Molecular Engineering of Functional Inorganic and ... - ACS Publications

Oct 2, 2013 - Gustavo PalácioSandra H. PulcinelliRachid MahiouDamien .... Vicente Rives , Raquel Trujillano , Miguel A. Vicente , Antonio Gil , Sophi...
0 downloads 0 Views 2MB Size
Review pubs.acs.org/cm

Molecular Engineering of Functional Inorganic and Hybrid Materials C. Sanchez,*,†,‡,§ C. Boissiere,†,‡,§ S. Cassaignon,†,‡,§ C. Chaneac,†,‡,§ O. Durupthy,†,‡,§ M. Faustini,†,‡,§ D. Grosso,†,‡,§ C. Laberty-Robert,†,‡,§ L. Nicole,†,‡,§ D. Portehault,†,‡,§ F. Ribot,†,‡,§ L. Rozes,†,‡,§ and C. Sassoye†,‡,§ †

UPMC Univ Paris 06, UMR 7574 Chimie de la Matière Condensée de Paris, Collège de France, 11 Place Marcelin Berthelot, 75231 Paris cedex 05, France ‡ CNRS, UMR 7574, Chimie de la Matière Condensée de Paris, Collège de France, 11 Place Marcelin Berthelot, 75231 Paris cedex 05, France § Collège de France, UMR 7574, Chimie de la Matière Condensée de Paris, Collège de France, 11 Place Marcelin Berthelot, 75231 Paris cedex 05, France ABSTRACT: For 25 years, Chemistry of Materials has been clearly providing a key forum to chemists, physicists, and engineers interested in materials preparation and characterization, in the search of unique physical properties and processing of innovative materials and devices. This short review presents a discussion on some recent advances in this field, illustrated with a few selected examples related to three of our main research areas: the synthesis of single nano-objects and the processing of porous and hierarchically structured materials and hybrid nanocomposite materials. A strong emphasis is also given to the need to realize a successful marriage between materials chemistry and smart processing. These cross-cutting approaches in the vein of bioinspired synthesis strategies are allowing the development of complex systems of various shapes with perfect mastery at different length scales of composition, structure, porosity, functionality, and morphology. These “integrative strategies”, where all aspects of materials science are coupled, open a land of opportunities to tailor-made advanced inorganic and hybrid materials. Some of them are already impacting numerous societal concerns and industrial applications. KEYWORDS: nanomaterials, hybrid, sol−gel, bottom-up, templated growth, self-assembly, processing, integrative chemistry



INTRODUCTION In 1989, Chemistry of Materials was the pioneer vector of a new multidisciplinary field that arose from the growing recognition of the importance of chemistry to materials science and the increasing involvement of chemists, physicists, and engineers in materials preparation and characterization, in the search of unique physical properties and processing of innovative materials and devices. At that time, Professor Leonard Interrante wrote “A major objective of Chemistry of Materials is to provide a forum for work in materials related chemistry and to highlight the pivotal role of chemistry as a source of new materials and approaches to materials processing”.1 Today the “Chemistry of Materials” field has achieved maturity and numerous smart scientific contributions are mushrooming from all parts of the planet2−64 where developments of numerous new nanotechnologies are able to spring at the interface between chemical and physical approaches. Indeed, achieving original nanomaterials, nanostructured or hierarchical hybrid architectures, involves cross-cutting synthetic strategies where all facets of chemistry (organic, polymers, soft matter, solid-state, physical, materials chemistry, biochemistry, etc.) and ingenious processing are synergistically coupled. Linking “soft chemistry” with the wide range of soft matter processing allows to rationally design and develop inorganic or hybrid nanomaterials in the form of fibers, thin or thick coatings, as well as powders, foams, and monoliths. These cross-cutting approaches are in © 2013 American Chemical Society

the vein of bioinspired synthesis strategies, where the integration of different areas of expertise allows the development of complex systems of various shapes with perfect mastery at different length scales of composition, structure, porosity, functionality, and morphology.9,27,65,66 These “integrative strategies” open a land of opportunities to create advanced inorganic and hybrid nanomaterials with organic−inorganic or bioinorganic character.9,27,65−67 Some general strategies associated to the molecular design and construction of functional inorganic and hybrid nanomaterials or nanostructured materials are cartooned in Figure 1. From top to bottom, this cartoon is split into two main parts corresponding respectively to chemical strategies and common processing methodologies. A summarized view of the different chemical approaches that allow to build functional materials by bottom up strategies can be split into five main physicochemical paths labeled A, B, C, D, E. Path A corresponds to the synthesis of well-defined nonaggregated single nano-objects (NBBs for NanoBuilding Special Issue: Celebrating Twenty-Five Years of Chemistry of Materials Received: July 26, 2013 Revised: October 1, 2013 Published: October 2, 2013 221

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

Figure 1. Cartoon of the general strategies associated to the molecular design and construction of functional inorganic and hybrid nanomaterials or nanostructured materials. Top: chemical pathways. Bottom: processing approaches and some examples of resulting materials: From left to right, sol− gel derived dense photochromic hybrid monoliths (Figure 1-1),100 monolithic hybrid foams for catalysis and biocatalysis (Figure 1-2),101 electrospinned hybrid ion conductive fibers and membranes (Figure 1-3 and 1-4),102 composite porous layers precursors of mesoporous and microporous quartz films (Figure 1-5),103 pillared mesoporous coatings for nanofluidics (Figure 1-6),104 silicon nanopatterns resulting from combination of surfactant templated sol−gel layers and RIE (Figure 1-7),105 ink jet processed micro-dot arrays constructed with mesoporous hybrid structures (Figure 1-8),106,107 aerosol processed mesoporous microspheres of γ-alumina (Figure 1-9).108,109

Blocks) that can be either clusters with strong molecular character or nanoparticles, nano-lamellar compounds (clays, layered double hydroxides, lamellar phosphates, oxides, or chalcogenides, etc.), the compositions of which can be fully inorganic or hybrid with organic moieties either included in the framework or located at the nano-object surface. The nanoobjects are obtained via the controlled reactivity of molecular precursors (metal alkoxides, metallo-organic or organo-metallic compounds, inorganic salts, etc.) that allows mastering of nucleation, growth, and aggregation processes. These syntheses can be performed, for example, in water, organic solvents, inorganic molten salts, ionic liquids, gels, and xerogels, etc. They are chemically triggered via different common reactions such as hydrolysis−condensation, nonhydrolytic condensation, reduction, sulfidation, phosphidation, boronization, etc. Some of these reactions can present a partial reversibility and therefore can be controlled (quenched or speed up) by tuning

chemical and physical parameters as water content, nature of the solvent, pH, dilution, temperature, chemical additives, etc. These reactions can be stimulated through diffusion or microwave heating, or via photonic or ultrasonic solicitations. Depending on precursors compositions (i.e., the chemical nature of nanocompounds targeted), the mixing mode, the solvent, and the excitation mode, the working temperature can range from room temperature to 900 °C. However, the 0−350 °C temperature range is the most extensively and commonly used. Following these strategies numerous clusters or nanosized metal-oxides, metals, metals chalcogenides, metal phosphides, etc., and their hybridized allied derivatives with controlled size structure, shape, and morphology have been produced and characterized during the last 25 years.2−19,21−60,62,68−74 Paths B−E correspond to strategies commonly used to generate nanostructured or hierarchically structured inorganic or hybrid materials. Path B corresponds to soft chemistry-based 222

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

These nanoporous materials present a high degree of order and their mesoporosity is available for further organic functionalization through surface grafting reactions.110 Another possibility corresponds to the combination of self-assembly and NBB approaches.22 These NBBs with tunable functionalities allowing to generate a large variety of hybrid organic−inorganic interfaces via covalent bonding, complexation, electrostatic interactions, can, through molecular recognition processes, permit the development of a kind of vectorial chemistry.111,112 The synthesis strategies (paths A−E) reported mainly offer the control of materials structures and textures from nanoscopic and mesoscopic to submicronic ranges. The micrometer range can be easily reached by including other templating strategies. Indeed, microtemplating methods using for example controlled phase separation phenomena, emulsion droplets, bigger latex beads, bacterial threads, colloidal templates, or organogelators have been developed. They allow to control the shape of complex objects in the micrometer scale. However, the construction, at all length scales9,27 (from nanometer to a few tens of centimeters), of hierarchically structured inorganic or hybrid materials that are integrated in real devices, needs the use of more integrative pathways where chemistry, soft matter, and processing are strongly coupled. Moreover, because of the tunable rheology of the colloidal dispersions, materials can be engineered through many processing methods used for organic polymers, among which are casting of monoliths or membranes, films deposition methods (dip-coating, spin coating, spray coating),113,114 fiber drawing (via extrusion,113 pulling,115 or electrospinning116−118), electrochemical deposition,113 (soft) or (hard) lithography based techniques (dip-pen, X-ray,119−121 two photon adsorption (TPA) lithography122,123), aerosol or spray, electro-spray processing,109,124 inkjet printing, electron beam writing, nanoimprinting, etc. Recently, the coupling of these synthesis procedures with a top-down approach such as RIE (reactive ion etching) plasma processing allowed building ever more sophisticated materials.105 All these integrative strategies are indeed opening new avenues for the tailored construction of hierarchically structured inorganic and hybrid materials. Some examples of homemade materials resulting from these approaches are illustrated in the bottom part of Figure 1.The very strong development and improvement of the “Chemistry of Materials” observed in the last 25 years is already impacting important societal concerns and domains of applications as those associated with energy, environment, and sustainable development, biology and medical sciences, sciences and techniques of the information, societal and personal comfort. This short review concerns the synthesis, processing, and properties of functional inorganic or hybrid materials. It describes and discusses some of the different facets of this important field of research. However, because this ocean of activities cannot be reviewed extensively, we will focus on some examples of materials that have been recently synthesized and processes in our research team. For the sake of clarity, this feature article has been split in three main topics: (i) single nano-objects that include nanoparticles of oxides, nanoheterostructures and nanohybrids, exotic nanoparticles (suboxides and nonoxide nanoparticles); (ii) porous and hierarchically structured materials, including templated growth approaches, liquid deposition processing of thin films, spraydrying of hierarchical powders, microemulsion derived hybrid foams, controlled processing of MOFs nanoparticles, etc.; (iii) hybrid nanocomposite materials including cluster-polymer-

routes including sol−gel chemistry (hydrolytic or non hydrolytic)60,75−77 performed from simple precursors or via the use of specific bridged and polyfunctional precursors. Conventional sol−gel pathways to hybrid networks are obtained through condensation of organically modified metal alkoxides or metal halides condensed with or without simple metallic alkoxides. These materials, exhibiting infinity of microstructures, are amorphous, generally transparent, and easily shaped as films or bulks. If tailored properties are sought for, a better control of the local and semilocal structure of the hybrid materials and their degree of organization is needed. This can be achieved by using bridged precursors as silsesquioxanes X3Si-R′-SiX3 (R′ being an organic spacer, X = Cl, Br, −OR)8,14,75 where the chemical tailoring of the organic bridge allows to improve supramolecular interactions yielding materials with a better degree of organization. Path C allows to obtain more often crystallized materials that are consequently more easy to characterize. It includes all aspect of hydrothermal or solvothermal synthesis generally performed in polar solvents.17,25,31,78−95 The presence of organic templates had given rise to numerous microporous hybrid materials as organically templated zeolites that have already led to an extensive number of applications in the domain of adsorbents or catalysts. More recently, a new generation of crystalline microporous hybrid solids (MOFs, metal organic frameworks) have been strongly developed by several research groups.17,25,31,78−94 These hybrid materials are coordination polymers built from telechelic or polyfunctional spacers that coordinate metallic centers or link in situ generated metal containing small oligomers such as metallic oxo-clusters. Path D corresponds to the association of well-defined nanobuilding blocks (NBBs) (via assembly or intercalation or intercalation and dispersion) to provide nanostuctured hybrid networks with better structural definition. Moreover, the stepby-step preparation of these materials usually allows for a high control over their semilocal structure. Indeed, the nanobuilding components being nanometric, monodispersed, and with a better definition of structures and shapes, their use as NBBs facilitates the characterization of the final materials. These NBBs selected to be able to keep their integrity in the final material can be capped with polymerizable ligands or connected through organic spacers, such as telechelic molecules, polymers, or functional dendrimers. The variety found in the nanobuilding blocks (nature, structure, and functionality) and links allows one to build an amazing range of different architectures with tunable organic−inorganic interfaces, allowing a large set of different assembling strategies. Path E corresponds to procedures allowing the organization or the texturation of growing inorganic or hybrid networks, the growth of which has been templated by micelles, lyotropic liquid crystals, latex, or silica beads. Since 1990, a large field, which corresponds to the construction of mesostructured inorganic or hybrid networks, has been explored.5,9,12,23,27,96,97 This growth process is based on the self-assembly of molecular or polymeric amphiphilic components coupled with sol−gel polycondensation reactions or with other colloidal chemistrybased approaches. More recent strategies consist in the use of alkoxysilylated surfactants98 or surfactant templated growths with bridged silsesquioxanes as precursors.99 This latter approach yields a new class of periodically organized mesoporous hybrid silicas with organic functionality within the walls. 223

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

Figure 2. Strategies for nanoparticle design from solvent-based approaches.

inorganic precursor nature, the reaction temperature and duration, the acidic and ionic strength conditions, the choice and concentration of organic or inorganic additives, the possibility to perform multistep reactions, etc.), outstandingly efficient materials could be obtained in the domains of catalysis and photocatalysis, for electronic, optical, and magnetic applications.145,153,154 Our group largely participated to that burst of experimental results by developing green processes in water (Figure 2A). It also continued its efforts in the understanding and the modeling of the impact of each synthetic parameter and its consequences on the properties. For instance, we pursued the tailoring of titanium dioxide nanoparticles in terms of structure, size, and, more recently, shape. Recent studies demonstrated the importance of exposing certain crystallographic faces in order to improve photocatalytic properties.155−157 We demonstrated that indeed, the morphology of anatase particles impacted the surface reactivity146,148 and, for instance, the property in dye sensitized solar cells (DSSC).158 Strong attention is also paid to the way photoactive nanoparticles may be packed.159−161 Using innovative processing techniques for anatase161 and/or rutile159 nanorods, we

based or nanoparticle−polymer based hybrid materials, new hybrid star-like macromolecular compounds, NBBs based hybrids with self-healing properties, photoactive MOFs, electrospinned hybrid organic−inorganic membranes for (PEMFC). Finally, some examples of applications113,125−136 will be shortly presented in a final conclusive outlook. This outlook will also summarize some of the trends that will allow to push further the limits of nanochemistry developed with inorganic or hybrid matter, associated with innovative processing techniques.



