Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Introduction to Sol-Gel Processing
Introduction to Sol-Gel Processing
Introduction to Sol-Gel Processing
Ebook1,433 pages13 hours

Introduction to Sol-Gel Processing

Rating: 0 out of 5 stars

()

Read preview

About this ebook

This book presents a broad, general introduction to the processing of Sol-Gel technologies.  This updated volume serves as a general handbook for researchers and students entering the field. This new edition provides updates in fields that have undergone rapid developments, such as Ceramics, Catalysis, Chromatropgraphy, biomaterials, glass science, and optics. It provides a simple, compact resource that can also be used in graduate-level materials science courses.

LanguageEnglish
PublisherSpringer
Release dateMar 10, 2020
ISBN9783030381448
Introduction to Sol-Gel Processing

Related to Introduction to Sol-Gel Processing

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Introduction to Sol-Gel Processing

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Introduction to Sol-Gel Processing - Alain C. Pierre

    © Springer Nature Switzerland AG 2020

    A. C. PierreIntroduction to Sol-Gel Processinghttps://doi.org/10.1007/978-3-030-38144-8_1

    1. General Introduction

    Alain C. Pierre¹  

    (1)

    allée des écureuils, Université Claude Bernard-Lyon 1, ROCHETAILLEE SUR SAONE, France

    Alain C. Pierre

    Keywords

    HistorySol-gel schemeApplicationsAdvantages and limitationsBook organization

    1.1 Short History

    1.1.1 Scientific Basis

    Sols and gels are two forms of matter which were known to exist naturally for a long time. They include various materials such as ink, clays, and a number of other substances such as the eye vitreous, blood, serum, and milk (Livage and Lemerle 1982). Sols and gels attracted the interest of scientist for a long time. The oldest sols prepared in a laboratory were synthesized with gold by Faraday (Faraday 1857). They were mentioned to be still stable nowadays (Matijevic 1987).

    1.1.2 Colloids

    Colloidal science is considered to have been founded by Graham (1861). Since then, the study of ceramic colloidal sols was slowly progressing, and laws governing the formation of sols and their behavior according to a number of different factors were progressively discovered. Later on, the size of colloid forming these sols could also be controlled, in particular by using inorganic salts to synthesize them (Roy 1956; Matijevic 1987). A good understanding of the nature of sols and of the laws explaining their behaviors was finally achieved and a major contribution in the understanding of sols’ physical chemistry came from the so-called Derjaguin, Landau, Verwey, and Overbeek (DLVO) theory (Derjaguin and Landau 1941; Verwey 1941; Verwey and Overbeek 1948), also known as electrostatic theory. This theory was particularly the first one to distinguish a precipitate from a kinetically stable colloidal suspension.

    1.1.3 Gels

    As for gels, apart from some silica gels which exist naturally (Iler 1979), their synthesis was only achieved since the nineteenth century. Ebelmen (1846, 1847) produced the first silica gels and Cossa (1870) synthesized the first alumina gels by an electrolytic method. Since then, all kinds of inorganic gels were progressively synthesized and a technique such as the supercritical drying of Kistler (1931) was introduced. This technique permitted to produce the first materials termed aerogels of silica, alumina, zirconia, stannic, and tungsten oxides. Still in ceramics but outside the field of oxides, Stock et al. (1921) synthesized the first silazanes, polymeric precursors to Si3N4. Some oxide gels were more recently synthesized, such as the first borate gels, by Tohge et al. (1984). Overall, the number of publications on gelatinous and colloidal forms of ceramics kept however steadily increasing and some of the properties and structure of gels were in turn elucidated. The first historical details were summarized by Iler (1979) for silica and by Gitzen (1970) for alumina.

    The theoretical basis needed in order to understand the network structure of a gel and the kinetics of its formation and gelation was developed by Flory (1941). His original work addressed organic gels but was soon extended to inorganic ones. Shortly after, the theory of percolation of Hammersley (1957) explained the gelation process as a special case of the critical phenomena occurring in physics. The lace geometry that characterizes a gel network and the porous ceramic obtained at the beginning of densification were then described by the fractal geometry introduced by Mandelbrot (1977).

