You are on page 1of 31

Applied Clay Science 8889 (2014) 239269

Contents lists available at ScienceDirect

Applied Clay Science


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

Review article

Intercalation of drugs in layered double hydroxides and their controlled release: A review
Vicente Rives , Margarita del Arco, Cristina Martn
GIR-QUESCAT, Departamento de Qumica Inorgnica, 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
The intercalation of different drugs in layered double hydroxides with the hydrotalcite-like structure is reviewed. The intercalation processes are carried out following different routes (direct synthesis, coprecipitation, anion exchange) and the advantages and disadvantages of these methods are described for the specic drug/LDH system studied. Characterisation of the intercalation compounds is also studied, to determine the way the guest molecules are intercalated between the layers of the layered double hydroxide. The controlled release (in some cases also the kinetics analysis) is also studied. We conclude that layered double hydroxides are very suitable materials to host different families of drugs and in the controlled release they show benetial properties, on comparing with the effect of the bulk drug. It should be also stressed that these are almost the unique materials (in addition to layered hydroxy salts) able to host drugs in the anionic form, so nicely completing the studies carried out so far on the suitability of cationic clays to host cationic or neutral drugs. 2013 Elsevier B.V. All rights reserved.

Article history: Received 16 August 2013 Received in revised form 28 November 2013 Accepted 2 December 2013 Available online 28 December 2013 Keywords: Layered double hydroxide Drug intercalation Controlled delivery Organicinorganic intercalation compounds Hydrotalcite

1. Introduction Among the few families of layered compounds with positively charge layers, the so-called layered double hydroxides (LDHs), also known very often as hydrotalcite-like systems, have deserved a lot of attention. Hydrotalcite is a natural occurring hydroxycarbonate of Mg and Al discovered in Sweden in 1842. The structure of hydrotalcite is similar to that of brucite, Mg(OH)2, an hexagonal close packing of hydroxyl groups, where magnesium cations ll all octahedral holes every two layers. Isomorphic, partial, Mg/Al substitution gives rise to development of a positive charge in the layers, which is balanced by hydrated carbonate anions located in the interlayers which originally were empty in brucite. A scheme is depicted in Fig. 1. The brucite-like layers can be stacked in different ways leading to different structures (Drits and Bookin, 2001; Evans and Slade, 2006), the most common ones being rhombohedral (3R symmetry) and hexagonal (2H symmetry). Probably, one proof of the increasing interest in these solids can be the continuous publication of reviews and monographs on their preparation, characterisation, properties and applications (Costantino et al., 2013; Duan and Evans, 2006; Forano et al., 2013; Rives, 2001; Wypych and Satyanarayana, 2004; Zumreoglu-Karan and Ay, 2012). The cations in the layers can be substituted by many others, such as Zn, Co, Ni, Mn, Fe (divalents), and Cr, Co, Fe, V, Y, Mn, Ga, lanthanides (trivalents) and also the interlayer anion can be substituted by a great variety of simple or complex anions, e.g., simple inorganic anions (Constantino and Pinnavaia, 1995), organic anions (Jaubertie et al.,
Corresponding author. Tel.: +34 923 294489; fax: +34 923 294574. E-mail address: vrives@usal.es (V. Rives). 0169-1317/$ see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.clay.2013.12.002

2006; Newman and Jones, 1998; San Romn et al., 2006), coordination compounds (Bhattacharjee and Anderson, 2006; Del Arco et al., 2003; Rives and Ulibarri, 1999), polyoxometalates (Carriazo et al., 2006a,b; Del Arco et al., 2004a,b; Hu and Li, 2004; Rives and Ulibarri, 1999), biomolecules or even DNA (Choy et al., 1999, 2001; Desigaux et al., 2006), etc. Concerning the layer cations, their ionic radii are always close to that of Mg2+ (0.72 ), except that of Al3+ (0.50 ), as distortions arise for larger cations. Regarding the anions, their size/charge ratio is important, as large anions with low charge are unable to balance homogeneously the positive charge of the layers. A compromise should be reached between the layer charge density and the dimensions of guests species in the interlayer. For non-spherical anions, and very specially when the anions contain long chains (e.g., carboxylates or sulfonates with long alkyl chains), several arrangements in the interlayer are possible, namely, a monolayer parallel to the layers, a parallel bilayer or tilted monolayers or bilayers. This versatility in the chemical composition leads to many and different potential applications. The current interest in LDHs is founded on several properties: They are basic materials and the mixed oxides formed upon thermal decomposition show even a larger basicity, related to the oxide anions; intercalation of acidic anions provides systems with unique acidbase properties. They show the so-called memory effect (Chibwe and Jones, 1989; Kwon and Pinnavaia, 1989), i.e., the ability to recover their original layered structure when mixed oxides (previously prepared by calcination of some LDHs at moderate temperatures) are put in contact with solutions containing anions. They also show anion exchange capacity (AEC), usually larger than that shown by cationic clays, ranging between 2 and 4 mEq/g.

240

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

Fig. 1. Idealised structure of a layered double hydroxide, with interlayer carbonate anions. Reprinted from Coordination Chemistry Reviews, Vol. 181. Vicente Rives and Maria Angeles Ulibarri, Layered double hydroxides (LDH) intercalated with metal coordination compounds and oxometalates, pp. 61120. Copyright (1999), with permission from Elsevier.

mixed oxides formed upon calcination can be used to scavenge anions from solutions on recovering the layered structure (Forano, 2004; Goh et al., 2008; Ulibarri and Hermosn, 2001). The intercalation of sodium alginate in ZnAl and MgAl LDHs gave rise to an increase in the adsorption capacity of the intercalated systems above those shown by alginate or the clay used individually for water treatment, such as removal of uorine ions or Orange II dye (Mandal et al., 2012). - Materials science mainly as additives to organic polymers; the composite formed exhibits enhanced mechanical properties, mainly related to the aspect ratio of the lamellar particles of the hydrotalcite (Kaluskova et al., 2004; Leroux and Besse, 2004; Wang and Zhang, 2004); they also act as ame retardants (Chen and Qu, 2003; Pereira et al., 2009) and possess barrier effects (Sorrentino et al., 2005). - Catalysis. Original LDHs exhibit strong basic properties and can be used as heterogeneous catalysts, to avoid environmental problems when using soluble basic catalysts (Jinesh et al., 2010; Rives et al., 2003). They can be also used as catalyst supports or catalyst precursors. Several reviews have reported on the catalytic properties of different families of LDHs (Albertazzi et al., 2004; Basile and Vaccari, 2001; Cavani et al., 1991; Centi and Perathoner, 2008; Costantino et al., 2013; Figueras, 2004; Monzn et al., 2001; Rives et al., 2010; Seis et al., 2001; Tichit and Coq, 2003). - Medicine. Biologically-active molecules are also among the different sorts of molecules which can be intercalated between the brucitelike layers of LDHs, a property which has opened their applications in Medicine and Pharmacy. One of the rst reviews on this subject was published in 2001 by Costantino and Nocchetti (2001), which has been followed by the works by Xu and Lu (2006), Choy et al. (2007), Ladewig et al. (2009), Jakubikov and Kovanda (2010), Chakraborty et al. (2010a,b), Cunha et al. (2010), Wang and O'Hare (2012) and Costantino et al. (2009, 2013). The applicability of LDHs in this eld is based mainly on three properties: increased solubility of the drug, basicity of the LDH matrix, and ability for drug controlled release. The addition of LDHs usually improves the solubility of the drugs (Perioli and Pagano, 2012; Perioli et al., 2013) without modifying their chemical structure and thus their pharmacological activity. On the other hand, the intrinsic basicity of the LDH structure provides these materials with antacid properties (Parashar et al., 2012.), which are dependent on the route followed to prepare the Mg,Al hydrotalcite. Finally, it is generally observed that on suspending a drugLDH intercalation compound in aqueous solutions, release of the drug follows a two step process. First a rapid release, and secondly a maintained, slow release, usually related to anionic exchange with anions of the medium. In this scenario, our aim in this work has been to provide a broad description of the interactions between several drugs and layered double hydroxides. We have excluded the interaction of LDHs with non-steroidal antiinammatory drugs, which has been recently reviewed elsewhere (Rives et al., 2013). Our work summarises the information published roughly from 2000 and we have organised the work attending to the pharmacological properties shown by the different drugs. We have given priority importance to the preparation procedures, as many of the papers reviewed have undoubtedly demonstrated that the ability for drug insertion and the release process is dependent on the preparation method followed, probably because of the different physicochemical properties (particle size, crystallinity, etc.) shown by the same compound, when prepared following different routes. Where available, data on the release of the drug are also included. 2. Antibiotics Encapsulation of penicillin in an inorganic host and the use of such encapsulation composites as sustained-release medications have been scarcely studied in the literature. Li et al. (2006) have reported the intercalation of phenoxymethylpenicillin (PMP), a member of the penicillin

Among the many ways proposed to prepare these solids (De Roy et al., 2001; Forano et al., 2013; He et al., 2006; Kanezaki, 2004), the main routes most frequently followed are: - Coprecipitation, consisting of the slow addition of a solution of the metal cations into a reactor containing the anion to be intercalated; increasing the pH by addition of a base or urea hydrolysis that leads to precipitation of the LDH. - Anionic exchange of anions originally existing in the interlayer of a LDH prepared usually by coprecipitation. Chloride or nitrate is preferred as the original anions, as the exchange is easier than for multicharged anions. The reaction is usually carried out by stirring the LDH precursor in a solution containing an excess of the anion to be intercalated; application of ultrasounds speeds up the exchange process (Kooli et al., 1997). - The reconstruction method is based on the memory effect (Chibwe and Jones, 1989; Kwon and Pinnavaia, 1989) shown by the product formed upon mild calcination (ca. 500 C under dynamic inert gas atmosphere (Del Arco et al., 1994; Rocha et al., 1999)) of a LDH (usually in the nitrate or carbonate form), i.e., the ability to recover a layered structure when the mixed oxide is immersed in a solution of the anions to be intercalated. - Hydrothermal and microwave treatments are often applied to improve the crystallinity and other properties of the LDHs (Benito et al., 2006a,b,c, 2008a,b, 2009; Herrero et al., 2007a,b, 2009). Layered double hydroxides or the homogeneously dispersed mixed oxides formed upon their calcination nd applications in different elds: - Water decontamination; based on the large AEC, hydrotalcites can be used to adsorb polluting anions from aqueous solutions, while the

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

241

family (Foye et al., 1995), in a Mg,Al hydrotalcite. First a reference Mg, AlCO3 material, with a molar Mg/Al ratio of 3, was prepared by coprecipitation at pH 9.0 and ageing at 100 C for 24 h; the composite (HTP) was prepared by reconstruction, starting from the parent Mg, AlCO3 hydrotalcite calcined in air at 500 C for 5 h, immersed in a PMP aqueous solution. These authors consider that for biomedical applications, the reconstruction method, based on the well known memory effect of hydrotalcites (Chibwe and Jones, 1989; Kwon and Pinnavaia, 1989), is much better than anion exchange from hydrotalcites containing chloride or nitrate in the interlayer, because of blood overpressure in the rst case, and because of reduction of nitrate anions to nitrite ones (able to cause many kinds of cancers) by microorganisms in the living bodies in the second one if exchange of chloride or nitrate anions was not complete. The composite consisted of thin akes (as seen by SEM, Fig. 2) with a rough thickness around 30 nm and size ranging from 100 to 300 nm; the PXRD pattern indicated a d spacing of 18.7 , thus suggesting formation of an interdigited monolayer of PMP between the hydrotalcite layers. The mass percentage of PMP was 33%. Upon incubation at 27 C overnight it was found that HTP was active against bacteria, while the hydrotalcite was inactive. Soaking of HTP at various pH values, followed by the addition of the solution to an Oxford cup placed on the surface of the culture medium (Staphyloloccus aureaus) in the Petri dish, and treated at 27 C overnight, clearly demonstrated an increase in the area of the inhibition zone with time and the decrease of the initial pH. The lowest pH tested was 2.9 (reaching nally a value of 5.5, thus further demonstrating also the ability of hydrotalcite systems to reduce acidity), but clearly this should lead to an almost complete dissolution and destruction of the basic layered structure. These authors concluded that oral administration of HTP would lead to major dissolution of the inorganic structure and release of the drug in the stomach (pH from 0.9 to 1.5), while in the small intestine (pH 67.5) release would arise from anion exchange with HCO 3 or Cl anions existing in the small intestine.

Fig. 2. Scanning electron micrograph of the HTP sample. Reprinted from Journal of Chemical Technology and Biotechnology, Vol. 81. Wen-Zhuo Li, Jun Lu, Jie-Sheng Chen, Guo-Dong Li, Yu-Sheng Jiang, Lian-Sheng Li, and Bai-Qu Huang, Phenoxymethylpenicillin-intercalated hydrotalcite as a bacteria inhibitor, pp. 89 93. Copyright (2005) Society of Chemical Industry.

Trikeriotis and Ghanotakis (2007) have studied the intercalation and controlled release of four antibiotics, two hydrophobic (gramicidin D and amphotericin B) and two hydrophilic (ampicillin and nalidixic acid) in LDHs. Intercalation was achieved by anion exchange from a parent Mg,Al-NO3 hydrotalcite (Mg/Al molar ratio 2) prepared under CO2free conditions and aged for 1 day at room temperature in the mother liquor. As gramicidin is hydrophobic and cannot be intercalated directly via anion exchange, it was rst solubilised in a solution of sodium cholate (a negatively charged micelle) and this was then intercalated via anion exchange. The d-spacing measured (35.3 ) for a sample containing 2.2% (w/w) of antibiotic showed the formation of a bilayer of cholate molecules in the interlayer space; this value was very similar to that measured for a hydrotalcitecholate system (33.9 ), and so it is probable that the gramicidin molecules were hosted in a bilayer of cholate molecules forming a sort of articial membrane in the interlayer space. Due to its chemical nature, gramicidin was not released in water, but was completely released in ethanol in merely 2 min; a slow release prole was measured in a solution containing dodecyl maltoside micelles; these authors concluded that using this intercalation compound in a controlled release system would lead to specic release of the drug in biological membranes. Amphotericin B was intercalated following two routes: rst same as gramicidin, via the micelle method (reaching an antibiotic content of 2.7%), or via direct anion exchange (9.7% antibiotic content), as the gramicidin molecule contains a carboxylate group. The behaviour observed for the rst sample was similar to that above commented for the LDHcholategramicidin system, i.e., the antibiotic molecules were hosted in the intercalated cholate bilayer (d-spacing 35.9 ) without aggregation of the antibiotic molecules, as conrmed by FT-IR and VisUV spectroscopy; in the second case (d-spacing 34 , 9.7% w/w drug) a strong contamination by unexchanged nitrate existed. Release was not observed in water or in several organic solvents, but the antibiotic was released specically in biological membranes and not in the bulk solution. Ampicillin and nalidixic acid were intercalated by mixing antibiotic solutions and LDH suspensions; almost complete exchange (corresponding to 51.7 and 40% w/w of the drug, respectively) was achieved. Despite the molecular size of ampicillin is larger than that of nalidixic acid, the d-spacing in the former case (20.6 ) was smaller than in the second one (22.3 ), pointing to formation of a bilayer in the second case. Spectroscopy measurements indicated that ampicillin was intercalated unaltered as an anion, while for nalidixic acid the UVvis spectrum suggested the simultaneous presence of anionic and neutral forms. Release was achieved by simply suspending the intercalation compounds in water containing NaCl or in plain water at pH 7, probably by anion exchange with chloride anions; the release was faster for ampicillin. These authors assumed that this was due to the large amount of intercalated ampicillin, and the fact that in plain water it was more extensive for nalidixic acid was claimed to be due to the presence of both neutral and anionic species. However, molecular size calculations to check the room available in the interlayer are lacking and so it is also possible that some of the antibiotic molecules were simply adsorbed on the external surface of the inorganic host, so being easily released in aqueous suspensions. San Romn et al. (2012) have reported the intercalation of chloramphenicol within the interlayer space of Zn,Al and Zn,Mg,Al LDHs, prepared by coprecipitation. The divalent/Al3+ molar ratio was close to 2 in both cases; a solution of the salts was dropwise added to a solution of the succinate salt of the drug in decarbonated water at pH 9 and the suspension stirred for 24 h at room temperature in a nitrogen atmosphere. The interlayer spacing, as measured by X-ray diffraction, was ca. 12 . Minor signals due to the sodium salt of chloramphenicol succinate (probably adsorbed on the external surface of the crystallites) were also recorded in the powder X-ray diffraction diagram; a minimum amount of a nitrate-containing phase was also detected, of which presence was conrmed by FT-IR spectroscopy. From the lattice parameters of the

242

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

solid, it was concluded that the molecules in the interlayer were oriented parallel to the brucite-like layers, as previously reported for lactate anions intercalated in a Mg,Al LDH (San Romn et al., 2006). Frunza et al. (2008) have incorporated chloramphenicol in Mg,Al and Zn,Al hydrotalcites by direct synthesis, anion exchange and reconstruction, Fig. 3. The PXRD diagrams recorded were typical of well crystallized hydrotalcites, and intercalation was further conrmed by FTIR spectroscopy, as bands of the hydrotalcite host and of the guest drug molecules were recorded in the spectrum of the intercalation compound. The thermal stability and the kinetic degradation of the drug were deeply dependent on the preparation method followed. Tammaro et al. (2007) have prepared a new polymer composite with chloramphenicol succinate intercalated in a Mg,Al hydrotalcite prepared by ion-exchange, and covered by a polymer lm (-polycaprolactone). The PXRD diagram conrmed intercalation of the drug without decomposition; intercalation also gave rise to an increase in the decomposition temperature of the drug, as concluded from thermal analysis studies (TG and DTA), combustion taking place around 440 C. Release of the drug showed a rather interesting behaviour, with a rst, fast step where a small amount of the drug was released, followed by a much slower process; probably the rst release corresponded to desorption from the surface of the crystallites, while the second one corresponded to release of intercalated chloramphenicol. Cefazolin is a broad spectrum -lactamic antibiotic with a high activity against gram positive and gram negative bacteria. Despite its advantages if compared to penicillin and other cefalosporins, it also shows several drawbacks, related to its chemical stability and medical use. It can be easily denaturalized by hydrolysis and decomposes at 65 C. Concerning its medical use, it needs to be supplied frequently to maintain required therapeutic levels. For these reasons, it is very important to develop matrices suitable to prepare controlled release composites to protect the drug and to increase its efcacy. Ryu et al. (2010) have used ionic exchange to intercalate cefazolin in a Zn,Al-nitrate LDH which had been previously prepared by coprecipitation at pH 7 under a nitrogen atmosphere to exclude carbon dioxide from the reaction medium. The intercalation compound was prepared by dissolving cefazolin (in its sodium form) in an

Fig. 3. Sketch of the possible location of chloramphenicol succinate anions in the interlayer space of ZnAl layered double hydroxides. Reprinted from Applied Clay Science, Vol. 55. Mara Soledad San Romn, Mara Jess Holgado, Beatriz Salinas, Vicente Rives, Characterisation of Diclofenac, Ketoprofen or Chloramphenicol Succinate encapsulated in layered double hydroxides with the hydrotalcite-type structure, pp. 158 163. Copyright (2012), with permission from Elsevier.

ethanol:water mixture (70:30), using a cefazolin:LDH ratio of 2:1, and stirring for 24 h under a nitrogen atmosphere, leading to 37% of cefazolin incorporation. Intercalation was assessed by PXRD, which showed a swelling of the layers from 8.9 for the nitrate precursor to 18 for the cefazolin-containing solid after exchange. From the gallery height (13.2 ) and the size of the intercalated drug, these authors concluded formation of a 62-tilted, intercalated, double layer of drug molecules. Release of the drug was studied by HPLC, both in water and in a 0.8% (w/w) NaCl aqueous solution; while in water only 25% of the drug was released after 90 min, 66% was released after merely 39 min in the NaCl solution, probably because the presence of chloride ions in the medium favoured the anion exchange; this could be checked also by performing the study in a basic medium, where drug/hydroxyl exchange could have taken place. It should be stressed that cefazolin was recovered unaltered in both cases. The antibacterial activity of cefazolin and of the intercalation compound has been studied in vitro by programmed diffusion from the disc, using distilled water and the same 0.8% NaCl solution used for the release studies; the microorganism to be inhibited was S. aureus ATCC 25923 (108 UFC/mL). Large differences were observed in the zones of bacterial inhibition for sodium cefazolin and the cefazolin LDH intercalation compound, which have been related to the fast dissolution of sodium cefazolin (both in water and the NaCl solution), vs. the time-dependent solubilisation of the intercalation compound, Fig. 4. From these results, the authors (Ryu et al., 2010) concluded that LDHs behave as very suitable inorganic matrices, also improving the antibacterial performance of cefazolin. Wang et al. (2009) have intercalated amoxicillin a hydrophylic antibiotic of the penicillin group used to treat many different bacterial infections in a Zn,Al-LDH by rehydration of a calcined LDH precursor, the so-called memory effect (Chibwe and Jones, 1989; Kwon and Pinnavaia, 1989). The parent LDH had been prepared by the coprecipitation method in a suspension containing Fe3O4 (Zn2+/Fe2+ molar ratio = 30) at pH 10 and at 65 C, to obtain a magneticallyfunctionalized LDH. This LDH was calcined at 400 C for 24 h and then it was added to an aqueous solution of sodium amoxicillin, and stirred for 24 h at room temperature under a nitrogen atmosphere. The powder X-ray diffraction diagram of the Zn,Al-LDH precursor showed only the signals due to the LDH, with a basal spacing of 7.5 for planes (003), suggesting that magnetite was undetectable because it was highly dispersed in the form of very small crystallites. Insertion of the amoxicillin anion in the interlayer was conrmed both from PXRD and FT-IR spectroscopy experiments. The former showed a swelling up to 12.35 for the basal spacing of planes (003), which corresponded to a gallery height of 7.55 , slightly shorter than the length of the amoxicillin molecule (9.41 ); the authors suggested that although a monolayer of drug molecules was formed, this was tilted 37. Intercalated amoxicillin was thermally more stable than in the bulk form, as thermogravimetric analysis has shown that its oxidative decomposition took place at 700 C instead of 562 C, probably because of the strong interactions developing between the intercalated molecules, on one side, and the interlayer water molecules and the brucite-like sheets, on the other. Release studies have been carried out at 37 C in a phosphate buffer (pH 7.4), comparing the results with those obtained with a physical mixture of 0.05 g amoxicillin and 0.08 g of Zn,Al-CO3 LDH. Complete dissolution was attained for the physical mixture after 30 min, while with the intercalated sample only 85% of drug was released after 2 h, due to the strong electrostatic interactions between the drug and the layers. These results further conrmed the suitability of these intercalated systems for controlled release of this drug. Magnetic studies revealed that the intercalation compound displayed a paramagnetic behaviour at room temperature, associated to the presence of magnetite particles dispersed between the hydrotalcite crystallites. These properties represented an important advantage of these compounds for transport and vectorization of drugs, as these intercalation compounds could be

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

243

Fig. 4. Zones of bacterial inhibition of (a) LDH, (b) cefazolin sodium, and (c) cefazolinLDH nanohybrid in (A) water and (B) a 0.8% NaCl solution. Reprinted from The Journal of Physics and Chemistry of Solids. Vol. 71. Seung-Jin Ryu, Hyun Jung, Jae-Min Oh, Jin-Kyu Lee, Jin-Ho Choy, Layered double hydroxide as novel antibacterial drug delivery system, pp. 685688. Copyright (2010), with permission from Elsevier.

rapidly concentrated in pathological cells by application of a magnetic eld, which then can be easily killed. Valarezo et al. (2013) encapsulated amoxicillin-intercalated Zn,Al LDHs into poly(-caprolactone) by electrospinning (Frunza et al., 2008; Russo and Lamberti, 2011); these authors have highlighted that the drug should be intercalated prior to encapsulation for controlled delivery of the drug, especially for low mass weight drugs. Xu et al. (2009) have reported a detailed study of the interaction of tetracycline (TC) with hydrotalcite, to remove the antibiotic from water. These authors have studied samples with different Mg/Al ratios (2, 3, and 4) prepared by coprecipitation, hydrothermal treatment at 120 C for 24 h, and then calcination; the results were compared to those obtained for uncalcined hydrotalcite and with MgO treated under similar conditions. Calcined HTs (CHT) showed markedly higher adsorption capacities for aqueous TC than hydrotalcites and were dependent on both reconstruction of layer structures (by the memory effect of CHT) and adsorption on MgO surfaces. As concluded from PXRD studies, the layered structure was only partially restored from CHT, and the TC molecules were oriented parallel to the brucite-like layers. At high TC equilibrium concentrations, higher TC adsorption capacities were related to higher Mg/Al ratios in CHT.