SINGLE NANO-OBJECTS The synthesis of nanoparticles from molecular precursors in solvents has reached33,41,68−70,137−148 an unprecedented degree of sophistication and precision as to the control of structural, morphological, and textural features (Figure 2). Some of the most relevant approaches are presented hereafter. The synthesis of metal oxide nanoparticles from molecular precursors in solvents is known as the most convenient way to finely tune the structure, size, and morphology of the prepared materials.41,138,141,149−152 By optimizing several experimental parameters in the colloids preparation (the solvent and the 224

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

more difficult to obtain as nanocrystals. It is also the case of even apparently simple spinel systems such as ferrochromites CrFe2O4. In order to improve the mixing of cations with seemingly different reactivities into ternary oxide nanoparticles, the use of original heating methods such as microwave-assisted syntheses may be useful (Figure 2D).68,69 This tool was used for the preparation of improved visible light photocatalyst bismuth mixed oxides such as Bi2WO6.144 Light emitting materials (laser, diode, etc.) have expanded greatly during the past ten years with major technological breakthrough in information and communications technology and of energy conversion devices. Recently, promising novel luminescent nanoparticles based on lanthanide- or transition metal-doped oxide have been synthesized combining sol−gel process and conventional solid state chemistry, opening a new field of research for in vivo optical imaging.175 In particular, we were pioneers for the development of long lasting fluorescent nanoprobes emitting in the wavelength therapeutic window (red to near-infrared) (Figure 2F) during more than one hour.176 Based on a better understanding of persistent luminescence mechanisms177 and of the influence of the chemical composition on the luminescence intensity,178,179 nanoparticle properties have been greatly improved suggesting amazing possibilities of such nanostructures in many other nanotechnology applications. Despite the important achievements in terms of nanofeatures tuning, many advances are yet to be done in order to ensure full understanding of precipitation mechanisms and to provide synthesis techniques suitable to a wide range of nanomaterials. In particular, most achievements in the community of nanoparticle design have been strikingly confined to the families of binary metal oxides, metal chalcogenides, and metals.33,137,139,140 Much less is known about nanostructures of more complex compounds, such as metal oxides bearing “uncommon” oxidation states (e.g., Ti(III)), which are usually poorly stable under ambient conditions, and metal−nonmetal alloys (e.g., metal carbides, nitrides, phosphides, borides), which often show strong reluctance to crystallization and require harsh synthesis conditions unsuitable to limit grain size. Because many peculiar mechanical, electrical and catalytic properties are only encountered within these “exotic” compounds, it is urgent to address their downscaling and to investigate its impact on the properties.70,147 Some fundamental leaps have been recently achieved in the nanodesign of these materials through the development of novel synthetic processes from molecular precursors.142,143,147,180−184 Metal borides and metal phosphides are a first case of such uncommon nanomaterials. They exhibit original properties such as high Tc superconductivity (MgB2, LiFeP), high catalytic activity (Ni−B, Ni2P, MoP), ultrahardness (ReB2), hard ferromagnetism (Nd2Fe14B), luminescence (GaP), and reactivity vs hydroxides or Li+ for batteries (Co−B, CoP3, MnP4). Most of these peculiar behaviors originate from the strong covalent character of the metal−metalloid and metalloid− metalloid bonds. This bonding scheme also results in high activation energy for heteroelement incorporation and/or crystallization from the amorphous state. High temperatures are therefore required to reach crystalline bulk phases but cannot be applied to the synthesis of intrinsically metastable nanostructured compounds. Borides and phosphides illustrate two complementary strategies to achieve crystalline nanostructures of such compounds, as recently reviewed.147 The first approach is “process-based” and was used for boron-containing

demonstrated the impact on photocatalytic properties of the relative nanoparticles orientation under polarized incident light. In a more fundamental approach, the combined use of experimental synthesis and first principle calculations were shown in the past decade to be a powerful tool to account for the morphology dependence on the additive used in the reacting medium. This was shown, for instance, in the case of anatase platelets tailored in the presence of fluoride ions that specifically stabilize {001} surfaces.156 In the same way, highly complex faceting of hematite (α-Fe2O3) particles, namely octodecahedral162 or truncated hexagonal bipyramidal163 morphologies, has been reported using fluoride and hydrazine as capping agent, respectively, while rhombohedral particles are formed without any additive.164 The relative surface stabilization of boehmite (γ-AlOOH) in the presence of bigger organic additives such as polyols (and more specifically xylitol C5H12O5) could be ascertained through experimental and calculation results.165,166 In the above-mentioned systems, the metallic salts used as precursors may differ in the coordination sphere of the cation (halides, nitrates, sulfates, and alkoxydes, for instance) but usually present the same oxidation state as that in the targeted oxide. It is interesting, in addition to other experimental parameters, to play on the oxido-reduction properties of the precursors to yield known structures but with reduced size and original morphology. This was used in the past decade for TiO2 using TiCl3 as titanium source167 and more recently with extremely small and well calibrated nanoparticles of 2 nm RuO2 obtained from Ru(III) salts.145 In this case, the fast addition of hydrogen peroxide triggers the oxidation of the precursor resulting in almost immediate hydrolysis/condensation and therefore immediate consumption of Ru species. Of course, some oxides with single or mixed valence correspond to oxidation states difficult to isolate in metal molecular precursors. It is the case of Mn(III) or Mn(IV) oxides that must be obtained in aqueous solution from the oxidation of Mn(II) salts, the reduction of Mn(VII) anions or the combination of those two former species. An accurate control of the redox reactions coupled with selected aging conditions (pH, T and counterions) allowed then the preparation of an interesting variety of manganese oxidebased nanostructures.168,169 Moreover, the simultaneous evolution in the reaction medium of the precursors’ concentrations, the equilibrium redox potential, and the solution acidity was demonstrated as a valuable toolbox for the design of nanoheterostructures, especially for manganese-based oxides. Heterogeneous nucleation was especially triggered to perform the one-pot (in situ seeding)170 synthesis of core−shell architectures or to achieve complex heterogeneous nanowires by a two-step (seeding) method.171 Self-assembly in water was especially assessed in the framework of the oriented attachment mechanism, which gives the possibility to achieve nanoparticle size control172 and unprecedented complex shapes173 in water through “chimie douce” routes. In most cases, the decrease in ternary oxides particles size down to the nanometer scale through sol−gel routes is not straightforward. Unlike in the ceramic approach, where high temperature may force the different cations in the same oxide structure, cations of ternary oxides in aqueous solution may precipitate separately unless specific experimental conditions are found to match their reactivity. For instance, certain perovskites such as BaTiO3 are quite easy to prepare from aqueous solutions,174 while others such as manganites are much 225

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

nanostructured pellets by spark plasma sintering. The resulting materials showed enhanced thermoelectric performances compared to bulk Magnéli phases, due to reduced thermal conductivity through phonon scattering at the carbon-nanoparticle interfaces. All in all, tremendous progress has occurred during the last ten years in the design of single nano-objects by solution-based chemistry. A rebirth for nanoparticles synthesis in solvents can now be foreseen, especially in “nano-foundry” laboratories where three axes can be highlighted: the use of alternative solvents and stimulation processes for targeting new nanomaterials is becoming a very active field; the design of nanoparticles with unprecedented complex compositions will open fascinating opportunities; and unveiling precipitation and crystallization mechanisms through innovative in situ experiments is another worthwhile investigation route in which many groups are now rushing. Taking into account the current urge for greener syntheses, there is no doubt that these solventbased approaches, especially those using environmentally benign solvents, will be further developed and will take a growing place in industrial processes. However, one has to keep in mind that most often, nanoparticles must be shaped and processed into nanostructured materials in order to achieve real life applications. The next sections clarify this point by providing a critical overview of state-of-the-art processing techniques.

systems (Figure 2C). It consists in enhancing the reaction kinetics to ensure crystallization in the intermediate temperature range 400−900 °C, in milder conditions than those usually required, above 1000 °C. To do so, the reaction between molecular precursors (metal chlorides and sodium borohydride) was triggered in a thermally stable and environmentally benign solvent, made of a eutectic mixture of alkali chlorides.142 This colloidal synthesis in molten salts affords nonpyrophoric nanocrystals with an unprecented variety in terms of compositions (Mn2B, FeB, NbB2, MoB4, CaB6, EuB6). It also provides for the first time the ability to tune the particle size between 5 and 20 nm. Noteworthy, molten salt syntheses were also successfully applied to metal-free boron carbon nitrides183−186 while others used the approach for the fabrication of Si, SiC, and Ge nanostructures.187,188 The second route toward nanostructures of compounds with strong activation energy for crystallization was applied to metal phosphides (Figure 2B). This “precursor-based” strategy uses suitable reactive precursors allowing a decrease in the energy input required for crystallization. The synthesis have been developed of metal phosphide nanocrystals by reacting below 300 °C M(0) complexes or nanoparticles with highly reactive P4 solubilized into a nonpolar solvent. The reaction occurs in a stoichiometric manner and leads to a wide range of metal phosphides (Ni2P, Cu3P, Pd5P2, PdP2, FeP, InP) with interesting properties for catalysis189 and Li batteries.190 Moreover, phase segregation can occur during the conversion of M(0) particles and lead to core−shell structures, as exemplified by Ni2P−Ni nanoparticles with tunable magnetic properties.181 In both borides and phosphides cases, many promising and exciting research areas are now within reach, including detailed mechanistic understanding, investigation of novel binary and ternary compositions, and electronic, magnetic, optical and mechanical properties. Titanium Magnéli phases TinO2n−1 (4 ≤ n ≤ 10) are a second family of exotic compounds to many nanomaterials chemists. Most of their properties of interest are based on their mixed valence Ti(IV)−Ti(III) and their rutile-derived crystal structure, providing high electron concentration and phonon scattering, respectively. Substoichiometric titanium oxides are a long known family of compounds that has experienced a strong revival during the last years due to surprising electronic and photonic properties,191−193 as well as the possibility to tune these behaviors by adjusting the Ti(III) to Ti(IV) ratio. Advances in the related application fields are mostly hindered by synthetic difficulties, especially because the production of Magnéli phases relies commonly on the reduction of stoichiometric TiO2 above 1000 °C, so that nanoscaling remained until now a somehow elusive target.156 To overcome this difficulty, we combined in a one-pot approach the sol−gel process with carbothermal reduction (Figure 2E).143 Interestingly, nanoparticles of Magnél i phases result from a combination of both precursor- and process-based strategies described above: the use of a polymer and of titanium ethoxide as sources of carbon and titanium, respectively, ensures the production at 500 °C of an intimate blend of a carbon matrix and TiO2 nanoparticles, while further “one-pot” heating between 900 and 1000 °C leads to the reduction of the titanium dioxide particles into Magnéli phases. This result showed the first ever reported ability to tune the particle size and the degree of substoichiometry through adjustment of the temperature and the carbon content. Moreover, the as-obtained Ti4O7 nanoparticles−carbon composites could be shaped into



POROUS AND HIERARCHICALLY STRUCTURED INORGANIC MATERIALS The construction of porous and hierarchically structured materials can be realized by the so-called Integrative Chemistry strategies consisting in the smart coupling between sol−gel chemistry, multiple templating, and advanced processing.9,67,126,194 The general concept relies on a solution containing the inorganic precursors and the templating agents (and other preformed nano-objects), which is processed in order to define the macroscopic shape of the final materials (films, powders, monoliths, fibers, etc.). In addition, a large variety of templating agents and strategies can be coupled with inorganic polymerization reactions for the design of ordered nanostructured hybrid phases that yield controlled porous and hierarchical materials.27 The processing conditions such as the evaporation of the volatile species or the condensation degree need to be controlled since they affect the kinetics of nanosegregation and formation of hybrid organic/inorganic interfaces, allowing the design of tailored porosity in the socalled “race towards order”.195 Different approaches can be used and combined. Microporosity can be intrinsically present in several crystalline materials as molecular cavities (as in metal organic frameworks or zeolites). In this category, zeolites are among the most valuable of microporous inorganic materials because of their thermal and mechanical stability, tunable porosity (from small to large micropores), and chemical versatility.196 The ability to control the synthetic process allowed recently the development of zeolites with new properties (e.g., hydrophobic structure) and architectures and particles size reduced to the nanometer scale.63,197 Remarkably, hierarchical mesoporous molecular sieves with microporous zeolithic aluminosilicate walls were directly generated by dualporogenic surfactant-driven synthetic process and used as efficient catalyst for various organic reactions involving bulky molecules.198 Similarly, catalytic bimodal micromesoporous amorphous aluminosilicate powders were obtained by aerosol 226