    1.1.4 The Traditional Sol-Gel Processing of Ceramics

    For a long time sols and gels were only of pure scientific interest. However, their extremely high specific area made them increasingly interesting in the field of catalysis. By the beginning of the 1960s, following the works on clays (Ford 1964) and nuclear fuel oxides in England, and those of Matijevic (1981) on the production of colloidal particles with controlled size and shape, new ceramics were also produced for more technical use, using inorganic sols and gels, giving name to the field as sol-gel processing. Since the 1970s, an increasing number of publications in the field of gels and inorganic colloids were published. This led to an increasing number of potentially interesting applications in high-technology ceramics and some of them begun to be commercialized. These new sol-gel processes were found to fulfill at least to some extent the current need for new and better products (Mazdiyasni et al. 1965; Wheat 1977). High-purity submicron powders, nuclear fuels, electronic and ionic conductors, and magnetic materials can now be produced by sol-gel techniques. Sol-gel processing is also very useful and important when the production of reproducible homogeneous complex ceramic materials is necessary.

    A domain of importance for ceramists concerns the synthesis of ceramics of complex composition and a good homogeneity was often difficult to achieve. In this case, colloidal particles can be synthesized by sol-gel processing in a first step, and further treated by conventional ceramic processing techniques (cold pressing, hot pressing, sintering) to make a ceramic part in a second step. The colloidal size of each component implied shorter diffusion distances to achieve a homogeneous distribution of the different cations, which was the main interest of the sol-gel process. To avoid segregation of the different oxide components, the colloidal particles could also be dispersed within a stable organic sol further transformed to a gel before heat treatment, such as in the Pechini method addressed in Chap. 2. Sols and gels could also be spun into fibers or applied using various techniques to coat a substrate.

    It must also be noted that sol-gel processing is not the only route to synthesize ceramic materials at low temperature (Ohya 2016). Other types of techniques such as chemical bath deposition (CBD), successive ionic layer adsorption and reaction (SILAR), liquid-phase deposition (LPD), electroless deposition (ED), and film deposition on organic self-assembled monolayers (SAMs) were reviewed by Niesen and DeGuire (2001).

    1.1.5 Recent Chemical Developments

    During the last two decades, the sol-gel field has experienced a tremendous expansion.

    In a first period the synthesis of perfect gel monoliths, particularly SiO2 hydrophobic aerogel monoliths obtained using the supercritical drying technique (Chap. 8), required a more involved chemical adaptation of the wet gels, initially by grafting hydrophobic organic ligands on them. For this purpose, chemists and particularly organic chemists become increasingly involved to design better ceramics through chemistry, which was the title of a number of specialized conferences. In a second step, the organic functionalization of SiO2 gels led to conceiving of quite new products comprising hybrid organic-inorganic materials and also mesoporous structured materials.

    In practice, the bonding of organic molecules or polymeric macromolecules to an inorganic counterpart was developed, and a very large span of hybrid organic-inorganic could be designed (Sanchez and Ribot 1994). Such materials are rather scarce in the nature: an example which can be mentioned is the material constituting the bones of living humans or animals where the inorganic part is hydroxyapatite, a hydrated calcium phosphate compound (Chap. 14). But, overall, these synthetic hybrid materials offered a very large variety of new properties considered as very interesting for a number of advanced technologic applications.

    To these, one must moreover add materials comprising an ordered mesoporous structure of spherical, cylindrical, or lamellar pores, obtained by using surfactants during the sol-gel synthesis (Corma 1997). Sol-gel was also recently extended to non-oxide ceramics, comprising chalcogenides, fluorides, and carbon aerogels, the latter ones obtained either by calcination of organic aerogels or by binding graphene monolayer particles or carbon nanotubes, as described in Chap. 4. Even regarding oxides, a complex range of new chemical synthesis processing was developed and is presented in Chaps. 2 and 3.

    1.2 Sols, Gels, and Gelation

    1.2.1 Sols

    A sol can be defined as a stable suspension of colloidal solid particles within a liquid (Hiemenz 1977). For a sol to exist, the solid particles denser than the surrounding liquid must be small enough, so that the forces responsible for its dispersion are greater than those of gravity. Moreover, these particles must include a number of atoms macroscopically significant. In fact, if the particles were too small, then it would be more accurate to speak of molecules in solution.