Sui et al. (2012) have intercalated noroxacin (NOR) in LDHs with different Mg/Al ratios (2, 3, and 4), different anions (carbonate, chloride) in the interlayer and after adding also Sn4 + to the brucite-like layers, in order to determine their effect on the adsorption ability of the matrix for the drug. Those samples with larger Al3+ content and chloride as the interlayer anion showed a larger adsorption capacity and rate of NOR release; however, the presence of Sn4 + decreased both parameters. A Freundlich model tted well with the adsorption equilibrium data. Following the ion exchange method, Wang and Zhang (2012) have intercalated several antibiotics within a Mg,Al hydrotalcite, namely, benzoate (BZ), succinate (SU), benzylpenicillin (BP), and ticarcillin (TCC) anions. Formation of the corresponding drugLDH systems has been conrmed by PXRD and FT-IR spectroscopy. SU and TCC anions were intercalated in a monolayer arrangement, whereas BZ and BP anions were accommodated in the interlayer region as a bilayer. Release studies indicated that an anion exchange process holds diffusion through the particles being the rate determining step. Costantino et al. (2012) have published a review collecting the patents recently published on the intercalation of anti-inammatory (diclofenac), antibacterial (chloramphenicol hemisuccinate), and antibrinolytic (tranexamic acid) agents in Mg,Al and Zn,Al LDHs.

244

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

3. Anticancer agents Camptothecin (CPT) is a pentacyclic indol alkaloid, with the terminal ring easily converting between the lactone (biologically active) under acidic conditions (pH b 5), to the carboxylate form at pH N 8. The lactone phase is only very slightly soluble in water, with the concomitant poor dispersion in physiological solutions. Tyner et al. (2004) have encapsulated CPT in an anionic micelle derived from a biocompatible surfactant, which negative charge allows intercalation of the composite between the layers of a LDH. A Mg,Al-NO3 (Mg/Al molar ratio ca. 2) LDH was prepared by coprecipitation at pH 10, aged for 25 days at 70 C to minimize crystal aggregates and kept as a slurry. Sucrose aspartate surfactant (SAS) or sodium cholate (SC) were dispersed in water or chloroform, respectively, and the LDH slurry was injected in the micelle mixture, Fig. 5. Carmine (a bright-red pigment used as a food additive) was directly intercalated in the LDH by mixing the slurry and an aqueous solution of carmine. The d-spacing increased from 8.8 for the nitrate precursor to 29.8 and 32.7 for intercalation compounds with CPT loaded micelles of SAS and SC, respectively; observation by TEM showed formation of hexagonal shaped particles ca. 500 10 nm. Intercalation compounds containing CPT inhibited growth of 9L glioma cells as efciently as the pure drug, while controls (e.g., surfactant, LDH, water) showed no inhibition. Inhibition was less effective with intercalation compounds based on SC than on SAS, probably because of the more efcient encapsulating ability of the latter, which also showed a larger CPT loading (5.6%). Carmine was completely released after 40 min at pH 4.8, but 70 days were required at pH 7.2. Such a difference was due to different release mechanisms: at acidic pHs the inorganic layers were dissolved, while at and above pH 7 release was through exchange with anions existing in the medium from the buffer solution; this conclusion was conrmed from TEM observation of solids exposed to different pH environments, which in addition showed a d-spacing of 8 , compatible with the presence of chloride (from the Trizma buffer used) in the interlayer. Contrary to carmine, CPT was released in several minutes both at pHs 4.8 and 7.2. The difference might be due to the neutral nature of CPT, which could be released at pH 7.2 without requiring an anion exchange process, while the anionic nature of carmine demanded such an exchange process in order to be released. Direct intercalation of CPT in a Mg,Al (molar ratio ca. 3) LDH by coprecipitation without encapsulation in micelles has been reported by Liu et al. (2008b). Nitrate anions from the original metal salts were also intercalated together with hydroxyl groups, and so these authors were able to study two samples, with 16.72 (sample W-1) and 26.03 (sample W-2) % (w/w) drug loadings. The diffraction maxima due to planes (003) were recorded at 7.9 (intense) and 34.9 for sample

W-1, and 7.9 (intense) and 23.9 for sample W-2. These authors claim that the three maxima were due to three different orientations of the CPT molecule in the interlayer space, namely, at and parallel to the inorganic layers (7.9 ), forming an end-to-end bilayer (34.9 ), and an interdigitized bilayer (23.9 ), stabilized through interactions between CPT molecules, on the basis of a value of 8.8 for the (003) spacing due to a nitrate-containing sample. However, it should be remembered that such a spacing was related to the presence of nitrate anions in an upward orientation, when a large concentration of nitrate is required in the interlayer to balance a large positive charge of the layer for a Mg/Al ratio around 2. However, for larger Mg/Al ratios, when the positive charge of the layer is lower, at-oriented nitrate anions led to a spacing of 7.87.9 . Moreover, molecular size calculations would permit to determine if the interlayer space was large enough to admit at-oriented CPT molecules, but these calculations were not reported in this study. CPT release at pH 7.5 showed also a rst, fast step, followed by a slower, second one, the release rate and amount released being somewhat lower for sample W-2 (t1/2 values were 12 min for sample W-1 and 52 min for sample W-2); release was completed in 10 min from a CPTLDH physical mixture. The differences were attributed to a restricted motion within the interlayer space and the druglayer electrostatic interactions. The release followed a pseudo-rst order kinetics, via anion exchange, the limiting step being diffusion through the particles. Pang et al. (2013) have intercalated a hydrophobic anticancer drug derived from camptothecin, 10-hydroxycamptothecin (HCPT), in a Zn, Al LDH. As HCPT was only very scarcely soluble in water, the LDH HCPT intercalation compounds were prepared by a secondary intercalation method, and the encapsulated HCPT could keep the biologically active lactone form. First, a Zn,Al LDH in its nitrate form was intercalated with sebacate anions by a co-precipitation method, and HCPT was then intercalated by hydrophobic interaction in an ethanol medium. The alkyl chains of the sebacate anions in the interlayer provided a hydrophobic medium, suitable for intercalation of the drug. Kinetics studies of HCPT release from the nanocomposites could be tted to a pseudo-second-order kinetic model, and the diffusion of HCPT through the LDH particles played an important role in controlling the drug release. Intercalation of 5 uorouracil (5-FU), an antimetabolic drug extensively used in cancer chemotherapy, but with adverse side effects due to its toxicity, in a Mg,Al (molar ratio 2) LDH prepared by the reconstruction method from a carbonate precursor calcined at 500 C for 4 h, has been reported by Wang et al. (2005), Fig. 6. Although the molecule was neutral and only weakly acidic, it became anionic under alkaline conditions, and thus it could be intercalated in a LDH host. Three samples were prepared, at 60 C (sample Z-1), at 70 C under hydrothermal conditions (Z-2), and adding glycerol to the synthesis medium

Fig. 5. Schematic for nanobiohybrid design. Reprinted from The Journal of Controlled Release. Vol. 95. Katherine M Tyner, Scott R Schiffman, Emmanuel P Giannelis, Nanobiohybrids as delivery vehicles for camptothecin, pp. 501514. Copyright (2004), with permission from Elsevier.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

245

(sample Z-3), to assist in swelling of the layers to facilitate intercalation of the organic molecule (Ulibarri et al., 1994). Sample Z-1 showed two (003) d-spacings of 8 (rather weak) and 10.6 ; only a maximum at 8 was recorded for sample Z-2, and a shift to 12.4 was observed for sample Z-3. Similarly to the results by Liu et al. (2008b) for CPT, three different orientations were claimed to explain these three spacing values, namely, at (8 ), monolayer (10.6 ) and bilayer (12.4 ), although again, no molecular size calculation was reported to support this ascription. However, element chemical analysis and FT-IR spectra ruled out the presence of other anions in the interlayer. FT-IR spectra also showed that 5-FU was intercalated, as expected, in its anionic form. These authors also reported a thermal analysis of the samples, but did not provide quantitative analysis results nor determined the gases evolved to conrm their assumptions about the decompositions processes: dehydration at ca. 240 C and 5-FU decomposition and host dehydroxylation between 300 and 600 C. Drug release was larger and faster at pH 4 than at pH 7, but never as fast as reported for other drugs. Again a dissolution mechanism was claimed at pH 4 and an anion exchange one (with phosphate anions from the medium) at pH 7. It should be noticed, however, that at equilibrium only 87% (pH 4) and 74% (pH 7) of drug has been released. A Mg,Al LDH has been also used to intercalate 5-FU to overcome its toxicity, inmunogenecity and poor integration capacity. The 5-FULDH system was prepared (Choi et al., 2010b) by coprecipitation at pH 9.5 under nitrogen purging. It showed favourable blood clearance proles compared to unintercalated 5-FU, including sustained drug release, prolonged drug half-life, and increased accumulation in target tumour tissue; the LDH nanoparticles were also rapidly excreted from the body after administration. Intercalation of methotrexate (MTX) in Mg,Al LDHs (Mg/Al molar ratio 2) was attained by ion exchange from a nitrate LDH precursor by stirring a suspension of the LDH in an aqueous solution of MTX at 60 C for 3 days under a nitrogen atmosphere (Choy and Son, 2004). Folinic acid was intercalated similarly. The basal spacings after intercalation were 19.1 and 19.9 (8.4 for the nitrate precursor) for the folinic and MTX -containing intercalation compounds, respectively, suggesting a tilted intercalation of the molecules in the interlayer space (tilting angle 50 and 46 for folinic acid and MTX, respectively). The FT-IR spectra for the exchanged samples showed complete removal of the bands due to nitrate species, and development of bands characteristic of the drug molecules, conrming that no denaturation of the drug molecules had taken place; on the other hand, UVvis spectra showed the bands expected for these molecules, conrming that stabilisation in the interlayer took place via electrostatic interactions with the layers. Thermal analyses showed that decomposition of the organic molecules took place at somewhat higher temperatures than the isolated drugs, probably because of the interaction with the layers. The efcacy of MTXLDH intercalation compounds and MTX on two different kinds of cancer cell culture lines has been studied, to verify the drug delivery efciency of the intercalated system. The intercalation

compounds were prepared by anion exchange at 60 C for 4 days under N2 atmosphere, from a Mg,Al-NO3 (Mg/Al molar ratio 2) precursor prepared by coprecipitation. Suppression of tumour cell (Saos-2) was achieved with MTX and MTXLDH (pure LDH was inactive) in both time and concentration dependent manners, and the intercalation compound showed a higher efciency: MTX showed 75% of cell viability at 500 g/mL dosage, whereas MTXLDH at about 0.01 g/mL. Such a difference was probably due to the easier permeation of the cell membrane by MTXLDH than MTX itself. Similar effects were observed in the MG-63 cell culture line. These ndings indicated that one of the main drawbacks of MTX for practical use (toxicity) could be overcome by using the intercalation compound instead of the bulk drug, as it permitted lower doses. On the other hand, the action of MTXLDH was faster, as MTX required an induction period of 48 h to reach the maximum value, probably also because of the difference in cell membrane permeation effects. It is generally assumed that the MTX molecules entered the cytosol via reduced folate carriers (RFC) mediated transport, while MTXLDH intercalation compounds did it through endocytosis, since the neutral charge of the inorganic surface did not induce any repulsive interaction against the negatively charged cell membrane. Intracellular LDHs (internalized via clathrin-mediated endocytosis) was highly located both with typical endocytic proteins, but also with transferrin (Oh et al., 2006a). The cellular uptake of MTX was higher in MTXLDH-treated cells than in MTX-treated ones; the inhibition of the cell cycle was greater for MTXLDH than for MTX only. The observed increase in the anticancer efcacy of MTX when intercalated in a LDH has been also studied varying the drug concentration and the incubation time (Oh et al., 2006b). The MTXLDH intercalation compound showed 5000 times better drug efcacy than MTX itself in terms of drug concentration, and moreover the former exhibited 48 h faster efcacy compared to the latter. In addition the study further conrmed that the LDH inorganic matrix is biocompatible so that it does not harm the healthy cells. The MTXLDH system is also able to overcome drug resistance, showing high effectivity in wild-type HOS cells and MTX-resistant HOS/MTX cells in terms of inhibition of cancer cell proliferation, cellular uptake, and intracelular retention (Choi et al., 2010a). Methotrexate was intercalated also by coprecipitation (Kim et al., 2007), without any observed change in the functionality of the drug. Its release was more effective under simulated intracellular lysosomal conditions (pH 4.5) than simulated body uid (pH 7.4), suggesting an effective release in lysosomes. The anticancer efcacy of the intercalation compound was larger than that of free MTX. Chakraborty et al. (2012) have studied the intercalation of MTX in a Mg,Al LDH prepared by coprecipitation. These authors found a dependence of the particle size on the pH during the synthesis: while at pH 9 the size was in the 10100 nm range, isotropic growth at pH close to the isoelectric point produced particles in the 25290 nm range. Up to 76% of loaded MTX was released at pH 7.4 at 27 C, and complete release was achieved within 120 h, following the Ritger

Fig. 6. Schematic illustration of the orientation of 5-FU intercalates. Reprinted from The Journal of Solid State Chemistry. Vol. 178. Zhongliang Wang, Enbo Wang, Lei Gao, Lin Xu, Synthesis and properties of Mg 2Al layered double hydroxides containing 5-uorouracil, pp. 736 741. Copyright (2005), with permission from Elsevier.

246

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

Peppas kinetics model, demonstrating a diffusion controlled process. When different synthetic routes were tested, namely, ion exchange, direct synthesis and hydrothermal treatment after exchange, a dependence of the particle size was also observed (Chakraborty et al., 2010a, 2011a). Samples prepared by ion exchange and submitted to hydrothermal treatment showed a particle size around 350 nm for, while particles with a narrow particles size distribution around 40 nm with a lower aspect ratio were obtained by coprecipitation; however, the drug loading in this sample was markedly lower than in the others. In any case, release proceeded in two steps, a rst fast step and a second, slower one, total release of the drug being achieved after 100150 min. The drug was successfully intercalated also in a Zn, Al LDH by ion exchange (Chakraborty et al., 2011b). Intercalation of drug molecules depended on many different factors, one of them being the formal charge of the layers. If this is too large, the strong electrostatic interactions between the layers and the interlayer anion inhibited swelling and anion/drug substitution, and then alternative routes (e.g., direct coprecipitation or reconstruction from calcined precursors) should be tested. So, Yang and Guo (2003) have shown that a cis-platin-DNA model adduct, cis-[Pt(NH3)2(5-GMP)]2 (5GMP = guanosine 5-monophosphate) can be intercalated in a Zn,AlNO3 (Zn/Al molar ratio 2) LDH, but not in a Li/Al LDH, with a larger layer charge density. Choi et al. (2008) have reported a comparative study to evaluate the potential of drug-LDH intercalation compounds as cancer chemotherapy agents. They have prepared a 5-uorouracil (5-FU)LDH system by coprecipitation and after characterisation, its effectivity in inhibition of cancer cells has been studied; they studied MTXLDH, doxorubicin (Dox)LDH and the bulk drugs. The performance decreased in the order MTXLDH N MTX N Dox N 5-FULDH N 5-FU in all cell lines. MTXLDH was the most effective, even higher than Dox, one of the most effective chemotherapy agents against a variety of cancers. This indicated not only the high efcacy of drugLDH intercalation compounds, but also the great potential of MTXLDH as a cancer chemotherapy agent. Floxuridine has been also intercalated by coprecipitation (Li et al., 2009b). The drug molecules were oriented in an upward orientation forming a bilayer, as concluded from the increase in the gallery height. Controlled release at pH 4.8 and 7.2 was much slower than from a drugLDH physical mixture, but decreasing when the oxuridine concentration was increased, and following a second-order kinetics. Doxiuridine (DFUR) shows more effective and less toxic properties than other uoropyrimidines, such as 5-uorouracil and oxuridine, but due to its side effects (diarrhoea, nausea and mucositis) by oral administration its application in cancer treatment is somewhat limited. Pan et al. (2010) have reported its intercalation by reconstruction from a Mg,Al-CO3 precursor calcined at 500 C for 4 h; different Mg/Al values (ranging from 1.7 to 2.9) and pH during ageing (from 7.2 to 10) were tested, Fig. 7. The d-spacing of the (003) planes decreased from 19.1 to 16.5 as the molar Mg/Al ratio was increased, but an additional phase with d(003) equal to 8.2 was identied, probably with a different arrangement of the DFUR species in the interlayer, and which

relative concentration increased as pH ageing did; at the highest pH tested, also an important contamination by a Mg,Al-CO3 phase (identied by a FT-IR band at 1365 cm 1) was observed; probably at the high pH values used absorption of atmospheric CO2 in the reaction medium was favoured. The FT-IR spectra indicate also that DFUR was intercalated in its conjugated, monoionized form, a conclusion also reached from solid UVvis spectroscopy studies. The different conguration of the DFUR molecules in the interlayer also determined the DFUR loading, which maximum value (42% w/w) was reached in the sample prepared at pH 7.2 with Mg/Al = 1.7. When the intercalation compounds were dispersed in deionized water, absorption bands at 208 and 269 nm conrmed the existence of pyrimidine ( transition) but, in addition, a band at 223 nm for the highest loaded sample suggested a possible conjugation of DFUR anions in the interlayer region. These data, together with molecular size calculations for the DFUR molecule, suggested a bilayer arrangement in the interlayer for the high loaded sample, but a monolayer for samples was prepared at high pH and larger Mg/Al ratio. Another effect of the preparation conditions was on the particle morphology: samples prepared at pH 7.2 and low (1.7) Mg/Al molar ratio showed a spherical shape (500700 nm diameter) with a wormlike morphology. This phenomenon was quite similar to that previously reported for LDHs with inorganic anions in the interlayer growing on a MgO/Al2O3 mixture upon hydrothermal treatment (Xu and Lu, 2005), suggesting a similar mechanism for these reconstructed samples. However, samples with lower Mg/Al ratios prepared at high pH showed much smaller irregular plate-like particles and discontinuous plate plate stacking, these differences probably also affecting the release rate of the drug from the interlayer. The in vitro release proles of DFUR in a solution simulating gastrointestinal and intestinal uid at pH 7.45 without pancreatine (phosphate buffered solution) indicated that release occurred in two consecutive steps, a fast one followed by a slower one. t1/2 values were 10 and 15 min, respectively, for the pH 7.2 and 10 samples, with 96% DFUR released after 6 and 2 h, respectively. The differences may also arise from the lower crystallinity of the latter sample. The release process could be tted to the Bhaskar equation (Bhaskar et al., 1986) corresponding to diffusion within the interlayer being the rate determining step, and the modied Freundlich model, which describes the release behaviour from the at surface from the heterogeneous sites by an ion exchange diffusion process. In order to model DFUR release in the colon, the LDHDFUR compounds were also encapsulated in Eudragit L100, with an encapsulating efciency of 88.6%. Dispersed microspheres with a particle size in the 1720 m range and a smooth surface were obtained. Only 6% DFUR was released at pH 1.2 in 2 h, indicating that the encapsulated system was stable under acidic conditions. However, 60% DFUR was released in 3 h at pH 6.8, and an equilibrium release of 72% was observed at pH 7.4. Nie and Hou (2012) have recently reported the preparation of ifosfamide-surfactantLDH intercalation compounds. Surfactants, such as dodecyl sulfate (SDS) or dodecylbenzene sulfonate (SDBS) in their sodium forms, were intercalated by anion exchange, ifosfamide (IFO)

Fig. 7. Synthesis strategy for DFURLDH hybrids. Reprinted from Chemical Engineering Science. Vol. 65. Dengke Pan, Hui Zhang, Ting Zhang, Xue Duan, A novel organicinorganic microhybrids containing anticancer agent doxiuridine and layered double hydroxides: Structure and controlled release properties, pp. 37623771. Copyright (2010), with permission from Elsevier.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

