You are on page 1of 12

Journal of Controlled Release 169 (2013) 28–39

Contents lists available at SciVerse ScienceDirect

Journal of Controlled Release


journal homepage: www.elsevier.com/locate/jconrel

Review

Layered double hydroxides as drug carriers and for controlled release of


non-steroidal antiinflammatory drugs (NSAIDs): A review
Vicente Rives ⁎, Margarita del Arco, Cristina Martín
GIR-QUESCAT, Departamento de Química Inorgánica, Universidad de Salamanca, 37008 Salamanca, Spain

a r t i c l e i n f o a b s t r a c t

Article history: Non-steroidal anti-inflammatory drugs constitute one of the groups most widely currently used, but show several
Received 14 February 2013 problems for administration due to low solubility and delivery control. For this reason, several matrices have been
Accepted 30 March 2013 tested to support them in order to overcome these drawbacks. Among them, layered double hydroxides have been
Available online 11 April 2013
used in recent years. The aim of this review is to update the current knowledge and findings on this hybrid system,
namely, layered double hydroxides intercalated with different NSAIDs. The basic nature of the matrix introduces
Keywords:
Layered double hydroxide
an additional advantage, i.e., to decrease ulceration damages. We have focused our review mostly on the prepara-
Drug intercalation tion procedures, as these control, define and determine the performance of the systems in vitro and also in living
Controlled delivery organisms.
Organic–inorganic nanohybrids © 2013 Elsevier B.V. All rights reserved.
NSAID

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2. Intercalation and controlled release of non-steroidal antiinflammatory drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3. Conclusions and further studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

1. Introduction forming infinite sheets; these layers are stacked on top of each other
and held together by weak hydrogen bonds. Some of the cations can be
Layered double hydroxides (LDHs) constitute a broad family of isomorphically substituted by others with similar size, but with higher
lamellar solids which in the last decades have deserved an increasing valence, developing a positive charge in the sheets. Charge balance is
interest because of their applications in different fields [1,2]. They are recovered by intercalating anions between the layers, Fig. 1, where
sometimes named as anionic clays due to the similarities shared with water molecules also exist. The brucite-like layers can be stacked in
cationic clays, or hydrotalcite-like materials, as derived from the natural different ways, leading to different structures [5], the most typical ones
hydroxycarbonate of Mg and Al discovered in Sweden in 1842. Their being rhombohedral (3R symmetry) and hexagonal (2H symmetry).
properties have been reviewed in recent years in different books and Divalent/divalent isomorphical substitution is also possible, as well as
monographs [2–4]. the trivalent/trivalent one; the most commonly found cations in the
Their structure is similar to that of brucite, Mg(OH)2, where hydroxyl layers are Mg2+, Zn2+, Co2+, Ni2+, Cu2+ or Mn2+, and Al3+, Cr3+,
anions are hexagonally close packed and magnesium cations are filling Co3+, Fe3+, V3+, Y3+ or Mn3+. With the exception of Al3+ (0.50 Å),
all octahedral sites every two layers; consequently, edge-sharing octahe- all these elements have similar ionic radii as that of Mg2+ (0.86 Å)
dra of magnesium cations are surrounded by six hydroxyl groups thus accounting for the isomorphic substitution without a substantial
distortion of the structure. Other LDHs containing monovalent and tetra-
valent metal cations have been also synthetised [6–9], as well as systems
containing three or even four different metal cations [10–17].
⁎ Corresponding author. Tel.: +34 923 294489; fax: +34 923 294574. There are no strict limitations to the nature of the interlayer anions,
E-mail address: vrives@usal.es (V. Rives). and systems with many different anionic species are known: simple

0168-3659/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jconrel.2013.03.034
V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39 29

not exceeding 500 °C) are put in contact with solutions containing an-
ions.
- A good anion exchange capacity (AEC), associated to the non-divalent
layer cation, usually higher than that shown by cationic clays, and
ranging between 2 and 4 mEq/g.
Many different synthetic procedures to prepare LDHs have been
described in the literature [37–39]. Selection of one or another greatly
depends on the specific material to be prepared. The most widely
used are the following:
- Coprecipitation, consists of the slow addition of a solution containing
salts of the metal cations in the required molar ratio into a reactor
containing water; simultaneous addition of an alkaline solution
keeps the pH within a selected, narrow range, to precipitate the
mixed hydroxide.
- Anionic exchange is largely applied to obtain LDHs intercalated with
anions of different nature. Strictly speaking this is not a proper syn-
thesis method, but a post-synthesis modification, since it is necessary
to have a LDH precursor, usually a carbonate-free LDH. The feasibility
of exchanging the anions in LDHs depends on the electrostatic inter-
actions between the layers and the interlayer anions. The equilibrium
constants increase as the ionic radius of the anion decreases. The
rate-determining step is the diffusion of the in-going anions within
Fig. 1. Idealised structure of a layered double hydroxide, with interlayer carbonate an- the interlayer. Exchange reactions are usually carried out by stirring
ions. the LDH precursor in a solution containing an excess of the anion to
Reprinted from Coordination Chemistry Reviews, Vol. 181, V. Rives and M.A. Ulibarri, be intercalated; ultrasounds have been also applied to speed up the
Layered double hydroxides (LDH) intercalated with metal coordination compounds
exchange process [40].
and oxometalates, pp. 61–120, Copyright (1999), with permission from Elsevier.
- The reconstruction method is based on the above mentioned memory
effect [32,36]. First a LDH with the desired metal cations in the
brucite-like layer, commonly intercalated with carbonate (which is
inorganic anions (carbonate, nitrate, halides, etc.) [18], organic anions released as CO2) is first prepared. Then the solid is calcined at or
(terephthalate, acrylate, lactate, etc.) [19–21], coordination compounds just below 500 °C (preferable under dynamic inert gas atmosphere
[22,23], polyoxometalates [24–28], biomolecules such as nucleoside to remove CO2) [41,42]. Finally, the mixed oxide formed is stirred
monophosphates (AMP, CMP, GMP or ATP) or even DNA fragments in an aqueous solution of the anions to be intercalated, usually in a
[29–31] have been successfully intercalated. concentration several times higher than that required for a stoichio-
The only restriction concerning the nature of the anions to be inter- metric reaction.
calated makes reference to their size/charge ratio, as large anions with - Hydrothermal and microwave treatments have been applied to pro-
low charge are unable to balance properly (homogeneously) the posi- cess LDHs, improving the crystallinity and other properties of the
tive charge in the layers. Kwon and Pinnavaia [32] claimed that LDHs, especially upon microwave treatment [43–51].
polyoxometalates (POMs) with the Keggin structure with a charge
The fields of applications of LDHs are very broad. Only a few of them
less than −4 are unable to be intercalated between brucite-like layers
will be outlined below.
with a M2+/M 3+ molar ratio close to 2, since these molecules are not
spatially able to balance the host layer charge. - Materials Science. LDHs are used as additives in polyvinyl chlo-
The M2+/M3+ molar ratio usually ranges between 2 and 4, both in all ride polymer (PVC), as they improve its strength and retard
natural occurring LDHs and in most of the synthetic ones. The synthesis darkening of the polymer [52,53]. LDHs enhance the mechanical
of these sorts of solids with molar ratios outside this range has been properties of polymer matrices and may provide them with
sometimes claimed [33], although it is difficult to know the accurate other properties such as colour [54], flame retardant [55,56] or
layer composition, since for such extreme ratios formation of dispersed barrier effect [57], and others [58,59]. Hydrotalcite-type mate-
amorphous oxides cannot be discarded. rials have been used as electrode surface modifiers. They are
These materials have deserved a great interest in recent decades more stable under high temperature and oxidizing conditions
in different fields because of their specific properties, namely: than organic polymers [60,61].
- Water decontamination. The large AEC together with the memory
- Acid–base properties. LDH solids are basic materials with surface effect exhibited by hydrotalcite-type solids, make them good adsor-
basic hydroxyl groups; the basicity of carbonate-intercalated bents for the removal of harmful species in anionic form; hydrotalcites
LDHs has been related to the electronegativity of the layer cations and their calcined products can be used at pH values close to those at
[34,35]. The mixed oxides formed upon thermal decomposition of which pollutants are usually found in the environment [62]. Layered
LDHs are more basic than the original LDHs, due to the presence of double hydroxides were proposed as adsorbents for the capture of
strong oxide basic sites. Moreover, the intercalation of different inorganic anions such as arsenate, chromate or phosphate from
species can give rise to the development of acid sites providing waste water [63–65], and in recent years their use has been extended
systems with unique acid–base properties. also for removal of organic toxic species, such as phenolic compounds,
- Homogeneous mixtures containing well dispersed elements in layer pesticides or even nuclear residues. This issue has been recently
and interlayer domains in a wide range of composition and ratios reviewed [66].
can be prepared, allowing to tailor the properties of these solids. - Separation processes. O'Hare and Lotsch [67] have reported the use of
- The so-called memory effect [32,36], i.e., the ability to recover their these materials for separation or purification of isomeric com-
original layered structure when mixed oxides, previously obtained pounds, due to the differences of affinity shown by the isomers for
upon calcination of some LDHs at moderate temperatures (usually the LDH. The intercalated isomer can be recovered by treatment
30 V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39