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

compositions are obtained through chemical quenching induced by rapid solvent evaporation of the precursors solution onto the substrate. The fundamental interest of this strategy is that, in such conditions, the composition of the final film is exactly the composition of the initial solution in nonvolatile species. Among the various techniques that can be utilized, dipcoating deposition (Figure 3a) gives the best control of all the evaporation-related parameters. It can be performed in controlled environment (temperature and atmosphere) and at extreme ultralow or fast withdrawal speeds allowing deposition from very diluted or aqueous solutions.214−216 Taking advantage of the latter control, mesoporous amorphous, hybrid, nanocrystalline films were produced by evaporation induced self-assembly with high control of the nanostructure and thus tunable functionalities.114,217−219 For example hybrid mesoporous organo-silica films functionalized with dibenzoylmethane were developed as an efficient and versatile sensor able to detect uranyl ions in aqueous solution220 or BF3 gases.221 A wide range of nanocrystalline based mesoporous films, such as TiO2,222 Al2O3,223 SrTiO3,224 ZnO,225 Sr-doped BaTiO3,226 etc., with tunable morphologies and thicknesses were also prepared. A remarkable case is displayed in Figure 3b and consists in RuO2-based mesoporous thin films, exhibiting very high capacitances (1000 F g−1) that can potentially find application as supercapacitors.227 This work opens the way to the design of cheaper nanostructured metal oxides supercapacitors.228,229 Controlling exactly the deposition conditions, ultrathin Inorganic Nanopatterns (INP) layers were developed on several substrates.108,230 These systems present morphological and chemical heterogeneities at the nanoscale and were exploited as nanoelectrodes arrays231,232 and as platform for optoelectronic233 and data storage devices.234−236 In addition, multifunctional hydrophobic, antireflective, antifogging, and photoactive coatings for solar cells were obtained by stacking TiO2 INP layers photoactive layers on hybrid mesoporous silica films.237 Using similar strategies mesoporous films with complex composition were synthesized and characterized for energy production based on hydrogen economy. Ceria mesoporous thin films have been integrated in Solid Oxide Fuel Cell as cathodic interface to enhance O2 transport.238 Better performances have been achieved for these devices.239 These approaches were extended to cermet Ni/Gd−doped Ceria mesoporous thin films as anode for micro-solid oxide fuel cell.240−243 These films exhibit a connected pore network that ensures good gas diffusion and good particle−particle contact for Gd-doped ceria and Ni, which gives satisfactory electrical properties (9.104 S m−1 at 500 °C in 10%H2/Ar). Because of their high specific surface area, mesoporous thin films have been also explored as photoelectrodes for water splitting. A photocurrent of about 100 μA cm−2 has been achieved for mesoporous α-Fe2O3 thin films combined with an electrocatalyst based on Co.244 Recently, nanostructured mesoporous indium−tin oxide electrodes exhibiting both high conductivities and optimized bicontinuous pore−solid network have been reported.245 In the presence of an electron mediator, photocurrents up to 50 μA cm−2 have been measured under visible light irradiation, demonstrating the potential of this new templated nano-ITO preparation for the construction of efficient photoelectrochemical devices.245 These “soft-chemistry” strategies were also recently coupled with solid-state chemistry processes for the preparation of macro- and nanoporous piezoelectric α-quartz thin films103 as displayed in the SEM photo in Figure 3c. Dense, macro- and

process, using tetrapropylammonium hydroxide (TPAOH) as the microstructure-directing agent.199 Finally, innovative preparation routes involving in certain cases dissolution− recrystallization equilibria allowed the formation of nanosized zeolites.200,201 Controlled mesoporosity is usually obtained through Evaporation Induced Self-Assembly by triggering micelle and mesophase formation23 or by packing preformed micelles. For example, a highly stable nanocrystalline γ-Al2O3 layer with contracted face-centered cubic (fcc) mesoporosity is obtained by block-copolymer micelles templating.203,204 Macroporosity can be formed by using condensed larger templates (latex for colloidal packing,19,205−207 organogelators,97,208 salt crystals,209,210 or by dynamic templating strategies such as microphase separation211,212 and breath figures.213 For instance, nanoparticles can be used as NBBs for breath figure-derived honeycomb-like macroporous membranes.202 In order to develop specific applications, adapted elaboration processes need to be coupled with the synthetic strategies yielding final materials with defined macroscopic shapes and functionalities. Starting from liquid solutions, the elaboration processes usually deal with the shaping of small quantities of liquid together with the control of the chemical and physicochemical phenomena taking place while processing. Several examples of elaboration of porous and hierarchically structured materials are illustrated in Figure 3 and will be detailed thereafter. Nanostructured sol−gel derived thin films are commonly prepared by chemical liquid deposition; layers with complex

Figure 3. Illustrations and examples of some processes used for the fabrication of porous and hierarchical materials. (a) Dip coating deposition used for the preparation of (b) RuO2227 and (c) quartz103 mesoporous films. (d) Aerosol technique toward (e) mesoporous γalumina246 and (f) microporous MoO3−SiO2−Al2O3247 powders. (g) Inkjet printing allowed fabrication of (h) arrays of functional silica dots with (i) defined mesoporosity. (j) Foaming process was used to fabricate (k) organo-silica based monoliths with (l) open-cell matrices.248 227

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

hybrid species yielding main advanced porous systems but also it permits the fine-tuning of specific functionalities (sensing, catalytic, adsorption activities, etc.). In the following, we will develop some other interesting aspects of hybrid materials by describing a few recent results that concern hybrid nanocomposites constructed with organic or hybrid polymers and inorganic or hybrid nanofillers.

mesoporous layers were grown directly on silicon by a novel mechanism based on the confined, heteroepitaxial devitrification of amorphous Sr- or Ba-doped mesoporous silica films at 900 °C. Following another approach, dip-coating deposition was also exploited for the preparation of optical quality microporous thin films based on different MOFs (MIL-101(Cr), ZIF-8, and flexible MIL-89(Fe)) from stable MOFs nanocrystals suspensions that could be processed as usual sol−gel solutions.249 Nanostructured inorganic and hybrid porous powders can be prepared by spray-drying process as illustrated in Figure 3d. In this case, the mother solution is not deposited onto a substrate but atomized to form microdroplets that after evaporation lead to nanostructured microspheres. The structuration mechanisms are similar to those described above for films.109,250,251 Hybrid, simple, or mixed metal-oxide based powders with amorphous or crystalline walls252 and various mesostructures could be obtained. Figure 3e shows mesostructured γ-alumina synthetized from aluminum chloride precursors and templated by micelles of block copolymer named KLE; the high decomposition temperature of the organic template (more than 300 °C) allowed formation and preservation of the nanocrystalline network at temperature as high as 900 °C.246 The spray drying technique is particularly suitable for the fabrication of efficient, high surface area, hierarchical porous catalysts since the rapid chemical quenching taking place during evaporation allows an excellent dispersion of the active sites. For instance, amorphous microporous MoO3−SiO2−Al2O3 powders shown in Figure 3f exhibited high specific surface areas and outstanding olefin metathesis activity.247 Recently, the spray-drying technique was also adapted for continuous, scalable and green synthesis of microporous and hierarchically structured MOF-based powders.253,254 Another possible strategy of fabrication hierarchical multifunctional structured materials is to combine EISA and inkjet printing, as illustrated in Figure 3g.255−257 The technique consists in making a microdots pattern on a substrate by depositing single microdroplets ejected via a nozzle. The precursor solution is based on the same chemical composition developed for preparing mesoporous silica thin films with adjusted surface tension and viscosity.258 In addition, a different organic functionalities can be added to each microdroplet opening the way to the fabrication of multiarrays of functional mesoporous dots to be utilized as highly sensitive miniaturized sensors (3h,i).107 Monolithic materials with controlled and accessible porosity and good mechanical properties were also developed as support for catalysis and chromatography. In particular, these systems could be made by foaming strategies that allowed synthesis of organo-silica foams with open-cell matrices and were fabricated by combining microemulsion and liquid crystal templating Figure 3j.248 These new monolithic porous matrices with tunable functionality present a hierarchical distribution of pore sizes from micronic to mesoscopic scales (Figure 3k,l). Function accessibility has been demonstrated through Pd heterogeneous nucleation258 or lanthanide trapping.259 Moreover, these materials present interesting properties as heterogeneous catalysts (for Susuki or Heck reactions) or as biocatalysts.260 We have shown via few examples (EISA, hydrid mesoporous thin films, MOFs, hybrid foams) the importance of hybridization in porous materials. Indeed, hybridization not only allows to template the growth of condensing inorganic or



HYBRID NANOCOMPOSITE MATERIALS Organic/inorganic hybrid materials are not simply physical mixtures. They can be defined as nanocomposites at the molecular scale, having at least one component, either the organic component or the inorganic component, with a characteristic length scale on the nanometer size. In the following examples, we emphasize that the properties of hybrid materials do not simply result from the sum of the individual contributions of their components but also from the strong synergy created by extensive hybrid interfaces. Numerous reviews and books have been devoted to this important topic; therefore, in this review, we will first report on recent progresses that have been made on hybrid nanocomposites based on polymers and well-defined inorganic or hybrid nanobuilding blocks (nanoparticles, metal-oxo clusters, MOFs). Then, taking PEMFC as an example, we will emphasize the importance of the coupling between processing and hybrid materials chemistry to realize highly efficient hybrid membranes in terms of conductivity, mechanical properties, and water management. Polymers−Nanoparticles Hybrid Materials. The usefulness of inorganic nanoparticles as additives to enhance polymer performances is well established,261−267 and now, such hybrids provide additional opportunities for many different commercial applications.135 Small amounts (1−10% volume) of isotropic nanoparticles, such as silica, titania, alumina, and silver, or anisotropic nanoparticles, such as layered silicates (nanoclays) and carbon nanotubes, provide properties enhancement, compared to the neat resin, which is comparable to the one achieved by larger conventional loadings (15−40%) of micrometer-scale inorganic fillers. Interestingly, with lower loadings, the processing is facilitated and the component weight reduced. However, most importantly, the unique properties of these hybrids are otherwise impossible with traditional fillers. For instance, reduced permeability, optical clarity, selfpassivation, as well as flammability, oxidation, and ablation resistance can be obtained.262 Actually, the properties of polymers based nanocomposites do not depend only on their composition (i.e., the nature of the organic and inorganic components and their ratio). Indeed, the synergetic effect caused by the association at the nanometer scale of a priori incompatible components, combined with an accurate monitoring of the hybrid interfaces, allow to design innovative architectures with improved transport, mechanical, or optical properties.270,271 As an example, the ability to finely tune the mechanical properties of polymer-based hybrid coatings, studied by nanoindentation,272,273 has been demonstrated by adjusting the interactions between iron oxides nanoparticles (goethite nanorods (α-FeOOH) or maghemite nanospheres (γ-Fe2O3)) and polymeric268 or polymeric/silica host matrices.269 The incorporation of iron oxides nanoparticles in a silica-PHEMA (poly(hydroxyethyl methacrylate)) matrix does not result in the expected improvement of the mechanical properties. In such a system, iron oxide nanoparticles interact preferentially with the silica network through the formation of 228

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

Figure 4. (a) Postfunctionalization of [Ti16O16(OEt)32] oxo-cluster with ATRP initiators allows the synthesis of hybrid star architecture.274 (b) The postfunctionalization of [Ti16O16(OEt)32] oxo-cluster with H-bond acceptors ligands allows the synthesis of supramolecular architecture when Hbond donor telechelic polymers are added.275 (c) Crystalline highly porous titanium(IV) dicarboxylate hybrids with photocatalytic properties.276

surface functionality to internal volume for cluster compounds also provides an easier characterization of the interface in such hybrid systems and makes them essential models for larger or less-defined systems, such as nanoparticles or surfaces. Over the past decades, many different associations, involving various nanobuilding blocks (NBBs) and polymeric matrices in a “Lego-like” approach, have been studied. Among them, those based on titanium oxo-clusters, which are appropriate NBBs for the build-up of advanced organic−inorganic materials models.278 In particular, one can take advantage of the lability of the organic ligands surrounding the titanium oxo-cores and the high reactivity of the metallic sites to accurately postfunctionalize nanobricks by a selective exchange of ligands.111 As an example, hybrid star-like macromolecular compounds have been designed from a titanium-based macroinitiator, Ti16O16(OR)26(OCH2CCl3)6, and the growth of n-butyl acrylate polymers by Atom Transfer Radical Polymerization (ATRP).279 Such polymerization process provides linear firstorder kinetics and the evolution of the experimental molecular weight is also linear with the conversion.274 (Figure 4a) It therefore leads to the formation of well-defined star-like clustermacromolecules assemblies that allow optimizing the dispersion and the affinities between the hybrid stars and the host polymer matrix. As a consequence, the tensile strength of elastomeric samples filled with such hybrid stars, evaluated by fracture tests, is significantly improved. Alternatively to this “grafting from” approach, a “grafting onto” strategy was also investigated by directly postmodifying a titanium oxo-cluster with preformed macromolecular chains.279 The inefficiency of this second approach, which yields only a

hydrogen bonds. Therefore, the mechanical properties of the hybrid coating are mainly governed by the interactions developed between the amorphous silica and PHEMA networks simultaneously formed. In the absence of strong interactions, the iron oxide nanoparticles could be considered as dissociated from the PHEMA:SiO2 hybrid matrix, leading to the formation of a porous structure through a phase separation mechanism that does not provide any mechanical strengthening, but induces a dramatic collapse of the mechanical properties.269 On the other hand, nanoindentation tests proved that the direct inclusion of nanoparticles into PHEMA leads to a strong reinforcement of the polymer. According to spectroscopic experiments, this efficient reinforcement was attributed to the existence of strong interactions at the iron oxide−PHEMA interface combined with a highly homogeneous dispersion of the nanoparticles.268 Moreover, the goethite nanorods-based hybrid coatings exhibited birefringence. This effect was explained by the stabilization inside the organic glassy matrix of a liquid crystal-type organization resulting from the self-assembly of the anisotropic particles.268 Polymer Clusters Hybrid Materials. Among the different possible strategies to design hybrid materials, the modular approach, which assembles metallic oxo-clusters into macroscopic networks, presents several advantages.22,277,278 First, the stepwise development of the material allows to better control its structure on the semilocal scale. Second, prefabricated nanoobjects often show reduced reactivity compared to inorganic molecular precursors. Moreover, their perfect monodispersity allows the construction of well-defined structures and leads to an enhanced quality of the final material. The high ratio of 229