    Originally, colloidal only referred to macroscopic particles that could not pass through a dialysis membrane. This definition, however, does not give accurate values of the size range of the particles which are concerned. Practically, particles in a colloidal sol must have a size comprised between 2 nm and 0.2 μm, which corresponds to 10³–10⁹ atoms per particle (Livage and Lemerle 1982). Particles in this size range can moreover be divided into three categories. They can be composed of subdivided parts of bulk matter (for example small particles of α-alumina), real macromolecules that are big enough to be colloidal (such as proteins), or small particles that can be considered as both macromolecules and tiny parts of macroscopic matter (such as lacey particles). In the case of subdivided parts of bulk material, two thermodynamic phases can be distinguished: one for the matter inside the particles, and another for the liquid in which they are dispersed. The sol can moreover be considered as either lyophobic (hence hydrophilic) or hydrophobic (hence lyophilic), depending on water, respectively, being or not being the main liquid-phase component. In the case of real macromolecules, only one thermodynamic phase is present and it is more appropriate to designate it as a hydrophilic or lyophilic solution. The small lacey particles are more difficult to characterize and can be observed, for instance, in some colloidal forms of silica of other lacey polymeric particles. Actually, the solvent most often used to disperse colloidal particles in a sol is either pure water or a solution composed mostly of water and an alcohol. Nevertheless, other liquid components may be used.

    1.2.2 Gels

    A gel can be defined as a porous three-dimensionally interconnected solid network that expands in a stable fashion throughout a liquid medium and is only limited by the size of the container. If the solid network is made of colloidal sol particles the gel is said to be colloidal. If the solid network is made of macromolecules initial in a polymeric solution, the gel is termed polymeric. A polymer, as defined by Flory (1974), is a macromolecule whose structure can be generated through repetition of one or a few elementary units. A great diversity of sols and gels exists and several classifications were proposed; the most simple of them maybe is the one given by Flory (1953) and presented in Chap. 8.

    The nature of gels is tied to an intricate contact and equilibrium coexistence between a very porous solid network and a fluid medium, either a liquid or a gas, filling its pores. In this type of equilibrium, the liquid impregnates the solid network mesh that composes the gel and constitutes the major part in volume%. Moreover it does not flow spontaneously out of this network which can reach a thermodynamic equilibrium with it. In Chap. 8, it is for instance described that a wet polymeric gel will undertake spontaneous and reversible swelling and shrinkage, depending on thermodynamics parameters such as the temperature or the liquid composition. If the liquid is mostly composed of water, the corresponding gel is often termed an aquagel or hydrogel. An aquagel is a soft material that can be easily cut with a knife. On the other hand, when the liquid phase is largely composed of an alcohol, the gel is then termed an alcogel. Finally, if most of the liquid is removed and only residual traces of it remain inside the gel, the dry form which is obtained, often a very brittle solid, is called either a xerogel or an aerogel, depending on the drying method.

    1.2.3 Gelation

    In practice, a gel is formed when the homogenous dispersion of colloidal particles or macromolecules present in the initial sol rigidifies the whole wet sol medium. This transformation, termed gelation, prevents the development of inhomogeneities within the material. The exact transition event from a sol or a solution to either a colloidal or a polymeric gel is known as the gel point (Flory 1974). Practically, at this point, the sol is abruptly transformed from a viscous liquid state to a gel , which behaves as a single solid monolith impregnated with liquid. This gel point and the properties of sols and gels near this point are now better characterized within the framework of the new theory of critical phenomena which is a branch of equilibrium thermodynamics.

    1.2.4 Xerogels and Aerogels

    The term xerogel is defined by IUPAC as an open network formed by the removal of all swelling agents from a gel (Alemán et al. 2007) but it was first introduced by Freundlich (1923) to designate shrinking (or swelling when shrinking is reversible) gels. As detailed in Chap. 8, drying a wet gel by evaporation of its mother liquid often induces an important contraction, for instance 30% or less of its initial volume (Brinker and Scherer 1990), under the action of capillary stresses. Hence, in this monography, the term xerogel is reserved for those gels which undertake an important shrinkage during drying.

    On the other hand, the term aerogel was first introduced by Kistler (1931, 1932) to designate gels in which the liquid was replaced with a gas, without any important shrinking of the gel solid network. As detailed in Chap. 8, Kistler achieved such a result by using a supercritical drying technique, according to which the liquid which impregnated the gels was evacuated after being transformed to a supercritical fluid, by heating above the fluid thermodynamic critical point. But this is not the only technique permitting to achieve such a result.