with an aqueous carbonate solution, due to the strong affinity of layers protect the intercalated molecules from degradation: once the
carbonate for LDH materials; the material can be regenerated LDH–DNA hybrids have entered into the cell through endocytosis, the
by calcining the carbonate-intercalated solid and hydrating the hydroxide layers are dissolved in the lysosome, with a slightly acid
resulting product. pH, and the encapsulated biomolecules are released inside the cell.
- Catalysis. LDHs are very widely used in catalysis, as they can include The use of these supports or matrices aims to improve the release
a wide number of elements homogeneously dispersed with con- of NSAIDs whichever the clinical stage of the patient or disease, as
trolled proportions, thus allowing the synthesis of tailored catalysts with its use we will be able to maintain high, constant levels of the
[68–70]. Hydrotalcites can be used in heterogeneous catalysis as drug because of an improved pharmacokinetics behaviour; as a result,
synthesised [71], as catalyst supports [72,73] and as catalyst precur- dosage will be more easily improved and the dosing intervals could
sors. Oxides prepared by calcination of hydrotalcites show wide be expanded.
compositional ranges, homogeneous dispersion of the elements, In this review we collect the information published, roughly from
preserved even after moderate thermal treatments or reduction, 2000, on the intercalation of non-steroidal antiinflammatory drugs
high specific surface areas (100–130 m 2/g) or memory effect, (AINEs) in different layered double hydroxides, as well as their con-
making them more attractive than other oxides obtained through trolled release. This is the family of drugs in which insertion in and
conventional methods. They have been used in basic catalysed controlled release from LDHs have been most widely studied recently.
reactions and to prepare fine chemicals and intermediates [74]. We concentrate our efforts on the experimental procedures, as it is
Other reactions studied include oxidative steam reforming of meth- clearly concluded from the literature here cited that minor changes
anol for selective production of hydrogen for fuel cells [75], as seeds in the experimental procedures may lead to important differences
for the catalytic synthesis of carbon nanotubes [17], gas phase hy- in the performance of these systems. On preparing this report, we
drogenation of acetonitrile [76] and selective hydrogenation of acet- have put together those studies concerning a given family of drugs
ylene to ethylene [13,14]. Catalysis by LDHs has been reviewed by for a specific illness or symptom.
Jacobs et al. [77] and more recently by Centi and Perathoner [78], One of the first reviews on this subject was published by
and basic catalysis processes by LDHs have been also summarised Costantino and Nocchetti in 2001 [102], and some other studies
by Tichit and Coq [79] and Figueras [80] for several reactions. The have appeared [112,113], highlighting the relevance of the subject,
application of LDHs with intercalated polyoxometalates in different and the need for further studies, both fundamental and applied.
catalytic processes has been recently reviewed by some of us [81]. Khan and O'Hare [114] published in 2002 a short review on the inter-
- Medicine. LDHs are able to intercalate biologically-active molecules, calation chemistry of LDHs, where they stressed the promising future
a property which has opened their applications in Medicine and of these systems, in particular the work on the intercalation of biolog-
Pharmacy. Due to their high adsorption ability, they are used in ically active guests, foreseeing “… that LDHs will move from being
cosmetics to remove inflammatory exudates, as found in acne used as simple low-cost ion exchange materials to be used as part
vulgaris, to encapsulate organic molecules, or for protection of a sophisticated drug or gene delivery system in patient care or
against solar damage of the skin, providing as well protection therapy”; a clear statement which ten years after has been demon-
against degradation by light, heat or oxygen of labile molecules. strated to be true. Xu and Lu [87] reviewed the use of LDHs as cellular
For instance, p-amino benzoic acid intercalated in Mg, Al or Zn, drug delivery agents; these authors concluded that in addition to the
Al hydrotalcites avoids skin reactions such as irritation, urticaria, huge information then available on the structure and preparation
contact dermatitis, etc., avoiding degradation to carcinogenic procedures of LDHs, as well as the first works on their ability to inter-
nitrosamines and extending its photostability, and increasing calate biomolecules, studies about the interaction of LDHs and cells
the protection in the UV-A range [82]. No cytotoxic results were still lacking, and need further exploration; a similar conclusion
were found when doses lower than 100 mg/kg were injected in was reached in a subsequent review on the same specific subject
Sprague–Dawley mice [83,84]; the safety profile of LDHs and [115].
their biocompatibility have led to the use of LDH–DNA hybrids for
gene transfer and storage [85–89]. LDHs have been also used as ma- 2. Intercalation and controlled release of non-steroidal
trices to store active principles with cytotoxic activity [90–95]. Due antiinflammatory drugs
to their low price, stability, versatility and adsorbing ability, LDH-
biosensors are also being developed to immobilise enzymes for Non-steroidal antiinflammatory drugs (NSAIDs) are aromatic or-
selective detection of molecules such as urea [96–99], ascorbic acid ganic compounds with easily ionizable carboxylic groups, thus permit-
[97], cyanide [100], nitrite [101], etc. They have been also used for ting their intercalation as anions between the layers of LDHs. Most of
treatment of peptidic ulcers [102,103]. On comparing different these compounds, however, are scarcely soluble in water, thus limiting
antacids, it has been shown that hydrotalcites show a larger their dissolution and absorption by living organisms. On intercalation of
neutralisation capacity than conventional antacids and a prolonged NSAIDs in inorganic matrices, such as hydrotalcite, their solubility is
buffer effect in an optimum pH range [104,105]; no dangerous in- increased, and their controlled release is also possible, while negative
crease of aluminium levels on serum was observed when using effects on the gastrointestinal tract are simultaneously diminished.
LDHs [105]. Hydrotalcite can be found in different commercial Fenbufen is used to improve the symptoms of arthritic rheumatism
drugs, such as Talcid© or Almax©. MgAl-LDHs have been also used and osteoarthritis, and is also applied for treatment of gout; however,
to adsorb intestinal phosphate to prevent hyperphosphatemia [106]. its use is limited because of its frequently observed negative effects on
the gastrointestinal tract and the central nervous system and because it
Hydrotalcite-type materials have been also proposed for the con- decreases the concentration of leucocytes and increases aminotransfer-
trolled release of intercalated drugs. LDHs have the advantage of ease ase. Controlled delivery of this drug may, however, decrease significantly
of preparation, low cost, good biocompatibility, low cytotoxicity, and these systemic negative effects.
full protection of the loaded drug. It has been reported that drug inter- Evans et al. [116] have intercalated fenbufen in Mg, Al and Al, Li
calation into LDH matrices does not only reduce stomach irritation but, hydrotalcites by coprecipitation from aqueous solutions of the metal
in some cases, an increase in drug solubility is also observed [107–110]. nitrates in a basic solution of the drug. These authors have pointed out
Choy et al. [111] proposed to use LDH materials as non-viral vectors for that the intercalation process depends on the chemical composition of
DNA release inside the cell. The ion exchange capability of LDHs allows the hydrotalcite and the pH of the reaction medium. If the reaction is
encapsulating functional negatively charged biomolecules such as anti- carried out at pH 8 a thermally stable phase is obtained; however,
sense DNA to form bio-LDH nanohybrids, a system where the hydroxide when the reaction is carried out at pH 13 the interlayer spacing is larger,
V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39 31

and these authors have claimed formation of a bilayer of intercalated intercalated drug was determined by ion exchange of the drug with car-
molecules weakly bonded to the brucite-like layers. Release of fenbufen bonate and by acid treatment with aqueous HCl to dissolve the
from hydrotalcite was studied at pH 7.8, as fenbufen is only dissolved hydrotalcite crystallites, the drug precipitating in its acidic form. Release
and completely absorbed at this pH value. Release from the Al, Li sample studies carried out at 37 °C with a phosphate buffer and at pH 4 or 7
is very large and fast in the initial stages of the process (10 min) and indicate that, except for gemfibrozil, for which release is identical at
reaches an almost constant release level after 20 min. The maximum both pH values, release is faster at pH 4 than at pH 7 for all other
amount of released fenbufen is 40% of the initially intercalated amount. drugs tested; the time needed for release of 50 and 90% of the initial
Release from the Mg, Al sample is also very fast during the first 15 min, amount of drug is reported.
but significantly slower afterwards; it increases linearly after the initial Duan et al. [120] have intercalated naproxen by ion exchange in a
step, reaching ca. 59% of the initially adsorbed amount after 120 min Mg, Al hydrotalcite containing nitrate in the interlayer, at pH 8 and
release. These results indicate that the LDH-Mg, Al-fenbufen system is 70 °C under a nitrogen atmosphere (to avoid carbonation), and the
more effective than the LDH-Al, Li-fenbufen system, and suggest that stability of the host–guest composite has been studied following differ-
the former system can be used for controlled release of this drug. ent techniques. According to these authors, the experimental conditions
Del Arco et al. [117] have studied also the intercalation of fenbufen in during exchange allow a complete nitrate/naproxen exchange, although
Mg, Al and Mg, Al, Fe hydrotalcites prepared following three different a small contamination by carbonate cannot be avoided. From PXRD
routes, namely, coprecipitation, ion exchange from hydrotalcites with experiments, the basal spacing corresponds to a gallery height of
chloride anions in the interlayer, and reconstruction from a starting 1.86 nm, with the naproxen molecules located with the naphthalene
hydrotalcite containing carbonate, but calcined at intermediate temper- double ring perpendicular to the brucite-like layers and the carbox-
atures to yield a mostly amorphous mixture of metal oxides. According ylate groups pointing alternatively to the top and bottom layers
to these authors, fenbufen is effectively intercalated between the layers forming an intercalated monolayer; water molecules are located be-
of the Mg, Al system following any of these three methods, although tween the naproxen and the brucite-like layers (Fig. 2).
more than a single crystalline phase is obtained; element chemical FT-IR spectroscopy, PXRD and thermogravimetric analysis are ap-
analysis data suggest the presence of small amounts of chloride or plied to check the stability of this system, and the results are com-
carbonate in the interlayer, where they are required for an electric pared to those measures for pure naproxen; intercalation increases
balance of the solid. However, when using the Mg, Al, Fe matrix the the thermal stability of naproxen, as decomposition starts at 250 °C
drug is intercalated only by coprecipitation or ion exchange, but not by while bulk naproxen decomposes at 170 °C.
reconstruction. The sample prepared by ion exchange also contains Berber et al. [121] have intercalated naproxen in a Mg, Al hydrotalcite
small amounts of chloride as an interlayer impurity. From the value of and have studied the effect of the inorganic matrix on the drug solubility
the interlayer spacing, it is concluded that the drug molecules form a at pH 2 (although under this strongly acidic condition dissolution of the
bilayer, somewhat tilted within the interlayer, with the carboxylate layers is strongly expected). The LDH–drug systems were prepared by
groups pointing towards the brucite layers. These authors have also coprecipitation, by adding the metal chlorides solution on a naproxen
studied the effect of hydrotalcite (well as an additive or as a hosting solution at pH 8, and also by reconstruction, starting from a calcined
matrix) on the solubility of fenbufen at different pH values. They find carbonate sample, at pH 8 in nitrogen flow at 80 °C. According to these
that the presence of Mg, Al-LDH as an additive increases the solubility authors, the coprecipitation method is better than the reconstruction
by 51, 62, and 21%, respectively, at pH 1.2, 4.5, and 6.8; the increases one, as the amount of naproxen fixed in the interlayer space was 52%
were 128, 99, and 98% of intercalated fenbufen when the LDH was following the first method, but merely 23% by reconstruction; this poor
used as a matrix. Studies on the release of fenbufen from the Mg, performance of the last sample can be due to the simultaneous
Al-LDH and Mg, Al, Fe-LDH systems show that the dissolution rate is (and competitive) presence of naproxen and carbonate in the reac-
slower when fenbufen is intercalated: while 2 h is required for a total tion medium; actually, the PXRD diagrams show the major presence
release from the Mg, Al matrix, only 93% is released from the Mg, Al, Fe of the LDH-carbonate phase, with only weaker diffraction maxima
matrix in a slower process. The presence of a residual amount of non- due to the LDH-naproxen phase. These differences can be also concluded
released fenbufen in the solid matrix is confirmed from analysis of from the thermogravimetric study reported, where the mass loss is
the solid residues after the release studies: two different residues much larger in the sample prepared by coprecipitation, due to the com-
are obtained from sample Mg, Al, Fe-LDH-fenbufen; a dense LDH- bustion of a larger amount of naproxen in this sample. From the position
phosphate (from the buffer used) phase, which precipitates at the of the basal maxima in the PXRD patterns, these authors have proposed
bottom of the reaction flask, while the other remains floating and corre- that naproxen molecules are located perpendicular to the brucite-like
sponds to the unreleased phase. However, only a single phase, corre- layers, with the carboxylate groups in a staggered interdigitated shape.
sponding to the LDH-phosphate, is obtained from the Mg, Al matrix. Intercalation in this matrix leads to an important increase of naproxen
Evans et al. [118] have studied the release of fenbufen from LDH solubility, from 8·10 − 3 to 26·10 − 3 g/L in simply 1 min, and from
systems covered with a Eudragit S-100® film; only 67% of the initial 30·10 −3 to 146·10 −3 g/L after 1 h, on comparing the values for pure
amount of drug is released, suggesting that the interaction between the and intercalated naproxen.
carboxylate groups of the polymer and the surface of the LDH crystallites Del Arco et al. [122] have intercalated naproxen in a Mg, Al
inhibits the release of the drug and the dissolution of the polymer. hydrotalcite following three different routes, namely (a) ion exchange
However, in similar studies carried out by del Arco et al. [117], where from the LDH in the nitrate form, (b) reconstruction, from a calcined
the samples are covered by simple dispersion and immobilization in carbonate LDH, and (c) coprecipitation, by addition of an aqueous solu-
microspheres, total release of the intercalated drug is observed; disper- tion of the metal cations to a basic solution of naproxen. The drug load-
sion is not enough for a total effective covering, while preparation of ings were rather similar for the exchange (44.2%) and coprecipitation
microspheres guarantees a total covering of the particles, suitable for (46.5%) samples, but somewhat lower (38.4%) for the reconstruction
delayed colonic release of the drug. sample, probably due to a minor contamination by carbonate, arising
O'Hare et al. [119] have intercalated naproxen, diclofenac, gemfibro- from incomplete removal during calcination or carbonation during
zil, 4-biphenylacetic acid, ibuprofen, 2-propylpentanoic acid and preparation of the sample. Such contamination is evident from the
tolfenamic acid in an Al, Li hydrotalcite by ion exchange at 60 °C, PXRD patterns, where maxima due to the LDH-carbonate phase are
starting from a sodium salt of the drug and a hydrotalcite containing undoubtedly recorded, and also from the 13C CP/MAS NMR spectra,
chloride as the interlayer anion. From the height of the interlayer where a peak at 170 ppm due to carbonate is recorded. The 13C CP/
space measured by PXRD, these authors conclude that the drug mole- MAS NMR spectra of the other two samples are similar to that for
cules are located in the interlayer forming bilayers. The amount of the bulk drug, with only minor shifts in the positions of the signals,
32 V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39