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

NH2-MIL-125(Ti), which is obtained by the substitution of terephtalate ligands by amino-terephtalate ligands.297 If NH2MIL-125(Ti) exhibits a moderate CO2 adsorption capacity (132 mg g−1), it exhibits an excellent selectivity over N2 (>27:1 at 298 K), with a low heat of adsorption. Therefore, adsorbent regeneration can be easily performed in a reversible manner during four consecutive adsorption−desorption cycles at 298 K for 550 min. NH2-MIL-125(Ti) demonstrated a higher catalytic activity in the cycloaddition of epichlorohydrin due to its higher amount of basic sites compared to MIL-125(Ti). However, MIL-125(Ti) showed a better catalytic performance in oxidative desulfurization of dibenzothiophene, likely due to a lower steric hindrance at the active sites. In batch mode, liquid phase competitive separation of isoprene from 2-methyl butane, MIL-125(Ti) exhibited a significantly higher isoprene selectivity against 2-methyl butane compare to NH2-MIL-125(Ti).298 In addition, hydrogen production from an aqueous medium with NH2-MIL-125(Ti) under visible light irradiation was demonstrated. Pt nanoparticles, deposited onto the MOF via a photodeposition process, were used as cocatalysts. NH2-MIL125(Ti) and Pt/NH2-MIL125(Ti) exhibit efficient photocatalytic activities, under visible light irradiation, for hydrogen production from an aqueous solution containing triethanolamine as a sacrificial electron donor. The results obtained from wavelength-dependent photocatalytic tests and photocurrent measurements, as well as in situ ESR measurements, demonstrate that the reaction proceeds through the light absorption by the organic linker and the subsequent electron transfer to the catalytically active titanium−oxo cluster.299 Finally, the possibility to tune the optical responses of MIL125(Ti) through a rational functionalization of the linking unit was recently investigated by a combined synthetic and computational approach.300 On one hand, MIL-125(Ti) derivatives, containing various amounts of amino-linkers, were synthesized and their optical responses measured, showing a shift of the absorption in the visible range as soon as 10% of amino-linkers are present. On the other hand, DFT calculations were performed to elucidate the role of the amino group in lowering the optical band gap of NH2-MIL-125(Ti) (ca. 2.6 eV). Hybrid Membrane for Polymer Electrolyte Membrane Fuel Cell (PEMFC). Progress in fuel cell technologies relies on the replacement of Nafion by hybrid organic−inorganic membranes.301−322 The mechanical properties, the swelling and the conduction in these membranes can be tuned through the control of their microstructure by using at least two components. Hybrid organic−inorganic membranes are often defined by discrete, intermingled hydrophilic and hydrophobic domains.301 Different concepts for hybrid organic−inorganic membranes have been developed.102,303,306,308,309,311,313,319−322 Currently, we are interested in the design of hybrid membranes for which conduction properties and mechanical properties have been separated into two main components (organic and inorganic components).308 The inorganic moieties support the proton conduction while the organic ones guarantee the mechanical and chemical robustness of the system. Accordingly, the idea is to produce composite materials, wherein a phase separation between hydrophilic and hydrophobic components exists at different scale from a few nanometers up to several micrometers.305,312 This phase separation contributes to its ion conduction performance. Combined effect of processing and chemistry allows a fine-tuning of this phase segregation from the nano- to the macroscale, giving rise to membranes with

small amount of hybrid star-like species, mixed with a large quantity of ungrafted polymer chains, was assessed by 1H DOSY NMR, which can discriminate free molecular species from those interacting with the surface of nano-object, on the basis of translational diffusion coefficients.280−286 More recently, a new approach, which combined inorganic nanobuilding blocks and supramolecular interactions, has been proposed. It is illustrated by Ti-oxo-clusters, postfunctionalized in a controlled manner by hydrogen bond acceptors and telechelic PDMS with hydrogen bond donors (Figure 4b).275 It shows that well-chosen supramolecular interactions, which cross-link the polymer chains, could be used to obtain new hybrid dynamers with improved mechanical properties.287 Strategies to build hybrid nanocomposites with controlled covalent (hard) and noncovalent (soft) interfaces open many possibilities to rationally design mendable hybrid materials.288−290 Among them, the preparation of elastomeric nanocomposites from cheap and common organic monomers and structurally well-defined nanobuilding blocks is especially appealing to yield dynamic materials that have the capacity to efficiently self-repair after strong damages. The controlled design of cross-linked poly(n-butyl acrylate) (pBuA) has been achieved by introducing a very low amount of a specific bifunctional tin oxo-cluster: the macrocation, [(BuSn)12O14(OH)6]2+, functionalized with two 2-acrylamido-2methyl-1-propanesulfonate anions (AMPS).291,292 The electrostatic interactions developed at the hybrid interface of such a system play a double role. They are strong enough to cross-link the polymeric network, which consequently exhibits rubber-like elasticity behavior, but they are labile enough to allow dynamic bond recombination, which contributes, after a severe mechanical damage, to an efficient self-healing process. Indeed, the small size of the inorganic component combined with the elastomeric behavior of the host matrix have been selected to direct mass transport toward the damage site and ease the subsequent redistribution of electrostatic bonds. A significant recovery of tensile strength at break and elongation at break (more than 75% of the original value) were observed on mended samples.291 Metallic−oxo clusters are also encountered as SBUs (secondary building units) in MOFs (metal organic frameworks).31,80,85,293−295 The MIL-125(Ti), constructed from dicarboxylate linkers and octameric titanium IV-oxo clusters, is the first highly porous and thermally robust titanium− carboxylate based MOF to be synthesized and characterized.276 The ability of the metal organic frameworks to absorb large quantities of organic molecules, associated with the photoinduced mixed valence character of titanium−oxo clusters, leads to an interesting reversible photochromic behavior (Figure 4c). A very high photonic sensitivity is actually also observed for interpenetrated hybrid networks made of titanium−oxo polymers and organic polymers.100,296 UV−visible irradiation promotes the formation of a Ti(III)−Ti(IV) mixed valence state. The confinement of the photoinduced electrons as titanium(III) centers, efficiently trapped inside the titanium− oxo domains, ensures a long-term stability of the mixed valence compound and provides a strong and stable coloration. Additionally, the results obtained on MOF models emphasize the photocatalytic properties of these hybrid materials. The first study dealt with the reduction of titanium(IV) centers and the concomitant oxidation of adsorbed alcohol molecules.276 Later, the photocatalytic reduction of CO2 to HCOO−, under visible light irradiation, was achieved over the photoactive isostructural 230

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

Figure 5. Processing, microstructure, and conductivity measurement. (a) Photograph of hybrid organic−inorganic membranes. (b) Top view of electrospun hybrid membrane system highlighting the fibers. (c) Comparison of proton conductivity for electrospun membrane (filled circles) and electrospun membrane impregnated with PDMS (polydimethylsiloxane) (open circles) and Nafion (dotted line) at 120 °C from ref 102.

mentals of nanointerfaces is indeed a key issue because it drives our ability to control the structure and the dynamics of hybrid interfaces in a large sense, and it impacts directly our capability to rationally design and construct nanostructured materials utilizing a performance-property driven methodology. This basic knowledge is necessary to identify and play with the important chemical and physical parameters governing the tuning of interfaces, to allow a better control of the final materials properties, the solubility, the colloidal behavior, the transferability, the stability of the nanostructured materials, etc. The multitude of functional materials produced via the “Chemistry of Materials” span a very large set of structures, textures, and chemical compositions. In addition to their high versatility with regard to shaping and chemical and physical properties, numerous materials produced today through the “Chemistry of Materials” present via hybridization the obvious advantage of facilitating both their integration and miniaturization. The synergetic coupling between soft matter, bottomup chemical strategies and tailored processing conditions has opened new avenues for the designed construction of new materials with numerous properties. The resulting materials are indeed relevant for a host of possible applications including optics, imaging, and photonic devices, nanoionics, and energy (nanostructured materials for fuel cells, PEMFC, SOFC, electrodes, photoelectrodes, batteries, etc.), environment (smart membranes, catalysts, biocatalysts, photocatalysts, chemical and biologi cal sensors, etc.), functional and green protective coatings, multifunctional biomaterials, and new hybrid carriers for medicine or cosmetics, environmentally responsive materials, microfluidic devices, new materials for construction and transport with an emphasis on new hybrid self-healing materials, and smart binders for construction, among others. Indeed, the field of the chemistry of materials represents a cornucopia nourished by imagination and creativity of chemists that allow the emergence of a large diversity of innovative materials of which strong applicative potential is clearly becoming a reality. Some of these materials have already shifted from research laboratories to the market but they only represent a small fraction of the tip of the iceberg.113,125,126 A few examples of materials that already entered the market, concerning automotive, construction, and packaging industries, protection of the environment, functional coatings (highly hydrophobic or highly hydrophilic, protective, anticorrosion, antiscratch, etc.), biomaterials and nanomedicine, micro-optics and information storage, energy saving, conversion, and storage, catalysis, and biocatalysis, are discussed in recent reviews.113,125−136,86,114−125

different bulk properties such as proton and water transport.102,319−322 The progresses that have been recently achieved in these fields could not have been possible without the advent of combined effect of processing and chemistry (Figure 5). To illustrate this aspect, hybrid organic−inorganic membranes were recently synthesized by electrospinning using a sol−gel based solution containing a thermostable polymer PVDF−HFP (polyvinylidenefluoride−hexafluoropropylene), functionalized or not with silicon alkoxides and additives.102 The electrospun membranes are constituted of bundle of fibers surrounded by a functionalized silica network. The bundle of fibers corresponds to the assembly of small polymer fibers surrounded by small anisotropic functionalized silica domains. Proton conduction measurements highlight that these hybrid membranes exhibit conductivity value between 100 and 150 mS cm−1 at 120 °C under 80% RH (relative humidity), value comparable to the best Nafion measured in the same conditions.323 More interestingly, these membranes have a proton conductivity humidity variation close to Nafion and modulus value (75 MPa compared to 2.9 MPa for Nafion at 120 °C) higher than Nafion above 80 °C.



CONCLUSION AND OUTLOOK Materials chemistry is a scientific domain that is still experiencing explosive growth as demonstrated by the large number of high impact journals, books, and international conferences devoted to this large topic. Today, targets and ambitions of most researchers working in the “Chemistry of Materials” field is to produce both fundamental and applied research at the frontier between nanosciences, biology, medicine, energy, and environment by pushing to its limits nanochemistry, together with inorganic or hybrid matter and innovative processing techniques. Their ultimate aim is to achieve a comprehensive view of fundamental synthesis and processing concepts in order to be able to tailor materials by fully controlling all involved mechanistic aspects. A deep understanding of construction mechanisms can only be reached by following in real time via modern spectroscopies and other physical characterization methods (use of synchrotron beamlines, 3D HRTEM associated with EELS, timeresolved TEM, etc.) advanced pulsed field gradient NMR (DOSY), environmental spectroscopic ellipsometry experiments,109,113 etc.)276,278 materials evolution during their chemical construction and processing. Complementary, DFT modelization of materials performed at the different steps of their formation will shed more light on materials structureproperties relationship. Moreover, understanding the funda231

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

nanoscopic materials should be performed in more exotic reaction media such as molten salts, ionic liquids, molten polymers, inorganic gels, or in spatially confined nanoreactors such as mesoporous cavities or immiscible solids. Confinement will not only restrain grain growth in the nanoregime but it might also in some cases trigger formation of metastable or uncommon inorganic phases.143,325,326 These reaction media combined with less conventional stimuli such as microwave heating and ultrasounds are expected to induce the formation of reactive intermediates that might change the reaction pathway and lead to novel structures, morphologies, and inorganic and hybrid assemblies. Following those strategies that could reduce energy consumption with recyclable media or solvent less procedures, a new nanochemistry is emerging. Systematic exploration in nanoscaled systems of phase diagrams or speciation diagrams and order−disorder transitions will open the possibility to obtain nanoscale multinary compositions both in terms of metallic and non metallic (O, F, S, P, B, N, C, etc.) elements, exotic mixed valence nanoscale objects, nanoheterocomposites, nano-objects with complex core−shell or Janus geometries, nanoscopic and micronic powders exhibiting gradients of composition and therefore interesting properties. Such a kind of nanomaterials is needed for the new generation of efficient catalysts (including photocatalysts, electrocatalysts) photocathodes and photoanodes, batteries, and supercapacitors, etc. These chemical approaches to new inorganic nanoscale objects will give access to multifunctional probes and carriers for biomedical imaging and hyperthermia. Combined with organic or/and biological functionalized mesoporous nanospheres, these nano-objects will provide new “protocellular” hybrid materials for bioapplications as active self-healing implants or smart responsive nanomaterials for theranostics. Moreover, tailored hybrid mesoporous nanospheres could allow better compaction and thus better transport of DNA or can be useful for the protection of fragile biocomponents. However, as far as research at nanoscale is concerned, it is important to be engaged with the nanotoxicology community in order to propose efficient and realistic health and safety policies for nanomaterials. One possible strategy will be to use adjustable pocket-sized mesocosms to develop an environmental risk assessment model to assess the nanomaterials life cycle. Alternatively, some multimodal 3D materials such as monolithic foams with hierarchical porous structures can be specifically designed for separation, cascading catalysis or for the study of the coupling between chemical and biochemical catalysts. Hybridized micro- and nanopatterns will provide labon-chip biosensors with interesting answers to fast and off-line biosensing for environmental problematics. The perfect control of the construction of very complex materials will be the next achievements of the chemistry of materials. Following this way, a first step already started concerns the development of integrative synthesis pathways, clearing the track to tailored materials and systems as those found in nature. Elucidating the basic components and building principles selected by evolution can permit to propose more reliable, efficient, and environment-respecting materials. Indeed, materials found in nature combine many wonderful features such as sophistication, miniaturization, hierarchical organizations, hybridization, resistance, and adaptability. Biomimetic or bioinspired approaches are strongly motivating the materials chemistry community. In the close future, original materials will be designed through the synthesis of new hybrid