    1.2.5 Gelatinous Precipitates

    When the solid particles in a sol are heavy enough to precipitate, more or less slowly, a wet viscous solid-liquid mixture is deposited on the lower end of the sol container. Some authors called such precipitates gel, but it would be more accurate to describe them as gelatinous precipitates. The border line between gels and gelatinous precipitates can however be ambiguous, because as explained in Chap. 6, a precipitate is the result of the progressive formation of bigger aggregates of colloidal particles, so that a sol will more or less rapidly precipitate. Actually, while a gel can be thermodynamically stable, a sol can only be kinetically stable. On a time scale which depends on the material nature a sol always tends to precipitate, unless the formation kinetics of a gel 3D network expanding through the entire sol volume happens to be faster than the kinetics for dense aggregation of the colloidal particles. In details, these aggregates can really have a texture which can range from densely packed to very open, including all possible intermediate texture.

    1.2.6 Sol-Gel Processes

    Many definitions of sol-gel processes exist. For instance, Dislich (1983) considered that the sol-gel procedure only concerns multicomponent oxides which must be homogenized at the atomic level. This definition therefore did not include single-oxide materials (e.g., TiO2 gels), nor colloidal coprecipitates of hydroxides and oxyhydrates since they become homogeneous only by reaction at high temperature. Moreover in this definition, the term sol-gel processing was restricted to the gels synthesized from alkoxides. On the other hand, Segal (1984) defined sol-gel processing as the production of inorganic oxides, either from colloidal dispersion or from metal alkoxides.

    Since this time, sol-gel processing is no longer limited to oxides but it also includes non-oxide materials such as the nitrides (Seyferth and Wiseman 1984) and sulfides (Chap. 4), organic gels such as the hydrogels used in biomaterials (Chap. 4), or hybrid organic-inorganic materials which constitute a big new subfield of sol-gel processing (Chap. 10). In this book, a more general definition of sol-gel processes was therefore adopted. It is considered that a sol-gel process is a colloidal route used to synthesize any material with an intermediate stage including a sol and/or a gel state.

    1.3 Outline of Sol-Gel Processing

    Many variations can be brought to the sol-gel synthesis of materials. In fact, sol-gel processing does designate not only a unique technique, but also a very broad type of procedures which can be described according to a global scheme presented in Fig. 1.1. The first step of any sol-gel process always consists of selecting initial chemical compounds, termed precursors, from which the desired material will be synthesized. In most cases, these precursors are more or less complex chemical molecules, and their transformation by chemical reactions leads to the formation of any possible type of structures, from dense colloidal particles to polymeric gels with a very open macromolecular network.

    ../images/94851_2_En_1_Chapter/94851_2_En_1_Fig1_HTML.png

    Fig. 1.1

    Simplified chart of sol-gel processes

    When the target material needs a combination of different chemical components, for instance several oxides of different cations (e.g., SiO2 and Al2O3), or an inorganic molecule plus an organic one, a variety of combinations of different precursors can be used and, depending on the chemical protocol selected, significantly different products in terms of phases formed, homogeneity, and texture can be obtained.

    A sol-gel synthesis scheme more in line with the recent developments must also comprise the possible use of non-hydrolytic solvents, or of colloidal carbon particle precursors: in detail carbon nanotubes and graphene or graphene oxide monolayer nanoparticles. The drying methods have also experienced a major development with ambigels, based on the grafting of hydrophobic groups on the pore surface of wet gels, which permit to dry them by simple evaporation of the solvent with minor drying shrinkage.

    In the last steps of a sol-gel processing scheme, the major progress concerned new good aerogel monoliths, either as ambigels or as aerogels dried by the supercritical method . On the other hand, gel coating or fiber drawing has not seen major new developments and the same is true regarding heat treatment and sintering: these steps were largely developed by ceramists before the introduction of more involved chemistry.

    1.4 Sol-Gel Processing Applications

    1.4.1 Materials

    The applications of sol-gel processing are in great expansion although their potentiality is still larger.