Fig. 3. Naproxen release evolution: (triangle) drug, (circle) physical mixture, (square)
intercalated drug.
(Left) Mg, Al-LDH, modified from Microporous and mesoporous materials, Vol. 130, D.
Carriazo, M. Del Arco, C. Martín, C. Ramos and V. Rives, Influence of the inorganic ma-
trix nature on the sustained release of naproxen, pp. 229–238, Copyright (2010), with
permission from Elsevier; (right) Mg, Al, Fe-LDH, modified from Applied Clay Science,
Vol. 42, M. Del Arco, A. Fernández, C. Martín, V. Rives, Release studies of different NSAIDs
encapsulated in Mg, Al, Fe-hydrotalcites, pp. 538–544, Copyright (2009), with permission
from Elsevier.

(mmol m −2), as calculated from the Langmuir fitting, increases as the


charge density of the LDH does, due to a large adsorption of naproxen,
much larger for the chloride-containing samples than for the carbonate-
containing samples, but decreases when the pH is increased. The release
rate decreases when the charge density of the structure is increased,
probably because the large electrostatic interactions between naproxen
anions and the hydroxyl layers inhibit the exchange between naproxen
anions and those originally existing in the reaction medium, HPO42− and
H2PO4−, from the buffer solution.
Fig. 2. (A) Structure of naproxen; (B) schematic representation of the possible arrange- Intercalation of flurbiprofen by coprecipitation and reconstruc-
ment for naproxen/LDH.
tion in a Mg, Al-LDH has been studied by Berber et al. [121]. The sol-
Reprinted from the Journal of Solid State Chemistry, Vol. 177, M. Wei, S. Shi, J. Wang, Y. Li, X.
Duan, Studies on the intercalation of naproxen into layered double hydroxide and its ther- ubility at pH 2 was compared to that for the pure, unintercalated
mal decomposition by in situ FT-IR and in situ HT-XRD, pp. 2534–2541, Copyright (2004), drug. The findings reported by these authors demonstrate that a
with permission from Elsevier. single crystalline phase is obtained by the coprecipitation method,
while by reconstruction the major phase contains carbonate anions
due to the ionic state of the drug upon intercalation. The dissolution in the interlayer. The amount of intercalated flurbiprofen corre-
rate of naproxen intercalated in the Mg, Al-LDH [123] or the Mg, Al, sponds to 49 and 21%, respectively, for the samples prepared by
Fe-LDH system [124] has been also studied by these authors, and the re- coprecipitation and reconstruction.
sults have been compared with dissolution results for pure naproxen The solubility of this hardly soluble drug is markedly increased
and for mixtures of naproxen and LDH-carbonate. The dissolution pro- upon intercalation. The initial dissolution rate is rather slow, due to
files of these two last systems are rather similar to each other, while the electrostatic interactions between the drug and the LDH layers,
the dissolution rate is much slower when the drug is intercalated in but after 4 min the solubility is four times larger when the drug is in-
any of both matrices. Both the dissolution rate and the total amount of tercalated than when it is not; it reaches a value five times larger after
released drug are larger from the Mg, Al-LDH than from the Mg, Al, 5 h and then it reaches a constant value for pure flurbiprofen,
Fe-LDH one (Fig. 3). 720 mg/L after 24 h (Fig. 4).
The data can be fitted to an order 1 process and also to the Weibull The improvement in naproxen and flurbiprofen solubilities and
model [125], with an important role played by interparticle and their dissolution rates depends on the properties of the LDH–drug
intraparticle diffusion in the Fe-free sample, but only interparticle diffu- composites; the high molecular arrangement of the drug within the
sion for the Mg, Al, Fe-LDH system. interlayer space prevents its recrystallization, while the hydrophilic
The intercalation of naproxen in a Zn, Al-LDH system has been nature of the LDH–drug composites favours water penetration into
studied by Hou and Jin [126]. These authors prepared their samples by these composites, as well as the acidic medium in which the protons
contacting solutions with different concentrations of the drug with a attack and destroy the layered materials, leading to the drug release.
given amount of Zn, Al-LDH, well with chloride or nitrate in the Perioli et al. [127] have studied the effect of the Mg, Al LDH on the
interlayer space, and studied the effect of different factors, namely, biopharmaceutical properties of flurbiprofen. The drug was inter-
contact time, composition of the matrix, charge density, specific surface calated by ion exchange at 60 °C for 7 days in a water–alcohol
area of the LDH sample, and pH of the solution, on the total amount of solution. Solubility studies at pH 1.2 (where the layered structure
intercalated naproxen and on its release rate. According to these has been completely destroyed, releasing the drug) show that the
authors, the time required to reach the adsorption equilibrium is concentration of flurbiprofen changes from 4.46 mg L − 1 for
160 min, whichever the specific LDH used, and the adsorption the pristine compound to 38.51 mg L − 1 when intercalated, after
isotherms can be fitted to a Langmuir model. The monolayer capacity 15 min. This result is related to the lack of crystallinity of the
V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39 33

acetone/ethanol mixture. Powder X-ray diffraction confirmed intercala-


tion of the drug, although a small amount of carbonate remained interca-
lated. From a d-optimal design these authors conclude that the main
factors affecting the quality of the encapsulated microparticles are the
addition order of the LDH and Eudragit® to the solution, the acetone/
ethanol solution volume and the LDH/polymer ratio; better results
were obtained at a high stirring speed (1000 rpm), adding the LDH be-
fore than Eudragit®, and when low LDH/polymer ratios and large vol-
umes of solution were used; Eudragit® L produced slightly better
results than the S form. SEM showed non aggregated spherical particles
with a smooth surface and a high homogeneity; the outer polymer shell
Fig. 4. Solubility data of flurbiprofen: (A) before intercalation, and (B) after intercala- was more compact, less porous and less irregular than the core. The
tion with hydrotalcite. gastroresistance of the microparticles was confirmed by treatment at
Reprinted from the European Journal of Pharmaceutical Sciences, Vol. 35, M.R. Berber,
37 °C with a pH 1.2 solution; no traces of Mg2+ nor diclofenac were ob-
K. Minagawa, M. Katoh, T. Mori and M. Tanaka, Nanocomposites of 2-arylpropionic acid
drugs based on Mg–Al layered double hydroxide for dissolution enhancement, pp. 354–360, served in the medium, indicating a good stability of the microparticle.
with permission from Elsevier. As expected, drug release was observed at pH 6.8 (above that for solu-
bilization of Eudragit® L) or 7.5 (Eudragit® S solubilizes at pH 7). For
the nanoparticles prepared with Eudragit® L no further diclofenac re-
intercalated flurbiprofen, which is straightforward released to the lease was observed when the pH was increased to 7.5, probably because
reaction medium in a “molecular” state due to the fast destruction of grafting of dihydrogen phosphate species (from the buffer solution),
of the layers in the acidic medium. This increase in the observed which caused difficulty in diffusion of diclofenac from the interlayer. Re-
solubility could be also related to the energy barrier for solvation lease from Eudragit® S was faster, probably because of its thinner coat-
of the drug molecules and the formation of a saturated solution, ing and the prevalence of monohydrogen phosphate, which does not
which has been identified as the dissolution limiting step. graft onto the brucite-like layers.
Studies of dissolution at pH values close to that of gastric fluid Perioli et al. [132] have intercalated diclofenac in nanosized Zn, Al
show that the percentage of released drug is approximately twice hydrotalcite by an ionic exchange procedure; the aim of the study
when the drug is intercalated than when it is in the crystallized, iso- was to analyse the effect of the particle size on the release rate of
lated, form. Permeability studies have demonstrated that the pres- diclofenac. The starting hydrotalcite material, with bromide anions
ence of hydrotalcite in some sort of way favours permeability of in the interlayer, was prepared by addition of a CTAB microemulsion
flurbiprofen through the gastric mucous, due to the increase of pH to an aqueous solution of Zn and Al nitrates; total exchange was
due to the basicity of the LDH layers. Release in the intestine takes achieved under these conditions, reaching a final diclofenac content
place through ion exchange with phosphate and carbonate anions of 41.8% (w/w).
existing in the medium, and is slower than for the corresponding The morphology of the nanohybrid particles was determined by
physical mixture (where the drug is not intercalated), but faster SEM and TEM; small hexagonal platelets were observed, with a diam-
than in Froben® during the first 5 h of the test; this is a consequence eter close to 200 nm, together with some larger platelets (around
of the increased solubility by ion exchange, as mentioned above. 350 nm), likely due to the formation of aggregates.
Ambrogi et al. [128] have intercalated diclofenac in a Mg, Al-LDH Release of the drug was carried out in three different media: phos-
in the chloride form by ion exchange in a water–alcohol dissolution phate buffer at pH 7.5 or 7.0 and at pH 7.0, but containing NaCl and
(50% vol/vol) of the sodium salt of diclofenac at 60 °C for 3 days in Na2CO3. The highest release rate and largest amount of drug released
an orbital incubator. The solid thus obtained contained up to 55% of (100% of drug was released) were reached at pH 7.5. The process was
drug. The gallery height, as measured from PXRD, is 1.88 nm, and slower at pH 7, when only 72% of the initial drug amount had been
these authors proposed that interdigitated bilayers of diclofenac are released after 24 h; no significant difference was observed after addi-
formed, the main symmetry axis of the molecules being perpendicu- tion of sodium chloride and carbonate to the reaction medium. These
lar to the layers. Release studies at pH 7 and 7.5 show that the process authors conclude that phosphate concentration is the most important
is much slower than for a drug and hydrotalcite physical mixture (for parameter on the release rate and on the amount of drug released, as
which complete dissolution is reached in 15 min) at both pH values. H2PO4− reacts with hydroxyl groups from the brucite-like layers,
The release rate and the total amount of drug released are larger at forming Zn and Al hydroxyphosphate by a solid grafting reaction.
pH 7.5 than at pH 7 (99 and 70%, respectively, of the initial amount The strong bonds between the grafted phosphates and layers are
of intercalated drug). Release of the drug actually occurs via an ion able to stop the ion exchange mechanism and to cause the obstruc-
exchange process, with phosphate anions of the buffer solution. The tion of HTlc galleries, entrapping the diclofenac anions into the inter-
solid residue isolated after the exchange at pH 7.5 corresponds to a nal part of interlayer region.
phosphate-intercalated hydrotalcite, while at pH 7 a mixture of Wang et al. [133] have studied the intercalation, by coprecipitation, of
diclofenac- and phosphate-intercalated hydrotalcites is obtained. diclofenac in a Zn, Al hydrotalcite; the nanohybrid formed was compared
These differences arise from the relative concentration of H2PO4− with sodium diclofenac on auricular tumescence induced by xylene in
and HPO42− in the reaction medium at the different pH values tested. mice.
The data can be fitted to the Higuchi model [129], confirming that Diclofenac was incorporated by ion exchange in the interlayer space
diffusion is the dissolution rate limiting step, and that the concentra- of a Mg, Al hydrotalcite by Sanmartino et al. [134], and the hybrid thus
tion of drug has no effect on the release rate. formed was incorporated into polycaprolactone, and processed as films
These authors have extended this study for Eudragit®-encapsulated of 0.15 mm thickness. The resulting compound showed better properties
LDH-diclofenac [130]. The main reason for this sort of studies is that than the original polymer. Release of the drug takes place in two steps, a
the LDH matrix is dissolved in the acidic medium on the stomach and first, fast, one releasing a small amount of drug, followed by a much
thus the drug cannot reach the colon. A Mg, Al carbonate LDH was slower second step.
prepared by homogeneous coprecipitation and the carbonate species Gunawan and Xu [135] have studied the possible relationship
exchanged by chloride; diclofenac was then intercalated [128]. The between the crystalline morphology and the aggregation state of
encapsulated microcapsules were prepared by an oil-in-oil solvent an ibuprofen–LDH system and their effect on the drug release. Ibuprofen
evaporation method [131] using Eudragit® S100 and L100 in an was intercalated in a Mg, Al LDH (Mg/Al molar ratio of 2) by
34 V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39