In this very active context, the emergence of new smart materials has already begun, and its continued growth seems assured due to the economic and technical limitations, which always crop up in current technology. Moreover, societal demands and environmental concerns cannot be ignored any more. Indeed, there is a strong wave of scientific and social thought and determination seeking for a harmonization between nature and human ingenuity. In this vein, materials and systems created by humans must become increasingly recyclable, environmentally friendly, energy efficient, reliable, and inexpensive. To conclude, lets comment shortly a few research lines that in our opinion need to be strengthen in the future. In the real world, materials performances are strongly dependent not only on intrinsic properties but also on devices optimization. Therefore, specific attention should be given to the fabrication by chemists, physicists and engineers in collaboration, of “full-device prototypes”, integrating multiscale materials with different morphologies, microstructures, and compositions. Understanding the properties of the materials, at work, into full devices is very important as their performances are affected by the working conditions (atmosphere, temperature, pH, light, stress, etc.). Indeed, performances of numerous devices (electrochemical-, photelectrochemical-, multiferroics-, thermoelectrics-, or metamaterials-based devices) depend very often on the “quality” of the interfaces. These interfaces can be numerous, and diverse (liquid/solid, solid/solid, solid/gas), and complex (2-dimensional or 3-dimensional, rough, smooth, fractal, etc.). Interfacial zones should be tuned not only by selection of the better physicochemical pathways and functions but also via the development of adapted processing techniques (spray-drying, electrospinning, soft lithography, etc.) and through the coupling between top-down and bottom-up approaches (sol−gel chemistry with nanoimprint lithography or RIE, single nano-objects with spark plasma sintering, etc.). Recent outcomes confirmed that merging these technologies will give access to novel materials or allow to simplify production protocols. Electrospinned membranes from hybrid colloidal dispersions exhibit better conduction, better mechanical properties, better water management, and less swelling than commercially available best NAFIONs.102 These approaches combined with templated growth will lead to intelligent permselective hybrid membranes having hierarchical structures and that can couple for example separation and catalytic processes. Because it allows performing thermal treatments with minimum grain growth and interdiffusion, Spark-Plasma Sintering will permit to produce original monoliths built from nanoparticles143,324 of different chemical compositions (p-type and n-type oxides, metallic and semiconductor metal oxides, metal boride, and metal oxides, etc.). Multiscale casting (at the nano-, meso-, and macro-scales) can be envisaged through synthetic (foams) or biological templates (butterfly wings, diatoms, etc.). Multimodal 3D hybrid nanocomposites can be designed via two photon absorption lithography.122,123 Multilayer liquid deposition (spray coating) of dispersions containing colloids of different sizes can be combined with top-down derived microor mesopatterns, made by example through RIE or lithography.105,114 Innovative chemical approaches should also be more developed. Water or common nontoxic organic solvents with low VOC are involved in most synthetic industrial pathways. However, basic research devoted to the synthesis of new 232

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

(2) livage, J.; Henry, M.; Sanchez, C. Prog. Solid State Chem. 1988, 18, 259. (3) Yanagisawa, T.; Shimizu, T.; Kuroda, K.; Kato, C. Bull. Chem. Soc. Jpn. 1990, 63, 988. (4) Schmidt, W. R.; Interrante, L. V.; Doremus, R. H.; Trout, T. K.; Marchetti, P. S.; Maciel, G. E. Chem. Mater. 1991, 3, 257. (5) Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Nature 1992, 359, 710. (6) Sanchez, C.; Ribot, F. New J. Chem. 1994, 18, 1007. (7) Schubert, U.; Huesing, N.; Lorenz, A. Chem. Mater. 1995, 7, 2010. (8) Loy, D. A.; Shea, K. J. Chem. Rev. 1995, 95, 1431. (9) Mann, S.; Burkett, S. L.; Davis, S. A.; Fowler, C. E.; Mendelson, N. H.; Sims, S. D.; Walsh, D.; Whilton, N. T. Chem. Mater. 1997, 9, 2300. (10) Levy, D. Chem. Mater. 1997, 9, 2666. (11) Corma, A. Chem. Rev. 1997, 97, 2373. (12) Goltner, C. G.; Antonietti, M. Adv. Mater. 1997, 9, 431. (13) Moller, K.; Bein, T. Chem. Mater. 1998, 10, 2950. (14) Corriu, R. C. R. Chim. 1998, 1, 83. (15) Forster, S.; Antonietti, M. Adv. Mater. 1998, 10, 195. (16) Antonietti, M.; Berton, B.; Goltner, C.; Hentze, H. P. Adv. Mater. 1998, 10, 154. (17) Yaghi, O. M.; Li, H. L.; Davis, C.; Richardson, D.; Groy, T. L. Acc. Chem. Res. 1998, 31, 474. (18) Zhao, D. Y.; Feng, J. L.; Huo, Q. S.; Melosh, N.; Fredrickson, G. H.; Chmelka, B. F.; Stucky, G. D. Science 1998, 279, 548. (19) Yang, P.; Deng, T.; Zhao, D.; Feng, P.; Pine, D.; Chmelka, B. F.; Whitesides, G. M.; Stucky, G. D. Science 1998, 282, 2244. (20) Brinker, C. J.; Lu, Y. F.; Sellinger, A.; Fan, H. Y. Adv. Mater. 1999, 11, 579. (21) Ozin, G. A. Chem. Commun. 2000, 419. (22) Sanchez, C.; Soler-Illia, G. J. A. A.; Ribot, F.; Lalot, T.; Mayer, C. R.; Cabuil, V. Chem. Mater. 2001, 13, 3061. (23) Schüth, F. Chem. Mater. 2001, 13, 3184. (24) Stein, A. Microporous Mesoporous Mater. 2001, 44, 227. (25) Ferey, G. Chem. Mater. 2001, 13, 3084. (26) Ryoo, R.; Joo, S. H.; Kruk, M.; Jaroniec, M. Adv. Mater. 2001, 13, 677. (27) Soler-Illia, G. J. A. A.; Sanchez, C.; Lebeau, B.; Patarin, J. Chem. Rev. 2002, 102, 4093. (28) Corma, A.; Garcia, H. Chem. Rev. 2003, 103, 4307. (29) Colfen, H.; Mann, S. Angew. Chem., Int. Ed. 2003, 42, 2350. (30) Parak, W. J.; Gerion, D.; Pellegrino, T.; Zanchet, D.; Micheel, C.; Williams, S. C.; Boudreau, R.; Le Gros, M. A.; Larabell, C. A.; Alivisatos, A. P. Nanotechnol. 2003, 14, R15. (31) Kitagawa, S.; Kitaura, R.; Noro, S. Angew. Chem., Int. Ed. 2004, 43, 2334. (32) Antonietti, M.; Ozin, G. A. Chem.−Eur. J. 2004, 10, 28. (33) Yin, Y.; Alivisatos, A. P. Nature 2005, 437, 664. (34) Colfen, H.; Antonietti, M. Angew. Chem., Int. Ed. 2005, 44, 5576. (35) Sanchez, C.; Arribart, H.; Giraud-Guille, M. M. Nat. Mater. 2005, 4, 277. (36) Niederberger, M.; Garnweitner, G. Chem.−Eur. J. 2006, 12, 7282. (37) Jun, Y. W.; Choi, J. S.; Cheon, J. Angew. Chem., Int. Ed. 2006, 45, 3414. (38) Kalsin, A. M.; Fialkowski, M.; Paszewski, M.; Smoukov, S. K.; Bishop, K. J. M.; Grzybowski, B. A. Science 2006, 312, 420. (39) Avnir, D.; Coradin, T.; Lev, O.; Livage, J. J. Mater. Chem. 2006, 16, 1013. (40) Hoffmann, F.; Cornelius, M.; Morell, J.; Froeba, M. Angew. Chem., Int. Ed. 2006, 45, 3216. (41) Park, J.; Joo, J.; Kwon, S. G.; Jang, Y.; Hyeon, T. Angew. Chem., Int. Ed. 2007, 46, 4630. (42) Niederberger, M. Acc. Chem. Res. 2007, 40, 793. (43) Vallet-Regi, M.; Balas, F.; Arcos, D. Angew. Chem., Int. Ed. 2007, 46, 7548.

nanosynthons (“hybridons”), selectively tagged with complementary connectivities, allowing for the coding of hybrid assemblies. These strategies, based on a more precise encoding of chemical information, are in the process of giving birth to a new “vectorial chemistry” allowing the assembly of various building blocks (molecules, polymers, nanoparticles, clusters, nanocomposites) into increasingly complex hierarchical and functional architectures. Moreover, the synthesis of materials using in synergy, self-assembly processes and chemical transformations in spatially restricted reaction f ields coupled with external solicitations such as gravity, electrical, magnetic, fields, mechanical stress will allow to generate materials with new compositions and structures and complex morphologies. Last but not least, the use of strong compositional flux variations of the reagents during the synthesis (operating open systems) combined with new synthetic approaches based on dynamic interactive systems is very promising. Indeed, dynamic selection327,328 can allow a given set of nano- or meso-objects reversibly exchanging and continuously organizing at the nanoor mesoscopic levels, to build self-optimized hybrid networks. This sort of biomimetic mode in assembling materials allows spatial/temporal and structural/functional adaptability in response to internal constitutional or to external stimulant factors. In conclusion, numerous scientific breakthroughs can be expected in the chemistry of materials field through a stronger implication of skilled chemists into innovative chemical approaches and original pathways of materials processing. Outlooking to the middle of the 21st century, the perspectives and expected developments of the Chemistry of Materials are immense. For sure, it will not only continue to contribute to a high level of scientific and technological developments but also it allows us to dream. Will we be able, in a close future, to build smart materials, which simultaneously respond to external stimuli (i.e., pH, solvent, light, external fields, temperature, etc.), adapt to their environment, self-replicate, self-repair, or self-destroy to form their former precursor elements at the end of their fleeting lives?



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]; [email protected]. Author Contributions

The manuscript was written through contributions of all authors. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors gratefully thank Pr. Jean-Pierre Jolivet and Pr. Jacques Livage for their remarkable contribution over the years to the fields of “chimie douce” and nanostructured materials. The many students and postdoctoral fellows, who participated to the work presented in this review, are also warmly thanked. UPMC, CNRS and the College de France are acknowledged for the financial supports provided to the Hybrid and Nanomaterials team of CMCP.



REFERENCES

(1) Interrante, L. V. Chem. Mater. 1989, 1, 1. 233

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

(86) Ferey, G. Dalton Trans. 2009, 4400. (87) Ferey, G.; Serre, C.; Devic, T.; Maurin, G.; Jobic, H.; Llewellyn, P.; De Weireld, G.; Vimont, A.; Daturi, M.; Chang, J. Chem. Soc. Rev. 2011, 40 (2), 550. (88) Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F. Acc. Chem. Res. 2005, 38, 217. (89) Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surble, S.; Margiolaki, I. Science 2005, 309, 2040. (90) Ferey, G.; Serre, C. Chem. Soc. Rev. 2009, 38, 1380. (91) Cheetham, A. K.; Rao, C. N. R. Mrs Bull. 2005, 30, 93. (92) Kitagawa, S.; Uemura, K. Chem. Soc. Rev. 2005, 34, 109. (93) Uemura, T.; Yanai, N.; Kitagawa, S. Chem. Soc. Rev. 2009, 38, 1228. (94) Ferey, G.; Serre, C.; Devic, T.; Maurin, G.; Jobic, H.; Llewellyn, P. L.; De Weireld, G.; Vimont, A.; Daturi, M.; Chang, J.-S. Chem. Soc. Rev. 2011, 40, 550. (95) Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Muller, J.-O.; Schlogl, R.; Carlsson, J. M. J. Mater. Chem. 2008, 18, 4893. (96) Stein, A.; Melde, B. J.; Schroden, R. C. Adv. Mater. 2000, 12, 1403. (97) van Bommel, K. J. C.; Friggeri, A.; Shinkai, S. Angew. Chem., Int. Ed. 2003, 42, 980. (98) Kuroda, K.; Shimoji, A.; Kawahara, K.; Wakabayashi, R.; Tamura, Y.; Asakura, Y.; Kitahara, M. Chem. Mater. 2013, accepted. (99) Wang, W. D.; Lofgreen, J. E.; Ozin, G. A. Small 2010, 6, 2634. (100) Kameneva, O.; Kuznestov, A. I.; Smirnova, L. A.; Rozes, L.; Sanchez, C.; Alexandrov, A.; Bityurin, N.; Chhor, K.; Kanaev, A. J. Mater. Chem. 2005, 15, 3380. (101) Brun, N.; Garcia, A. B.; Deleuze, H.; Achard, M. F.; Sanchez, C.; Durand, F.; Oestreicher, V.; Backov, R. Chem. Mater. 2010, 22, 4555. (102) Maneeratana, V.; Bass, J. D.; Azais, T.; Patissier, A.; Valle, K.; Maréchal, M.; Gébel, G.; Laberty-Robert, C.; Sanchez, C. Adv. Funct. Mater. 2013, 23 (22), 2872. (103) Carretero-Genevrier, A.; Gich, M.; Picas, L.; Gazquez, J.; Drisko, G. L.; Boissiere, C.; Grosso, D.; Rodriguez-Carvajal, J.; Sanchez, C. Science 2013, 340, 827. (104) Faustini, M.; Vayer, M.; Marmiroli, B.; Hillmyer, M.; Amenitsch, H.; Sinturel, C.; Grosso, D. Chem. Mater. 2010, 22, 5687. (105) Faustini, M.; Drisko, G. L.; Letailleur, A. A.; Montiel, R. S.; Boissiere, C.; Cattoni, A.; Haghiri-Gosnet, A. M.; Lerondel, G.; Grosso, D. Nanoscale 2013, 5, 984. (106) Fousseret, B.; Mougenot, M.; Rossignol, F.; Baumard, J. F.; Soulestin, B.; Boissiere, C.; Ribot, F.; Jalabert, D.; Carrion, C.; Sanchez, C.; Lejeune, M. Chem. Mater. 2010, 22, 3875. (107) De Los Cobos, O.; Fousseret, B.; Lejeune, M.; Rossignol, F.; Dutreilh-Colas, M.; Carrion, C.; Boissière, C.; Ribot, F.; Sanchez, C.; Cattoën, X.; Wong Chi Man, M.; Durand, J.-O. Chem. Mater. 2012, 24, 4337. (108) Kuemmel, M.; Allouche, J.; Nicole, L.; Boissière, C.; Laberty, C.; Amenitsch, H.; Sanchez, C.; Grosso, D. Chem. Mater. 2007, 19, 3717. (109) Boissiere, C.; Grosso, D.; Chaumonnot, A.; Nicole, L.; Sanchez, C. Adv. Mater. 2011, 23, 599. (110) Corriu, R. J. P.; Mehdi, A.; Reye, C.; Thieuleux, C. New J. Chem. 2003, 27, 905. (111) Fornasieri, G.; Rozes, L.; Le Calve, S.; Alonso, B.; Massiot, D.; Rager, M. N.; Evain, M.; Boubekeur, K.; Sanchez, C. J. Am. Chem. Soc. 2005, 127, 4869. (112) Steunou, N.; Forster, S.; Florian, P.; Sanchez, C.; Antonietti, M. J. Mater. Chem. 2002, 12, 3426. (113) Aegerter, M. A.; Mennig, M. Sol-Gel Technologies for Glass Producers and Users; Kluwer Academic Publishers: Boston, 2004. (114) Sanchez, C.; Boissière, C.; Grosso, D.; Laberty, C.; Nicole, L. Chem. Mater. 2008, 20, 682. (115) Matsuzaki, K.; Arai, D.; Taneda, N.; Mukaiyama, T.; Ikemura, M. J. Non-Cryst. Solids 1989, 112, 437. (116) Li, D.; Xia, Y. N. Adv. Mater. 2004, 16, 1151.