247

being intercalated in a second step. Formation of the intercalation compounds after the second exchange process was denitively conrmed by different experimental techniques. The release rate was much slower than from samples prepared by simply mixing the drug and the inorganic matrix at pH 7.5. Qin et al. (2010) have intercalated the anticancer drug podophyllotoxin (PPT) and have investigated the in vitro cytotoxicity to tumour cells, the cellular uptake and in vivo antitumor inhibition of PPTLDHs. The intercalation compounds were prepared by a two-step method: rst a Mg,Al-LDHtyrosine intercalation compound was prepared by coprecipitation in a basic medium and inert atmosphere, and secondly the freshly prepared TyrLDH suspension was added to a 0.1 M podophyllotoxin solution (pH previously adjusted to 12 with 1.0 M NaOH) for a further exchange process to intercalate the drug. The particle size was in the 8090 nm range and the zeta potential measured was 20.3 mV. The in vitro cytotoxicity experiments indicated that PPTLDH nanoparticles showed better anti-tumour efcacy than PPT and were more readily taken up by HeLa cells. PPTLDH showed a long-term suppression effect on the tumour growth, and enhanced the apoptotic process of tumour cells. Studies in vivo showed a lower toxicity than PPT alone. Pharmacokinetics studies showed a prolonged circulation time and an increased bioavailability of PPT LDH than PPT. These studies led to the conclusion that LDH nanoparticles can be considered as one of the ideal carriers for anti-tumour drugs, and drugLDHs can be widely applied in future anti-tumour chemotherapy. A short review on the intercalation and delivery of anticancer drugs in and from LDHs has been recently published by Riaz and Ashaf (2013). 4. Vitamins The use of vitamins as active components of cosmetics, food, drugs, etc., is rather limited because of their easy degradation under light, high temperature, the presence of oxygen or small amounts of alkaline metal cations. In order to diminish this lack of stability, several studies to encapsulate or to immobilise them using liposomes, microemulsions (water-in-oil or oil-in-water) and liquid crystals, have been carried out (Austria et al., 1997; Gallarte et al., 1999; Semenzato et al., 1995; Spiclin et al., 2001; Yamamoto et al., 2002). The human body needs only minor amounts of vitamins for physiological functioning, but both an excess or a shortage of vitamins leads to undesired effects on the living organism; consequently, it is very important to develop matrices for protecting vitamins from degradation, while permitting also their controlled delivery. Layered double hydroxides being compatible with living organisms may full a dual role: on one side, to host the vitamins in the interlayer space, protecting them from degradation and, on the other hand, to facilitate their slow delivering by anionic exchange with the anions in the medium. Hwang et al. (2001) and Choy et al. (2004b) have intercalated retinoic acid (vitamin A), ascorbic acid (vitamin C) and -tocophenol acid succinate (vitamin E) in the interlayer space of Zn,Al hydrotalcites. A sample of the hydrotalcite in its nitrate form, with a nominal Zn/Al molar ratio of 2.0, was used as the precursor to incorporate vitamin C by anion exchange (0.001 mol LDH/0.003 mol ascorbic acid) at pH = 7 and constant stirring for 48 h at 0 C (sample named by these authors as LDHVC). On the contrary, the systems with intercalated vitamin E (LDHVE) and vitamin A (LDHVA) were prepared by direct synthesis (Zn/Al molar ratio 3, pH = 7.5 and 25 C, [vitamin] / [Al3+] = 2), and washing the precipitate thus formed with water and ethanol. The mass loadings of vitamin were 23.1 and 60.8%, respectively, for LDHVC and LDHVE. Intercalation of the vitamins in the interlayer space was conrmed by the values measured for the basal spacing by X-ray diffraction, 10.5, 37.8, and 53.8 , for vitamins A, C and E, respectively (Choy and Son, 2004; Hwang et al., 2001); monolayers of vitamins parallel to the basal planes of the LDH were intercalated in the rst case, while for

LDHVE a bilayer, tilted ca. 62 with respect to the brucite-like layers, was formed; this value has been calculated from the shift of the absorption maxima in the UVvis spectra of free and intercalated vitamin E, by applying the Kasha formula. Concerning sample LDHVA, these authors suggested that the height of the gallery, 33 , exceeds the molecular length of retinoate and the hydroxide layer has its limit in accommodating guest species; swing to its anionic charge density bulk molecules like retinoate adopted a parafne-like bilayer stacking between the hydroxide layers in order to overcome large steric restrictions, resulting in a large interlayer separation. The role played by the layer charge density on the values calculated for the basal spacing of systems with intercalated vitamins has been studied by Hwang et al. (2001) on comparing the d(003) values for LDHvitamin samples and those for the layered solids prepared with basic zinc salts containing the same vitamins, nding larger interlayer spacing values in these last cases. Thermogravimetric analysis revealed that in both samples (LDHVE and LDHVC) the total mass loss reached a value of 50% upon calcination at 600 C, although a lower value was obtained at lower temperatures for sample LDHVE because of the hydrophobic properties of the vitamin E molecule; decomposition of the anion took place around 350 C. Deintercalation studies were carried out under different conditions, depending on the specic nature of the intercalated vitamin; an aqueous carbonate solution was used for sample LDHVC, and a 50% ethanol solution for sample LDHVE (no deintercalation studies were reported for vitamin A); in both cases the release of the vitamin was slow, following an ion exchange process (type L curve), although it was hardly difcult to measure exactly the amount of vitamin released, due to its immediate oxidation upon release by oxygen dissolved in the carbonate solution. These authors pointed out the integrity and stability of the vitamin molecules in the interlayer space, while the release rate could be controlled via the nature of the solvent, the type of anion and the layer charge density. Aisawa et al. (2007) have intercalated L-ascorbic acid (VC) into MgAl, MgFe, and ZnAl LDHs (M2+/M3+ molar ratio 3) by coprecipitation and by reconstruction. The method followed determined the amount of VC intercalated and the orientation of the molecules in the interlayer space. So, when using the reconstruction method, a VC/Zn coordination compound was formed for the Zn,Al system, while up to 40% VC was incorporated into the Mg,Al and Mg,Fe matrices. However, when the coprecipitation method was used, only nitrate was intercalated (while VC remained adsorbed on the external surface of the crystallites) into the MgFe matrix, but intercalation was observed for MgAl and ZnAl (the amount of VC intercalated being larger for ZnAl than for MgAl). In these samples prepared by coprecipitation the molecules were vertically oriented in the interlayer, while they were parallel to the brucite-like layers in the samples prepared by coprecipitation, Fig. 8. No explanation was, however, provided for this different behaviour. As VC was easily oxidized in air under light or heat, stability tests against these factors have been also carried out by these authors (Aisawa et al., 2007). They concluded that the interlayer space provided a stabilizer effect and thus the molecule was mostly in the reduced state. As the matrices containing Mg2+ are stronger bases than the ZnAl one, the former facilitated the oxidation of the intercalation molecule. Deintercalation has been studied by exchange with carbonate; up to 8292% of the amount initially intercalated was released after 24%, mostly in the reduced state. Gasser (2009) has intercalated vitamin C by adsorption in the chloride forms of Zn,Fe and Mg,Fe LDHs; the experimental conditions were as follows: 0.1 g LDH, 10 mL of vitamin C aqueous solution (200 mg/L), stirring at 75 rpm and at room temperature; the reaction variables studied were pH of the starting solution, relative amount of LDH, and vitamin C concentration in the solution. The basal spacings measured by X-ray diffraction were 10.8 (Zn,FeVC) and 11.5 (Mg,FeVC), these values suggesting that the reaction proceeded by anionic exchange without damaging the layered structure. The amount of vitamin

248

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

5. Amino acids and peptides Intercalation of amino acids is important since these molecules are amphoteric, and can be protonated to cations (thus being intercalated in clays) or deprotonated to anions, so being able to be intercalated in LDHs. Many pharmaceutical agents are chiral, and some of their optical activity is readily lost by racemization under even relatively mild conditions (Patel et al., 1997). Moreover, it is very common that while one enantiomer is therapeutically active, the other has no activity or has even negative effects; the case of thalidomide is unfortunately well known. In order to maintain the efcacy of drugs, it is necessary to develop strategies and systems to inhibit the rate of racemization. L-Dopa (L dyhydroxy phenyl alanine), derived from amino acid L-tyrosine, is one of the major pharmaceuticals for treatment of the main symptoms of Parkinson's disease (Lai and Yu, 1997); however, its oxidation, decomposition and racemization occur even under rather mild conditions, and its enantiomer D-dopa has toxic properties, so its presence in the drug or its formation should be strongly avoided. L-Tyrosine (L-Tyr) is a non-essential amino acid normally synthesized in the body from phenylalanine; its deciency has been associated to depression and diet supplements are usually needed; it is also used for treatment of dementia (Growdon et al., 1982), vitiligo (Chakraborty et al., 1996), and stress (Kelly, 1999). Adverse properties of L-Tyr include racemization and oxidation to quinone (Luthra et al., 1994), and so its intercalation in LDHs has been studied, in order to reduce the rate of racemization and its oxidation (Wei et al., 2005). The intercalation of L-tyrosine (L-Tyr) (Fudala et al., 1999a,b) and L-phenylalanine (L-Phe) was reported by Fudala et al. (1999b). A Zn,AlNO3 LDH was prepared by addition of zinc nitrate solution to a slightly acidic (pH 6.06.5) solution of aluminium nitrate containing also sodium carbonate; exchange was achieved at pH 8 and at 60 C. The (003) basal spacings measured were 17.5 and 18.0 , respectively, for the L-Tyr and L-Phe samples. These authors assumed formation of a bilayer of organic guest anions in the interlayer, with the benzenic ring parallel to the brucite-like layers and the carboxylate groups directly pointing to them. These authors did not report results concerning release of the amino acids from the interlayer space. Intercalation of L-Tyr in M,Al LDHs (M = Mg, Ni, Zn) has been studied, in order to reduce the rate of racemization and its oxidation (Wei et al., 2005). The samples were prepared by coprecipitation, a solution of Al3 + and M2 + nitrates (M2 +/Al3 + molar ratio 2:1) was slowly added to a solution of NaOH and L-Tyr, with vigorous agitation under a nitrogen atmosphere and ageing at 70 C for 24 h. The basal spacings were 17.1, 16.9, and 17.2 for Ni,Al-LDH, Mg,Al-LDH, and Zn,AL-LDH, respectively, in good agreement with the values reported by Fudala et al. (1999b) and slightly different to those determined by Newman et al. (2002) by molecular dynamics, 18.0 ; the differences were probably due to the different metal molar ratios, different water content in the interlayer and the presence of traces of intercalated nitrate; in any case, the similarities suggested an arrangement corresponding to an interdigitated bilayer of the monovalent anion (since ionization of the phenolic proton is not favoured at the pH of synthesis), with strong interactions of the carboxylate groups with the layers. Such an incomplete intercalation was in agreement with the data previously reported by Whilton et al. (1997) and by Aisawa et al. (2001). FT-IR spectroscopy conrmed intercalation of the monovalent anion. The technique has been also applied to follow the thermal decomposition of the intercalation compound, showing that dehydration and dehydroxylation were complete at ca. 450 C, as well as a change in the interaction between the guest moiety and the layers around 150 C (the symmetry of the carboxylate group changes), this process being more evident at 250 C. On heating the Ni,Al-LDH sample, the d(003) spacing decreased in two steps, from 17.1 to 15.2 at 150 C (related to destruction of hydrogen bonding as a result of dehydration), and then to 13.8 at 350 C, related to dehydroxylation; the diffraction lines of NiO were evidenced at 400 C. Racemization was almost

Fig. 8. Schematic illustrations of ASA/LDHs. (a) reconstruction method, and (b) coprecipitation method. Reprinted with permission from Applied Clay Science. Vol. 35. S. Aisawa, N. Higashiyama, S. Takahashi, H. Hirahara, D. Ikematsu, H. Kondo, H. Nakayama, E. Narita, Intercalation behavior of L -ascorbic acid into layered double hydroxides, pp. 146 154. Copyright (2007), with permission from Elsevier.

C adsorbed increased with pH and the initial concentration of vitamin C; the maximum concentration tested was 500 mg/L (for which a larger loading of vitamin C was reached). The adsorption equilibrium was reached merely after 1 h, incorporating 98% of the initial amount of vitamin C when using 0.2 or 0.14 g Zn,FeCl and Mg,FeCl, respectively, and 10 mL of the vitamin C solution. The adsorption isotherms followed a Langmuir adsorption model, with Q0 values of 18.87 and 28.41 mg/g for Zn,FeCl and Mg,FeCl, respectively, following a pseudo-rst order mechanism. The rate constant slightly increased when the reaction temperature was increased from 15 to 40 C (at pH 7.5 and an initial vitamin C concentration of 200 mg/L). However, application of the Weber and Morris models did not lead to good ttings for any of the samples at the temperatures tested. Deintercalation studies were followed by UVvis spectroscopy ( = 265 nm) and carried out in 0.5 M aqueous solutions of carbonate at pH 9; 40 and 53% of the initial vitamin amount were released from Zn,FeVC and Mg,FeVC, respectively, after 180 min; the release rate was slower for sample Mg,FeVC, these authors thus concluding that this sample resulted more effective for this sort of application. The strength of the interactions developing between the drug and the inorganic matrix has been claimed as the responsible for the different release rate and amount of drug released. Khan et al. (2009) have reported an extensive study on the intercalation of key drugs, including agrochemicals, fragrances, a dye and a colour xant, as well as vitamins L1 (anthranilic acid) and B5 (pantothenic acid) in a Li,Al-LDH. The samples were prepared by ion exchange, from the chloride precursor. Molecules of vitamin L1 were oriented as an intertwined bilayer; however, due to its exibility, it was difcult to assess the orientation of the B5 molecules. Release was studied in physiologic serum; it was very fast in both cases, with t90 values of 4 and 10 min, respectively, for the B5 and L1 samples, suggesting that this matrix was not adequate to prepare controlled release systems with these vitamins, although they could be used for storage purposes, as the intercalated species could be recovered very rapidly. Chakraborty et al. (2010b) have also reported intercalation of different active molecules in different LDHs, namely, drugs, vitamins, fragances, dyes, etc.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

249

unappreciable after 50 h under sunlight, 25 h under UV radiation or thermal treatment at 180 C, the specic optical rotation decreasing less than 0.5 from the value for pristine L-Tyr (10.38), while a decrease of 45 was observed when unintercalated. The hostguest interactions in the galleries of the LDH in some sort of way inhibited racemization, in addition to the shielding effect of the layers to radiation. Wei et al. (2008) have studied the intercalation of L-dopa, in a Mg,AlLDH to assess the suitability of this intercalation compound to avoid the problems above-mentioned, and have also reported density functional theoretical calculations to prove the mechanism of racemization and the effect of LDH on its rate. First, they prepared a Mg,Al-NO3 LDH by a conventional method at 70 C and pH 10 under a nitrogen atmosphere; L-dopa was intercalated by two consecutive anion exchange steps at pH 7.7 for 48 h at 20 C. Despite the cautions taken, the experimental results found showed cointercalation of a small amount of carbonate. Incorporation of the drug led to an increase of the (003) spacing from 8.25 (for the nitrate precursor) to 13.12 , without altering the layers, as the position of the signal due to diffraction by (110) planes remained unaltered. UVvis, FT-IR and 13C NMR spectroscopies were used to check the stability of the intercalated drug. Oxidation of pristine L-dopa under alkaline conditions (i.e., those also existing in an LDH environment) formed a dark-coloured compound (Findrik et al., 2006), originated by a shift of an original absorption band at 280 nm (due to the phenolic structure) for the unaltered drug to 303 nm, and developing of an additional band at 483 nm (both due to the quinone form); the shift was not observed upon intercalation. The FT-IR spectrum showed typical bands of L-dopa, the absence of bands due to nitrate (indicating a total anion exchange), but a weak band at 1363 cm 1 due to the antisymmetric mode of intercalated carbonate, despite the cautions taken during synthesis. The 13C NMR spectrum also remained mostly unaltered, and only a downeld shift of ca. 8 ppm was observed for the C\OH carbon atoms, which may be attributed to the effect of hydrogen bonding between the hydroxyl group and the interlayer water molecules; the presence of carbonate was conrmed by a weak, sharp resonance at 170 ppm, absent in the spectrum of pure L-dopa. All these results conrmed that the integrity of the L-dopa molecule was preserved in the interlayer space. Moreover, racemization did not occur, as the specic optical rotation (measured at 589.3 nm and 18 C in 1 M HCl solution) changed only from 11.8 to 12.2 upon intercalation. The formula thereof calculated from these results and element chemical analysis was [Mg2.06Al(OH)6.12](C9H11NO4)0.60(CO3)0.201.73H2O. The gallery height, as measured from the PXRD diagram, was 8.32 , very close to the length of the L-dopa anion (8.4 ); from the pKa's values for L-dopa and bearing in mind the experimental conditions during synthesis (pH 7.7), it was concluded that the drug was mostly in its monovalent state, vertically orienting the carboxylate group of alternate anions to the upper and lower hydroxide layers. In this way, the host guest interactions corresponded to electrostatic attractions between the negative anions and the positively charged layers, and hydrogen bonding between the layers, the guest anions and the interlayer water molecules. The thermal stability of intercalated L-dopa has been also investigated by these authors (Wei et al., 2008) from the variable temperature PXRD patterns; it has been found that a steady shrinkage exists up to 200 C, nally amounting 2.3 , Fig. 9, which was ascribed to removal of interlayer water molecules, as conrmed by mass spectrometry monitoring of the evolved gases. The (003) spacing was reduced to 8.4 at 300 C, due to dehydroxylation of the layers, and decomposition (combustion) of the interlayer guest; collapsing of the layered structure was observed at 400 C, with simultaneous evolution of H2O and CO2 (MS signals at m/e 18 and 44, respectively). On comparing the decomposition pattern of LDHL-dopa and of the pristine drug, it was concluded that the thermal stability is signicantly enhanced upon intercalation, as decomposition takes place at a ca. 100 C higher temperature. Further insight on the stability of intercalated L-dopa was concluded from racemization studies under sunlight, UV light and heat treatment,

Fig. 9. Relationship between d(003) basal spacing of L-dopaLDH and temperature. Reprinted with permission from Intercalation of L-Dopa Into Layered Double Hydroxides: Enhancement of Both Chemical and Stereochemical Stabilities of a Drug Through HostGuest Interactions, Min Wei, Min Pu, Jian Guo, Jingbin Han, Feng Li, Jing He, David G. Evans, and Xue Duan, Chemistry of Materials, Vol. 20, pp. 51695180. Copyright (2008) American Chemical Society.

following the changes in the specic optical rotation (). The results indicated a dramatic difference: Concerning sunlight effect, decreased ca. 40% after 22 h for pristine L-dopa, but only 6% if intercalated; even a smaller decrease (less than 4%) was observed under UV light after 80 h exposure, while the decrease for pristine L-dopa was ca. 38%; the decrease after treatment up to 200 C was 40% for pristine L-dopa, but only 11% for the intercalated phase. 13C NMR studies revealed that the chemical structure of L-dopa remained unaltered after these treatments, and so racemization was left as the only origin for the changes observed; in other words, racemization rate is markedly inhibited upon intercalation, the LDH acting as a sort of molecular container, as dened by these authors, providing an effective way to avoid racemization. Computational studies showed that racemization of L-dopa in the solid state took place via an enol intermediate formed by hydrogen transfer from the chiral carbon atom to the nearby carboxylate group. As upon intercalation the carboxylate group became involved in strong interactions with the hydroxyl layers, it was no longer available to act as a hydrogen receptor and thus racemization was inhibited. Both at pH 7.6 and 6.4 (typical values for small intestine and ascending colon, respectively), release of the drug from the L-dopaLDH intercalation compound took place with a fast release during the rst 15 min (reaction in the external parts of the particle, anion exchange with phosphate anions from the buffer solution), and a slower release (up to 92 and 65% of the initial amount of drug, respectively, at pH 6.4 and 7.6), due to diffusion through the particle, which became the rate determining step (Wang et al., 2005; Yang et al., 2007). The dipeptide glycinetyrosine (GlyTyr) is an important source of tyrosine for animal and humans, applied to treat renal failure, and smoothing the adverse effects of stress, hypertension and dementia (Peter, 1998). However, it can undergo oxidation, racemization or decomposition even under rather mild conditions (Luthra et al., 1994), thus restricting its use and encouraging to look for appropriate vehicles for its delivery. Kong et al. (2010a) have intercalated GlyTyr in a Mg,Al LDH aiming to disclose the precise nature of the hostguest interactions; they have also prepared GlyTyrLDH lms by a solvent evaporation method to explore the potential of this alternative controlled release formulation for drugs. A reference Mg,Al-NO3 LDH was rst prepared by the hydrothermal method (Bontchev et al., 2003); the GlyTyrLDH was prepared by coprecipitation under a nitrogen atmosphere and ageing at 65 C for 24 h at pH 9.0. GlyTyrLDH lms were prepared by suspending the GlyTyrLDH in water and ultrasound-treatment under a nitrogen

250

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

atmosphere; after ltration, the suspension was dropped onto quartz substrates and dried under vacuum at 70 C for 24 h. Intercalation of GlyTyr led to layered solids with the hydrotalcitelike structure, with a (003) basal spacing of 13.3 (8.8 for the nitrate-containing sample), but the (110) reection showed no shift, indicating a swelling of the layers without signicant change in the nature of the LDH host layers. However, some high angle reections (h, k 0) were absent for the lm sample, evidencing a well c-oriented assembly of LDH platelets in the lm (Adachi-Pagano et al., 2000); the (003) reection was recorded at a somewhat higher diffraction angle, probably because a difference in the interlayer water content between the powder and the lm samples existed. The FT-IR and FT-IRATR spectra conrmed that the chemical structure of GlyTyr remained unaltered in the GlyTyrLDH powder and in the lm samples. So, characteristic FT-IR GlyTyr bands were recorded for the powder sample at 1668, 1420, and 1208 cm 1, with no band which could be ascribed to nitrate; the amide bands at 1653 and 1543 cm 1 were recorded in the FT-IRATR spectrum of the lm sample. Further conrmation of the unaltered intercalation was concluded from solid state 13C NMR: compared to pristine GlyTyr, all the resonances in the spectrum of GlyTyrLDH were shifted only ca. 4 ppm downeld, as a result of the hostguest interactions (electrostatic ones and hydrogen bonding). The specic optical rotation of GlyTyr did not signicantly change upon intercalation, indicating a lack of racemization. According to the species existing at the experimental conditions used to prepare these solids, the distribution coefcient of the monovalent anion was calculated to be 77.3%, indicating its major presence in the interlayer; therefore, these authors (Kong et al., 2010a) proposed its intercalation as a monolayer, with the carboxylate groups of individual anions alternatively pointing to the upper and lower hydroxide layers (interdigitation), giving rise to electrostatic interactions between the interlayer anions and the positively charged layers, as well as hydrogen bonding among the layers, the anions and the interlayer water molecules. In situ PXRD patterns were recorded at increasing temperatures to insight in the thermal stability of the samples. Removal of interlayer water (and the corresponding shrinkage of the layers) was observed at 185 C; a further decrease up to 305 C has been attributed to a condensation reaction of GlyTyr anions, total collapsing being observed at 465 C, with developing of diffraction signals due to MgO; unfortunately, analysis of evolved gases has not been reported. The decomposition process was also followed by FT-IR spectroscopy; together with changes due to the decomposition processes, it should be noticed that bands at 1330 and 1230 cm 1 due to condensation were recorded at 185 C (an exothermic DTA effect was recorded at 172 C) for pristine GlyTyr, but at 205 C (a gradual mass loss is observed in the TG curve) for GlyTyrLDH, indicating an improvement in the thermal stability of intercalated GlyTyr. In vitro release in PBS solution (pH 7.4) from the lm samples has been studied (Kong et al., 2010a); a fast release (up to 30%) was observed in the rst 25 min, and 91% was released after ca. 5 h, Fig. 10. The rst one can be attributed to release of GlyTyr species adsorbed on the external surface of the crystallites, while the slow second process should be due to an anion exchange process, as conrmed from the analysis of the solid after the study, which conrmed that the layered structure had not been destroyed, but the interlayer space of the (003) planes was 8.8 , in agreement with the presence of intercalated phosphate from the buffer solution (Costantino et al., 1997). Whilton et al. (1997) reported some years ago the intercalation of aspartate and glutamate (anions of aspartic and glutamic amino acids) in a Mg,Al LDH by coprecipitation, aiming to study the polymerization of the intercalated species. D,L-Aspartic and D-glutamic acids were dissolved in a NaOH solution under nitrogen, to which another solution containing Mg2+ and Al3+ nitrates was added, at pH 11.512, and the suspension was aged for 24 h at 65 C and the ltered solid dried at 70 C. The PXRD patterns showed broad (003) diffraction peaks at

Fig. 10. Release prole of Gly Tyr from Gly TyrLDH lms in PBS solution (pH 7.4). Reprinted from Chemical Engineering Journal. Vol. 157. Xianggui Kong, Shuxian Shi, Jingbin Han, Fengjie Zhu, Min Wei, Xue Duan, Preparation of Glycy- L-Tyrosine intercalated layered double hydroxide lm and its in vitro release behavior, pp. 598 604. Copyright (2010), with permission from Elsevier.