coprecipitation using two different solvents (water, and a 1:1 ethylene


glycol:water mixture) under two reaction conditions (ambient pressure
or hydrothermal treatment for 18 h at 150 °C). PXRD patterns and FT-IR
spectra showed that coprecipitation and ageing (both under ambient or
hydrothermal conditions) lead to intercalation of the drug, although a
small amount of carbonate is simultaneously intercalated; ethylene
glycol is not intercalated. From the molecular size of ibuprofen and the
gallery height measured by PXRD, it is concluded that the ibuprofen
anions are tilted with respect to the brucite-like layers, forming an
antiparallel bilayer, with the carboxylate groups bonded to the inorganic
layers. The crystallinity of the samples is not dependent on the nature of
the solvent used, but is highly dependent on the ageing conditions; a
larger crystallinity is observed after hydrothermal treatment at 150 °C
for 18 h than after ambient pressure ageing at 70 °C for 3 days. This
effect is clearly concluded from TEM and SEM images, Fig. 5, which
show somewhat rounded platelets and irregularly shaped particles
with an average diameter of 200 nm, when the samples were aged at
ambient pressure. The surface of the hydrothermally treated crystallites
is smoother and the nanoplatelets are regularly oriented with a larger
packing density (face-to-face and border-to-border); however, for the
ambient pressure-aged samples the compacity of the crystals is lower
and the surface is porous and shows a certain degree of roughness.
These differences can be related to the presence of ethylene glycol,
which would act as a dispersing agent, limiting the agglomeration of
the particles during the hydrothermal treatment. It can be also a conse-
quence of an Ostwald ripening during hydrothermal treatment.
Studies on release of the drug at pH 7 show that release is complete
in all samples, although the release rate depends on the crystallinity
and aggregation degrees; the slowest release rate is observed for the
hydrothermally treated samples prepared in ethylene glycol. The diffu-
sion rate of the anions from the hydrotalcite aggregates can thus be con-
trolled by the characteristics of the solid, namely, porosity, aggregation
degree and crystallinity of the LDH platelets. For a dense powder formed
by well oriented particles, as obtained in the presence of ethylene glycol
and submitted to hydrothermal treatment, the length of the path
followed by diffusing anions and the resistance to diffusion is much
larger, due to the large size of the aggregates and their rigid structure.
As a result, the release rate of ibuprofen is slower.
Diclofenac and ketoprofen have been intercalated in LDHs containing
Zn or Mg and Zn, together with Al, in the brucite-like layers [136] by
coprecipitation from addition of a metal salt solution into an ethanolic
solution of the drug at pH 9 under nitrogen atmosphere at room
temperature and stirring for 24 h. The divalent/Al3+ molar ratio was in
all cases slightly lower than the expected value of 2, probably by a partial
dissolution of the divalent cations and formation of coordination
compounds with the organic anions. However, it should to be noticed
that the presence of Mg leads to values closer to 2.0 than the exclusive
presence of Zn2+ as the divalent cation. The interlayer spacings were
ca. 23 Å in both cases; from the molecular sizes of the drug it can be
concluded that a slightly tilted bilayer perpendicular to the brucite-like
layers is formed. The ketoprofen sample contained a small impurity of
nitrate, as concluded from element chemical analysis and the X-ray
diffraction patterns, as well as from the FT-IR spectrum, which displays
a sharp band at 1367 cm−1 (mode ν3) and another at 1066 cm−1, due
to mode ν1, inactive for the free anion (D3h symmetry), but activated in
the restricted symmetry situation within the interlayer. Thermal decom-
position proceeds at a higher temperature than for the free drugs, prob-
ably because of the enhanced stabilization due to the strong electrostatic
interactions with the layers.

Fig. 5. TEM images of as-prepared Mg2Al-Ibp-LDH samples using two different types
of solvent systems (EG/W = ethylene glycol and water; W = water) at atmospheric
(C) or hydrothermal (H) conditions: (A) EG/W–H; (B) W–H; (C) EG/W–C; (D) W–C.
Reprinted from the Journal of Pharmaceutical Sciences, Vol. 97, P. Gunawan and R. Xu,
Direct control of drug release behaviour from layered double hydroxides through
particle interactions, pp. 4367–4378, with permission from Wiley Interscience.
V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39 35