(44) Robinson, R. D.; Sadtler, B.; Demchenko, D. O.; Erdonmez, C. K.; Wang, L.-W.; Alivisatos, A. P. Science 2007, 317, 355. (45) Meldrum, F. C.; Coelfen, H. Chem. Rev. 2008, 108, 4332. (46) Skrabalak, S. E.; Chen, J. Y.; Sun, Y. G.; Lu, X. M.; Au, L.; Cobley, C. M.; Xia, Y. N. Acc. Chem. Res. 2008, 41, 1587. (47) Wan, Y.; Shi, Y.; Zhao, D. Chem. Mater. 2008, 20, 932. (48) Mutin, P. H.; Vioux, A. Chem. Mater. 2009, 21, 582. (49) Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J. M.; Domen, K.; Antonietti, M. Nat. Mater. 2009, 8, 76. (50) Coti, K. K.; Belowich, M. E.; Liong, M.; Ambrogio, M. W.; Lau, Y. A.; Khatib, H. A.; Zink, J. I.; Khashab, N. M.; Stoddart, J. F. Nanoscale 2009, 1, 16. (51) Ruiz-Hitzky, E.; Darder, M.; Aranda, P.; Ariga, K. Adv. Mater. 2010, 22, 323. (52) Grzelczak, M.; Vermant, J.; Furst, E. M.; Liz-Marzan, L. M. ACS Nano 2010, 4, 3591. (53) Ye, X.; Collins, J. E.; Kang, Y.; Chen, J.; Chen, D. T. N.; Yodh, A. G.; Murray, C. B. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 22430. (54) Deka, S.; Miszta, K.; Dorfs, D.; Genovese, A.; Bertoni, G.; Manna, L. Nano Lett. 2010, 10, 3770. (55) González, E.; Arbiol, J.; Puntes, V. F. Science 2011, 334, 1377. (56) Carbone, L.; Cozzoli, P. D. Nano Today 2010, 5, 449. (57) Miszta, K.; de Graaf, J.; Bertoni, G.; Dorfs, D.; Brescia, R.; Marras, S.; Ceseracciu, L.; Cingolani, R.; van Roij, R.; Dijkstra, M.; Manna, L. Nat. Mater. 2011, 10, 872. (58) Shi, Y.; Wan, Y.; Zhao, D. Chem. Soc. Rev. 2011, 40, 3854. (59) Duguet, E.; Desert, A.; Perro, A.; Ravaine, S. Chem. Soc. Rev. 2011, 40, 941. (60) Nassif, N.; Livage, J. Chem. Soc. Rev. 2011, 40, 849. (61) Antonietti, M.; Fechler, N.; Fellinger, T.-P. Chem. Mater. 2013, DOI: 10.1021/cm402239e. (62) Chung, I.; Kanatzidis, M. G. Chem. Mater. 2013, DOI: 10.1021/ cm401737s. (63) Davis, M. E. Chem. Mater. 2013, DOI: 10.1021/cm401914u. (64) Rousse, G.; Tarascon, J.-M. Chem. Mater. 2013, DOI: 10.1021/ cm4022358. (65) Prouzet, E.; Ravaine, S.; Sanchez, C.; Backov, R. New J. Chem. 2008, 32, 1284. (66) Brun, N.; Ungureanu, S.; Deleuze, H.; Backov, R. Chem. Soc. Rev. 2011, 40, 771. (67) Backov, R. Soft Matter 2006, 2, 452. (68) Baghbanzadeh, M.; Carbone, L.; Cozzoli, P. D.; Kappe, C. O. Angew. Chem., Int. Ed. 2011, 50, 11312. (69) Bilecka, I.; Niederberger, M. Nanoscale 2010, 2, 1358. (70) Giordano, C.; Antonietti, M. Nano Today 2011, 6, 366. (71) Kim, J. Y.; Voznyy, O.; Zhitomirsky, D.; Sargent, E. H. Adv. Mater. 2013, n/a. (72) Rogez, G.; Massobrio, C.; Rabu, P.; Drillon, M. Chem. Soc. Rev. 2011, 40, 1031. (73) Ruiz-Hitzky, E.; Aranda, P.; Darder, M.; Ogawa, M. Chem. Soc. Rev. 2011, 40, 801. (74) Schubert, U. Chem. Soc. Rev. 2011, 40, 575. (75) Boury, B.; Corriu, R. J. P. Chem. Commun. 2002, 795. (76) Debecker, D. P.; Mutin, P. H. Chem. Soc. Rev. 2012, 41, 3624. (77) Lebeau, B.; Innocenzi, P. Chem. Soc. Rev. 2011, 40, 886. (78) Yaghi, O. M.; M, O. K.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Nature 2003, 423, 705. (79) Ockwig, N. W.; Delgado-Friedrichs, O.; O’Keeffe, M.; Yaghi, O. M. Acc. Chem. Res. 2005, 38, 176. (80) Tranchemontagne, D. J.; Mendoza-Cortes, J. L.; O’Keeffe, M.; Yaghi, O. M. Chem. Soc. Rev. 2009, 38, 1257. (81) Chae, H. K.; Siberio-Perez, D. Y.; Kim, J.; Go, Y.; Eddaoudi, M.; Matzger, A. J.; O’Keeffe, M.; Yaghi, O. M. Nature 2004, 427, 523. (82) Furukawa, H.; Miller, M. A.; Yaghi, O. M. J. Mater. Chem. 2007, 17, 3197. (83) Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998. (84) Ferey, G. Nature 2005, 436, 187. (85) Ferey, G. Chem. Soc. Rev. 2008, 37, 191. 234

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

(117) Madhugiri, S.; Zhou, W.; Ferraris, J. P.; Balkus, K. J. Microporous Mesoporous Mater. 2003, 63, 75. (118) Macias, M.; Chacko, A.; Ferraris, J. P.; Balkus, K. J. Microporous Mesoporous Mater. 2005, 86, 1. (119) Innocenzi, P.; Kidchob, T.; Falcaro, P.; Takahashi, M. Chem. Mater. 2008, 20, 607. (120) Xia, Y. N.; Whitesides, G. M. Angew. Chem., Int. Ed. 1998, 37, 551. (121) Gates, B. D.; Xu, Q.; Stewart, M.; Ryan, D.; Willson, C. G.; Whitesides, G. M. Chem. Rev. 2005, 105, 1171. (122) Houbertz, R.; Frohlich, L.; Popall, M.; Streppel, U.; Dannberg, P.; Brauer, A.; Serbin, J.; Chichkov, B. N. Adv. Eng. Mater. 2003, 5, 551. (123) Houbertz, R. Appl. Surf. Sci. 2005, 247, 504. (124) Okuyama, K.; Abdullah, M.; Lenggoro, I. W.; Iskandar, F. Adv. Powder Technol. 2006, 17, 587. (125) Sanchez, C.; Julian, B.; Belleville, P.; Popall, M. J. Mater. Chem. 2005, 15, 3559. (126) Nicole, L.; Rozes, L.; Sanchez, C. Adv. Mater. 2010, 22, 3208. (127) Schottner, G.; Kron, J.; Deichmann, A. J. Sol-Gel Sci. Technol. 1998, 13, 183. (128) Schottner, G.; Rose, K.; Posset, U. J. Sol-Gel Sci. Technol. 2003, 27, 71. (129) Schmidt, H. J. Sol-Gel Sci. Technol. 2006, 40, 115. (130) Arkles, B. MRS Bull. 2001, 26, 402. (131) Mahltig, B.; Haufe, H.; Bottcher, H. J. Mater. Chem. 2005, 15, 4385. (132) Mahltig, B.; Swaboda, C.; Roessler, A.; Bottcher, H. J. Mater. Chem. 2008, 18, 3180. (133) Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; SchierleArndt, K.; Pastre, J. J. Mater. Chem. 2006, 16, 626. (134) Czaja, A. U.; Trukhan, N.; Muller, U. Chem. Soc. Rev. 2009, 38, 1284. (135) Sanchez, C.; Belleville, P.; Popall, M.; Nicole, L. Chem. Soc. Rev. 2011, 40, 696. (136) Schottner, G. Chem. Mater. 2001, 13, 3422. (137) Jolivet, J.-P. Metal Oxide Chemistry and Synthesis: From Solution to Solid State; Wiley: Chichester, 2000. (138) Burda, C.; Chen, X. B.; Narayanan, R.; El-Sayed, M. A. Chem. Rev. 2005, 105, 1025. (139) Niederberger, M.; Pinna, N. Metal Oxide Nanoparticles in Organic Solvents; Springer: London, 2009. (140) Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Angew. Chem., Int. Ed. 2009, 48, 60. (141) Jolivet, J.-P.; Cassaignon, S.; Chanéac, C.; Chiche, D.; Durupthy, O.; Portehault, D. C. R. Chim. 2010, 13, 40. (142) Portehault, D.; Devi, S.; Beaunier, P.; Gervais, C.; Giordano, C.; Sanchez, C.; Antonietti, M. Angew. Chem., Int. Ed. 2011, 50, 3262. (143) Portehault, D.; Maneeratana, V.; Candolfi, C.; Oeschler, N.; Veremchuk, I.; Grin, Y.; Sanchez, C.; Antonietti, M. ACS Nano 2011, 5, 9052. (144) Saison, T.; Chemin, N.; Chaneac, C.; Durupthy, O.; Ruaux, V.; Mariey, L.; Mauge, F.; Beaunier, P.; Jolivet, J.-P. J. Phys. Chem. C 2011, 115, 5657. (145) Sassoye, C.; Muller, G.; Debecker, D. P.; Karelovic, A.; Cassaignon, S.; Pizarro, C.; Ruiz, P.; Sanchez, C. Green Chem. 2011, 13, 3230. (146) Ali Ahmad, M.; Prelot, B.; Razafitianamaharavo, A.; Douillard, J. M.; Zajac, J.; Dufour, F.; Durupthy, O.; Chaneac, C.; Villieras, F. J. Phys. Chem. C 2012, 116, 24596. (147) Carenco, S.; Portehault, D.; Boissière, C.; Mézailles, N.; Sanchez, C. Chem. Rev. 2013, DOI: 10.1021/cr400020d. (148) Ali Ahmad, M.; Prelot, B.; Dufour, F.; Durupthy, O.; Razafitianamaharavo, A.; Douillard, J. M.; Chaneac, C.; Villieras, F.; Zajac, J. J. Phys. Chem. C 2013, 117, 4459. (149) Fernandez-Garcia, M.; Martinez-Arias, A.; Hanson, J. C.; Rodriguez, J. A. Chem. Rev. 2004, 104, 4063. (150) Jana, N. R.; Chen, Y.; Peng, X. Chem. Mater. 2004, 16, 3931.

