Q. Adams, Anal. Chem. 43 (1971) 1245.
` A. Vaccari, Catal. Today 11 (1991) 173.
[31] F. Cavani, F. Trifiro, [53] D.M. Grant, E.G. Paul, J. Am. Chem. Soc. 86 (1964) 2984.
[32] K. El Malki, M. Guenane, C. Forano, A. de Roy, J.P. Besse, [54] Y.A. Shahab, H.A. Al-Wahab, J. Chem. Soc., Perkin Trans. II
Materials Science Forum 91–93 (1992) 171. (1988) 255.
[33] A. Stein, S.W. Keller, T.E. Mallouk, Science 259 (1993) [55] D.H. Williams, I. Fleming, Spectroscopic Methods in Or-
1558. ganic Chemistry, 1st ed., McGraw-Hill, Maidenhead, 1980,
[34] C. Kato, M. Ogawa, M. Hirata, K. Kuroda, Chem. Lett. p. 37.
(1992) 365. [56] T. Sato, A. Okuwaki, Solid State Ionics 45 (1991) 43.

You might also like