Ambrogi et al. [137] have also studied the ion exchange intercalation the other drugs tested in this study; although the differences were not
of ibuprofen in a Mg, Al LDH with chloride anions in the interlayer. The as sharp as for indomethacin, it was found that the simultaneous
drug loading was very high, 50%, and the molecules are forming a mono- presence of the inorganic matrix enhances the drug solubility, much
layer in the interlayer with the main molecular axis perpendicular larger when intercalated than when merely mixed. In addition, these
to the inorganic layers, interaction taking place through the carbox- authors also highlight the beneficial role of hydrotalcite to improve the
ylate groups. These authors relate their release results with those of stability of the drug, delaying its hydrolysis and decarbonation.
commercial Neo-Mindol® and a physical mixture of the Del Arco et al. [107] have also intercalated indomethacin in a Mg, Al
hydrotalcite and ibuprofen. The release rate at pH 7.5 is much LDH, and have reported in vivo studies using Swiss mice (25 g weight),
slower from the interlayer (in the hybrid) than in the commercial to whom indomethacin, the physical mixture and the intercalation
product and the physical mixture, as in the first case it proceeds product were supplied; the aim of the study was to check the effect of
via ion exchange with phosphate anions of the buffer solution. the simultaneous presence of the inorganic matrix on ulcer formation,
DeLeon et al. [138] have prepared and characterised polymer as measured from the ulcerated area of mice stomach. Intercalation
nanocomposites with poly-L-lactic acid (PLLA) and LDH–ibuprofen, was carried out by coprecipitation or reconstruction in a basic medium.
studying the effect of the presence of the LDH and the LDH–ibuprofen/ Coprecipitation yields a hybrid with 60% drug loading, indomethacin
PLLA ratio on the release rate of the drug. This was incorporated into molecules forming a bilayer in the interlayer space with the carboxylate
the interlayer space of the hydrotalcite by ionic exchange in a Zn, groups pointing towards the brucite-like layers; however, a biphasic
Al-nitrate hydrotalcite, and observed a practically complete exchange system is obtained by reconstruction, with components LDH-carbonate
(only 0.06% N was identified in the final solid), final ibuprofen loading and LDH–indomethacin, where an intercalated monolayer is formed,
being 40.9% (w/w). The composites with PLLA were prepared by spin with the molecules somewhat tilted from the perpendicular orientation;
casting and the results are compared with a similar composite, but moreover, only 18% drug loading was reached; for this reason, the
formed exclusively by PLLA and ibuprofen, without the inorganic phase. pharmacological studies were carried out with the high-loaded sample,
Intercalation was confirmed by PXRD, as a basal spacing of 24.7 Ǻ prepared by coprecipitation. Oral doses of 80 mg kg−1 of pure indo-
was measured for the hydrotalcite–ibuprofen sample, with a weak methacin led to gastric hemorrhagic damage in 88% of the mice, while
swelling to 26.59 Ǻ after dispersion in the PLLA matrix. using the LDH–indomethacin system ulcer formation was observed
The release studies were carried out in a phosphate buffer (pH = 7.4) only in 70% of the mice, Fig. 6; the ulcerated surface also decreased
at 37 °C for 10 days. In the absence of hydrotalcite, only 5% release was from 0.40 to 0.106% when using indomethacin or the hybrid, respectively.
observed and in the very early stages of the process; probably it corre- This value was 0.215% for the physical mixture, confirming not only the
sponds to ibuprofen adsorbed on the surface of the matrix. However, a positive role of the use of hydrotalcite because of its basic nature, but
two-step release process was observed from the nanocomposites, a first also the additional positive effect of the intercalated drug.
fast one lasting 15 h, followed by a slower second step along 50 h. In Coprecipitation, reconstruction and ion exchange were used by del
this second step, both the release rate and the amount of drug released Arco et al. to intercalated mefenamic and meclofenamic acids in a Mg,
increase as the amount of LDH–ibuprofen in the PLLA matrix increases Al hydrotalcite [143]. Total chloride exchange was achieved, with drug
(a maximum loading of 70% of LDH–ibuprofen was tested). Changes in loadings of 52 and 58%, respectively, for mefenamic and meclofenamic
pH along the release process were also studied: pH increased during the acids. PXRD patterns suggest formation of a bilayer between the LDH
first 10 h and then slowly decreased until a limiting equilibrium value layers, with the carboxylate groups pointing towards the brucite-like
was reached for the LDH-containing composites. However, such a pH layers. These authors have also studied [144] the effect of the LDH on
decrease was larger in the LDH-free samples, probably due to dissolution the solubility of mefenamic acid. Solubility is highly increased, especially
of ibuprofen from the polymer surface, without a maintained release. when the acid is intercalated, the best results being achieved at pH values
Ay et al. [139] have studied the intercalation of ibuprofen and glucu- 1.2 and 6.8, but being low at pH 4.5. The inorganic matrix is dissolved
ronic acid in a Mg, Al-LDH, supported on a magnesium ferrite core; this under acidic conditions and the residue is the pure drug, while at
had been prepared by calcination of a non-stoichiometric Mg, Fe-CO3 pH 6.8 a fast ion exchange takes place with the buffer anions. The release
LDH. The drugs were intercalated by ion exchange with nitrate-LDH rate is very high for the pure drugs and their physical mixtures with
precursors, once supported on the magnetic core. Drug loadings of 45% hydrotalcite. As expected, the release rate decreases upon intercalation,
(ibuprofen) and 25% (glucuronate) were achieved. The nanocomposites and up to 90 min is required, despite the release rate decreases with
would be hopefully suitable for magnetic targeted delivery, i.e., magnetic time. The release is fast in the first 30 min, and then it reaches an almost
field-controlled delivery of a drug to the targeted organ and its subse- constant value; total release of mefenamic acid was not, however,
quent release, as previously reported by other authors with other differ- achieved.
ent drugs: Duan et al. [140] intercalated 5-amino salicylic acid in a Zn, When a partial Al3+/Fe3+ substitution is made in the composition of
Al-LDH, and diclofenac on a Mg, Al-LDH coated onto a magnesium ferrite the brucite-like layers, forming a Mg, Al, Fe-LDH, a decrease in the release
core [141], while Carja et al. [142] intercalated aspirin in a Mg, Al-LDH rate and in the total amount of drug released is observed [124], in com-
supported on FeOx/Fe-LDH. parison with the results obtained for the Mg, Al-LDH. The release data
Ketoprofen, tiaprofenic acid and indomethacin are scarcely soluble can be fitted to order 1 and Weibull models [125], interparticle and
drugs. Ambrogi et al. [110] have intercalated these drugs in a Mg, Al intraparticle diffusion predominating in the iron-free systems, but only
LDH by ion exchange of originally intercalated chloride or hydroxide intraparticle diffusion for the iron-containing systems.
anions, and have studied the effect of the inorganic matrix on their solu-
bility at gastric pH (1.2). All data (PXRD, thermogravimetric analysis, 3. Conclusions and further studies
FT-IR spectroscopy and SEM) confirm that the drug has been effectively
intercalated. Solubility studies with the pure drug, a LDH–drug physical The amount of information concerning intercalation of NSAIDs into
mixture and the hybrid samples were carried out at pH 1.2 and 37 °C. LDHs is very large; probably, this is the family of drugs in which in-
Values below 2 mg mL−1 were determined for pure indomethacin and tercalation has been more widely studied. From the literature here
its physical mixture with the drug, while for intercalated indomethacin summarised some conclusions and proposals for further studies
the value was 12 mg mL−1 after 1 min, and 14 mg mL−1 after 15 min. can be reached.
This much larger solubility has been related by these authors to the Among the different procedures applied to intercalate NSAID
non-crystalline nature of indomethacin in the interlayer, so being imme- molecules in the interlayer space of LDHs, larger drug loadings are
diately released in an ionic form by the easy dissolution of the inorganic reached when using coprecipitation (direct synthesis) or anionic ex-
matrix in the strongly acidic medium. Similar results were obtained for change; in this last case, however, a small amount of the anion (generally
36 V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39

The NSAID loading in the interlayer depends on several factors,


namely, chemical composition of the matrix, method and synthesis
experimental conditions, charge layer density, etc. All of them are able
to modify the orientation of the NSAID molecule in the interlayer, its
release rate and the amount of released drug. In most of the cases the
drug molecules form bilayers in the interlayer inorganic space, forming
antiparallel bilayers, sometimes slightly tilted, with the carboxylate
groups pointing towards the inorganic layers. In some cases, an orienta-
tion of the molecules perpendicular to the layers has been claimed, with
the water molecules located between the drug molecules and the
layers.
The LDH, well as an additive or as a matrix for the drug, gives rise to
an increase in the usually poor NSAID solubility, in a larger extent when
the drug is intercalated. These differences are related to the crystalline
nature of the intercalated NSAID. Release is fast at low pH because of
dissolution of the matrix, while at pH values around 7 release takes
place via anionic exchange with the anions existing in the liquid
medium. The basic nature of the LDH might be also responsible for the
larger solubility of the drug, even when they are simply mixed and
not intercalated.
The release rate depends on the particle size, crystalline morphology
and aggregation state of the particles, which can be modified or even
tuned by using different solvents during preparation. The release rate
and the amount of drug released depend on the chemical composition
of the matrix and on the synthesis variables mentioned above. Release
is slow when the drug is intercalated, as it proceeds via anionic exchange
with the anions in the medium, and depends on the pH as well, release
being usually faster for low pH values. When release takes place at pH
provided by phosphate buffers, the release rate is determined by the con-
centration of H2PO4− and HPO42− species, as the former is irreversibly
grafted to the layers via reaction with its hydroxyl groups, obstructing
the galleries and preventing the drug release.
When the LDH–NSAID systems are covered by polymers in which
stability depends on pH, the quality of the encapsulated particles de-
pends on the precise preparation procedure (for instance, order of the
reactants has been added, type and concentration of solvents used,
and LDH/polymer ratio). Formation of microspheres guarantees a better
covering than simple dispersion.
Nevertheless, despite the numerous advantages that intercalation of
NSAIDs in LDHs exhibits, these cannot be used as matrices for controlled
release in oral administration, as they are very sensitive to the acidic
medium, unless they are covered with a polymer film which preserves
them from being acidically degraded. So, this is a field where new in-
sights are still required, although some studies are already available in
the literature; for instance, several papers have already reported the
Fig. 6. Photographs of extracted stomachs after treatment with (A) pure indomethacin use of Eudragit® or polylactic to cover and to protect the LDH–drug
(In), and (B) sample LDH–indomethacin (InD) prepared by coprecipitation. (C) Ulcer- hybrid from degradation. This encapsulation has been also applied to
ation degree (%) after administration of 80 mg active ingredient/kg weight of mouse other systems containing antibiotics or antihypertensive drugs, which
(n = 17).
hybrids have been covered with poly(ε-caprolactone) or xyloglucan.
Reprinted from the Journal of Pharmaceutical Sciences, Vol. 93, M. Del Arco, E.
Cebadera, S. Gutiérrez, C. Martín, M.J. Montero, V. Rives, J. Rocha, M.A. Sevilla, Mg, Al Even the use of mixed natural polymers has been proposed, such as
layered double hydroxides with intercalated indomethacin: synthesis, characterization sodium alginate and zein, a prolamine protein found in maize; quite
and pharmacological study, pp. 1649–1658, with permission from Wiley Interscience. surprisingly these two last polymers are unable to act as release control-
lers when combined with the drug, but by using the ternary system
drug–LDH–polymer release systems at a controlled pH will be prepared.
chloride or nitrate) in the original precursor usually remains, so it is Another very interesting field where developments are expected
recommended to use a chloride precursor to avoid the presence of toxic are those related to the use of LDH–drug composites, together with
nitrate species. The reconstruction method also permits intercalation of a magnetic nucleus for precise delivery in locations of the body
NSAIDs, but the effectiveness of the method depends on the nature of where the action is required. This procedure will permit a faster
the layer cations, and a biphasic system is usually obtained (LDH–carbon- action of the drug, as well as decreasing the total dose, thus avoiding
ate and LDH–NSAID) or lower drug loadings are achieved, due to incom- undesired side-effects.
plete removal of carbonate during calcination of the precursor or
carbonation during the preparation due to the strong competition be-
tween carbonate and the anionic drug, because of the high basicity of Acknowledgements
the calcined solids. Intercalation leads to an increase in the thermal stabil-
ity of the drug, due to the strong electrostatic interactions developing be- Financial support from MICINN (grant MAT2009-08526) is greatly
tween the drug and the layers. acknowledged.
V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39 37

References [33] E. López-Salinas, M. García-Sánchez, J.A. Montoya, D.R. Acosta, J.A.