(151) Jolivet, J.-P.; Froidefond, C.; Pottier, A.; Chanéac, C.; Cassaignon, S.; Tronc, E.; Euzen, P. J. Mater. Chem. 2004, 14, 3281. (152) Chen, X.; Mao, S. S. Chem. Rev. 2007, 107, 2891. (153) Magne, C.; Cassaignon, S.; Lancel, G.; Pauporté, T. ChemPhysChem 2011, 12, 2461. (154) Okal, J.; Zawadzki, M.; Tylus, W. Appl. Catal. 2011, 101, 548. (155) Han, X.; Kuang, Q.; Jin, M.; Xie, Z.; Zheng, L. J. Am. Chem. Soc. 2009, 131, 3152. (156) Yang, H. G.; Sun, C. H.; Qiao, S. Z.; Zou, J.; Liu, G.; Smith, S. C.; Cheng, H. M.; Lu, G. Q. Nature 2008, 453, 638. (157) Liao, D. L.; Liao, B. Q. J. Photochem. Photobiol., A 2007, 187, 363. (158) Magne, C.; Dufour, F.; Labat, F.; Lancel, G.; Durupthy, O.; Cassaignon, S.; Pauporté, T. J. Photochem. Photobiol., A 2012, 232, 22. (159) Dessombz, A.; Chiche, D.; Davidson, P.; Panine, P.; Chanéac, C.; Jolivet, J.-P. J. Am. Chem. Soc. 2007, 129, 5904. (160) Xu, X.; Zhai, T.; Shao, M.; Huang, J. Phys. Chem. Chem. Phys. 2012, 14, 16371. (161) Kinadjian, N.; Le Bechec, M.; Pigot, T.; Dufour, F.; Durupthy, O.; Bentaleb, A.; Prouzet, E.; Lacombe, S.; Backov, R. Eur. J. Inorg. Chem. 2012, 2012, 5350. (162) Lv, B.; Liu, Z.; Tian, H.; Xu, Y.; Wu, D.; Sun, Y. Adv. Funct. Mater. 2010, 20, 3987. (163) Van, T. K.; Cha, H. G.; Nguyen, C. K.; Kim, S. W.; Jung, M. H.; Kang, Y. S. Cryst. Growth Des. 2012, 12, 862. (164) Rodriguez, R. D.; Demaille, D.; Lacaze, E.; Jupille, J.; Chanéac, C.; Jolivet, J.-P. J. Phys. Chem. C 2007, 111, 16866. (165) Chiche, D.; Chizallet, C.; Durupthy, O.; Chanéac, C.; Revel, R.; Raybaud, P.; Jolivet, J.-P. Phys. Chem. Chem. Phys. 2009, 11, 11310. (166) Chiche, D.; Chaneac, C.; Revel, R.; Jolivet, J.-P. Phys. Chem. Chem. Phys. 2011, 13, 6241. (167) Cassaignon, S.; Koelsch, M.; Jolivet, J.-P. J. Phys. Chem. Solids 2007, 68, 695. (168) Portehault, D.; Cassaignon, S.; Baudrin, E.; Jolivet, J.-P. J. Mater. Chem. 2009, 19, 2407. (169) Portehault, D.; Cassaignon, S.; Baudrin, E.; Jolivet, J.-P. Cryst. Growth Des. 2010, 10, 2168. (170) Portehault, D.; Cassaignon, S.; Nassif, N.; Baudrin, E.; Jolivet, J.-P. Angew. Chem., Int. Ed. 2008, 47, 6441. (171) Portehault, D.; Cassaignon, S.; Baudrin, E.; Jolivet, J.-P. Chem. Commun. 2009, 674. (172) Portehault, D.; Cassaignon, S.; Baudrin, E.; Jolivet, J.-P. Chem. Mater. 2007, 19, 5410. (173) Portehault, D.; Cassaignon, S.; Baudrin, E.; Jolivet, J.-P. Cryst. Growth Des. 2009, 9, 2562. (174) Mosset, A.; Gautier-Luneau, I.; Galy, J.; Strehlow, P.; Schmidt, H. J. Non-Cryst. Solids 1988, 100, 339. (175) Mialon, G.; Gohin, M.; Gacoin, T.; Boilot, J.-P. ACS Nano 2012, 2, 2505. (176) le Masne de Chermont, Q.; Chanéac, C.; Seguin, J.; Pellé, F.; Maitrejean, S.; Jolivet, J.-P.; Gourier, D.; Bessodes, M.; Scherman, D. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 9266. (177) Bessiere, A.; Lecointre, A.; Priolkar, K. R.; Gourier, D. J. Mater. Chem. 2012, 22, 19039. (178) Rosticher, C.; Chaneac, C.; Viana, B.; Bessiere, A. Luminescence 2012, 27, 157. (179) Maldiney, T.; Lecointre, A.; Viana, B.; Bessiere, A.; Bessodes, M.; Gourier, D.; Richard, C.; Scherman, D. J. Am. Chem. Soc. 2011, 133, 11810. (180) Carenco, S.; Demange, M.; Shi, J.; Boissière, C.; Sanchez, C.; Le Floch, P.; Mézailles, N. Chem. Commun. 2010, 46, 5578. (181) Carenco, S.; Le Goff, X. F.; Shi, J.; Roiban, L.; Ersen, O.; Boissière, C.; Sanchez, C.; Mézailles, N. Chem. Mater. 2011, 23, 2270. (182) Carenco, S.; Resa, I.; Le Goff, X.; Le Floch, P.; Mézailles, N. Chem. Commun. 2008, 2568. (183) Lei, W.; Portehault, D.; Dimova, R.; Antonietti, M. J. Am. Chem. Soc. 2011, 133, 7121. (184) Lei, W.; Portehault, D.; Liu, D.; Qin, S.; Chen, Y. Nat. Comm. 2013, 4, 1777. 235

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

(185) Bojdys, M. J.; Müller, J.-O.; Antonietti, M.; Thomas, A. Chem.−Eur. J. 2008, 14, 8177. (186) Lei, W.; Qin, S.; Liu, D.; Portehault, D.; Liu, Z.; Chen, Y. Chem. Commun. 2013, 49, 352. (187) Liu, X.; Antonietti, M.; Giordano, C. Chem. Mater. 2013, DOI: 10.1021/cm303727g. (188) Liu, X.; Giordano, C.; Antonietti, M. J. Mater. Chem. 2012, 22, 5454. (189) Carenco, S.; Leyva-Pérez, A.; Concepción, P.; Boissière, C.; Mézailles, N.; Sanchez, C.; Corma, A. Nano Today 2012, 7, 21. (190) Carenco, S.; Surcin, C.; Morcrette, M.; Larcher, D.; Mézailles, N.; Boissière, C.; Sanchez, C. Chem. Mater. 2012, 24, 688. (191) Harada, S.; Tanaka, K.; Inui, H. J. Appl. Phys. 2010, 108, 083703. (192) Kwon, D.-H.; Kim, K. M.; Jang, J. H.; Jeon, J. M.; Lee, M. H.; Kim, G. H.; Li, X.-S.; Park, G.-S.; Lee, B.; Han, S.; Kim, M.; Hwang, C. S. Nat. Nanotechnol. 2010, 5, 148. (193) Ohkoshi, S.-I.; Tsunobuchi, Y.; Matsuda, T.; Hashimoto, K.; Namai, A.; Hakoe, F.; Tokoro, H. Nat. Chem. 2010, 2, 539. (194) Innocenzi, P.; Kidchob, T.; Falcaro, P.; Takahashi, M. Chem. Mater. 2007, 20, 607. (195) Soler-Illia, G. J. A. A.; Innocenzi, P. Chem.−Eur. J. 2006, 12, 4478. (196) Davis, M. E. Nature 2002, 417, 813. (197) Moliner, M.; Martínez, C.; Corma, A. Chem. Mater. 2013, DOI: 10.1021/cm4015095. (198) Na, K.; Jo, C.; Kim, J.; Cho, K.; Jung, J.; Seo, Y.; Messinger, R. J.; Chmelka, B. F.; Ryoo, R. Science 2011, 333, 328. (199) Pega, S.; Boissière, C.; Grosso, D.; Azaïs, T.; Chaumonnot, A.; Sanchez, C. Angew. Chem., Int. Ed. 2009, 48, 2784. (200) Valtchev, V.; Majano, G.; Mintova, S.; Perez-Ramirez, J. Chem. Soc. Rev. 2013, 42, 263. (201) Valtchev, V.; Tosheva, L. Chem. Rev. 2013, 113, 6734. (202) Sakatani, Y.; Boissière, C.; Grosso, D.; Nicole, L.; Soler-Illia, G. J. A. A.; Sanchez, C. Chem. Mater. 2007, 20, 1049. (203) Kuemmel, M.; Grosso, D.; Boissière, C.; Smarsly, B.; Brezesinski, T.; Albouy, P. A.; Amenitsch, H.; Sanchez, C. Angew. Chem., Int. Ed. 2005, 44, 4589. (204) Soler-Illia, G. J. A. A.; Crepaldi, E. L.; Grosso, D.; Sanchez, C. Curr. Opin. Colloid Interface Sci. 2003, 8, 109. (205) Holland, B. T.; Blanford, C. F.; Stein, A. Science 1998, 281, 538. (206) Velev, O. D.; Jede, T. A.; Lobo, R. F.; Lenhoff, A. M. Nature 1997, 389, 447. (207) Arsenault, A. C.; Clark, T. J.; von Freymann, G.; Cademartiri, L.; Sapienza, R.; Bertolotti, J.; Vekris, E.; Wong, S.; Kitaev, V.; Manners, I.; Wang, R. Z.; John, S.; Wiersma, D.; Ozin, G. A. Nat. Mater. 2006, 5, 179. (208) Llusar, M.; Roux, C.; Pozzo, J. L.; Sanchez, C. J. Mater. Chem. 2003, 13, 442. (209) Jiang, X.; Brinker, C. J. J. Am. Chem. Soc. 2006, 128, 4512. (210) Malfatti, L.; Falcaro, P.; Marongiu, D.; Casula, M. F.; Amenitsch, H.; Innocenzi, P. Chem. Mater. 2009, 21, 4846. (211) Nakanishi, K. J. Porous Mater. 1997, 4, 67. (212) Nakanishi, K.; Tanaka, N. Acc. Chem. Res. 2007, 40, 863. (213) Bunz, U. H. F. Adv. Mater. 2006, 18, 973. (214) Faustini, M.; Louis, B.; Albouy, P. A.; Kuemmel, M.; Grosso, D. J. Phys. Chem. C 2010, 114, 7637. (215) Grosso, D. J. Mater. Chem. 2011, 21, 17033. (216) Krins, N.; Faustini, M.; Louis, B.; Grosso, D. Chem. Mater. 2010, 22, 6218. (217) Angelomé, P. C.; Fuertes, M. C.; Soler-Illia, G. J. A. A. Adv. Mater. 2006, 18, 2397. (218) Innocenzi, P.; Malfatti, L.; Soler-Illia, G. J. A. A. Chem. Mater. 2011, 23, 2501. (219) Brinker, C. J.; Hurd, A. J.; Schunk, P. R.; Frye, G. C.; Ashley, C. S. J. Non-Cryst. Solids 1992, 147−148, 424. (220) Nicole, L.; Boissiere, C.; Grosso, D.; Hesemann, P.; Moreau, J.; Sanchez, C. Chem. Commun. 2004, 0, 2312.

(221) Banet, P.; Legagneux, L.; Hesemann, P.; Moreau, J. J. E.; Nicole, L.; Quach, A.; Sanchez, C.; Tran-Thi, T. H. Sens. Actuators, B 2008, 130, 1. (222) Soler-Illia, G. J. A. A.; Angelome, P. C.; Fuertes, M. C.; Grosso, D.; Boissiere, C. Nanoscale 2012, 4, 2549. (223) Pidol, L.; Grosso, D.; Soler-Illia, G. J. A. A.; Crepaldi, E. L.; Sanchez, C.; Albouy, P. A.; Amenitsch, H.; Euzen, P. J. Mater. Chem. 2002, 12, 557. (224) Grosso, D.; Boissiere, C.; Smarsly, B.; Brezesinski, T.; Pinna, N.; Albouy, P. A.; Amenitsch, H.; Antonietti, M.; Sanchez, C. Nat. Mater. 2004, 3, 787. (225) Lepoutre, S.; Julian-Lopez, B.; Sanchez, C.; Amenitsch, H.; Linden, M.; Grosso, D. J. Mater. Chem. 2010, 20, 537. (226) Ferreira, P.; Hou, R. Z.; Wu, A.; Willinger, M.-G.; Vilarinho, P. M.; Mosa, J.; Laberty-Robert, C.; Boissière, C.; Grosso, D.; Sanchez, C. Langmuir 2012, 28, 2944. (227) Sassoye, C.; Laberty, C.; Le Khanh, H.; Cassaignon, S.; Boissière, C.; Antonietti, M.; Sanchez, C. Adv. Funct. Mater. 2009, 19, 1922. (228) Brezesinski, T.; Wang, J.; Tolbert, S. H.; Dunn, B. Nat. Mater. 2010, 9, 146. (229) Brezesinski, T.; Wang, J.; Polleux, J.; Dunn, B.; Tolbert, S. H. J. Am. Chem. Soc. 2009, 131, 1802. (230) Fisher, A.; Kuemmel, M.; Järn, M.; Linden, M.; Boissière, C.; Nicole, L.; Sanchez, C.; Grosso, D. Small 2006, 2, 569. (231) Laberty-Robert, C.; Kuemmel, M.; Allouche, J.; Boissiere, C.; Nicole, L.; Grosso, D.; Sanchez, C. J. Mater. Chem. 2008, 18, 1216. (232) Fontaine, O.; Laberty-Robert, C.; Sanchez, C. Langmuir 2012, 28, 3650. (233) Lockwood, D. J.; Rowell, N. L.; Berbezier, I.; Amiard, G.; Ronda, A.; Faustini, M.; Grosso, D. J. Electrochem. Soc. 2010, 157, H1160. (234) Allouche, J.; Lantiat, D.; Kuemmel, M.; Faustini, M.; Laberty, C.; Chanéac, C.; Tronc, E.; Boissière, C.; Nicole, L.; Sanchez, C.; Grosso, D. J. Sol-Gel Sci. Technol. 2010, 53, 551. (235) Faustini, M.; Capobianchi, A.; Varvaro, G.; Grosso, D. Chem. Mater. 2012, 24, 1072. (236) Grobis, M.; Schulze, C.; Faustini, M.; Grosso, D.; Hellwig, O.; Makarov, D.; Albrecht, M. Appl. Phys. Lett. 2011, 98, 192504. (237) Faustini, M.; Nicole, L.; Boissière, C.; Innocenzi, P.; Sanchez, C.; Grosso, D. Chem. Mater. 2010, 22, 4406. (238) Hierso, J.; Sel, O.; Ringuede, A.; Laberty-Robert, C.; Bianchi, L.; Grosso, D.; Sanchez, C. Chem. Mater. 2009, 21, 2184. (239) Hierso, J.; Boy, P.; Vallé, K.; Vulliet, J.; Blein, F.; LabertyRobert, C.; Sanchez, C. J. Solid State Chem. 2013, 197, 113. (240) Muller, G.; Boissiere, C.; Grosso, D.; Ringuede, A.; LabertyRobert, C.; Sanchez, C. J. Mater. Chem. 2012, 22, 9368. (241) Muller, G.; Vannier, R.-N.; Ringuede, A.; Laberty-Robert, C. J. Phys. Chem. C 2013, 117, 162972. (242) Muller, G.; Vannier, R.-N.; Baldinozzi, G.; Ringuede, A.; Laberty-Robert, C.; Sanchez, C. J. Phys. Chem. B 2013, DOI: 10.1039/ C3TA11175J. (243) Baldinozzi, G.; Muller, G.; Laberty-Robert, C.; Gosset, D.; Simeone, D.; Sanchez, C. J. Phys. Chem. C 2012, 116, 7658. (244) Hamd, W.; Cobo, S.; Fize, J.; Baldinozzi, G.; Schwartz, W.; Reymermier, M.; Pereira, A.; Fontecave, M.; Artero, V.; LabertyRobert, C.; Sanchez, C. Phys. Chem. Chem. Phys. 2012, 14, 13224. (245) Hamd, W.; Chavarot-Kerlidou, M.; Fize, J.; Muller, G.; Leyris, A.; Matheron, M.; Courtin, E.; Fontecave, M.; Sanchez, C.; Artero, V.; Laberty-Robert, C. J. Mater. Chem. A 2013, 1, 8217. (246) Boissière, C.; Nicole, L.; Gervais, C.; Babonneau, F.; Antonietti, M.; Amenitsch, H.; Sanchez, C.; Grosso, D. Chem. Mater. 2006, 18, 5238. (247) Debecker, D. P.; Stoyanova, M.; Colbeau-Justin, F.; Rodemerck, U.; Boissière, C.; Gaigneaux, E. M.; Sanchez, C. Angew. Chem., Int. Ed. 2012, 51, 2129. (248) Ungureanu, S.; Birot, M.; Laurent, G.; Deleuze, H.; Babot, O.; Julián-López, B.; Achard, M.-F.; Popa, M. I.; Sanchez, C.; Backov, R. Chem. Mater. 2007, 19, 5786. 236