11.19 and 11.9 , corresponding to gallery heights of 6.3 and 7.1 , respectively, for aspartate and glutamate. The values were in agreement with an arrangement of the intercalated amino acids with the carboxylate anions bridging adjacent layers of the inorganic host; a minor amount of nitrate had been also intercalated. Element chemical analysis and FT-IR spectroscopy results suggested that the guest molecules existed as dianions, consistently with the high alkaline conditions used during the synthesis (pH 12). Despite the initial Mg/Al molar ratio was 2, the experimental values were 1.9 and 1.8 for the AspLDH and GluLDH samples, respectively, probably because of a partitioning between the solid and the Mg2 +and Al3 + complexes of the amino acid dianions. Aisawa et al. (2001) reported a classical study on the intercalation of amino acid phenylalanine (Phe) within the interlayer space of LDHs differing in the nature of the layer cations, namely, Mg,Al; Mn,Al; Ni,Al; Zn, Al; and Zn,Cr. The samples were prepared by coprecipitation from the corresponding nitrates at 40 C under a nitrogen atmosphere; the pH of the solution was xed according to Ksp for the monohydroxides of the corresponding metal cations. Coprecipitation of Phe was maximum for the Ni,Al sample in the pH range 610. Two kinds of conguration for Phe in the interlayer space were concluded from PXRD data for basal spacings: a vertical orientation for Mn,Al; Zn,Al, and Zn,Cr LDHs, but a horizontal one for the Mg,Al LDH. Nakayama et al. (2004) have studied the intercalation of several amino acids and peptides (oligoglycine) into a Mg,Al-LDH. Except for Lys, Arg and the different peptides tested, the other amino acids have been intercalated exclusively by the reconstruction method, from the mixture of amorphous Mg and Al oxides; intercalation by anion exchange seems to be extremely difcult, probably because the amino acids existed as zwitterions in the interlayer of LDH and were electrically neutral at pH 7. The intercalation process has been followed by PXRD and solid state NMR. The results showed that the swelling observed upon intercalation of Ala, Gly, Ser, Thr, Pro, Asn, Gln, Asp and Glu was due to formation of a bilayer structure in the interlayer space; however the other amino acids and oligoglycine formed a single layer parallel to the brycite-like layers. Electroneutrality was attained by simultaneous intercalation of carbonate and/or hydroxyl groups. The amount of amino acid (or oligoglycine) incorporated was in the 0.92.7 mmol/g range. Release studies were carried out in an aqueous solution of K2CO3 and it was a very easy process because the interaction between the intercalated zwitterion and the positive LDH layers was only by hydrogen bonding, rather weak. The authors concluded that hydrotalcites

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

251

were not suitable matrices for a controlled release of these substances, although they could be used as reservoir and adsorbents. 6. Anticardiovascular agents Khan et al. (2001) reported in a pioneering work the ability of LDHs to intercalate and readily release gembrozil, a drug used to reduce lipid levels. In this case they used a Li,Al layered double hydroxide, formally derived from Al(OH)3, but where ordered inclusion of Li+ cations in the layers that required intercalation of charge-balancing anions. The solid was used in its chloride form, and gemobrozil was incorporated by ion exchange, forming a rst stage intercalation compound (all originally empty interlayers occupied by anions). Swelling of the layers, as concluded from PXRD interlayer spacing (23.2 ), indicated formation of a bilayer arrangement of the organic molecules. Treatment with aqueous M2CO3 (M = Li or Na) led to quantitative release of the drug. When a phosphate buffer was used (pH 7 or 4) the release curve showed that complete release was achieved within ca. 30 min. Statin family drugs such as pravastatin (prava) and uvastatin (uva) are highly effective for reducing cholesterol level in bloodstream; elevated cholesterol level is a primary risk factor for coronary artery disease, a major problem in developed countries. Unfortunately, these drugs are generally unstable because of their hygroscopic nature, and stabilize undergoing a conformation change, which affects their solubility and can potentially modify their bioavailability. Instability leading to pharmaceutically unactive forms (lactones) is also a problem. Overcoming these problems is attained by dispersing them in biodegradable polymers, but this increases their price. Panda et al. (2009) have reported a study on the intercalation of these two drugs in a Mg, Al LDH (2:1 molar ratio) by direct coprecipitation, from the corresponding nitrates under a nitrogen atmosphere at pH 8.59; the suspension was aged at 65 C for 24 h and the precipitate repeatedly centrifuged and ltered. The measured spacings for the (003) basal planes were

14.9 and 15.5 , respectively, for the prava and uva LDHs. On comparing with the molecular dimensions of these drugs, it was concluded that pravastatin adopted a tilted conguration, while the conformation of intercalated uvastatin molecules was similar as that in the free state. The FT-IR spectra showed a carboxylate-related band at 1576 or 1561 cm 1, respectively, for LDHprava and LDHuva, respectively, shifted from the values recorded for the corresponding sodium salts, 1583 and 1576 cm 1. The authors did not mention the other expected band for the carboxylate group (both a symmetric and an antisymmetric related bands should be recorded). This shift has been related to a decrease of the C\O bond because of the electronegativity effect of Mg and Al cations. The C_O stretching vibration was not altered upon intercalation, conrming that the guesthost interaction took place mostly through the carboxylate groups, and that the structure of the drug was retained upon intercalation; this was further conrmed by UVvis spectroscopy studies, which showed bands for pravastatin (238 nm) and uvastatin (236 and 305 nm) in the same positions as for the pristine drugs. Thermal analysis studies showed that adsorbed and crystalline water were released below 150 C, and degradation took place at 160300 C, together with layer dehydroxylation around 340 C. Element chemical analysis showed that the drugs content were 30% for both drugs. On dispersing in water, the uvaLDH solid exhibited a hydrophobic behaviour, while the pravaLDH one was hydrophilic. AFM micrographs showed that the particle diameter was around 65 and 16 nm, respectively, for pravaLDH and uvaLDH, Fig. 11. These values corresponded to single monolayers (single sandwich) for uvaLDH, but multilayers for pravaLDH. Contact angles of water suspensions were also measured; while the value was 24 for a reference Mg,AlCO3 LDH, showing its hydrophilic nature, the value was 86 for uva LDH, indicating its hydrophobic properties, probably because of steric repulsions between adjacent bulky hydrophobic tails. It should be noticed that the PXRD pattern of uvaLDH remained unchanged (diffraction lines in the same positions) even after keeping the sample for

Fig. 11. Tapping mode AFM image of (a) 2D image of Mg/Aluva LDHs. (b) 3D image of panel a. (c) Section analysis spectra of image a, and (d) 2D image of Mg/Alprava LDHs. Reprinted with permission from In Vitro Release Kinetics and Stability of Anticardiovascular Drugs-Intercalated Layered Double Hydroxide Nanohybrids, H. S. Panda, R. Srivastava, and D. Bahadur, The Journal of Physical Chemistry B, Vol. 113, pp. 15,09015,100. Copyright (2009) American Chemical Society.

252

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

12 days at 37 C and 75 10% relative humidity; probably the hydrophobic tails enhanced water repulsion, accounting for the stability of the system (the pattern changed dramatically for sodium uvastatin salt under the same experimental conditions); the TG analysis of the samples kept under these humid conditions further conrmed no change in their structure nor stability. Similar results were observed for pravaLDH. Summarising, the observed stable behaviour was due to the electrostatic interactions with the host (as concluded from the FT-IR studies) and the inter- and intra-molecular interactions between the drug molecules. The chemical stability of the hostguest systems was studied by UV vis spectroscopy upon dispersing the solids in water at pH 7.4 (phosphate buffer) at 37 C. Despite statin drugs are unstable in acidic media and stable in alkaline media (pH 813), major decomposition was observed, the byproduct being a lactone; the stability was considerably improved upon their intercalation in the LDH host. Probably the layers provided the alkaline medium necessary to stabilize the drug. Release studies were carried out at pH 7.4 (phosphate buffer), 4.5 (HCl buffer) and simulated intestinal body uids. An early fast release (4% prava and 20% uva) was observed, followed by a relatively slower release for uvastatin at pH 7.4, probably because of its hydrophobic character, in contrast with the hydrophilic character of pravastatin; ionic exchange with phosphate anions of the buffer was observed. The fast release at pH 7.4 permitted the establishment of a therapeutic dose in a short period of time, while the subsequent sustained release allowing maintenance of this dose over a long period of time. The experimental data tted a dissolutiondiffusion kinetics model; probably heterogeneous diffusion via anionic exchange played a major role for the hydrophobic intercalation compound, while for the hydrophilic system it was interparticle diffusion. Fibrates such as bezabrate (BZF) and clobric acid (CF) are a class of lipid-regulating drugs that have been used in the therapy in many forms of hyperlipoproteinemia (Remick et al., 2008). BZF (2-[4-(2-[4chlorobenzamindo] ethyl)-phenoxy]-2-methyl-propanoic acid) is a well-known antihyperlipidaemic agent that lowers elevated blood serum lipid concentrations (cholesterol and triglycerides) (Ayaori et al., 2008). CF [2-(4-Chloro phenoxy)-2-methylpropanoic acid] is the active metabolite of the blood lipid regulator colbrate. Berber et al. (2010) have used the coprecipitation method to intercalate BFZ and CF in a Mg,Al hydrotalcite; drug loadings of 54 and 45%, respectively, were reached. The interlayer spaces calculated from the PXRD diagrams were 23.5 and 16.3 , respectively, for BZFLDH and CFLDH; from these values these authors claimed that the drug molecules were located in the interlayer as a monolayer in a staggered inter-digitated arrangement through aromatic ring interaction. Studies in solution on these systems as well as on the bula drugs have been carried out at different pHs (2, 6, and 8). Solubilisation of the drugs was very limited at pH 2, but increased as the pH did. The solution rates for the BZF LDH and CFLDH systems were smaller than for the corresponding bula drugs, and were determined by the different afnity of the LDH for the anions existing in the buffer medium. 7. Diabetes Gliclazide (GLI) is a sulfonylurea derivative widely used for treatment of type II diabetes mellitus; it shows a good tolerability and low hypoglycaemia. Its effectiveness requires a rapid gastrointestinal absorption. The slow absorption rate shown by GLI has been related to the poor permeability across the gastrointestinal membrane, and also to its low solubility in the gastric uid because of its weak acidity (pKa = 5.8) and its hydrophobic nature (Winters et al., 1994). To overcome these problems, the intercalation of GLI in a Zn,Al,Cr-LDH has been studied by Ambrogi et al. (2009). These authors chose these cations because Cr3+ and Zn2+ are directly involved in the performance of insulin: Cr3+ is an essential nutrient, but it has been also suggested that in enhances insulin binding (Cefalu and Hu, 2004), and clinical studies

have demonstrated that Zn2+ is involved in insulin synthesis, storage, secretion and signalling (Hajo and Wolfgang, 2005), while both function as antioxidants (Rostan et al., 2002). A Zn,Al,Cr-LDH in the carbonate form was rst prepared by urea hydrolysis and then it was changed to the nitrate form by treatment with 1 N HNO3 (pH 5); the solid was ltered, washed and degassed under nitrogen atmosphere. GLI was simultaneously dissolved in degassed acetone/water and converted to its anionic form by adding 0.1 M NaOH; a portion of the solid was added to this solution and stirred at room temperature for 24 h and washed under a nitrogen atmosphere. The formula determined for the solid was [Zn0.762Al0.231Cr0.007(OH)2][GLI]0.2380.66H2O (ca. 43% drug loading). The amount of chromium was calculated to be equivalent to the chromium daily supplement. Physical mixtures of GLI and the LDH in its nitrate or chloride forms were also prepared (Ambrogi et al., 2009). The PXRD pattern of the solid prepared showed a basal spacing increase from 8.9 (nitrate precursor) to 15.0 , conrming the intercalation. The structural model, Fig. 12, suggested that the GLI anions form a monolayer of partially superimposed species; with this arrangement the interactions between two benzene rings of GLI and the electrostatic interactions between the C_O (formally supporting the negative charge) and the positively charged LDH layers are enhanced. Thermogravimetric studies showed a mass loss at ca. 180 C due to removal of intercalated water and two further mass losses due to thermal degradation of the drug. The DSC trace of pristine GLI showed an endothermic effect around 170 C due to melting of the drug, which was absent for the LDHGLI sample; the patterns for the physical mixtures were simply the superimposition of the curves for the single components. The FT-IR spectra did not show absorption bands due to nitrate, indicating a complete anionic exchange; the NH ureidic group band at 3265 cm 1 was also absent, further conrming formation of the GLI anion; the shifts observed for important groups of the molecule, namely, carbonyl (from 1710 to 1665 cm 1), and sulfonyl (from 1335 and 1164 cm 1 for the antisymmetric and symmetric modes, respectively, to 1371 and 1146 cm 1) were in agreement with the presence of the GLI anion. Solubility studies were carried out at pH 1.2 (gastric juice pH) in the presence of NaCl, and at pH 3, both at 37 C. Concentration of GLI at pH 1.2 after 5 min was ve times larger than that of the crystalline powder; the layers were dissolved and the molecularly dispersed anion was solubilised much faster than crystalline pristine GLI. When the tests were carried out at pH 3, a larger concentration of GLI was measured during the rst hour, with respect to the crystalline powder, then decreasing to the same values reached on dissolving crystalline GLI. The higher apparent solubility during the rst 20 min is due, as it was observed at pH 1.2, to the liberation of the drug in a molecular form,

Fig. 12. Computer generated model of HTlcGLI (some Al ions are substituted by Cr ions in the ratio of HTlcGLI formula). Reprinted from The European Journal of Pharmaceutics and Biopharmaceutics. Vol. 73. V. Ambrogi, L. Perioli, V. Ciarnelli, M. Nocchetti, C. Rossi, Effect of gliclazide immobilization into layered double hydroxide on drug release, pp. 285291. Copyright (2009), with permission from Elsevier.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

253

whereas the following decrease in the apparent solubility is due to destruction of the LDH matrix, which increased the pH from 3 to 5.2, as experimentally measured. In vitro release was studied also at different pH values. Drug release was observed at pH 1.2, faster than from crystalline GLI or a physical mixture. A similar effect was observed at pH 3 in the presence of NaCl, followed by a slower release; in this case, the fast release in the rst minutes of the reaction was due to dissolution of the LDH layers and free release of the drug molecules; however, once an equilibrium pH was reached (pH 5) the layers were no longer dissolved, and the drug was released via anion exchange with chloride anions of the solution. These authors reported a release of 72% under these experimental conditions, but only 59% at the same pH in the absence of NaCl; the larger chloride concentration when NaCl was added favours release by an ion exchange process. When pH was further increased to 6.8 and 7.5, only the slow ion exchange (with phosphate anions of the buffer solution) was observed, and complete release was never reached. Obviously, both Cr3 + and Zn2 + were also released to the solution, very fast at pH 1.2 (only 5 or 10 min was needed for total release of Cr3 + and Zn2 +, respectively), and also in the rst 30 min at pH 3, but in this case only 3035% was released, probably because the pH increase due to layers' dissolutions avoided further destruction of the layers. 8. Antibrinolytic agents Tranexamic acid, Trx (trans-4-(aminoethyl)cyclohexane carboxylic acid), is a synthetic derivative of lysine, with properties as antibrinolytic and haemostatic, acting also against breakdown of clots, and is used in surgery and dental extractions for people suffering from haemophilia. Its intercalation in a Zn,Al LDH has been studied by Tammaro et al. (2009), as well as the dispersion of the intercalation compound thus formed in poly(-caprolactone), PCL, an aliphatic polyester which is biodegradable and biocompatible, and is used as delivery for biomedical applications (Hu and Dorset, 1990). The LDH was prepared in its carbonate form by the urea method, (Costantino et al., 1998) and titrated with HCl 0.1 M at pH 5 to convert it to the chloride form, from which the nitrate one was obtained upon equilibration for 24 h at room temperature with a NaNO3 solution. Further equilibration with a solution of the tranexamic sodium salt led to the nal product, LDHTrx, with the formula [Zn0.65Al0.35(OH)2] Trx0.24 (CO3)0.0551.5H2O, containing a small amount of intercalated carbonate anions. Powder composites (PCLLDHTrx) were prepared by milling mixtures of PCL and LDHTrx at 850 rpm for 1 h, and lms were prepared by moulding the powder at 70 C. The TG curve of sample LDHTrx showed two mass losses centred around 50 and 280 C, due to removal of intercalated water, and water derived from dehydroxylation of the layers, together with release of carbon dioxide from intercalated carbonate species. Combustion of Trx led to a third mass loss between 280 and 800 C; ZnO and the ZnAl2O4 spinel were identied as the solid residues after thermal treatment at 800 C. However, the TG curves for the nanocomposites showed a single mass loss effect which midpoint is some 100 C lower than that for PCL (402 C), probably because the base catalysed hydrolysis of the ester bonds. The basal PXRD spacing for the intercalation compound was 17.1 , decreasing to 15.1 upon equilibration with P4O10, probably because of partial removal of intercalated water molecules. This shrinkage underwent with simultaneous development of a diffraction maximum due to intercalated carbonate, from atmospheric CO2. In addition to diffractions at 2 = 21.3 and 23.7, due to crystalline PCL, the composites, quite surprisingly, showed a double diffraction maxima, corresponding to basal spacings of 17.1 and 15.1 , suggesting the presence of both, the hydrated and dehydrated forms. However, it should be also stressed that the development of a broad diffraction line centred at ca. 28.5 , which presence suggested intercalation of the polymeric chains in the galleries of the inorganic ller.

The melting point of PCL (62 C) decreased to 60 C for the intercalation compounds, and the crystallinity degree of PCL, calculated from the DSC curves, was in the 5662% range for pristine PLC and for the composites, i.e., it did not change upon formation of the intercalation compounds. As a nal prove of formation of the intercalation compound, the FTIR spectrum of LDHTrx showed a band at 1535 cm 1 due to the antisymmetric mode of the COO group, conrming Trx was intercalated in its anionic form; the band did not shift upon formation of the nanocomposites. Due to the lack of chromophores in Trx, FTIR spectroscopy was used to monitor Trx release from the PCL matrix as a function of time. An initial fast release, due to drug molecules anchored in the inorganic lamellae lying on the surface of the lm, was followed by a slow release extending up to 200 days (only 43% drug had been released), Fig. 13, due to drugs from inside of the matrix.

9. Antihypertensives Another family of drugs which have been encapsulated in anionic clays to diminish their side effects has been antihypertensive ones. Zhang et al. (2006) have chosen captopril (Cpl), an angiotensinconverting enzyme inhibitor as a model drug to be intercalated in a Mg,Al LDH by a coprecipitation method. From a chemical point of view it is a mercapto derivative of proline, with the formula C9H15NO3S. In water it is deprotonated in two steps related to release of the proton from the carboxylic group (pKa1 = 3.7) and from the thiol group (pKa2 = 9.8). A basic solution of Cpl was added to an aqueous solution of Mg2+ and Al3+ nitrates (molar Mg/Al ratio 2.06) under nitrogen at pH 10, and the slurry aged at 25 C for 48 h. The PXRD data demonstrated that the interlayer height after intercalation was almost twice the size of the drug molecule (5.26 ), and, according to FT-IR and Raman spectra, it was concluded that Cpl formed four disulde metabolites with a S\S bond, vertically oriented between the layers, with the carboxylate groups hydrogen bonded to the brucite-like layers. The thermal stability of Cpl was improved upon intercalation, if compared to that of pristine Cpl. Release studies in vitro have been carried out at pH 4.6 and 7.45. The release rate and the amount of released drug decreased as the pH was increased, probably because of a change in the release mechanism: dissolution took place at pH 4.6 at the rst stages of the studies, followed by anion exchange, decreasing the release rate. At pH 7.45 the release was very slow, and was ascribed to anionic exchange between Cpl and phosphate anions of the buffer solution; nevertheless, up to 95% of the total amount of intercalated drug could be released at both pHs.

Fig. 13. Fraction of drug released in the physiological solution from the polymeric matrix of PCLHTrx10, as a function of time (days). Reprinted from Applied Clay Science. Vol. 43. L. Tammaro, U. Costantino, M. Nocchetti, V. Vittoria, Incorporation of active nano-hybrids into poly(-caprolactone) for local controlled release: Antibrinolytic drug, pp. 350356. Copyright (2009), with permission from Elsevier.