Abasolo, I. Schifter, Structural characterization of synthetic hydrotalcite-
like [Mg1 − x Ga x (OH)2 ](CO3 ) x/2 ·m H 2O, Langmuir 13 (1997) 4748–4753.
[1] F. Cavani, F. Trifirò, A. Vaccari, Hydrotalcite-type anionic clays: preparation, [34] J.S. Valente, F. Figueras, M. Gravelle, P. Kumbhar, J. López, J.P. Besse, Basic prop-
properties and applications, Catal. Today 11 (1991) 173–301. erties of the mixed oxides obtained by thermal decomposition of hydrotalcites
[2] In: V. Rives (Ed.), Layered Double Hydroxides: Present and Future, Nova Sci. Pub. containing different metallic compositions, J. Catal. 189 (2000) 370–381.
Inc., New York, 2001. [35] I. Rousselot, C. Taviot-Gueho, F. Leroux, P. Leone, P. Palvadeau, J.P. Besse, Insights
[3] In: X. Duan, D.G. Evans (Eds.), Layered Double Hydroxides, Springer, Berlin, 2006. on the structural chemistry of hydrocalumite and hydrotalcite-like materials:
[4] In: F. Wypych, K.G. Satyanarayana (Eds.), Clay Surfaces: Fundamentals and investigation of the series Ca2M3+ (OH)6Cl·2H2O (M3+: Al3+, Ga3+, Fe3+, and
Applications, Elsevier, Amsterdam, 2004. Sc3+) by X-ray powder diffraction, J. Solid State Chem. 167 (2002) 137–144.
[5] V.A. Drits, A.S. Bookin, Crystal structure and X-ray identification of layered [36] K. Chibwe, W. Jones, Intercalation of organic and inorganic anions into layered
double hydroxides, in: V. Rives (Ed.), Layered Double Hydroxides: Present and double hydroxides, J. Chem. Soc. Chem. Commun. (1989) 926–927.
Future, Nova Sci. Pub. Inc., New York, 2001, pp. 39–92. [37] A. de Roy, C. Forano, J.P. Besse, Layered double hydroxides: synthesis and
[6] A.M. Fogg, A.G. Freij, F.M. Parkinson, Synthesis and anion exchange chemistry of post-synthesis modification, in: V. Rives (Ed.), Layered double hydroxides: pres-
rhombohedral Li/Al layered double hydroxides, Chem. Mater. 14 (2002) 232–234. ent and future, Nova Sci. Pub. Inc., New York, 2001, pp. 1–37.
[7] D. Tichit, N. Das, B. Coq, R. Durand, Preparation of Zr-containing layered double [38] E. Kanezaki, Preparation of layered double hydroxides, in: F. Wypych, K.G.
hydroxides and characterization of the acido-basic properties of their mixed Satyanarayana (Eds.), Clay Surfaces: Fundamentals and Applications, Elsevier,
oxides, Chem. Mater. 14 (2002) 1530–1538. Amsterdam, 2004, pp. 345–373.
[8] F. Basile, G. Fornasari, M. Gazzano, A. Vaccari, Synthesis and thermal evolution of [39] J. He, M. Wei, B. Li, Y. Kang, D.G. Evans, X. Duan, Preparation of double layered
hydrotalcite-type compounds containing noble metals, Appl. Clay Sci. 16 (2000) hydroxides, in: X. Duan, D.G. Evans (Eds.), Layered Double Hydroxides, Springer,
185–220. Berlin, 2006, pp. 89–119.
[9] I. Dobrosz, K. Jiratova, V. Pitchon, J.M. Rynkowski, Effect of the preparation of [40] F. Kooli, W. Jones, V. Rives, M.A. Ulibarri, An alternative route to
supported gold particles on the catalytic activity in CO oxidation, J. Mol. Catal. polyoxometalate-exchanged layered double hydroxides: the use of ultra-
A: Chem. 234 (2005) 187–197. sounds, J. Mater. Sci. Lett. 16 (1997) 27–29.
[10] V. Rives, S. Kannan, Layered double hydroxides with the hydrotalcite-type struc- [41] M. del Arco, V. Rives, R. Trujillano, Surface and textural properties of hydrotalcite-like
ture containing Cu2+, Ni2+ and Al3+, J. Mater. Chem. 10 (2000) 489–495. materials and their decomposition products, Stud. Surf. Sci. Catal. 87 (1994) 507–515.
[11] V. Rives, A. Dubey, S. Kannan, Synthesis, characterization and catalytic hydroxyl- [42] J. Rocha, M. del Arco, V. Rives, M.A. Ulibarri, Reconstruction of layered double
ation of phenol over CuCoAl ternary hydrotalcites, Phys. Chem. Chem. Phys. 3 hydroxides from calcined precursors: a powder XRD and 27Al MAS NMR study,
(2001) 4826–4836. J. Mater. Chem. 9 (1999) 2499–2503.
[12] S. Morpugo, M. Jacono, P. Porta, Copper–zinc–cobalt–aluminum–chromium [43] P. Benito, F.M. Labajos, V. Rives, Microwaves and layered double hydroxides: a
hydroxicarbonates and mixed oxides, J. Solid State Chem. 122 (1996) smooth understanding, Pure Appl. Chem. 81 (2009) 1459–1471.
324–332. [44] M. Herrero, F.M. Labajos, V. Rives, Size control and optimisation of intercalated
[13] A. Monzón, E. Romeo, C. Royo, R. Trujillano, F.M. Labajos, V. Rives, Use of layered double hydroxides, Appl. Clay Sci. 42 (2009) 510–518.
hydrotalcites as catalytic precursors of multimetallic mixed oxides. Application [45] P. Benito, M. Herrero, C. Barriga, F.M. Labajos, V. Rives, Microwave-assisted ho-
in the hydrogenation of acetylene, Appl. Catal. A Gen. 185 (1999) 53–63. mogeneous precipitation of hydrotalcites by urea hydrolysis, Inorg. Chem. 47
[14] V. Rives, F.M. Labajos, R. Trujillano, E. Romeo, C. Royo, A. Monzón, Acetylene hydroge- (2008) 5453–5463.
nation on Ni–Al–Cr oxide catalysts: the role of added Zn, Appl. Clay Sci. 13 (1998) [46] P. Benito, I. Guinea, F.M. Labajos, V. Rives, Microwave-assisted reconstruction
363–379. of Ni–Al hydrotalcite-like compounds, J. Solid State Chem. 181 (2008)
[15] A. Monzón, E. Romeo, C. Royo, R. Trujillano, F.M. Labajos, V. Rives, Desarrollo de 987–996.
óxidos mixtos de Ni como catalizadores de hidrogenación selectiva (in Spanish), [47] M. Herrero, P. Benito, F.M. Labajos, V. Rives, Nanosize cobalt oxide-containing
Av. Ing. Quim. 8 (1998) 24–27. catalysts obtained through microwave assisted methods, Catal. Today 128
[16] E. Romeo, C. Royo, A. Monzón, R. Trujillano, F.M. Labajos, V. Rives, Preparation (2007) 129–137.
and characterisation of Ni–Mg–Al hydrotalcites as hydrogenation catalysts, [48] M. Herrero, P. Benito, F.M. Labajos, V. Rives, Stabilization of Co2+ in layered dou-
Stud. Surf. Sci. Catal. 130 (2000) 2099–2104. ble hydroxides (LDHs) by microwave-assisted ageing, J. Solid State Chem. 180
[17] P. Benito, M. Herrero, F.M. Labajos, V. Rives, N. Royo, N. Latorre, A. Monzón, Pro- (2007) 873–884.
duction of carbon nanotubes from methane: use of Co–Zn–Al catalysts prepared [49] P. Benito, F.M. Labajos, V. Rives, Microwave-treated layered double hydroxides
by microwave-assisted synthesis, Chem. Eng. J. 149 (2009) 455–462. containing Ni2+ and Al3+: the effect of added Zn, J. Solid State Chem. 179
[18] V.R.L. Constantino, T.J. Pinnavaia, Basic properties of Mg2+1 − xAl3+x layered (2006) 3784–3797.
double hydroxides intercalated by carbonate, hydroxide, chloride and sulfate [50] P. Benito, F.M. Labajos, J. Rocha, V. Rives, Influence of microwave radiation on
anions, Inorg. Chem. 34 (1995) 883–892. the textural properties of layered double hydroxides, Microporous Mesoporous
[19] S.P. Newman, W. Jones, Synthesis, characterization and applications of layered Mater. 94 (2006) 148–158.
double hydroxides containing organic guests, New J. Chem. 22 (1998) 105–111. [51] P. Benito, F.M. Labajos, V. Rives, Uniform fast growth of hydrotalcite like com-
[20] M.S. San Román, M.J. Holgado, C. Jaubertie, V. Rives, Synthesis, characterisation pounds, Cryst. Growth Des. 6 (2006) 1961–1966.
and delamination behaviour of lactate-intercalated Mg, Al-hydrotalcite-like [52] X. Wang, Q. Zhang, Effect of hydrotalcite on the thermal stability, mechanical
compounds, Solid State Sci. 10 (2008) 1333–1341. properties, rheology and flame retardance of poly(vinyl chloride), Polym. Int.
[21] C. Jaubertie, M.J. Holgado, M.S. San Román, V. Rives, Structural characterisation 53 (2004) 698–707.
and delamination of lactate-intercalated Zn, Al-layered double hydroxides, [53] R. Kaluskova, M. Novota, Z. Vymazal, Investigation of thermal stabilization of
Chem. Mater. 18 (2006) 3114–3121. poly(vinyl chloride) by lead stearate and its combination with synthetic
[22] S. Bhattacharjee, J.A. Anderson, Comparison of the epoxidation of cyclohexene, hydrotalcite, Polym. Degrad. Stab. 85 (2004) 903–909.
dicyclopentadiene and 1,5-cyclooctadiene over LDH hosted Fe and Mn [54] C. Taviot-Gueho, A. Illaik, C. Vuillermoz, S. Commereuc, V. Vernay, F. Leroux, LDH–
sulfonato-salen complexes, J. Mol. Catal. A: Chem. 249 (2006) 103–110. dye hybrid material as coloured filler into polystyrene: structural characterization
[23] M. del Arco, S. Gutiérrez, C. Martín, V. Rives, Intercalation of [Cr(C2O4)3]3-complex in and rheological properties, J. Phys. Chem. Solids 68 (2007) 1140–1146.
Mg, Al layered double hydroxides, Inorg. Chem. 42 (2003) 4232–4240. [55] W. Chen, B.J. Qu, Structural characteristics and thermal properties of
[24] V. Rives, M.A. Ulibarri, Layered double hydroxides (LDH) intercalated with metal co- PE-g-MA/Mg, Al-LDH exfoliation nanocomposites synthetized by solution inter-
ordination compounds and oxometalates, Coord. Chem. Rev. 181 (1999) 61–120. calation, Chem. Mater. 15 (2003) 3208–3213.
[25] M. del Arco, D. Carriazo, S. Gutiérrez, C. Martín, V. Rives, An FT-IR study of the [56] C.M.C. Pereira, M. Herrero, F.M. Labajos, A.T. Marques, V. Rives, Preparation and
6−
adsorption and reactivity of ethanol on systems derived from Mg2Al-W7O24 properties of new flame retardant unsaturated polyester nanocomposites based
layered double hydroxides, Phys. Chem. Chem. Phys. 6 (2004) 465–470. on layered double hydroxides, Polym. Degrad. Stab. 93 (2009) 939–946.
[26] M. del Arco, D. Carriazo, S. Gutiérrez, C. Martín, V. Rives, Synthesis and charac- [57] A. Sorrentino, G. Gorrasi, M. Tortora, V. Vittoria, U. Costantino, F. Marmottini, F.
terization of new Mg2Al-paratungstate layered double hydroxides, Inorg. Padella, Incorporation of Mg–Al hydrotalcite into a biodegradable poly(ε-
Chem. 43 (2004) 375–384. caprolactone) by high energy ball milling, Polymer 46 (2005) 1601–1608.
[27] D. Carriazo, C. Martín, V. Rives, Thermal evolution of a MgAl hydrotalcite-like mate- [58] S. Martínez-Gallegos, M. Herrero, V. Rives, Preparation of composites by in situ
rial intercalated with hexaniobate, Eur. J. Inorg. Chem. 22 (2006) 4608–4615. polymerisation of PET-hydrotalcite using dodecylsulphate, Mater. Sci. Forum
[28] D. Carriazo, C. Domingo, C. Martín, V. Rives, Structural and texture evolution with 587–588 (2008) 568–571.
temperature of layered double hydroxides intercalated with paramolybdate anions, [59] S. Martínez-Gallegos, M. Herrero, V. Rives, In situ microwave-assisted polymer-
Inorg. Chem. 45 (2006) 1243–1251. ization of polyethylene terephthalate in layered double hydroxides, J. Appl.
[29] J.H. Choy, S.Y. Kwak, J.S. Park, J.Y. Jeong, Cellular uptake behavior of [γ-32P] Polym. Sci. 109 (2008) 1388–1394.
labeled ATP-LDH nanohybrids, J. Mater. Chem. 11 (2001) 1671–1674. [60] E. Scavetta, B. Balalrin, M. Berettoni, I. Caprani, M. Giorgetti, D. Tonelli, Electro-
[30] J.H. Choy, S.Y. Kwak, J.S. Park, J.Y. Jeong, J. Portier, Intercalative nanohybrids of chemical sensors based on eletrodes modified with synthetic hydrotalcites,
nucleoside monophosphates and DNA in layered metal hydroxide, J. Am. Electrochim. Acta 51 (2006) 2129–2134.
Chem. Soc. 121 (1999) 1399–1400. [61] D. Shan, C. Mousty, S. Cosnier, Subnanomolar cyanide detection at polyphenol
[31] L. Desigaux, M. Ben Belkacem, P. Richard, J. Cellier, P. Leone, L. Cario, F. Leroux, C. oxidase/clay biosensors, Anal. Chem. 76 (2004) 178–183.
Taviot-Gueho, B. Pitard, Self-assembly and characterization of layered double [62] F. Kovanda, E. Kovacsova, D. Kolousek, Removal of anions from solution by cal-
hydroxides/DNA hybrids, Nano Lett. 6 (2006) 199–204. cined hydrotalcite and regeneration of used sorbent in repeated calcination–
[32] T. Kwon, T.J. Pinnavaia, Pillaring of a layered double hydroxide by polyoxometalates rehydration–anion exchange processes, Collect. Czech. Chem. Commun. 64
with Keggin-ion structure, Chem. Mater. 1 (1989) 381–383. (1999) 1517–1528.
38 V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39