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

(249) Horcajada, P.; Serre, C.; Grosso, D.; Boissière, C.; Perruchas, S.; Sanchez, C.; Férey, G. Adv. Mater. 2009, 21, 1931. (250) Lu, Y.; Fan, H.; Stump, A.; Ward, T. L.; Rieker, T.; Brinker, C. J. Nature 1999, 398, 223. (251) Pang, J.; Stuecker, J. N.; Jiang, Y.; Bhakta, A. J.; Branson, E. D.; Li, P.; Cesarano, J.; Sutton, D.; Calvert, P.; Brinker, C. J. Small 2008, 4, 982. (252) Tsung, C.-K.; Fan, J.; Zheng, N.; Shi, Q.; Forman, A. J.; Wang, J.; Stucky, G. D. Angew. Chem., Int. Ed. 2008, 47, 8682. (253) Carné-Sánchez, A.; Imaz, I.; Cano-Sarabia, M.; Maspoch, D. Nat. Chem. 2013, 5, 203. (254) Garcia Marquez, A.; Horcajada, P.; Grosso, D.; Ferey, G.; Serre, C.; Sanchez, C.; Boissiere, C. Chem. Commun. 2013, 49, 3848. (255) Fan, H.; Lu, Y.; Stump, A.; Reed, S. T.; Baer, T.; Schunk, R.; Perez-Luna, V.; Lopez, G. P.; Brinker, C. J. Nature 2000, 405, 56. (256) Liu, X.; Shen, Y.; Yang, R.; Zou, S.; Ji, X.; Shi, L.; Zhang, Y.; Liu, D.; Xiao, L.; Zheng, X.; Li, S.; Fan, J.; Stucky, G. D. Nano Lett. 2012, 12, 5733. (257) Mougenot, M.; Lejeune, M.; Baumard, J. F.; Boissiere, C.; Ribot, F.; Grosso, D.; Sanchez, C.; Noguera, R. J. Am. Ceram. Soc. 2006, 89, 1876. (258) Ungureanu, S.; Deleuze, H.; Sanchez, C.; Popa, M. I.; Backov, R. Chem. Mater. 2008, 20, 6494. (259) Brun, N.; Julián-López, B.; Hesemann, P.; Laurent, G.; Deleuze, H.; Sanchez, C. m.; Achard, M.-F.; Backov, R. n. Chem. Mater. 2008, 20, 7117. (260) Brun, N.; Babeau-Garcia, A.; Achard, M.-F.; Sanchez, C.; Durand, F.; Laurent, G.; Birot, M.; Deleuze, H.; Backov, R. Energy Environ. Sci. 2011, 4, 2840. (261) Kickelbick, G. Prog. Polym. Sci. 2003, 28, 83. (262) Vaia, R. A.; Maguire, J. F. Chem. Mater. 2007, 19, 2736. (263) Alexandre, M.; Dubois, P. Mater. Sci. Eng., R 2000, 28, 1. (264) Giannelis, E. P. Adv. Mater. 1996, 8, 29. (265) Althues, H.; Henle, J.; Kaskel, S. Chem. Soc. Rev. 2007, 36, 1454. (266) Zou, H.; Wu, S. S.; Shen, J. Chem. Rev. 2008, 108, 3893. (267) Hussain, F.; Hojjati, M.; Okamoto, M.; Gorga, R. E. J. Compos. Mater. 2006, 40, 1511. (268) Chemin, N.; Rozes, L.; Chaneac, C.; Cassaignon, S.; Le Bourhis, E.; Jolivet, J.-P.; Spalla, O.; Barthel, E.; Sanchez, C. Chem. Mater. 2008, 20, 4602. (269) Chemin, N.; Rozes, L.; Chaneac, C.; Cassaignon, S.; Le Bourhis, E.; Jolivet, J.-P.; Barthel, E.; Sanchez, C. Eur. J. Inorg. Chem. 2012, 2675. (270) Sanchez, C.; Lebeau, B.; Chaput, F.; Boilot, J. P. Adv. Mater. 2003, 15, 1969. (271) Mammeri, F.; Le Bourhis, E.; Rozes, L.; Sanchez, C. J. Mater. Chem. 2005, 15, 3787. (272) Malzbender, J.; den Toonder, J. M. J.; Balkenende, A. R.; de With, G. Mater. Sci. Eng., R 2002, 36, 47. (273) Mammeri, F.; Rozes, L.; Sanchez, C.; Le Bourhis, E. J. Sol-Gel Sci. Technol. 2003, 26, 413. (274) Perineau, F.; Hu, G. J.; Rozes, L.; Ribot, F.; Sanchez, C.; Creton, C.; Bouteiller, L.; Pensec, S. J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 2636. (275) Perineau, F.; Pensec, S.; Sanchez, C.; Creton, C.; Rozes, L.; Bouteiller, L. Polym. Chem. 2011, 2, 2785. (276) Dan-Hardi, M.; Serre, C.; Frot, T.; Rozes, L.; Maurin, G.; Sanchez, C.; Ferey, G. J. Am. Chem. Soc. 2009, 131, 10857. (277) Schubert, U. Chem. Mater. 2001, 13, 3487. (278) Rozes, L.; Sanchez, C. Chem. Soc. Rev. 2011, 40, 1006. (279) Périneau, F.; Pensec, S.; Sassoye, C.; Ribot, F.; Van Lokeren, L.; Willem, R.; Bouteiller, L.; Sanchez, C.; Rozes, L. J. Mater. Chem. 2011, 21, 4470. (280) Johnson, C. S. Prog. Nucl. Magn. Reson. Spectrosc. 1999, 34, 203. (281) Ribot, F.; Escax, V.; Martins, J. C.; Biesemans, M.; Ghys, L.; Verbruggen, I.; Willem, R. Chem.−Eur. J. 2004, 10, 1747. (282) Van Lokeren, L.; Willem, R.; van der Beek, D.; Davidson, P.; Morris, G. A.; Ribot, F. J. Phys. Chem. C 2010, 114, 16087.

(283) Ribot, F.; Escax, V.; Roiland, C.; Sanchez, C.; Martins, J. C.; Biesemans, M.; Verbruggen, I.; Willem, R. Chem. Commun. 2005, 1019. (284) Van Lokeren, L.; Maheut, G.; Ribot, F.; Escax, V.; Verbruggen, I.; Sanchez, C.; Martins, J. C.; Biesemans, M.; Willem, R. Chem.−Eur. J. 2007, 13, 6957. (285) Coppel, Y.; Spataro, G.; Pages, C.; Chaudret, B.; Maisonnat, A.; Kahn, M. L. Chem.−Eur. J. 2012, 18, 5384. (286) Hens, Z.; Martins, J. C. Chem. Mater. 2013, 25, 1211. (287) Lehn, J. M. Prog. Polym. Sci. 2005, 30, 814. (288) Rowan, S. J.; Cantrill, S. J.; Cousins, G. R. L.; Sanders, J. K. M.; Stoddart, J. F. Angew. Chem., Int. Ed. 2002, 41, 898. (289) Bergman, S. D.; Wudl, F. J. Mater. Chem. 2008, 18, 41. (290) Hager, M. D.; Greil, P.; Leyens, C.; van der Zwaag, S.; Schubert, U. S. Adv. Mater. 2010, 22, 5424. (291) Potier, F.; Guinault, A.; Delalande, S.; Sanchez, C.; Ribot, F.; Rozes, L., Submitted. (292) Ribot, F.; Veautier, D.; Guillaudeu, S. J.; Lalot, T. J. Mater. Chem. 2005, 15, 3973. (293) Long, J. R.; Yaghi, O. M. Chem. Soc. Rev. 2009, 38, 1213. (294) Rowsell, J. L. C.; Yaghi, O. M. Microporous Mesoporous Mater. 2004, 73, 3. (295) O’Keeffe, M.; Yaghi, O. M. J. Solid State Chem. 2005, 178, V. (296) Kuznetsov, A. I.; Kameneva, O.; Bityurin, N.; Rozes, L.; Sanchez, C.; Kanaev, A. Phys. Chem. Chem. Phys. 2009, 11, 1248. (297) Fu, Y. H.; Sun, D. R.; Chen, Y. J.; Huang, R. K.; Ding, Z. X.; Fu, X. Z.; Li, Z. H. Angew. Chem., Int. Ed. 2012, 51, 3364. (298) Kim, S. N.; Kim, J.; Kim, H. Y.; Cho, H. Y.; Ahn, W. S. Catal. Lett. 2013, 204, 85. (299) Horiuchi, Y.; Toyao, T.; Saito, M.; Mochizuki, K.; Iwata, M.; Higashimura, H.; Anpo, M.; Matsuoka, M. J. Phys. Chem. C 2012, 116, 20848. (300) Hendon, C. H.; Tiana, D.; Fontecave, M.; Sanchez, C.; D’arras, L.; Sassoye, C.; Rozes, L.; Mellot-Draznieks, C.; Walsh, A. J. Am. Chem. Soc. 2013, 135 (30), 10942−10945. (301) Kreuer, K. D.; Paddison, S. J.; Spohr, E.; Schuster, M. Chem. Rev. 2004, 104, 4637. (302) Hickner, M. A.; Ghassemi, H.; Kim, Y. S.; Einsla, B. R.; McGrath, J. E. Chem. Rev. 2004, 104, 4587. (303) Jones, D. J.; Roziere, J. Fuel Cells I 2008, 215, 219. (304) Stimming, U.; Jones, D.; Bele, P. Fuel Cells 2013, 13, 3. (305) Kreuer, K. D. J. Membr. Sci. 2001, 185, 29. (306) Mauritz, K. A. Mater. Sci. Eng., C 1998, 6, 121. (307) Alberti, G.; Casciola, M.; Costantino, U. J. Membr. Sci. 1983, 16, 137. (308) Laberty-Robert, C.; Valle, K.; Pereira, F.; Sanchez, C. Chem. Soc. Rev. 2011, 40, 961. (309) Di Noto, V.; Piga, M.; Giffin, G. A.; Vezzu, K.; Zawodzinski, T. A. J. Am. Chem. Soc. 2012, 134, 19099. (310) Di Noto, V.; Boaretto, N.; Negro, E.; Pace, G. J. Power Sources 2010, 195, 7734. (311) Kundu, P. P.; Sharma, V.; Shul, Y. G. Crit. Rev. Solid State Mater. Sci. 2007, 32, 51. (312) Diat, O.; Gebel, G. Nat. Mater. 2008, 7, 13. (313) Mauritz, K. A.; Stefanithis, I. D.; Davis, S. V.; Scheetz, R. W.; Pope, R. K.; Wilkes, G. L.; Huang, H. H. J. Appl. Polym. Sci. 1995, 55, 181. (314) Gummaraju, R. V.; Moore, R. B.; Mauritz, K. A. J. Polym. Sci., Part B: Polym. Phys. 1996, 34, 2383. (315) Jiang, R. C.; Kunz, H. R.; Fenton, J. M. J. Membr. Sci. 2006, 272, 116. (316) Apichatachutapan, W.; Moore, R. B.; Mauritz, K. A. J. Appl. Polym. Sci. 1996, 62, 417. (317) Shao, P. L.; Mauritz, K. A.; Moore, R. B. Chem. Mater. 1995, 7, 192. (318) Jannasch, P. Curr. Opin. Colloid Interface Sci. 2003, 8, 96. (319) Sel, O.; Soules, A.; Ameduri, B.; Boutevin, B.; Laberty-Robert, C.; Gebel, G.; Sanchez, C. Adv. Funct. Mater. 2010, 20, 1090. 237

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238

Chemistry of Materials

Review

(320) Niepceron, F.; Lafitte, B.; Galiano, H.; Bigarre, J.; Nicol, E.; Tassin, J. F. J. Membr. Sci. 2009, 338, 100. (321) Sel, O.; Laberty-Robert, C.; Azais, T.; Sanchez, C. Phys. Chem. Chem. Phys. 2009, 11, 3733. (322) Sel, O.; Azais, T.; Marechal, M.; Gebel, G.; Laberty-Robert, C.; Sanchez, C. Chem.−Asian J. 2011, 6, 2992. (323) Laberty-Robert, C.; Sanchez, C. Private Communication 2013. (324) Scheele, M.; Oeschler, N.; Meier, K.; Kornowski, A.; Klinke, C.; Weller, H. Adv. Funct. Mater. 2009, 19, 3476. (325) Gich, M.; Roig, A.; Taboada, E.; Molins, E.; Bonafos, C.; Snoeck, E. Faraday Discuss. 2007, 136, 345. (326) Popovici, M.; Gich, M.; Niznansky, D.; Roig, A.; Savii, C.; Casas, L.; Molins, E.; Zaveta, K.; Enache, C.; Sort, J.; de Brion, S.; Chouteau, G.; Nogues, J. Chem. Mater. 2004, 16, 5542. (327) Barboiu, M. Chem. Commun. 2010, 46, 7466. (328) Barboiu, M.; Lehn, J.-M. Isr. J. Chem. 2013, 53, 9.

238

dx.doi.org/10.1021/cm402528b | Chem. Mater. 2014, 26, 221−238