254

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

Xia et al. (2008) have intercalated a series of antihypertensive drugs (also known as angiotensin-converting enzyme inhibitors, ACEIs), namely, enalapril, lisinopril, captopril and ramipril, in a Zn,Al-NO3 LDH (molar Zn/Al ratio 2.03). Enalapril was incorporated by coprecipitation, similarly to the method described by Zhang et al. (2006), but using zinc nitrate instead of magnesium nitrate; for the other drugs, anionic exchange at pH 5 from the nitrate form of the LDH was used. Both methods demonstrated to be effective to intercalate these drugs, as concluded from PXRD and FT-IR studies; FT-IR was very useful to assess that the layerdrug interaction was through hydrogen bonding, in addition to electrostatic interactions, between the layer cations and the intercalated anions. As for captopril studies reported by Zhang et al. (2006), deprotonation of the thiol group was also observed, and also the thermal stability was enhanced upon intercalation. However, depending on the precise nature of the drug, its intercalation took place in different orientations: ramipril and captopril anions were forming a tilted bilayer between the hydrotalcite layers, while enalapril and lisinopril formed a single vertically oriented layer; this last orientation enhanced the thermal stability in a larger extent than for the two rst drugs. Release studies at pH 4.5 and 7.45 showed a steady release of the drug anions. An important release was observed, at both pH values, during the rst minutes; afterwards a maintained slower release was observed, especially at pH 7.45; probably the faster release at pH 4.5 was related to the instability of the hydrotalcite-type structure at this rather low pH value; in other words, the fast release at this pH should be related to a (at least) partial destruction of the layered structure. Complete release of enalapril and lisinopril (forming a monolayer in the interlayer) required longer periods of time than release of captopril and ramipril, which were forming a tilted bilayer; kinetics data follow the Higuchi and rst order models, respectively. One of the drawbacks of anionic clays to be used as matrices for controlled drug release is that they are basic compounds and thus, able to be decomposed under strong acidic conditions (pH 1.2) in the stomach. To overcome this problem, several authors have proposed to encapsulate the LDHdrug intercalation compounds. Ribeiro et al. (2009) have used xyloglucan, a polysaccharide found in the primary cell walls of non-graminaceous (monocotyledons) and in the cotyledon of some dicotyledon seeds. They studied the intercalation of enalaprilate (the pharmacologically active metabolite of enalapril maleate) in a Mg,Al hydrotalcite, by anionic exchange. Xyloglucan was extracted by an exhaustive aqueous process at 25 C of pooled and milled seeds of Hymenaea courbaril. Two aqueous solutions of xyloglucan were obtained, containing respectively 0.5 and 3% of the polysaccharide, the rst one behaving as a typical liquid and the second one showing a viscoelastic behaviour. Encapsulation was attained by adding the enalaprilateLDH intercalation compound to both solutions. Release of the drug was measured at pH 5 (phosphate buffer) and 37 C. Release was very fast in the absence of xyloglucan. When using the 0.5% solution, the encapsulation did not improve the protection of the intercalation compounds, due to the liquid behaviour of xyloglucan. However, encapsulation with the 3% solution gave rise to a protection of 40% during at least 8 h, and thus this sample could be used for a slow release of the drug, especially when submitted to the action of the gastrointestinal tract, Fig. 14. These ndings are of paramount importance, especially taking into account the large amounts of drugs usually taken by hypertensive individuals. By using this controlled release system, individuals can decrease the amount of ingested drugs, reducing the stress factor and improving their quality life. Nakayama et al. (2008) have studied the controlled release of prazosin (a positively-charged, non-soluble antihypertensive drug) from LDHcyclodextrin intercalation compounds. Cyclodextrin (CD) is the only oligosaccharide with a hydrophobic cavity where organic molecules can be introduced and stabilized. On interacting an anionic clay and a cyclodextrin a balance between the positive charge of the former and the hydrophobicity of the latter is attained, so leading to a ne tuning of the release of the drug encapsulated in the cyclodextrin.

Fig. 14. Release proles of enalaprilate from simulation of the gastrointestinal tract: () LDHEnal, () LDHEnalXG(3), and () EnalXG(3). Reprinted from The International Journal of Pharmaceutics. Vol. 367. C. Ribeiro, G.G.C. Arizaga, F. Wypych, M.-R. Sierakowski, Nanocomposites coated with xyloglucan for drug delivery: In Vitro studies, pp. 204210. Copyright (2009), with permission from Elsevier.

These authors have intercalated sulfobutyl ether -cyclodextrin (SBE7--CD) into a Mg,Al LDH (Mg/Al molar ratio 2) by anionic exchange from an LDH in the nitrate form. Total nitrate/cyclodextrin exchange was reached after 24 h, the interlayer spacing increasing from 8.8 to 21.5 . This value suggested location of the cyclodextrin molecules forming a monolayer, parallel to the brucite-like layers, whichever the concentration used. The drug could be hosted well before or after intercalation of the CD. By controlling the concentration of CD in the intercalation compound, the release rate of the drug could be controlled. Prolonged release of prazosin was observed when the CD uptake was 0.7 mmol/g; probably, the interlayer was highly crowded and a steric barrier prevented a fast prazosin release. 10. Liposomes As mentioned elsewhere, one of the main advantages of supporting drugs in LDHs is that the unstability of the drug in physiological media can be avoided. Liposomes present an important drawback for drug delivery, as they are unstable in physiological media, but this can be overcome by intercalating them in LDHs, specially in Mg,Al ones, which are stable under such pH conditions (the Zn,Al analogues are stable at more acidic pH). Bgu et al. (2009) have studied the intercalation of lipophilic drug-loaded phospholipid bilayers between the layers of a Mg,Al (Mg/Al molar ratio 2) LDH, thus obtaining a new class of intercalation compounds affording protection and stabilisation of the bilayer and of the drug over a long period of time, and thus can be potentially used for sustained release. Lipid bilayers originating from liposomes were made from 1,2-dimyristoyl-sn-glycero-3-phosphate monosodium salt (DMPG), loaded with a neutral lipophilic uorescence probe (1,6diphenyl-1,3,5-hexatriene, DPH), mimicking a lipophilic drug. The sustained release after subcutaneous application was also studied. Intercalation was attempted by anionic exchange (an originally containing nitrate LDH was used) with a chloroform solution of phospholipids, anionic exchange with the liposome suspension, and reconstruction of the calcined (450 C) LDH precursor with the liposome suspension. In the last case the reconstructed layered material contained intercalated chloride anions (from the reaction medium) and so this method was discarded. X-ray diffraction lines at low angles were recorded for the samples prepared by anionic exchange, and the method using an aqueous medium was preferred, to avoid the use of organic solvents. As the amount of DMPG was increased, the diffraction lines below 5 were enhanced, and those

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

255

due to the phase with intercalated nitrate weakened. The spacing for the (003) diffraction of the DMPG intercalation compound was 46.4 , corresponding to a gallery height of 41.6 . This value was lower that the repeat distances (47.9 and 47.5 ) reported for two different conformations of DMPG bilayers, below and above their transition temperature, 23.5 C. The difference could be explained assuming a tilting of the chains in the constrained interlayer space of the LDH host, with a tilting angle between 29 and 31, but these values were lower than those reported for dodecyl sulfate intercalated in LDHs (ca. 50) (Guo et al., 2005; Zhao and Nagy, 2004). As the exchange process was carried out at 40 C, above a high-order transition at 31.7 C (Tajima et al., 2000), a conformational change, corresponding to a rotation of the terminal glycerol groups of DMPG, took place, leading to a stretched conformation forming a bilayer with a thickness of 55.5 . This would t with the experimental height of the gallery assuming a tilting angle of 41.5, very close to the values reported for analogous systems. Unfortunately, the insertion mechanism is not known, although it may correspond to a multistep process or adsorption of the liposome on the external brucite-like layers of the LDH particles. It should be noted that, as the compounds were prepared at pH 7.4, where the Mg,Al is very stable, the Mg/Al molar ratio did not change upon intercalation of DMPG, but the C/P ratio was in the range of 34.1240.00, larger than in pristine DMPG (34), suggesting the presence of cointercalated carbonate, in addition to nitrate. Unfortunately, no infrared spectroscopy data, which could help to ascertain the nature of the intercalated anions, was reported in this study. A maximum of 87% of the AEC is reached for DMPG, probably because of sterical hindrance, more related to surface charge of the layers than to conventional volumetric problems. Contrary to the usual increase in thermal stability of several biopolymers (Darder et al., 2005), drugs (Choy and Son, 2004; Del Arco et al., 2004c, 2009; Wang et al., 2005) and dyes (Tian et al., 2007) upon intercalation in the LDH galleries, intercalated DMPG decomposed at a lower temperature than pristine DMPG (325 C), and this temperature decreased as the DMPG loading increased; this could be related to different conformations of the molecule in the pristine or intercalated phase, Fig. 15. Fluorescence experiments showed a less compact arrangement of the lipids, well because of the high hydrophobic character of DPH or well because a large disorder of the alkyl chain, but, in any case, data conrmed that a single phospholipid bilayer was intercalated. Regarding release studies, the presence of DMPG was followed spectrophotometrically after leaving the solid in suspension with a physiological medium (NaCl 0.15 M, pH 7.4). A depolarization curve was observed only after 24 h, conrming the organisation of the phospholipids into bilayers. A QLS (quasi-elastic light scattering) analysis revealed the presence of spherical particles (average diameter 154 nm) after 24 h, decreasing to 90 nm after 36 h, corresponding to the spontaneous formation of liposomes, i.e., the intercalated phospholipids rearranged forming liposomes after being released, but carrying the originally encapsulated lipophilic drug. Only 1.5% of the originally existing DMPG was extracted after 7 days, and 6% after 14 days, probably by anionic exchange with phosphate anions of the buffer solution. 11. Antimycotic agents The antifungal drug 5-uorocytosine (5-FC) was originally developed as an anticancer drug, but was found to be valuable against Candida spp., Cryptococcus neoformans and Chromomycosis spp.; to decrease its potential toxicity a controlled release system should be used. The drug being a neutral weak acid, it can be hardly intercalated in LDHs. Liu et al. (2008a) have, however, prepared the anionic conjugate base in an alkaline medium, which has been intercalated in a Zn,Al LDH by coprecipitation, Fig. 16. They started from Zn2+ and Al3+ nitrates (molar ratio 2:1) under a N2 atmosphere; the drug was dissolved in aqueous ammonia and the pH adjusted using 2 M NaOH or 2 M HCl. The precipitate thus formed was aged at 40 C in the mother liquour

Fig. 15. DTG proles in the temperature range from 30 to 800 C of (a) DMPG, (b) DMPG 13.3 Mg/Al, (c) DMPG 20 Mg/Al, (d) DMPG 100 Mg/Al, and (e) LDH. Reprinted with permission from New Layered Double Hydroxides/Phospholipid Bilayer Hybrid Material with Strong Potential for Sustained Drug Delivery System, Sylvie Bgu, Anne Aubert- Poussel, Ramona Polexe, Eliska Leitmanova, Dan A. Lerner, Jean-Marie Devoisselle, and Didier Tichit, Chemistry of Materials, Vol. 21, pp. 26792687. Copyright (2009) American Chemical Society.

and then peptized at 60 C for 24 h. The actual Zn/Al molar ratio decreased as the pH during synthesis was increased, and a small amount of nitrate was also intercalated. PXRD patterns showed the presence of nitrate (gallery height 0.41 nm); intercalation of 5-FC led to splitting or shift of the (003) diffraction, a new maximum at 0.76 nm developing. A new maximum at 0.92 nm developed when the intercalated amount was increased. These changes were assumed to demonstrate intercalation of 5-FC. As the molecular dimensions of 5-FC are 0.30 0.48 nm (Wang et al., 2005), these authors concluded formation of a biphasic material for low drug loadings, one with nitrate anions and another with 5-FC anions parallel to the brucite-like layers; on increasing the amount of intercalated drug, nitrate was expelled and another biphasic system, with 5-FC anions parallel (gallery 0.28 nm, basal spacing 0.76 nm) or perpendicular (gallery 0.48 nm, basal spacing 0.96 nm) to the brucite-like layers. The FT-IR spectra of the intercalated systems showed the disappearance of the band at 1682 cm 1 due to the C_O stretching mode of the neutral 5-FC molecule upon intercalation. Quite surprisingly, the morphology of the particles (TEM) changed with the drug/LDH ratio, from hexagonal plates exclusively to hexagonal plates mixed with threadlike particles and unitary threadlike particles, as the drug/LDH ratio was increased from 0.33 to 0.50 and to 0.67, respectively; however, the mechanism of this change remains unknown. Release of the drug under controlled conditions was also studied. Both pristine 5-FC and a 5-FC mixture with the nitrate form of the LDH released 5-FC quickly (complete after 5 min) at pH 7.5 in a Na2HPO4NaH2PO4 buffer solution. As expected, release was slower from the intercalation compound and it was also pH-dependent, being lower at pH 7.5 than at pH 4.8. Almost 100% was released at pH 4.8 in 60 min. The differences found between the release behaviour at different pHs could be due to a partial dissolution of the layers under acidic conditions, but a restricted motion of the drug entities at pH above 7, and their electrostatic interactions with the layers. Release took place through ion exchange with the phosphate anions from the buffer solution, although the nature of the solid after release was not investigated. Kinetic studies revealed the importance of diffusion through the LDH particle in controlling drug release.

256

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

Fig. 16. Schematic structure representation of 5-FC/LDHs nanohybrids with different intercalated amounts of 5-FC. Reprinted from The Journal of Solid State Chemistry. Vol. 181. Chunxia Liu, Wanguo Hou, Lifang Li, Yan Li, Shaojie Liu, Synthesis and characterization of 5-uorocytosine intercalated ZnAl layered double hydroxide, pp. 17921797. Copyright (2008), with permission from Elsevier.

12. Anticoagulants According to WHO (World Health Organization), cardiovascular diseases have been the rst cause of death globally; so, many studies have dealt with the use of anticoagulants, among other drugs, for treating from a broad point of view this sort of diseases. Gu et al. (2008) have studied the intercalation and release of low molecular weight heparin (LMWH) into a Mg,Al LDH. The main interest was on overcoming the pharmaceutical limitations of heparin, namely, short half-life, low efciency of cellular delivery and lack of oral absorption, thus requiring usually two daily injections. In addition, the multinegative charge of LMWH permitted an easy and strong interaction with the brucite-like layers. The authors prepared a reference Mg, AlCl LDH by coprecipitation from the metal chloride salts. Samples with different LMWH (10 to 100% substitution of interlayer chloride) were also prepared by coprecipitation under nitrogen atmosphere. The PXRD diagrams showed a swelling of the layers from 7.7 to 14.0 for the chloride and LMWH forms, respectively, although the diffraction maxima for the LMWH samples were rather broad, indicating a lack of crystallinity. The gallery height for the 100% LMWH containing sample was 9.2 , slightly smaller than the thickness of the LMWH (11 ) molecules when they adopt a helix oriented parallel to the basal planes of the LDH; the difference could be attributed to the strong electrostatic interactions between the LMWH and the layers. The FT-IR spectra showed the bands corresponding to the anionic form of LMWH, mainly: 1420 cm 1 (symmetric C_O stretching), and 1230 and 1040 cm 1 (antisymmetric and symmetric S_O stretching, respectively), as in the LMWH sodium salt. The difference between the positions of the S_O stretching bands (190 cm 1) was very close to that in the sodium salt (210 cm 1), implying the presence of a quasi free-like anion in the interlayer space. SEM images showed formation of compact nonporous crystallites with a hexagonal plate-like shape; when the LMWH content was increased the particles became less regularly shaped, the lateral dimension decreasing from ca. 190 to ca. 90 nm, meeting the size for administration by injection. SEM images showed a change from hexagonal to ellipse-shaped particles, Fig. 17. Release in a buffer phosphate solution (pH 7.4) showed a biphasic release shape, with ca. 25% of the initial amount of drug released after 24 h and 4045% after 120 h, depending on the drug loading (the larger the loading, the lesser the amount released). This release rate was much slower than that from a physical drugLDH mixture (80% in 30 min), probably because of the strong druglayers interactions when intercalated, due to the multianionic nature of the drug (20 sulfonate and 10 carboxylate groups per molecule). This was undoubtedly an important nding (as in most of the systems here reported), as the initial fast release quickly allowed establishment of a therapeutic dose, and the subsequent sustained release allowed maintenance of this dose over a long period of time.

Analysis of the solid residue after drug release showed intercalation of phosphate species from the buffer solution; the release kinetics model could be adjusted to surface diffusion and bulk diffusion of the loaded LMWH.

13. Osteoporosis Biphosphonates containing P\C\P bonds inhibit bone resorption and are used for the treatment of osteoporosis (Lin, 1996). They disappear rapidly in plasma and so large doses are usually administrated, leading to undesirable side-effects (e.g., abdominal discomfort, nausea and diarrhoea). Nakayama et al. (2003) have reported intercalation and release studies of 1-hydroxyethylidene-1,1-diphosphonic acid (HEDP) in Mg,Al LDHs. Contrary to usual reports, these authors have used commercial hydrotalcites containing chloride (LDH-Cl) or carbonate (LDH-CO3) for ion exchange preparations, and the calcined carbonate for reconstruction studies at pH 6. As expected, the uptake of HEDP by the carbonate LDH was rather low. Concerning the chloride form, exchange at 25 C or above gave rise to partial dissolution of the layers (the Mg/Al molar ratio was markedly lower than in the original solid), due to the chelating ability of HEDP; thus, intercalation at low temperature was essential to avoid dissolution of the inorganic host and for obtaining the intercalation compound as a single phase. The uptake rate was lower through the reconstruction method, and so the exchange method starting from the LDH-Cl was identied as the most suitable one. Uptake at pH b 3 was low, because of dissolution of the layers, and above 7 the uptake also decreased, because upon deprotonation an increased chelating ability of HEDP developed, Fig. 18. Intercalation was thus carried out at pH 46, where HEDP existed mostly as the divalent anion, for which a maximum uptake of 1.8 mmol/g (ca. 50% of the LDH ionic exchange capacity, IEC) was observed; however, the experimental results indicated an uptake of ca. 3 mmol/g, a value rather close to that determined from the Langmuir plot for the exchange (3.5 mmol/g), suggesting that HEDP existed as a monovalent anion in the interlayer. The discrepancy probably came from an effect caused by the conned environment within the brucite-like layers. These authors also suggested the cointercalation of Na+ cations to explain this discrepancy, but no chemical analysis of sodium was reported. 31P MAS NMR spectra showed intercalation without decomposition, and a shift and splitting of the signals due to solid state effects were merely observed. Molecular packing modelling conrmed the orientation of the molecules within the interlayer space, as determined experimentally from the swelling of the layers upon intercalation, with the long axis of the molecule perpendicular to the layers; this was also in agreement with the space requirements (48 2/molecule for parallel arrangement and

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

257

Fig. 17. SEM images of (a) ClLDH, (b) LMWH20LDH, (c) LMWH50LDH, (d) LMWH100LDH. Reprinted with permission from In Vitro Sustained Release of LMWH from MgAl-layered Double Hydroxide Nanohybrids, Zi Gu, Anita C. Thomas, Zhi Ping Xu, Julie H. Campbell, and Gao Qing (Max) Lu, Chemistry of Materials, Vol. 20, pp. 37153722. Copyright (2008) American Chemical Society. 20, 50 and 100 in the names of the samples stand for the percentage of LMWH exchanged for chloride.

24 2/molecule for perpendicular arrangement) and electric density charge of the layers (26 2/charge). Drug release (30% of the drug loading) was complete within 2 h in an aqueous K2CO3 solution and nil in degassed water, conrming the anion exchange mechanism for exchange. However, only 1% release was observed in buffer solutions irrespective of their pH, probably because of the lower afnity of the buffer anions (monovalent chloride

and acetate) than HEDP for the interlayer space; a 6% release was observed in articial intestinal juice (according to Japanese Pharmacopoeia JP XIII 2nd uid, pH = 6.8) solutions; in other words, the results were rather negative as the release rates and the amounts released were not at satisfactory levels. Immediate release was observed only from HEDP and LDH physical mixtures. 14. Antioxidants Free radicals are able to cause damage to nucleic acids, proteins and membrane lipids, as well as to accelerate the ageing process of the cells and trigger several diseases. These problems can be diminished by using antioxidants such as carnosine and gallic acid (3,4,5-trihydroxybenzoic acid). However, both can be oxidized under mild conditions and are pHsensitive, decreasing their biofunctionality; carnosine is easily hydrolyzed in the blood and gallic acid has a short half-life in plasma. Kong et al. (2010b) have intercalated these molecules in Mg,Al LDHs. Carnosine was intercalated by ion exchange with a LDH-nitrate precursor at pH 9.5 under nitrogen atmosphere for 48 h at 70 C; gallic acid was intercalated by coprecipitation by simultaneous addition of aqueous NaOH, a mixed solution of Mg2+ and Al3+ nitrates (molar ratio ca. 2.0) to a gallic solution under nitrogen atmosphere and vigorous stirring at pH ca. 8; the suspension was protected from sunlight and aged for 24 h at 40 C. The powder X-ray diffraction patterns indicated a swelling of the LDH layers from 8.8 for the nitrate precursor to 15.1 and 10.1 for the carnosine and gallic acid derivatives, respectively. FT-IR and UVvis spectra showed the stability of the anionic forms of both species upon intercalation. The N/C ratios for both samples were larger than for both pristine drugs, indicating the simultaneous presence of a small amount of intercalated nitrate; the molar percentages of carnosine and gallic acid were 80.6 and 95.6%, respectively, remaining up to

Fig. 18. The pH dependence of () the uptake of HEDP in LDH(Cl) at 0 C, and () the solubility of LDH(Cl). One milligramme of Al corresponds to 9.36 mg of LDH(Cl). Reprinted from The Journal of Pharmaceutical Sciences. Vol. 92. Hirokazu Nakayama, Koji Takeshita, Mitsutomo Tsuhako, Preparation of 1-hydroxyethylidene-1,1-diphosphonic acidintercalated layered double hydroxide and its physicochemical properties, pp. 24192426. Copyright (2003) Wiley-Liss, Inc.