[63] S. Miyata, Anion-exchange properties of hydrotalcite-like compounds, Clays [96] S. Vial, C. Forano, D. Shan, C. Mousty, H. Barhoumi, C. Martelet, N. Jaffrezic,
Clay Miner. 31 (1983) 305–311. Nanohybrid-layered double hydroxides/urease materials: synthesis and appli-
[64] M. del Arco, D. Carriazo, C. Martin, A.M. Pérez-Grueso, V. Rives, Acid and redox cation to urea biosensors, Mater. Sci. Eng. C Biomim. 26 (2006) 387–393.
properties of mixed oxides prepared by calcinations of chromate-containing [97] L. Fernández, H. Carrero, Electrochemical evaluation of ferrocene carboxylic
layered double hydroxides, J. Solid State Chem. 178 (2005) 3571–3580. acids confined on surfactant-clay modified glassy carbon electrodes: oxidation
[65] D. Carriazo, M. del Arco, C. Martin, V. Rives, A comparative study between chlo- of ascorbic acid and uric acid, Electrochim. Acta 50 (2005) 1233–1240.
ride and calcined carbonate hydrotalcites as adsorbents for Cr(VI), Appl. Clay [98] S. Vial, V. Prevot, F. Leroux, C. Forano, Immobilization of urease in ZnAl layered
Sci. 37 (2007) 231–239. double hydroxides by soft chemistry routes, Microporous Mesoporous Mater.
[66] K.H. Goh, T.T. Lim, Z. Dong, Application of layered double hydroxides for removal 107 (2007) 190–201.
of oxyanions: a review, Water Res. 42 (2008) 1343–1368. [99] H. Barhoumi, A. Maaref, M. Rammah, C. Martelet, N. Jaffrezic, C. Moustly, S. Vial,
[67] D.M. O'Hare, B. Lotsch, EP Patent 20010954187 (2001). C. Forano, Urea biosensor based on Zn3Al-urease layered double hydroxides
[68] A. Vaccari, Preparation and catalytic properties of cationic and anionic clays, nanohybrid coated on insulated silicon structures, Mater. Sci. Eng. C Biomim.
Catal. Today 41 (1998) 53–71. 26 (2006) 328–333.
[69] D. Carriazo, C. Martín, V. Rives, A. Popescu, B. Cojocaru, I. Mandache, V.I. [100] D. Shan, S. Conier, C. Moustly, HRP/[Zn–Cr–ABTS] redox clay-based biosensor: design
Parvulescu, Hydrotalcites composition as catalysts: preparation and their be- and optimization for cyanide detection, Biosens. Bioelectron. 20 (2004) 390–396.
havior on epoxidation of two bicycloalkenes, Microporous Mesoporous Mater. [101] H. Chen, C. Mousty, S. Conier, C. Silveira, J.J.G. Moura, M.G. Almeida, Highly sen-
95 (2006) 39–47. sitive nitrite biosensor based on the electric wiring of nitrite reductase by [ZnCr–
[70] D. Carriazo, S. Lima, C. Martín, M. Pillinger, A.A. Valente, V. Rives, AQS] LDH, Electrochem. Commun. 9 (2007) 2240–2245.
Metatungstate and tungstoniobate-containing LDHs: preparation, characteri- [102] U. Costantino, M. Nocchetti, Layered double hydroxides and their intercalation
sation and activity in epoxidation of cyclooctene, J. Phys. Chem. Chem. Solids compounds on photochemistry and in medicinal chemistry, in: V. Rives (Ed.),
68 (2007) 1872–1880. Layered Double Hydroxides: Present and Future, Nova Sci. Pub. Inc., New York,
[71] V. Rives, O. Prieto, A. Dubey, S. Kannan, Synergistic effect in the hydroxylation of 2001, pp. 383–411.
phenol over CoNiAl ternary hydrotalcites, J. Catal. 220 (2003) 161–171. [103] A.S. Tarnawsky, M. Tomikawa, M. Ohta, I.J. Sarfeh, Antacid talcid activates in gas-
[72] J.S. Yoo, A.A. Bhattacharyya, C.A. Radiowski, De-SOx catalysts: an XRD study of tric mucosa genes encoding for EGF and its receptor. The molecular basis for its
magnesium aluminate spinel and its solid solutions, Ind. Eng. Chem. Res. 30 ulcer healing action, J. Physiol. Paris 94 (2000) 93–98.
(1991) 1444–1448. [104] A. Schmassmann, A. Tarnawski, H.A. Gerber, F. Flogerzi, M. Sanner, L. Varga, F.
[73] A.A. Nikolopoulos, B.W.L. Jang, J.J. Spivey, Acetone condensation and selective to Halter, Antacid provides better restoration of glandular structures within the
MIBK on Pd and Pt hydrotalcite-derived Mg–Al mixed oxide catalysts, Appl. gastric ulcer scar than omeprazole, Gut 35 (1994) 896–904.
Catal. A Gen. 296 (2005) 128–136. [105] G.B. Van der Voet, F.A. Wolf, Intestinal absorption of aluminium from antacids: a
[74] C.M. Jinesh, V. Rives, D. Carriazo, C.A. Antonyraj, S. Kannan, Influence of copper comparison between hydrotalcite and algeldrate, J. Toxicol. Clin. Toxicol. 24
on the isomerization of eugenol for as-synthetized NiCuAl ternary hydrotalcites: (1986) 545–553.
an understanding through physicochemical study, Catal. Lett. 134 (2010) [106] A. Ookubo, K. Ooi, H. Hayashi, Hydrotalcites as potential absorbents of intestinal
337–342. phosphate, J. Pharm. Sci. 81 (1992) 1139–1140.
[75] S. Velu, K. Suzuki, M. Okazaki, M.P. Kapoor, T. Osaki, F. Ohashi, Oxidative steam [107] M. del Arco, E. Cebadera, S. Gutiérrez, C. Martín, M.J. Montero, V. Rives, J.
reforming of methanol over CuZnAl(Zr)-oxide catalysts for the selective produc- Rocha, M.A. Sevilla, Mg, Al layered double hydroxides with intercalated indo-
tion of hydrogen for fuel cells: catalyst characterization and performance evalu- methacin: synthesis, characterization and pharmacological study, J. Pharm.
ation, J. Catal. 194 (2000) 373–384. Sci. 93 (2004) 1649–1658.
[76] B. Coq, D. Tichit, S. Ribet, Co/Ni/Mg/Al layered double hydroxides as precursors [108] J.H. Yang, Y.S. Han, M. Park, T. Park, S.J. Hwang, J.H. Choy, New inorganic-based
of catalysts for the hydrogenation of nitriles: hydrogenation of acetonitrile, J. drug delivery system of indole-3-acetic acid-layered double hydroxide
Catal. 189 (2000) 117–128. nanohybrids with controlled release rate, Chem. Mater. 19 (2007) 2679–2685.
[77] B.F. Sels, D.E. de Vos, P.A. Jacobs, Hydrotalcite-like anionic clays in catalytic or- [109] D.M. O'Hare, European Patent 1341556 (2002).
ganic reactions, Catal. Rev. Sci. Eng. 43 (2001) 443–488. [110] V. Ambrogi, G. Fardella, G. Grandolini, M. Nocchetti, L. Perioli, Effect of hydrotalcite-
[78] G. Centi, S. Perathoner, Catalysis by layered materials: a review, Microporous like compounds on the aqueous solubility of some poorly water-soluble drugs, J.
Mesoporous Mater. 107 (2008) 3–15. Pharm. Sci. 92 (2003) 1407–1418.
[79] D. Tichit, B. Coq, Catalysis by hydrotalcites and related materials, CATTECH 7 [111] J.H. Choy, S.Y. Kwak, J.Y. Jeong, J.S. Park, Inorganic layered double hydroxides as
(2003) 206–217. nonviral vectors, Angew. Chem. Int. Ed. Engl. 39 (2000) 4041–4045.
[80] F. Figueras, Base catalysis in the synthesis of fine chemicals, Top. Catal. 29 (2004) [112] B. Jakubikova, F. Kovanda, Utilization of layered double hydroxides in medicinal
189–196. applications, Chem. List. 104 (2010) 906–912.
[81] V. Rives, D. Carriazo, C. Martín, Heterogeneous Catalysis by Polyoxometalate- [113] V.R.R. Cunha, A.M.C. Ferreira, V.R.L. Constantino, Hidróxidos duplos laminares:
intercalated Layered Double Hydroxides, in: A. Gil, S.A. Korili, M.A. Vicente, R. Nanopartículas inorgânicas para armazenamento e liberaçao de espécies de
Trujillano (Eds.), Springer, New York, 2010, pp. 319–397. interesse biológico e terapêutico (in Portuguese), Quim. Nova 33 (2010) 159–171.
[82] L. Perioli, V. Ambrogi, B. Bertini, M. Ricci, M. Nocchetti, L. Latteeini, C. Rossi, An- [114] A.I. Khan, D.M. O'Hare, Intercalation chemistry of layered double hydroxides:
ionic clays for sunscreen agent safe use: photoprotection, photostability and recent developments and applications, J. Mater. Chem. 12 (2002) 3191–3198.
prevention of their skin penetration, Eur. J. Pharm. Biopharm. 62 (2006) [115] K. Ladewig, Z.P. Xu, G.Q. (Max) Lu, Layered double hydrocapability oxide
185–193. nanoparticles in gene and drug delivery, Expert Opin. Drug Deliv. 6 (2009) 907–922.
[83] S.Y. Kwak, W.M. Kriven, M.A. Wallin, J.H. Choy, Inorganic delivery vector for in- [116] B. Li, J. He, D.G. Evans, X. Duan, Inorganic layered double hydroxides as a drug
travenous injection, Biomaterials 25 (2004) 5995–6001. delivery system — intercalation and in vitro release of fenbufen, Appl. Clay Sci.
[84] W.M. Kriven, S.Y. Kwak, M.A. Walling, J.H. Choy, Bio-resorbable nanoceramics 27 (2004) 199–207.
for gene and drug delivery, MRS Bull. 29 (2004) 33–37. [117] M. del Arco, A. Fernández, C. Martín, V. Rives, Solubility and release of fenbufen
[85] S.Y. Kwak, Y.J. Jeong, J.S. Park, J.S. Choy, Bio-LDH nanohybrid for gene therapy, intercalated in Mg, Al and Mg, Al, Fe layered double hydroxides (LDH): the effect
Solid State Ionics 151 (2002) 229–234. of Eudragit© S 100 covering, J. Solid State Chem. 183 (2010) 3002–3009.
[86] J.H. Choy, J.S. Jung, J.M. Oh, M. Park, K.M. Sohn, J.W. Kim, Inorganic-biomolecular [118] B. Li, J. He, D.G. Evans, X. Duan, Enteric-coated layered double hydroxides as a
hybrid nanomaterials as a genetic molecular code system, Adv. Mater. 16 (2004) controlled release, Int. J. Pharm. 287 (2004) 89–95.
1181–1184. [119] A.I. Khan, L. Lei, J.N. Norquist, D.M. O'Hare, Intercalation and controlled release of
[87] Z.P. Xu, G.Q. (Max) Lu, Layered double hydroxide nanomaterials as potential cel- pharmaceutically active compounds from a layered double hydroxide, Chem.
lular drug delivery agents, Pure Appl. Chem. 78 (2006) 1771–1779. Commun. (2001) 2342–2343.
[88] Z.P. Xu, Q.H. Zeng, G.Q. Lu, A.B. Yu, Inorganic nanoparticles as carriers for effi- [120] M. Wei, S. Shi, J. Wang, Y. Li, X. Duan, Studies on the intercalation of naproxen
cient cellular delivery, Chem. Eng. Sci. 61 (2006) 1027–1040. into layered double hydroxide and its thermal decomposition by in situ FT-IR
[89] S.K. Dey, R. Sistiabudi, Ceramic nanovector based on layered double hydroxide: and in situ HT-XRD, J. Solid State Chem. 177 (2004) 2534–2541.
attributes, physiologically relevant compositions and surface activation, Mater. [121] M.R. Berber, K. Minagawa, M. Katoh, T. Mori, M. Tanaka, Nanocomposites of
Res. Innov. 11 (2007) 108–117. arylpropionic acid drugs based on Mg–Al layered double hydroxide for dissolu-
[90] K.M. Tyner, E.P. Giannelis, Nanobiohybrids: novel gene and drug delivery sys- tion enhancement, Eur. J. Pharm. Sci. 35 (2008) 354–360.
tems, Arch. Appl. Biomater. Biomol. Mater. 1 (2004) 449–451. [122] M. del Arco, S. Gutiérrez, C. Martín, V. Rives, J. Rocha, Synthesis and character-
[91] K.M. Tyner, S.R. Schiffman, E.P. Giannelis, Nanobiohybrids as delivery vehicles isation of layered double hydroxides (LDH) intercalated with non-steroidal
for camptothecin, J. Control. Release 95 (2004) 501–514. anti-inflammatory drugs (NSAID), J. Solid State Chem. 177 (2004) 3954–3962.
[92] J.H. Choy, J.S. Jung, J.M. Oh, M. Park, J. Jeong, Y.K. Kang, O.J. Han, Layered double [123] D. Carriazo, M. del Arco, C. Martín, C. Ramos, V. Rives, Influence of the inorganic
hydroxide as an efficient drug reservoir for folate derivatives, Biomaterials 25 matrix nature on the sustained release of naproxen, Microporous Mesoporous
(2004) 3059–3064. Mater. 130 (2010) 229–238.
[93] J.M. Oh, J.W. Kim, J.W. Jung, Y.K. Kang, J.H. Choy, Efficient delivery of anticancer drug [124] M. del Arco, A. Fernández, C. Martín, V. Rives, Release studies of different NSAIDs
MTX through MTX–LDH nanohybrid system, J. Phys. Chem. Solids 67 (2004) encapsulated in Mg, Al, Fe-hydrotalcites, Appl. Clay Sci. 42 (2009) 538–544.
1024–1027. [125] P. Costa, J.M. Sousa, Modeling and comparison of dissolution profiles, Eur. J.
[94] Z.L. Wang, E.B. Wang, L. Gao, L. Xu, Synthesis and properties of Mg2Al layered Pharm. Sci. 13 (2001) 123–133.
double hydroxides containing 5-fluorouracil, J. Solid State Chem. 178 (2005) [126] W.G. Hou, Z.L. Jin, Synthesis and characterization of naproxen intercalated Zn–Al
736–741. layered double hydroxides, Colloid Polym. Sci. 285 (2007) 1449–1454.
[95] J.Y. Kim, S.J. Choi, J.M. Oh, T. Park, J.H. Choy, Anticancer drug-inorganic nanohybrid [127] L. Perioli, V. Ambrogi, L. Di Nauta, M. Nocchetti, C. Rossi, Effects of
and its cellular interaction, J. Nanosci. Nanotechnol. 7 (2007) 3700–3705. hydrotalcite-like nanostructured compounds on biopharmaceutical properties
V. Rives et al. / Journal of Controlled Release 169 (2013) 28–39 39

and release of BCS class II drugs: the case of flurbiprofen, Appl. Clay Sci. 51 [137] V. Ambrogi, G. Fardella, G. Grandolini, L. Perioli, Intercalation compounds of
(2011) 407–413. hydrotalcite-like anionic clays with antiinflammatory agents, I: intercalation
[128] V. Ambrogi, G. Fardella, G. Grandolini, L. Perioli, M.C. Tiralti, Intercalation com- and in vitro release of ibuprofen, Inter. J. Pharm. 220 (2001) 23–32.
pounds of hydrotalcite-like anionic clays with antiinflammatory agents. 2. Up- [138] V.H. DeLeon, T.D. Nguyen, M. Nar, N.A. D'Souza, T.D. Golden, Polymer
take of diclofenac for a controlled release formulation, AAPS PharmSciTech 3 nanocomposites for improved drug delivery efficiency, Mater. Chem. Phys. 132
(2003), (article 26). (2012) 409–415.
[129] T. Higuchi, Mechanism of sustained-action medication. Theoretical analysis of [139] A.N. Ay, B. Zümreoglu-Karan, A. Temel, V. Rives, Bioinorganic magnetic
rate of release of solid drugs dispersed in solid matrices, J. Pharm. Sci. 52 core-shell nanocomposites carrying antiarthritic agents: intercalation of ibupro-
(1963) 1145–1149. fen and glucuronic acid into Mg–Al layered double hydroxides supported on
[130] V. Ambrogi, L. Perioli, M. Ricci, L. Pulcini, M. Nocchetti, S. Giovagnoli, C. Rossi, magnesium ferrite, Inorg. Chem. 48 (2008) 8871–8877.
Eudragit® and hydrotalcite-like anionic clay composite system for diclofenac [140] H. Zhang, K. Zou, H. Sun, X. Duan, A magnetic organic–inorganic composite:
colonic delivery, Microporous Mesoporous Mater. 115 (2008) 405–415. synthesis and characterization of magnetic 5-amino salicylic acid intercalated
[131] M.L. Lorenzo-Lamosa, C. Remunan-Lopez, Y.L. Vita-Joto, M.J. Alonso, Design of layered double hydroxides, J. Solid State Chem. 178 (2005) 3485–3493.
microencapsulated chitosan microspheres for colonic drug delivery, J. Control. [141] H. Zhang, D. Pan, K. Zou, J. He, X. Duan, A novel core–shell structured magnetic
Release 52 (1998) 109–118. organic–inorganic nanohybrid involving drug-intercalated layered double hy-
[132] L. Perioli, T. Posati, M. Nocchetti, F. Bellezza, U. Costantino, A. Cipiciani, Interca- droxides coated on a magnesium ferrite core for magnetically controlled drug
lation and release of antiinflammatory drug diclofenac into nanosized ZnAl release, J. Mater. Chem. 19 (2009) 3069–3077.
hydrotalcite-like compound, Appl. Clay Sci. 53 (2011) 374–378. [142] G. Carja, H. Chiriac, N. Lupu, New magnetic organic–inorganic composites based
[133] Y.J. Wang, F. Cao, Q.N. Pin, Preparation and characterisation of diclofenac sodium on hydrotalcite-like anionic clays for drug delivery, Magn. Magn. Mater. 311
layered double hydroxide nanohybrids, J. China Pharm. Univ. 40 (2009) 321–326. (2007) 26–30.
[134] G. Sammartino, G. Marenzi, L. Tammaro, A. Bolognese, A. Calignano, U. [143] M. del Arco, F. Fernández, C. Martín, V. Rives, Intercalation of mefenamic and
Costantino, L. Califano, V. Vittoria, Anti-inflammatory drug incorporation into meclofenamic acid anions in hydrotalcite-like matrixes, Appl. Clay Sci. 36
polymeric nano-hybrids for local controlled release, Int. J. Inmunopathol. (2007) 133–140.
Pharmacol. 18 (2005) 55–62. [144] M. del Arco, A. Fernández, C. Martín, M.L. Sayalero, V. Rives, Solubility and re-
[135] P. Gunawan, R. Xu, Direct control of drug release behaviour from layered double lease of fenamates intercalated in layered double hydroxides, Clay Miner. 43
hydroxides through particle interactions, J. Pharm. Sci. 97 (2008) 4367–4378. (2008) 255–265.
[136] M.S. San Román, M.J. Holgado, B. Salinas, V. Rives, Characterisation of diclofenac,
ketoprofen or chloramphenicol succinate encapsulated in layered double hy-
droxides with the hydrotalcite-like structure, Appl. Clay Sci. 55 (2012) 158–163.

You might also like