258

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

100% corresponding to nitrate. According to the dissociation constants of these polyacids and the experimental conditions during synthesis, the presence of the monovalent anions was expected in both cases, in agreement with the chemical formula calculated. From the height of the interlayer space and the molecular dimensions of the drugs, it was concluded that carnosine anions form bilayers and gallate anion monolayers, with an electrostatic interaction of the carboxylate groups with the layers. Release into buffer (phosphate) solutions did not show the usual burst release phenomenon. Carnosine was released in a twostep process at pH 7.4, releasing ca. 43% in the rst 110 min and a relatively slow second step (71% released in 750 min); a similar trend was observed for gallic acid, Fig. 19. This behaviour was rather interesting from the therapeutical point of view, as the initial fast release rapidly provides a therapeutic dose, and the subsequent sustained release maintains this dose over a long period of time. The drug release in the whole release range could not be satisfactorily tted to a single kinetics model; however, on taking both steps separately, the rst one could be adjusted to a modied Freundlich model and the second, much slower, step, to a parabolic diffusion process. The solid residues after release showed PXRD patterns typical of LDHs with intercalated phosphate anions, and so the process could be described as follows: surface diffusion was the rate determining step in the rst stage; drug anions on the external surface of the LDH particles diffused into the solution via anion exchange; and the rate determining step in the second stage would be the diffusion of the drug anions from the interlayer space to the external surface of the crystallites. The scavenging of the free radical DPPH (2,2diphenyl-1-picrylhydrazyl) followed a similar trend as that of drug release, i.e., a fast stage followed by a slow one; therefore, these LDH drug intercalation compounds could be successfully used for storing the pharmaceutical agents, maintaining their bioactivity and achieving a controlled release of the antioxidant drugs. Silion et al. (2012) have intercalated gallic acid (GA) and vanillic acid (4-hydroxy-3methoxy benzoic acid, hereafter VA) by coprecipitation and anion exchange in Zn,Al LDHs (Zn/Al molar ratio 2). The amount of antioxidant incorporated into the LDH did not depend on the preparation method followed and was in the 36.743.7 (w/w) range, a small amount of nitrate being detected by PXRD in the samples prepared by anion exchange. From the height of the gallery, these authors concluded that the VA molecules form a monolayer perpendicular to the brucitelike layers, with the carboxylate groups alternatively pointing to one brucite layer or another. However, the GA molecules are oriented

parallel to the layers, the carboxylate groups interacting electrostatically with the Al3+ cations. In vitro release at pH 2 and 7.4 indicated a larger release at pH 2 (probably because an easy dissolution of the layers), although no complete release was attained; the process followed a nonFickian diffusion model. Intercalation of vanillic acid had been previously reported by Hong et al. (2008) by direct synthesis in a Zn,Al LDH under nitrogen atmosphere, centrifugation and washing with decarbonated water. The d(003) spacing for a reference nitrate sample prepared under similar conditions was 0.85 nm, which expanded to 1.51, 1.53, and 1.52 nm, respectively, for VA-containing LDHs and Zn/Al molar ratios of 2, 3, and 4, respectively. These values were in agreement with a zigzag orientation of the VA molecules, maximizing intermolecular interactions. The FT-IR spectra (compared to those of VA and its sodium salt) conrmed development of electrostatic interactions between anionic VA and the positively charged layers. Release in water and 0.9% NaCl aqueous media showed a rapid burst until 10 min and partial dissolution of the layers in a neutral aqueous medium. After the rst, fast, step, the release was higher in a NaCl medium, indicating that it proceeded via an anion exchange process, and this effect was even more pronounced for the Zn/Al = 2/1 sample, probably because of the larger afnity of the exchangeable anions in the media (hydroxyl or chloride) for solids with a higher positive charge. The release data tted a Freundlich mechanism, but again the Zn/Al = 2/1 sample approached also a rst order diffusion process. Antioxidant 3-(3,5-di-tert-butyl-4 hydroxy-phenyl)-propionic acid (BHPPA) has been intercalated by Lonkar et al. (2013) in a Mg,Al LDH following the reconstruction method, obtaining a sample (slightly contaminated with carbonate species) with 55% (w/w) loading of the drug. PXRD diagrams conrmed insertion between the layers, with d(003) = 2.8 nm, in agreement with formation of a tilted bilayer. SEM studies conrmed that after reconstruction the hexagonal shape of the particles is preserved, although the large size of the anions modied the growing pattern of the hydroxide layers. Intercalation increased the thermal stability of the drug by ca. 100 C and the antioxidant ability of the intercalated species, as studied by UVvis spectroscopy, was similar to that shown by the pure drug. 15. Immunosuppressant corticosteroids Prednisone (PNS) is a steroidal antiinammatory drug, but it is scarcely soluble in water, and so it is poorly dispersed in physiological uids, leading to problems in efcient delivery; encapsulation in micelles, microspheres or microemulsions has been tested as alternative routes to overcome these problems (Brown et al., 2007; Karpela et al., 2007; Liu et al., 2001). F. Li et al. (2009) and Y. Li et al. (2009a) have carried out a study on the encapsulation of prednisone in micelles and their intercalation in Mg,Al LDHs as an alternative delivery route using two different encapsulating agents. The use of micelles (combining hydrophilic and hydrophobic segments) has some advantages, namely, an increase in water solubility, and protection of the drug, thus preserving its activity during transport to the precise target. Prednisone was encapsulated in cholate (F. Li et al., 2009) or PTBEM an anionic micelle derived from a biocompatible surfactant, (poly(tert-butyl acrylate-co-ethyl acrylate-co-methacrylic acid) (Y. Li et al., 2009a), and then the PNSloaded micelle was intercalated in a Mg,Al LDH by coprecipitation. Prednisone was dispersed in acetone (Y. Li et al., 2009a) or dissolved in chloroform (F. Li et al., 2009), added to a PTBEM or sodium cholate solution, respectively, and stirred under nitrogen until the organic solvent was evaporated. Once the micelles were formed (as conrmed by a uorescence probe technique), a solution containing Mg2 + and Al3 + nitrates (Mg/Al molar ratio 2) was added under nitrogen atmosphere at pH 10 (NaOH aqueous solution); the slurry was aged above room temperature for several hours and the product ltered and washed. The spacing of the (003) basal reection of the PNS/PTBEMLDH system (Y. Li et al., 2009a) was 29.33 , suggesting that incorporation of

Fig. 19. In vitro release curves of (a) carnosine from carnosineLDH and (b) GA from GA LDH. Reprinted from Applied Clay Science. Vol. 49. Xianggui Kong, Lan Jin, Min Wei, Xue Duan, Antioxidant drugs intercalated into layered double hydroxide: Structure and in Vitro release, pp. 324329. Copyright (2010), with permission from Elsevier.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

259

the micelle had been achieved. Concerning the cholate system, the spacing was 39.5 , slightly larger than that for a cholateLDH (i.e., a LDH sample with intercalated micelle, but without the PNS molecule), which was 38.4 (F. Li et al., 2009) Broadening of the other diffraction lines has been attributed to the disordered stacking of the layers. Several spectroscopies were used to check if encapsulation and intercalation had led to any change in the chemical nature of PNS. The UVvis spectrum of prednisone showed an absorption band at 244 nm due to a transition of the drug; neither sodium cholate nor PTBEM showed any absorption band in the 200400 nm range. While encapsulation in cholate, nor its intercalation in the LDH, led to no shift in the position of the PNS band, encapsulation in PTBEM led to a red shift up to 260 nm, probably because of J-aggregation of PNS in the micelle core; upon intercalation the band was recorded at 249 nm. Upon release from the PNS/PTBEMLDH nanocomposite, the band was recorded very close to that of pristine PNS, indicating that, in both cases, encapsulation and intercalation did not change the properties of the prednisone molecule. FT-IR spectroscopy was also applied for further characterisation of these systems. Quite surprisingly, assignation of some bands was different in one paper or another (F. Li et al., 2009; Y. Li et al., 2009a). So, a band at 1660 cm 1 was assigned to the (C_C) mode (F. Li et al., 2009), while in (Y. Li et al., 2009a) the band was recorded at 1612 cm 1. The (C_O) was related to bands at 1700 (F. Li et al., 2009) or 1712 and 1657 (Y. Li et al., 2009a) cm 1. The PNS bands were hardly recorded in the spectra of the intercalation compounds, probably because of the restricted movements of the drug within the micelle core, or to its low loading. Based on the basal spacings of the (003) planes and the molecular size of the intercalated species, a schematic supramolecular structure of the PNScholateLDH composite was proposed (F. Li et al., 2009), Fig. 20. In vitro drug release was studied for both systems at three different pHs, close to those existing in the small intestine (7.5 0.4), in the ascending colon (6.4 0.6), and in the distal colon (7.0 0.7) (Li et al., 2004). Drug release was dependent on the nature of the micelle and on the pH. For the PTBEM system, release was faster in the rst 10 h, thereafter slowering down and equilibrium was reached after ca. 90 h. The percentage of PNS released was 40, 65, and 100% at pH 4.8, 6.4, and 7.2, respectively (F. Li et al., 2009). Concerning the cholate system, release was fast during the rst 40 min, and then slowered, equilibrium being achieved after ca. 150 min at pH 7.6, but after 105 min at pH 6.8

or 4.8. For neither of the system, nor none of the pH tested, a burst release was observed at low pH, as in other cases described in the literature (Gu et al., 2008; Hussein et al., 2002; Khan et al., 2001; Tronto et al., 2004; Wei et al., 2008;). Although this can be attributed to the double protecting action of the micelle and of the LDH, it is more probably due to the fact that the lowest pH tested in these studies (4.8) is not low enough to dissolve the basic LDH layers, thus releasing the intercalated anion. Mathematical modelling of the drug release has shown that for the TPBEM system the mechanism could be described by a Fickian diffusion at pH 7.2, anomalous transport at pH 6.4, and combination of diffusion through the matrix and degradation of the micelles at pH 4.8, while for the cholate system the release prole tted to the RigterPeppas (RP) equation (Scott and Peppas, 1999; Serra et al., 2006). Differences in the value of the n parameter were found depending on the pH, which were related to the release mechanism: at pH 7.6 release was controlled by drug diffusion, while at pH 6.8 and 4.8 the drug release mechanism depended on the combination behaviour control, including dissolution of the composite, ion exchange and diffusion of prednisone. The authors concluded about the suitability of the two systems tested, taking advantage of the applicability in a wide pH range and almost complete release. 16. Nucleosides Choy et al. (2001) described for the rst time the intercalation of nucleoside monophosphates and DNA between the layers of a Mg,Al LDH by ionic exchange at pH 7, starting from the nitrate form (Mg/Al ratio 2.0) which had been prepared at pH 10. The biomolecules were intercalated at pH 7 by dispersing the nitrate LDH in a deaereated aqueous solution containing an excess of adenosine-5-monophosphate (AMP), guanosine-5-monophosphate (GMP), cytidine-5-monophosphate (CMP) and herring testis DNA, and stirring for 48 h. The PXRD pattern of the nitrate precursor indicated a basal spacing of 8.5 for the (003) planes, corresponding to a gallery height of 3.7 (which increased to 12.1, 13.6 and 9.7 for AMP, GMP, and CMP, respectively), corresponding to a thickness of a monolayer, probably with the phosphate groups pointing towards the hydrotalcite layers to maximize the electrostatic interactions. The interlayer space (19.1 ) for the LDHDNA intercalation compound corresponded to the thickness of the DNA molecule adopting a double helix shape parallel to the basal plane of the LDH. Circular dichroism spectra showed a weakening of the CD band at ca.

Fig. 20. Schematic representation for the drug release processes of PNS from PNS/PTBEMLDH nanocomposite in a buffer solution. Reprinted from Chemical Engineering Journal. Vol. 151. Yong Li, Hong Li, Min Wei, Jun Lu, Lan Jin, pH-Responsive composite based on prednisone-block copolymer micelle intercalated inorganic layered matrix: Structure and in vitro drug release, pp. 359366. Copyright (2009), with permission from Elsevier.

260

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

275 nm, similar to that observed for DNA in a highly concentrated salt solution, indicating an efcient neutralization (by the positively charged layers) of the DNA phosphate groups. The infrared spectra showed rather broad (because of the biomoleculelayer interactions) bands due to phosphate species, and the UVvis spectra showed the expected absorption bands due to the adenosine, guanine and cytosine groups, indicating that the molecules preserved their integrity upon intercalation. DNA could be released without any degradation by suspending the LDH DNA intercalation compound in an aqueous solution at pH below 2, indicating that the system is pH-sensitive (a so low pH simply destroys the mostly basic layers) and can be used as a gene reservoir. These authors have also studied the transfer of LDHbiomolecules systems into mammalian cells (Choy et al., 2000). Herring testis DNA, adenosine triphosphate (ATP), uoresceine-5-isothiocyanate (FITC) and c-myc antisense oligonucleotide (As-myc) were intercalated into a Mg, Al LDH by anion exchange following a method similar to that described above. These solids showed a gallery height (from PXRD diagrams) of 19.1, 14.6, 14.0, and 12.3 , respectively, in agreement with the molecular dimensions, and with the molecules (except for DNA, see above) oriented towards the layer planes to maximize the electrostatic interactions. Cellular uptake of the LDHFITC intercalation compound was concluded from laser scanning confocal microscopy experiments. Dispersions of LDHFITC were added to NIH3T3. The uorophores were detected in the cells after 1 h incubation and the uorescence intensity increased continuously up to 8 h incubation (longer times were not tested). More intense uorescences were observed when the concentration of LDHFITC was increased, but no uorescence was observed when pristine FITC was used, as cells could not take FITC even at high concentrations. FITC could be partially released by ion exchange under physiological salt conditions. HL-60 cells in a culture treated with LDH As-myc exhibited time dependent inhibition of proliferation, with 65% growth inhibition compared to the untreated cells, a further proof of antisense oligonucleotides entering into the cells and participating in cell division processes (Kwak et al., 2002). Ladewig et al. (2010) have reported the interaction of DNA plasmids (larger than DNA fragments which intercalation into LDHs had been previously reported) with LDHs, aiming to use these intercalation compounds as gene delivery vehicles. The Mg,Al LDH (molar ratio 3) was prepared by coprecipitation and hydrothermally treated at 100 C for 16 h, obtaining a transparent, homogeneous suspension containing ca. 3 mg/mL of LDH nanoparticles; other samples were prepared by changing the synthesis temperature and the time and temperature of the hydrothermal treatment. The particle size distribution showed a maximum at 100 nm with low dispersity index. TEM showed formation of regularly shaped hexagonal particles; the zeta potential of the particles was 36 mV. When lower synthesis temperatures were used smaller particles were formed, while an increase in hydrothermal time and/or temperature increased the peak particle size. Probably this was a consequence of the lower solubility at low temperature, i.e., supersaturation was readily reached and crystal nucleation occurred rapidly, forming a larger number of smaller crystals. Hydrothermal treatment deaggregated the product and tailor the size via an Ostwald ripening effect. An aliquot of the LDH suspension was mixed with a plasmid (pDNA) stock solution, the mixture incubated at 25 or 60 C with light agitation (350 rpm) and then centrifuged (15,000 rpm) for 30 min. LDH and plasmids were mixed in various mass ratios, water added and incubated at 37 C for 10 min with light agitation (350 rpm), Fig. 21. As the size of the plasmid increased, less plasmid was associated to the LDH particles. The DNA:LDH ratio never approached the ion exchange capacity of the LDH. The plasmids were wrapped around LDH individual particles, leading to their aggregation. Even so, the particles of the intercalation compounds showed considerable transfection efciency when

Fig. 21. Plasmid size vs. uptake and how variation of conditions can improve the uptake. Reprinted from Applied Clay Science. Vol. 48. Katharina Ladewig, Marcus Niebert, Zhi Ping Xu, Peter P. Gray, Gao Qing (Max) Lu, Controlled preparation of layered double hydroxide nanoparticles and their application as gene delivery vehicles, pp. 280289. Copyright (2010), with permission from Elsevier.

transfecting adherent cell lines, but not CHO-S cells, probably because of the formation of LDH:DNA aggregates which deposited on the adherent cells, but not on the CHO-S cells. DNA/LDH ultrathin lms have been also prepared via a layer-bylayer method using exfoliated LDH nanosheets (Zhao et al., 2012) aiming to use them in lm sensors. 17. Conclusions and further studies The review here reported collects the information in open literature on the interaction of more than 15 families of drugs with layered double hydroxides. The versatility of LDHs to intercalate different drugs has been shown. It can be also concluded that the preparation method followed modied the nal pharmacological behaviour of the drug LDH system, probably because of the different physicochemical properties (mainly related to particle size, way of intercalation, loading, etc.) shown by these systems. As Khan and O'Hare anticipated (Khan and O'Hare, 2002), LDHs to be used as part of a sophisticated drug or gene delivery system in patient care or therapy. In the authors' opinion, despite the drugLDH systems have been being studied for more than fteen years, their importance has not decreased, but increased on developing new systems with improved properties. Probably in the next years research will continue along these lines (some papers have actually been published on this same topic while this manuscript was under review) and we will be surely witness of the application of other materials as matrices to support anionic drugs; in this sense we can point out the use of layered hydroxide salts (LHSs), which structure is also a modication of the brucite structure (Carbajal Arizaga et al., 2007); by using LDHs or LHSs the opportunities given widens and provides a quite interesting scenario for this research. For an easy location of the systems studied, Table 1 summarises most of the papers here reviewed, indicating the family of drugs and the specic drug inserted, the chemical composition of the layers (metal cations), the preparation method followed and an indication of the availability of release data. Acknowledgements Financial support from MICINN (grant MAT2009-08526) and ERDF is greatly acknowledged.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269 Table 1 Summary of studies on insertion of drugs in layered double hydroxides. Drug or action Amino acids and peptides Amino acids: Ala, Gly, Ser, Tyr, Pro, Formula LDH matrix Preparation Release studies Yes Reference

261

Mg,Al

Coprecipitation

Nakayama et al. (2004)

Aspartate

Mg,Al

Coprecipitation

Whilton et al. (1997)

Glutamate

Mg,Al

Coprecipitation

Whilton et al. (1997)

L-Dopa

Mg,Al

Ion exchange

Yes

Wei et al. (2008), Yang et al. (2007), Wang et al. (2005)

L-Phenylalanine

Zn,Al; Mn,Al; Ni, Al; Zn,Cr

Ion exchange coprecipitation

Fudala et al. (1999b), Aisawa et al. (2001), Wei et al. (2005)

L-Tyrosine

Mg,Al; Ni,Al; Zn,Al

Coprecipitation ion exchange

Fudala et al. (1999a,b)

Oligoglycine Peptides: GlyTyr

Mg,Al Mg,Al

Coprecipitation Coprecipitation

Yes Yes

Nakayama et al. (2004) Costantino et al. (1997), Kong et al. (2010a,b)

Antibiotics Amoxicillin

Zn,Al

Reconstruction

Yes

Wang et al. (2009)

Amphotericin B

Mg,Al

Micelle method ion exchange

Yes

Trikeriotis and Ghanotakis (2007)

Ampicillin

Mg,Al

Ion exchange

Yes

Trikeriotis and Ghanotakis (2007)

Benzilpenicillin

Mg,Al

Ion exchange

Yes

Wang and Zhang (2012)

Benzoate

Mg,Al

Ion exchange

Yes

Wang and Zhang (2012)

Cefazolin

Zn,Al

Coprecipitation

Yes

Ryu et al. (2010)

(continued on next page)

262 Table 1 (continued) Drug or action Antibiotics Chloramphenicol Formula

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

LDH matrix

Preparation

Release studies Yes

Reference

Zn,Al; Zn,Mg,Al; Mg,Al

Coprecipitation ion exchange reconstrution Micelle method Ion exchange

San Romn et al. (2012), Frunza et al. (2008), Tammaro et al. (2007) Trikeriotis and Ghanotakis (2007) Trikeriotis and Ghanotakis (2007)

Gramicidin D Nalidixic acid

Protein C99H140N20O17

Mg,Al Mg,Al

Yes Yes

Noroxacin

Mg,Al,Sn

Ion exchange

Yes

Sui et al. (2012)

Phenoxymethylpenicillin

Mg,Al

Reconstruction

Yes

Li et al. (2006)

Succinate

Mg,Al

Ion exchange

Yes

Wang and Zhang (2012)

Tetracycline

Mg,Al

Reconstruction

Xu et al. (2009)

Ticarcillin

Mg,Al

Ion exchange

Yes

Wang and Zhang (2012)

Anticancer agents 10-Hydroxy Camptothecin

Zn,Al

A secondary intercalation method

Yes

Pang et al. (2013)

5-Fluorouracil

Mg,Al

Reconstruction coprecipitation Micelle method coprecipitation

yes

Wang et al. (2005), Choi et al. (2008, 2010a,b) Tyner et al. (2004), Liu et al. (2008a,b)

Camptothecin

Mg,Al

Yes

cis-Platin-DNA Doxiuridine

Zn,Al Mg,Al

Coprecipitation reconstruction Reconstruction hydrotermal treatment Yes

Yang and Guo (2003) Pan et al. (2010), Xu and Lu (2005)

Floxuridine

Mg,Al

Coprecipitation

Yes

F. Li et al. (2009), Y. Li et al. (2009a,b)

Ifosfamide

Mg,Al

A secondary intercalation method

Yes

Nie and Hou (2012)

V. Rives et al. / Applied Clay Science 8889 (2014) 239269 Table 1 (continued) Drug or action Anticancer agents Methotrexate Formula LDH matrix Preparation Release studies Yes Reference

263

Mg,Al; Zn,Al

Ion-exchange coprecipitation and hidrothermal treatment

Choy and Son (2004), Kim et al. (2007), Chakraborty et al. (2010a,b, 2011a,b, 2012)

Podophyllotoxin

Mg,Al

Ion exchange

Yes

Qin et al. (2010)

Anticardiovascular Bezabrate

Mg,Al

Coprecipitation

Yes

Berber et al. (2010)

Clobric acid

Mg,Al

Coprecipitation

Yes

Berber et al. (2010)

Fluvastatin

Mg,Al

Coprecipitation

Yes

Panda et al. (2009)

Gembrozil

Li,Al

Ion exchange

Yes

Khan et al. (2001)

Pravastatin

Mg,Al

Coprecipitation

Yes

Panda et al. (2009)

Anticoagulants Heparin

Mg,Al

Coprecipitation

Yes

Gu et al. (2008)

(continued on next page)

264 Table 1 (continued) Drug or action Antibrinolytic Tranexamic acid Formula

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

LDH matrix

Preparation

Release studies Yes

Reference

Zn,Al

Ion exchange

Tammaro et al. (2009)

Antihypertensives Captopril

Mg,Al; Zn,Al

Coprecipitation ion exchange

Yes

Zhang et al. (2006), Xia et al. (2008)

Enalapril

Zn,Al; Mg,Al

Coprecipitation ion exchange

Yes

Xia et al. (2008), Ribeiro et al. (2009)

Lisinopril

ZnAl

Ion exchange

Yes

Xia et al. (2008)

Prazosin

Mg,Al

Ion exchange

Yes

Nakayama et al. (2008)

Ramipril

Zn,Al

Ion exchange

Yes

Xia et al. (2008)

Antimycotic 5-Fluorocytosine

Zn,Al

Coprecipitation

Yes

Liu et al. (2008a,b)

Antioxidant BHPPA

Mg,Al

Reconstruction

Lonkar et al. (2013)

Carnosine

Mg,Al

Ion exchange

Yes

Kong et al. (2010a,b)

Gallic acid

Zn,Al Mg,Al

Ion exchange coprecipitation

Yes

Silion et al. (2012), Kong et al. (2010a, 2010b)

Vanillic acid

Zn,Al

Ion exchange coprecipitation

Yes

Silion et al. (2012)

V. Rives et al. / Applied Clay Science 8889 (2014) 239269 Table 1 (continued) Drug or action Diabetes Gliclazide Formula LDH matrix Preparation Release studies Yes Reference

265

Zn,Al,Cr

Ion exchange

Ambrogi et al. (2009)

Immunosuppressant corticosteroids Prednisone

Mg,Al

Coprecipitation

Yes

F. Li et al. (2009), Y. Li et al. (2009a)

Liposomes DMPG

Mg,Al

Ion exchange

Yes

Bgu et al. (2009)

Nucleosides Adenosine-5-monophosphate

Mg,Al

Ion exchange

Yes

Choy et al. (2001)

Adenosine triphosphate

Mg,Al

Ion exchange

Choy et al. (2000)

Cytedine-5-monophosphate

Mg,Al

Ion exchange

Choy et al. (2001)

c-Myc antisense oligonucleotide Fluorsceine-5-isothiocyanate

Mg,Al Mg,Al

Ion exchange Ion exchange

Choy et al. (2000) Choy et al. (2000)

Nucleosides Guanosine-5-monophosphate

Mg,Al

Ion exchange

Choy et al. (2001)

DNA

Mg,Al

Ion exchange hydrothermal coprecipitation

Choy et al. (2000, 2001), Ladewig et al. (2010) (continued on next page)

266 Table 1 (continued) Drug or action Osteoporosis 1-Hydroxyethylidene-1,1diphosphonic acid Formula

V. Rives et al. / Applied Clay Science 8889 (2014) 239269

LDH matrix

Preparation

Release studies Yes

Reference

Mg,Al

Ion exchange reconstruction

Nakayama et al. (2003)

Vitamin Vitamins A, C, E

Zn,Al; Mg,Al; Zn, Fe; Mg,Fe

Ion exchange coprecipitation reconstruction

Yes

Choy et al. (2004), Hwang et al. (2001), Gasser (2009), Aisawa et al. (2007)

Vitamin L1

Li,Al

Ion exchange

Yes

Khan et al. (2009)

Vitamin B5

Li,Al

Ion exchange

Yes

Khan et al. (2009)

References
Adachi-Pagano, M., Forano, C., Besse, J.P., 2000. Delamination of layered double hydroxides by use of surfactants. Chem. Commun. 1, 9192. Aisawa, S., Takahashi, S., Ogasawara, W., Umetsu, Y., Narita, E., 2001. Direct intercalation of amino acids into layered double hydroxides by coprecipitation. J. Solid State Chem. 162, 5262. Aisawa, S., Higashiyama, N., Takahashi, S., Hirahara, H., Ikematsu, D., Kondo, H., Nakayama, H., Narita, E., 2007. Intercalation behavior of L-ascorbic acid into layered double hydroxides. Appl. Clay Sci. 35, 146154. Albertazzi, S., Basile, F., Vaccari, A., 2004. Catalytic properties of hydrotalcite-type anionic clays. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay Surfaces: Fundamentals and Applications. Elsevier, Amsterdam, pp. 496546. Ambrogi, V., Perioli, L., Ciarnelli, V., Noccheti, M., Rossi, C., 2009. Effect of gliclazide immobilization into layered double hydroxide on drug release. Eur. J. Pharm. Biopharm. 73, 285291. Austria, R., Semenzato, A., Bettero, A., 1997. Stability of vitamin C derivatives in solution and topical formulations. J. Pharm. Biomed. Anal. 15, 795801. Ayaori, M., Momiyama, Y., Fayad, Z.A., Yonemura, A., Ohmori, R., Kihara, T., Tanaka, N., Nayaka, K., Ogura, M., Sawada, S., Taniguchi, H., Kusuhara, M., Nagata, M., Nakamura, H., Ohsuzu, F., 2008. Effect of bezabrate therapy on atherosclerotic aortic plaques detected by MRI in dyslipidemic patients with hypertriglyceridemia. Atherosclerosis 196, 425433. Basile, F., Vaccari, A., 2001. Applications of hydrotalcite-type anionic clays (layered double hydroxides) in catalysis. In: Rives, V. (Ed.), Layered Double Hydroxides: Present and Future. Nova. Sci. Pub., Inc., New York, pp. 285321. Bgu, S., Aubert-Poussel, A., Polexe, R., Leitmanova, E., Lerner, D.A., Devoisselle, J.M., Tichit, D., 2009. New layered double hydroxides/phospholipid bilayer hybrid material with strong potential for sustained drug delivery system. Chem. Mater. 21, 26792687. Benito, P., Labajos, F.M., Rives, V., 2006a. Microwave-treated layered double hydroxides containing Ni2+ and Al3+: the effect of added Zn. J. Solid State Chem. 179, 37843797. Benito, P., Labajos, F.M., Rives, V., 2006b. Uniform fast growth of hydrotalcite-like compounds. Cryst. Growth Des. 6, 19611966. Benito, P., Labajos, F.M., Rocha, J., Rives, V., 2006c. Inuence of microwave radiation on the textural properties of layered double hydroxides. Microporous Mesoporous Mater. 94, 148158. Benito, P., Guinea, I., Labajos, F.M., Rives, V., 2008a. Microwave-assisted reconstruction of NiAl hydrotalcite-like compounds. J. Solid State Chem. 181, 987996. Benito, P., Herrero, M., Barriga, C., Labajos, F.M., Rives, V., 2008b. Microwave-assisted homogeneous precipitation of hydrotalcites by urea hydrolysis. Inorg. Chem. 47, 54535463. Benito, P., Labajos, F.M., Rives, V., 2009. Microwaves and layered double hydroxides: a smooth understanding. Pure Appl. Chem. 81, 14591471. Berber, M.R., Hafez, I.H., Minagawa, K., Mori, T., Tanaka, M., 2010. Nanocomposite formulation system of lipid-regulating drugs based on layered double hydroxide: synthesis, characterization and drug release properties. Pharm. Res. 27, 23942401. Bhaskar, R., Murthy, R.S.R., Miglani, B.D., Viswanathan, K., 1986. Novel method to evaluate diffusion controlled release of drug from resinate. Int. J. Pharm. 28, 5966.

Bhattacharjee, S., Anderson, J.A., 2006. Comparison of the epoxidation of cyclohexene, dicyclopentadiene and 1,5-cyclooctadiene over LDH hosted Fe and Mn sulfonato salen complexes. J. Mol. Catal. A: Chem. 249, 103111. Bontchev, R.P., Liu, S., Krumhansl, J.L., Voigt, J., Nenoff, T.M., 2003. Synthesis, characterization, and ion exchange properties of hydrotalcite Mg6Al2(OH)16(A)x(A)2 x4H2O (A, A = Cl, Br, I, and NO 3 , 2 x 0) derivatives. Chem. Mater. 15, 36693675. Brown, E.S., Frol, A.B., Khan, D.A., Larkin, G.L., Bret, M.E., 2007. Impact of levetiracetam on mood and cognition during prednisone therapy. Eur. Psychiatry. 22, 448452. Carbajal Arizaga, G.G., Satyanarayana, K.G., Wypych, F., 2007. Layered hydroxide salts: synthesis, properties and potential applications. Solid State Ionics 178, 11431162. Carriazo, D., Domingo, C., Martn, C., Rives, V., 2006a. Structural and texture evolution with temperature of layered double hydroxides intercalated with paramolybdate anions. Inorg. Chem. 45, 12431251. Carriazo, D., Martn, C., Rives, V., 2006b. Thermal evolution of a MgAl hydrotalcite-like material intercalated with hexaniobate. Eur. J. Inorg. Chem. 22, 46084615. Cavani, F., Trir, F., Vaccari, A., 1991. Hydrotalcite-type anionic clays: preparation, properties and applications. Catal. Today 11, 173301. Cefalu, W.T., Hu, F.B., 2004. Role of chromium in human health and in diabetes. Diabetes Care 27, 27412751. Centi, G., Perathoner, S., 2008. Catalysis by layered materials: a review. Microporous Mesoporous Mater. 107, 315. Chakraborty, D.P., Roy, S., Chakraborty, A.K., 1996. Vitiligo, psoralen, and melanogenesis: some observations and understanding. Pigment Cell Res. 9, 107116. Chakraborty, M., Bose, P., Mandal, T., Data, B., Tarak, D., Shilpa, P., Chakaborty, J., Manoj, M., Basu, D., 2010a. Effect of process variatons on anticancerous drug intercalation in ceramic based delivery system. Trans. Ind. Ceram. Soc. 69, 229234. Chakraborty, M., Dasgupta, S., Sengupta, S., Chakraborty, J., Basu, D., 2010b. Layered double hydroxides based ceramic nanocapsules as reservoir and carrier of functional anions. Trans. Ind. Ceram. Soc. 69, 153163. Chakraborty, M., Dasgupta, S., Bose, P., Misra, A., Mandal, T.K., Mitra, M., Chakraborty, J., Basu, D., 2011a. Layered double hydroxide: inorganic organic conjugate nanocarrier for methotrexate. J. Phys. Chem. Solids 72, 779783. Chakraborty, M., Dasgupta, S., Soundrapandian, C., Chakraborty, J., Ghosh, S., Mitra, M.J., Basu, D., 2011b. Methotrexate intercalated Zn, Al-layered double hydroxide. J. Solid State Chem. 184, 24392445. Chakraborty, M., Dasgupta, S., Sengupta, S., Chakraborty, J., Ghosh, S., Ghosh, J., Mitra, M.K., Mishra, A., Mandal, T.K., Basu, D., 2012. A facile synthetic strategy for MgAl layered double hydroxide material as nanocarrier for methotrexate. Ceram. Int. 38, 941949. Chen, W., Qu, B.J., 2003. Structural characteristics and thermal properties of PE-g-MA/Mg, Al-LDH exfoliation nanocomposites synthetized by solution intercalation. Chem. Mater. 15, 32083213. Chibwe, K., Jones, W., 1989. Intercalation of organic and inorganic anions into layered double hydroxides. J. Chem. Soc., Chem. Commun. 926927. Choi, S.J., Oh, J.M., Choy, J.H., 2008. Anticancer drug-layered hydroxide nanohybrids as potent cancer chemotherapy agents. J. Phys. Chem. Solids 69, 15281532. Choi, S.J., Choi, G.E., Oh, J.M., Oh, Y.J., Park, M.C., Choy, J.H., 2010a. Anticancer drug encapsulated in inorganic lattice can overcome drug resistance. J. Mater. Chem. 20, 94639469.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269 Choi, S.J., Oh, J.M., Choy, J.H., 2010b. Biocompatible nanoparticles intercalated with anticancer drug for target delivery: pharmacokinetic and biodistribution study. J. Nanosci. Nanotechnol. 10, 29132916. Choy, J.H., Son, Y.H., 2004. Intercalation of vitamer into LDH and their controlled release properties. Bull. Korean Chem. Soc. 25, 122126. Choy, J.H., Kwak, S.Y., Park, J.S., Jeong, Y.J., Portier, J., 1999. Intercalative nanohybrids of nucleoside monophosphates and DNA in layered metal hydroxide. J. Am. Chem. Soc. 121, 13991400. Choy, J.H., Kwak, S.Y., Jeong, Y.J., Park, S.J., 2000. Inorganic layered double hydroxide as nonviral vectors. Angew. Chem. Int. Ed. 39, 40414045. Choy, J.H., Kwak, S.Y., Park, S.J., Jeong, Y.J., 2001. Cellular uptake behavior of [gamma-P-32] labeled ATPLDH nanohybrids. J. Mater. Chem. 11, 16711674. Choy, J.H., Jung, J.S., Oh, J.M., Park, M., Jeong, J., Kang, Y.K., Han, O.J., 2004. Layered double hydroxide as an efcient drug reservoir for folate derivatives. Biomaterials 25, 30593064. Choy, J.H., Choi, S.J., Oh, J.M., Park, T., 2007. Clay minerals and layered double hydroxides for novel biological applications. Appl. Clay Sci. 36, 122132. Constantino, V.R.L., Pinnavaia, T.J., 1995. Basic properties of Mg2+1 xAl3+x layered double hydroxides intercalated by carbonate, hydroxide, chloride and sulphate anions. Inorg. Chem. 34, 883892. Costantino, U., Nocchetti, M., 2001. Layered double hydroxides and their intercalation compounds in photochemistry and in medicinal chemistry. In: Rives, V. (Ed.), Layered Double Hydroxides: Present and Future. Nova. Sci. Pub., Inc., New York, pp. 383411. Costantino, U., Casciola, M., Massinelli, L., Nocchetti, M., Vivani, R., 1997. Intercalation and grafting of hydrogen phosphates and phosphonates into synthetic hydrotalcites and a.c.-conductivity of the compounds there by obtained. Solid State Ionics 97, 203212. Costantino, U., Marmottini, F., Nocchetti, M., Vivani, R., 1998. New synthetic routes to hydrotalcite-like compoundscharacterisation and properties of the obtained materials. Eur. J. Inorg. Chem. 10, 14391446. Costantino, U., Nocchetti, M., Sisani, M., Vivani, R., 2009. Recent progress in the synthesis and application of organically modied hydrotalcites. Z. Kristallogr. 224, 273281. Costantino, U., Nocchetti, M., Tammaro, L., Vittoria, V., 2012. Modied hydrotalcite-like compounds as active llers of biodegradable polymers for drug release and food packaging applications. Recent Pat. Nanotechnol. 6, 218230. Costantino, U., Leroux, F., Nocchetti, M., Mousty, C., 2013. LDH in physical, chemical, biochemical and life sciences, In: Bergaya, F., Lagaly, G. (Eds.), Handbook of Clay Science, 2nd edition. Part B: Techniques and Applications, Developments in Clay Science, vol. 5. Elsevier, Amsterdam, pp. 765791. Cunha, V.R., Ferreira, A.M.C., Constantino, V.R.L., Tronto, J., Valim, J.B., 2010. Hidrxidos duplos lamelares: Nanpartculas inorgnicas para armazenamento e liberaao de espcies de interesse biologico e teraputico (in Portuguese). Quim. Nova 33, 159171. Darder, M., Lopez-Blanco, M., Aranda, P., Leroux, F., Ruiz-Hitzky, E., 2005. Bionanocomposites based on layered double hydroxides. Chem. Mater. 17, 19691977. De Roy, A., Forano, C., Besee, J.P., 2001. Layered double hydroxides: synthesis and postsynthesis modication. In: Rives, V. (Ed.), Layered Double Hydroxides: Present and Future. Nova. Sci. Pub., Inc., New York, pp. 137. Del Arco, M., Rives, V., Trujillano, R., 1994. Surface and textural properties of hydrotalcitelike materials and their decomposition products. Stud. Surf. Sci. Catal. 87, 507515. Del Arco, M., Gutirrez, S., Martn, C., Rives, V., 2003. Intercalation of [Cr(C2O4)3]3 complex in Mg, Al layered double hydroxide. Inorg. Chem. 42, 42324240. Del Arco, M., Carriazo, D., Gutirrez, S., Martn, C., Rives, V., 2004a. An FT-IR study of the adsorption and reactivity of ethanol on systems derived from Mg2Al-W7O6 24 layered double hydroxides. Phys. Chem. Chem. Phys. 6, 465470. Del Arco, M., Carriazo, D., Gutirrez, S., Martn, C., Rives, V., 2004b. Synthesis and characterization of new Mg2Al-paratungstate layered double hydroxides. Inorg. Chem. 43, 375384. Del Arco, M., Cebadera, E., Gutierrez, S., Martn, C., Montero, M.J., Rives, V., Rocha, J., Sevilla, M.A., 2004c. Mg, Al layered double hydroxides with intercalated indomethacin: synthesis, characterization and pharmacological study. J. Pharm. Sci. 93, 16491658. Del Arco, M., Fernndez, A., Martn, C., Rives, V., 2009. Release studies of different NSAIDs encapsulated in Mg, Al, Fe-hydrotalcites. Appl. Clay Sci. 42, 538544. Desigaux, L., Belkacem, M.B., Richard, P., Cellier, J., Lone, P., Cario, L., Leroux, F., TaviotGuho, C., Pitard, B., 2006. Self-assembly and characterization of layered double hydroxide/DNA hybrids. Nano Lett. 6, 199204. Drits, V.A., Bookin, A.S., 2001. Crystal structure and X-ray identication of layered double hydroxides. In: Rives, V. (Ed.), Layered Double Hydroxides: Present and Future. Nova Sci. Pub. Inc., New York, pp. 3992. Duan, X., Evans, D.G. (Eds.), 2006. Layered Double Hydroxides, Structure and Bonding, 119. Springer, Berlin, pp. 1234. Evans, D.G., Slade, R.C.T., 2006. Structural aspects of layered double hydroxides. In: Duan, X., Evans, D.G., Duan, X., Evans, D.G. (Eds.), Layered Double Hydroxides, Structure and Bonding, 119. Springer, Berlin, pp. 187. Figueras, F., 2004. Base catalysis in the synthesis of ne chemicals. Top. Catal. 29, 180196. Findrik, Z., Geueke, B., Hummel, W., Vasc-Racki, D., 2006. Modelling of L-DOPA enzymatic oxidation catalyzed by L-amino acid oxidase from Crotalus adamanteus and Rhodococcus opacus. Biochem. Eng. J. 27, 275286. Forano, C., 2004. Environmental remediation involving layered double hydroxides. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay Surfaces: Fundamentals and Applications. Elsevier, Amsterdam, pp. 425458. Forano, C., Costantino, U., Prvot, V., Taviot Gueho, C., 2013. Layered double hydroxides (LDH), In: Bergaya, F., Lagaly, G. (Eds.), Handbook of Clay Science, 2nd edition. Part A: Fundamentals, Developments in Clay Science, vol. 5. Elsevier, Amsterdam, pp. 745782. Foye, W.O., Lemke, T.L., Williams, D.A., 1995. Principles of Medicinal Chemistry, 4th ed. Williams & Wilkins, Baltimore.

267

Frunza, M., Lisa, G., Popa, M.I., Miron, N.D., Nistor, D.J., 2008. Thermogravimetric analysis of layered double hydroxides with chloramphenicol and salicylate in the interlayer space. J. Therm. Anal. Calorim. 93, 373379. Fudala, A., Palinko, I., Hrivnak, B., Kiricsi, I., 1999a. Amino acid-pillared layered double hydroxide and montmorillonite thermal characteristics. J. Therm. Anal. Calorim. 56, 317322. Fudala, A., Palinko, I., Kiricsi, I., 1999b. Preparation and characterization of hybrid organic inorganic composite materials using the amphoteric property of amino acids: amino acid intercalated layered double hydroxide and montmorillonite. Inorg. Chem. 38, 46534658. Gallarte, M., Carlotti, M.E., Trotta, M., Bovo, S., 1999. On the stability of ascorbic acid in emulsied systems for topical and cosmetic use. Int. J. Pharm. 188, 233241. Gasser, M.S., 2009. Inorganic layered double hydroxides as ascorbic acid (vitamin C) delivery systemintercalation and their controlled release properties. Colloids Surf., B 73, 103109. Goh, K.H., Lim, T.T., Dong, Z., 2008. Application of layered double hydroxides for removal od oxyanions: a review. Water Res. 42, 13431368. Growdon, J.H., Melamed, E., Logue, M., Hefti, F., Wurtman, R.J., 1982. Effects of oral L -tyrosine administration on CSF tyrosine and homovanillic acid levels in patients with Parkinson's disease. Life Sci. 30, 827832. Gu, Z., Thomas, A.C., Xu, Z.P., Campbell, J.H., Lu, G.Q., 2008. In vitro sustained release of LMWH from MgAl-layered double hydroxide nanohybrids. Chem. Mater. 20, 37153722. Guo, Y., Zhang, H., Zhao, L., Li, G.D., Chen, J.S., Xu, L., 2005. Synthesis and characterization of CdCr and ZnCdCr layered double hydroxides intercalated with dodecyl sulphate. J. Solid State Chem. 178, 18301836. Hajo, H., Wolfgang, M., 2005. Protein tyrosine phosphatases as targets of the combined insulinomimetic effects of zinc and oxidants. BioMetals 18, 333338. He, J., Wei, M., Li, B., Kang, Y., Evans, D.G., Duan, X., 2006. Preparation of double layered hydroxides. In: Duan, X., Evans, D.G. (Eds.), Layered Double Hydroxides. Springer, Berlin, pp. 89119. Herrero, M., Benito, P., Labajos, F.M., Rives, V., 2007a. Nanosize cobalt oxide-containing catalysts obtained through microwave assisted methods. Catal. Today 128, 129137. Herrero, M., Benito, P., Labajos, F.M., Rives, V., 2007b. Stabilization of Co2+ in layered double hydroxides (LDHs) by microwave-assisted ageing. J. Solid State Chem. 180, 873884. Herrero, M., Labajos, F.M., Rives, V., 2009. Size control and optimisation of intercalated layered double hydroxides. Appl. Clay Sci. 42, 510518. Hong, M.M., Oh, J.M., Choy, J.H., 2008. Encapsulation of avor molecules, 4-hydroxy-3methoxy benzoic acid, into layered inorganic nanoparticles for controlled release of avor. J. Nanosci. Nanotechnol. 8, 50185021. Hu, H., Dorset, D.L., 1990. Crystal structure of poly(-caprolactone). Macromolecules 23, 46044607. Hu, C., Li, D., 2004. Polyoxometalate complexes of layered double hydroxides. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay Surfaces: Fundamentals and Applications. Elsevier, Amsterdam, pp. 374402. Hussein, M.Z.B., Zainal, Z., Yahaya, A.H., Foo, D.W., 2002. Controlled release of a plant growth regulator, -naphthaleneacetate from the lamella of ZnAl-layered double hydroxide nanocomposite. J. Control. Release 82, 417427. Hwang, S.H., Han, Y.S., Choy, J.H., 2001. Intercalation of functional organic molecules with pharmaceutical, cosmeceutical and nutraceutical functions into layered double hydroxides and zinc basic salts. Bull. Korean Chem. Soc. 22, 10192021. Jakubikov, B., Kovanda, F., 2010. Utilization of layered double hydroxides in medical applications. Chem. List. 104, 906912. Jaubertie, C., Holgado, M.J., San Romn, M.S., Rives, V., 2006. Structural characterisation and delamination of lactate-intercalated Zn, Al-layered double hydroxides. Chem. Mater. 18, 31143121. Jinesh, C.M., Rives, V., Carriazo, D., Antonyraj, C.A., Kannan, S., 2010. In uence of copper on the isomerization of eugenol for as-synthetized NiCuAl ternary hydrotalcites: an understanding through physicochemical study. Catal. Lett. 134, 337342. Kaluskova, R., Novota, M., Vymazal, Z., 2004. Investigation of thermal stabilization of poly(vinyl chloride) by lead stearate and its combination with synthetic hydrotalcite. Polym. Degrad. Stab. 85, 903909. Kanezaki, E., 2004. Preparation of layered double hydroxides. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay Surfaces: Fundamentals and Applications. Elsevier, Amsterdam, pp. 345373. Karpela, J.P., Nayakb, A., Lumryc, W., Craigd, T.J., Kerwine, E., Fishf, J.E., Lutskyg, B., 2007. Inhaled mometasone furoate reduces oral prednisone usage and improves lung function in severe persistent asthma. Respir. Med. 101, 628637. Kelly, G.S., 1999. Nutritional and botanical interventions to assist with the adaptation to stress. Altern. Med. Rev. 4940, 249265. Khan, A.I., Lei, L., Norquist, A.J., O'Hare, D., 2001. Intercalation a controlled release of pharmaceutically active compounds from a layered double hydroxide. Chem. Commun. 23422343. Khan, A.I., O'Hare, D., 2002. Intercalation chemistry of layered double hydroxides: recent development and applications. J. Mater. Chem. 12, 31913198. Khan, A.I., Ragavan, A., Fong, B., Markland, C.H., O'Brien, M., Dunbar, T.G., Williams, T.G., O'Hare, D., 2009. Recent developments in the use of layered double hydroxides as host materials for the storage and triggered release of functional. Ind. Eng. Chem. Res. 48, 1019610205. Kim, J.Y., Choi, S.J., Oh, J.M., Park, T., Choy, J.H., 2007. Anticancer drug-inorganic nanohybrid and its cellular interaction. J. Nanosci. Nanotechnol. 7, 37003705. Kong, X., Shi, Sh., Han, J., Zhu, F., Wei, M., Duan, X., 2010a. Preparation of glycy-l-tyrosine intercalated layered double hydroxide lm and its in vitro release behavior. Chem. Eng. J. 157, 598604. Kong, X., Jin, L., Wei, M., Duan, X., 2010b. Antioxidant drugs intercalated into layered double hydroxide: structure and in vitro release. Appl. Clay Sci. 49, 324329.

268

V. Rives et al. / Applied Clay Science 8889 (2014) 239269 Pereira, C.M.C., Herrero, M., Labajos, F.M., Marques, A.T., Rives, V., 2009. Preparation and properties of new ame retardant unsaturated polyester nanocomposites based on layered double hydroxides. Polym. Degrad. Stab. 93, 939946. Perioli, L., Pagano, C., 2012. Inorganic matrices: an answer to low drug solubility problem. Expert Opin. Drug Deliv. 9, 15591572. Perioli, L., Mutascio, P., Pagano, C., 2013. The inuence of the nanocomposite MgAlHTlc on gastric absorption of drugs: in vitro and ex vitro studies. Pharm. Res. 30, 156166. Peter, F., 1998. Old and new substrates in clinical nutrition. J. Nutr. 128, 789796. Qin, L., Xue, M., Wang, W., Zhu, R., Wang, S., Sun, J., Zhang, R., Sun, X., 2010. The in vitro and in vivo anti-tumor effect of layered double hydroxides nanoparticles as delivery for podophyllotoxin. Int. J. Pharm. 388, 223230. Remick, J., Weintraub, H., Setton, R., Offenbacher, J., Fisher, E., Schwartzbard, A., 2008. Fibrate therapy: an update. Cardiol. Rev. 16, 129141. Riaz, U., Ashaf, S.M., 2013. Double layered hydroxides as potential anti-cancer drug delivery agents. Mini. Rev. Med. Chem. 13, 522529. Ribeiro, C., Arizaga, G.G.C., Wypych, F., Sierakowski, M.R., 2009. Nanocomposites coated with xyloglucan for drug delivery: in vitro studies. Int. J. Pharm. 367, 204210. Rives, V. (Ed.), 2001. Layered Double Hydroxides: Present and Future. Nova Sci. Pub., Inc., New York. Rives, V., Ulibarri, M.A., 1999. Layered double hydroxides (LDH) intercalated with metal coordination compounds and oxometalates. Coord. Chem. Rev. 181, 61120. Rives, V., Prieto, O., Dubey, A., Kannan, S., 2003. Synergistic effect in the hydroxylation of phenol over CoNiAl ternary hydrotalcites. J. Catal. 220, 161171. Rives, V., Carriazo, D., Martn, C., 2010. Heterogeneous Catalysis by PolyoxometalateIntercalated Layered Double Hydroxides. In: Gil, A., Korili, S.A., Vicente, M.A., Trujillano, R. (Eds.), Springer, New York, pp. 319397. Rives, V., del Arco, M., Martn, C., 2013. Layered double hydroxides as drug carriers and for controlled release of non-steroidal antiinammatory drugs (NSAID): a review. J. Control. Release 169, 2839. Rocha, J., del Arco, M., Rives, V., Ulibarri, M.A., 1999. Reconstruction of layered double hydroxides from calcined precursors: a powder XRD and 27Al MAS NMR study. J. Mater. Chem. 9, 24992503. Rostan, E.F., DeBuys, H.V., Madey, D.L., Pinnell, S.R., 2002. Evidence supporting zinc as an important antioxidant for skin. Int. J. Dermatol. 41, 606611. Russo, G., Lamberti, G., 2011. Electrospinning of drug-loaded polymer systems: preparation and drug release. J. Appl. Polym. Sci. 122, 35513556. Ryu, S.J., Jung, H., Oh, J.M., Lee, J.K., Choy, J.H., 2010. Layered double hydroxide as novel antibacterial drug delivery system. J. Phys. Chem. Solids 71, 685688. San Romn, M.S., Holgado, M.J., Jaubertie, C., Rives, V., 2006. Synthesis, characterisation and delamination behaviour of lactate-intercalated Mg,Al hydrotalcite-like compounds. Solid State Sci. 10, 13331341. San Romn, M.S., Holgado, M.J., Salinas, B., Rives, V., 2012. Characterisation of diclofenac, ketoprofen or chloramphenicol succinate encapsulated in layered double hydroxides with the hydrotalcite-type structure. Appl. Clay Sci. 55, 158163. Scott, R.A., Peppas, N.A., 1999. Kinetics of copolymerization of PEG containing multiacrylates with acrylic acid. Macromolecules 32, 61496158. Seis, B.F., de Vos, D.E., Jacobs, P.A., 2001. Hydrotalcite-like anionic clays in catalytic organic reactions. Catal. Rev. Sci. Technol. 43, 443488. Semenzato, A., Austria, R., Dall'Aglio, C., Bettero, A., 1995. High-performance liquid chromatographic determination of ionic compounds in cosmetic emulsions: application to magnesium ascorbyl phosphate. J. Chromatogr. A 705, 385389. Serra, L., Domenech, J., Peppas, N.A., 2006. Drug transport mechanisms and release kinetics from molecularly designed poly (acrylic acid-g-ethylene glycol) hydrogels. Biomaterials 27, 54405451. Silion, M., Hritcu, D., Lisa, G., Popa, M.I., 2012. New hybrid materials based on layered double hydroxides and antioxidant compounds. Preparation, characterization and release kinetic studies. J. Porous. Mater. 19, 267276. Sorrentino, A., Gorrasi, M., Tortora, M., Vittoria, V., Costantino, U., Marmottini, F., Padella, F., 2005. Incorporation of MgAl hydrotalcite into a biodegradable poly(-caprolactone) by high energy ball milling. Polymer 46, 16011608. Spiclin, P., Gasperlin, M., Kmetec, V., 2001. Stability of ascorbyl palmitte in topical microemulsions. Int. J. Pharm. 222, 271279. Sui, M.H., Zhou, Y.F., Sheng, L., Duan, B.B., 2012. Adsorption of noroxacin in aqueous solution by MgAl layered double hydroxides with variable metal composition and interlayer anions. Chem. Eng. J. 210, 451460. Tajima, K., Imai, Y., Nakamura, A., Koshinuma, M., 2000. Thermal transitions in the bilayer assembly of dimyristoylphosphatidylglycerol sodium salt dispersion in water: a new phase transition through an intermediate state. Adv. Colloid Interface Sci. 88, 7997. Tammaro, L., Costantino, U., Bolognese, A., Sammartino, G., Marenzi, G., Calignanoe, A., Tet, S., Mastrangelo, F., Califano, L., Vittoria, V., 2007. Nanohybrids for controlled antibiotic release in topical applications. Int. J. Antimicrob. Agents 29, 417423. Tammaro, L., Costantino, U., Nocchetti, M., Vittoria, V., 2009. Incorporation of active nanohybrids into poly(-caprolactone) for local controlled release: antibrinolytic drug. 43, 350356. Tian, Y., Wang, G., Li, F., Evans, D.G., 2007. Synthesis and thermo-optical stability of omethyl red-intercalated NiFe layered double hydroxide material. Mater. Lett. 61, 16621666. Tichit, D., Coq, B., 2003. Catalysis by hydrotalcites and related materials. CATTECH 7, 206217. Trikeriotis, M., Ghanotakis, D.F., 2007. Intercalation of hydrophilic and hydrophobic antibiotics in layered double hydroxides. Int. J. Pharm. 332, 176184. Tronto, J., Dos Reis, M.J., Silvrio, F., Balbo, V.R., Marchetti, J.M., Valim, J.B., 2004. In vitro release of citrate anions intercalated in magnesium aluminium layered double hydroxides. J. Phys. Chem. Solids 65, 475480. Tyner, K.M., Schiffmanb, S.C., Giannelis, E.P., 2004. Nanobiohybrids as delivery vehicles for camptothecin. J. Control. Release 95, 501514.

Kooli, F., Jones, W., Rives, V., Ulibarri, M.A., 1997. An alternative route to polyoxometalateexchanged layered double hydroxides: the use of ultrasounds. J. Mater. Sci. Lett. 16, 2729. Kwak, S.Y., Jeong, Y.J., Park, J.S., Choy, J.H., 2002. Bio-LDH nanohybrid for gene therapy. Solid State Ionics 151, 229234. Kwon, T., Pinnavaia, T.J., 1989. Pillaring of a layered double hydroxide by polyoxometalates with Keggin-ion structures. Chem. Mater. 1, 381383. Ladewig, K., Xu, Z.P., Lu, G.Q., 2009. Layered double hydroxide nanoparticles in gene and drug delivery. Expert Opin. Drug Deliv. 6, 907922. Ladewig, K., Niebert, M., Xu, Z.P., Gray, P.P., Lu, G.Q., 2010. Controlled preparation of layered double hydroxide nanoparticles and their application as gene delivery vehicles. Appl. Clay Sci. 48, 280289. Lai, C.T., Yu, P.H., 1997. Dopamine- and L-beta-3,4-dihydroxyphenylalanine hydrochloride (L-Dopa)-induced cytotoxicity towards catecholaminergic neuroblastoma SH-SY5Y cells. Effects of oxidative stress and antioxidative factors. Biochem. Pharmacol. 53, 363372. Leroux, F., Besse, J.P., 2004. Layered double hydroxide/polymer nanocomposites. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay Surfaces: Fundamentals and Applications. Elsevier, Amsterdam, pp. 459495. Li, B.X., He, J., Evans, D.G., Duan, X., 2004. Enteric-coated layered double hydroxides as a controlled release drug delivery system. Int. J. Pharm. 287, 8995. Li, W.Z., Lu, J., Chen, J.S., Li, G.D., Jiang, Y.S., Li, L.S., Huang, B.Q., 2006. Phenoxymethylpenicillinintercalated hydrotalcite as a bacteria inhibitor. J. Chem. Technol. Biotechnol. 81, 8993. Li, F., Jin, L., Han, J., Wei, M., Li, C., 2009. Synthesis and controlled release properties of prednisone intercalated MgAl layered double hydroxide composite. Ind. Eng. Chem. Res. 48, 55905597. Li, Y., Li, H., Wei, M., Lu, J., Jin, L., 2009a. pH-Responsive composite based on prednisoneblock copolymer micelle intercalated inorganic layered matrix: structure and in vitro drug release. Chem. Eng. J. 151, 359366. Li, Y., Xu, J., Zhang, S.J., Li, D.X., Zheng, B., Hou, W.G., 2009b. Synthesis and characterization of oxuridine-layered double hydroxide nanohybrid. Chin. J. Inorg. Chem. 25, 21242128. Lin, J.H., 1996. Biphosphonates: a review of their pharmacokinetics properties. Bone 18, 7585. Liu, M.C., Proud, D., Lichtenstein, L.M., Hubbard, W.C., Bochner, B.S., Stealey, B.A., Breslin, L., Xiao, H.Q., Freidhoff, L.R., Schroeder, J.T., Schleimer, R.P., 2001. Effects of prednisone on the cellular responses and release of cytokines and mediators after segmental allergen challenge of asthmatic subjects. J. Allergy Clin. Immunol. 108, 2938. Liu, C., Hou, W., Li, L., Li, Y., Liu, S., 2008a. Synthesis and characterization of 5uorocytosine intercalated ZnAl layered double hydroxide. J. Solid State Chem. 181, 17921797. Liu, Ch.X., Hou, W.G., Li, Y., Li, L.F., 2008b. Synthesis and characterization of camptothecin intercalated into Mg/Al layered double hydroxide. Chin. J. Chem. 26, 18061810. Lonkar, S.P., Kutlu, B., Leuteritz, A., Heinrich, G., 2013. Nanohybrids of phenolic antioxidant intercalated into MgAl-layered double hydroxide clay. Appl. Clay Sci. 71, 814. Luthra, M., Ranganathan, D., Ranganathan, S., Balasubramanian, D., 1994. Racemization of tyrosine in the insoluble protein fraction of brunescent aging human lenses. Biol. Chem. 269, 2267822682. Mandal, S., Patil, V.S., Mayadevi, S., 2012. Alginate and hydrotalcite-like anionic clay composite systems: synthesis, characterization and application studies. Microporous Mesoporous Mater. 158, 241246. Monzn, A., Romeo, E., Marchi, A.J., 2001. Hydrogenation catalysys by mixed oxides prepared from LDHs. In: Rives, V. (Ed.), Layered Double Hydroxides: Present and Future. Nova. Sci. Pub., Inc., New York, pp. 323382. Nakayama, H., Takeshita, K., Tsuhako, M., 2003. Preparation of 1-hydroxyethylidene-1,1diphosphonic acid-intercalated layered double hydroxide and its physicochemical properties. J. Pharm. Sci. 92, 24192426. Nakayama, H., Wada, N., Tsuhako, M., 2004. Intercalation of amino acids and peptides into MgAl layered double hydroxide by reconstruction method. Int. J. Pharm. 269, 469478. Nakayama, H., Kuwano, K., Tsuhako, M., 2008. Controlled release of drug from cyclodextrinintercalated layered double hydroxide. J. Phys. Chem. Solids 69, 15521555. Newman, S.P., Jones, W., 1998. Synthesis, characterization and applications of layered double hydroxides containing organic guests. New J. Chem. 22, 105111. Newman, S.P., Cristina, T.D., Coveney, V., Jones, W., 2002. Molecular dynamics simulation of cationic and anionic clays containing amino acids. Langmuir 18, 29332939. Nie, H.Q., Hou, W.G., 2012. Synthesis and characterization of ifosfamide intercalated layered double hydroxides. J. Dispersion Sci. Technol. 33, 339345. Oh, J.M., Choi, S.J., Kim, S.T., Choy, J.H., 2006a. Cellular uptake mechanism of an inorganic nanovehicle and its drug conjugates: enhanced efcacy due to clathrin-mediated endocytosis. Bioconjugate Chem. 17, 14111417. Oh, J.M., Park, M., Kim, S.T., Jung, J.Y., Kang, Y.G., Choy, J.H., 2006b. Efcient delivery of anticancer drug MTX through MTXLDH nanohybrid system. J. Phys. Chem. Solids 67, 10241027. Pan, D., Zhang, H., Zhang, T., Duan, X., 2010. A novel organicinorganic microhybrids containing anticancer agent doxiuridine and layered double hydroxides: structure and controlled release properties. Chem. Eng. Sci. 65, 37623771. Panda, H.S., Srivastava, R., Bahadur, D., 2009. In-vitro release kinetics and stability of anticardiovascular drugs-intercalated layered double hydroxide nanohybrids. J. Phys. Chem. B 113, 1509015100. Pang, X.J., Ma, X.M., Li, D.X., Hou, W.G., 2013. Synthesis and characterization of 10hydroxycamptothecin sebacate layered double hydroxide nanocomposites. Solid State Sci. 16, 7175. Parashar, P., Sharma, V., Agarwal, D.D., Richhariya, N., 2012. Rapid synthesis of hydrotalcite with high antacid activity. Mater. Lett. 74, 9395. Patel, R.N., Hanson, R.L., Banerjee, A., Szarka, L.J., 1997. Biocatalytic synthesis of some chiral drug intermediates by oxidoreductases. J. Am. Oil Chem. Soc. 74, 13451360.

V. Rives et al. / Applied Clay Science 8889 (2014) 239269 Ulibarri, M.A., Hermosn, M.C., 2001. Layered double hydroxides in water decontamination. In: Rives, V. (Ed.), Layered Double Hydroxides: Present and Future. NOVA Sci. Pub., Inc., New York, pp. 251284. Ulibarri, M.A., Labajos, F.M., Rives, V., Trujillano, R., Kagunya, W., Jones, W., 1994. Comparative study of the synthesis and properties of vanadate-exchanged layered double hydroxides. Inorg. Chem. 33, 25922599. Valarezo, E., Tammaro, L., Gonzalez, S., Malagn, O., Vittoria, V., 2013. Fabrication and sustained release properties of poly (-caprolactone) electrospum bers loaded with layered double hydroxide nanoparticles intercalated with amoxicillin. Appl. Clay Sci. 72, 104109. Wang, Q., O'Hare, D., 2012. Recent advances in the synthesis and application of layered double hydroxide (LDH) nanosheets. Chem. Rev. 112, 41244155. Wang, X., Zhang, Q., 2004. Effect of hydrotalcite on the thermal stability, mechanical properties, rheology and ame retardance of poly(vinyl chloride). Polym. Int. 53, 698707. Wang, Y., Zhang, D., 2012. Synthesis, characterization, and controlled release antibacterial behaviour of antibiotic intercalated MgAl layered double hydroxides. Mater. Res. Bull. 47, 31853194. Wang, Z., Wang, E., Gao, L., Xu, L., 2005. Synthesis and properties of Mg2Al layered double hydroxides containing 5-uorouracil. J. Solid State Chem. 178, 736741. Wang, J., Liu, Q., Zhang, G., Li, Z., Yang, P., Jing, X., Zhang, M., Liu, T., Jiang, Z., 2009. Synthesis, sustained release properties of magnetically functionalized organicinorganic materials: amoxicillin anions intercalated magnetic layered double hydroxides via calcined precursors at room temperature. Solid State Sci. 11, 15971601. Wei, M., Yuan, Q., Evans, D.G., Wang, Z., Duan, X., 2005. Layered solids as a molecular container for pharmaceutical agents: l-tyrosine-intercalated layered double hydroxides. J. Mater. Chem. 15, 11971203. Wei, M., Pu, M., Guo, J., Han, J.B., Li, F., He, J., Evans, D.G., Duan, X., 2008. Intercalation of l-dopa into layered double hydroxides: enhancement of both chemical and stereochemical stabilities of a drug through hostguest interactions. Chem. Mater. 20, 51695180. Whilton, N.T., Vickers, P.J., Mann, S., 1997. Bioinorganic clays: synthesis and characterization of amino- and polyamino acid intercalated layered double hydroxides. J. Mater. Chem. 7, 16231629.

269

Winters, C.S., Shields, L., Timmins, P., York, P., 1994. Solid-state properties and crystal structure of gliclazide. J. Pharm. Sci. 33, 3003004. Wypych, D., Satyanarayana, K.G. (Eds.), 2004. Clay Surfaces: Fundamentals and Applications. Interface Science and Technology, vol. 1. Elsevier Academic Press, Amsterdam, pp. 1553. Xia, S.J., Ni, Z.M., Xu, Q., Hu, B.X., Hu, J., 2008. Layered double hydroxides as supports for intercalation and sustained release of antihypertensive drugs. J. Solid State Chem. 181, 26102619. Xu, Z.P., Lu, G.Q., 2005. Hydrothermal synthesis of layered double hydroxides (LDHs) from mixed MgO and Al2O3: LDH formation mechanism. Chem. Mater. 17, 10551062. Xu, Z.P., Lu, G.Z., 2006. Layered double hydroxide nanomaterials as potential cellular drug delivery agents. Pure Appl. Chem. 78, 17711779. Xu, Z., Fan, J., Zheng, Sh., Ma, F., Yin, D., 2009. On the adsorption of tetracycline by calcined magnesiumaluminum hydrotalcites. J. Environ. Qual. 38, 13021310. Yamamoto, I., Tai, A., Fujinami, Y., Sasaki, K., Okazaki, S., 2002. Synthesis and characterization of a series of novel monoacylated ascorbic acid derivatives, 6-O-acyl-2-O-alphaD-glucopyranosyl-L-ascorbic acids, as skin antioxidants. J. Med. Chem. 45, 462468. Yang, Z., Guo, Z.J., 2003. Selection of inorganic layered double hydroxide materials for intercalation of a cisplatin-guanosine adduct. Chin. J. Inorg. Chem. 19, 673677. Yang, J.H., Han, Y.S., Park, M., Park, T., Hwang, S.J., Choy, J.H., 2007. New inorganic-based drug delivery system of indole-3-acetic acid-layered metal hydroxide nanohybrids with controlled release rate. Chem. Mater. 19, 26792685. Zhang, H., Zou, K., Guo, S., Xue Duan, X., 2006. Nanostructural drug-inorganic clay composites: structure, thermal property and in vitro release of captoprilintercalated MgAl-layered double hydroxides. J. Solid State Chem. 179, 17921801. Zhao, H., Nagy, K.L., 2004. Dodecyl sulfatehydrotalcite nanocomposites for trapping chlorinated organic pollutants in water. J. Colloid Interface Sci. 274, 613624. Zhao, Z., Qi, Y., Wei, M., Zhang, F., Xu, S., 2012. Layer-by-layer assembly and morphological characterizations of DNA/layered double hydroxide thin lms. Mater. Lett. 78, 6265. Zumreoglu-Karan, B., Ay, A.N., 2012. Layered double hydroxides multifunctional nanomaterials. Chem. Pap. 66, 110.

You might also like