    Major fields of application concern thermal insulation and optical applications. SiO2 aerogel occupies a major place with products comprising translucent aerogel flakes, as well as transparent aerogel slabs and optically transparent insulating windows. Many optical sol-gel applications concern coating, such as in antireflective coatings with index gradation and optical or infrared absorbing coatings. To these, one must add electrically conductive coatings and protection coatings that work against scratch, oxidation, and erosion on all types of materials, including plastic and steel.

    Noteworthy success was also achieved in advanced fine-grain ceramics with ferroelectric, dielectric, piezoelectric, optical, or electro-optical properties. This success is due to the feasibility of better ultrafine ceramic powders (Mazdiyasni 1982). New interesting glasses and glass ceramics can now be synthesized from sol-gel monoliths as well as by the sintering or melting of sol-gel powders. Those glasses could not be obtained by any conventional processing. Moreover, sols and gels can themselves have particular properties which make them interesting for some specific applications, such as the coating of photographic films. Other examples of noteworthy recent achievements are the design of new organic-inorganic gels used in the embedment of photochromic and laser dyes (Sanchez and Ribot 1994), or new catalysts based on the synthesis of ordered pore structures from micellar surfactant solutions (Corma 1997).

    1.4.2 Advantages and Limitations of Sol-Gel Processing

    Sol-gel processing presents many advantages. First, it permits to synthesize materials with any oxide composition, but also new hybrid organic-inorganic materials which do not exist naturally. Another significant advantage concerns the purity of materials, which can be made very good given the nature of the chemical precursors which can be purified by distillation, crystallization, or electrolysis. Moreover, the chemical processes of the first steps are always carried out at low temperatures. By comparison with the classical high-temperature synthesis of conventional ceramics, this minimizes considerably the chemical interactions between the material and the container walls. Another advantage is the association of the solid colloidal state with a liquid medium, which permits to avoid any pollution by the eventual dispersion of dust. This explains why the biggest industrial application of sol-gel synthesis of ceramics is for nuclear fuels where control of the pollution is very critical.

    There are other more fundamental advantages to sol-gel processing. For instance, the kinetics of the various chemical reactions can be easily controlled by the low processing temperature, the precursor nature and concentration of solution, and the solvent nature. The nucleation and growth of primary colloidal particles can also be controlled in order to obtain particles with a given shape, size, and size distribution. Typically, this size is submicronic which in turn permits sintering at temperatures currently lower by several hundred degrees, than by conventional processing (Zelinski and Uhlmann 1984). A lower sintering temperature can also be critical to form certain metastable phases, including some glasses, to avoid or limit fiber-matrix reaction in fiber composites or to densify some hybrid organic-inorganic materials.

    The structure of sol-gel ceramics can more easily be controlled by sol-gel processing than by conventional processing, in part due to the smaller primary sol-gel particles. An amorphous or semivitreous state is also easier to obtain and many new glass compositions could be synthesized by sol-gel, which were not feasible by a conventional quench technique. This is due to the fact that two problems encountered with the conventional quench techniques could be avoided by using sol-gel processing. The first one is the very high fluidity of the material liquidus which requires a very fast cooling rate to form a glass. The second one is the frequent existence of a miscibility gap in some domains of composition and temperature making it impossible to obtain a homogeneous glass by melt quenching. The distribution of pores and crystalline or amorphous phases in a crystalline material texture can also be tailored to a large extent. This is particularly true when a segregated phase can be controlled to be either present at grain boundaries or completely eliminated. Eliminating this phase can be important in some cases, such as for the ionic conductors composed of sodium and silicon (Nasicon) for which the segregation of ZrO2 at grain boundaries must be avoided. But the presence at grain boundaries only of segregated phase can be important to avoid abnormal grain growth and to permit full densification in some ceramics, such as Al2O3.

    Sol-gel processing offers the most outstanding advantages for mixed-oxide systems in which the chemical homogeneity of the various elements can be controlled down to the atomic level. This is the case of some lead lanthanum zirconium titanates (PLZT) (Haertling and Land 1972). This condition is virtually impossible to achieve when solid powders are mechanically mixed such as in the processing of bigger conventional powders, so that the optical transparency is not as high as that obtained by sol-gel processing.

    The greatest limitation to the synthesis of ceramic by sol-gel processing remains the cost of the precursors, especially that of alkoxides. Most of these alkoxides are nonetheless quite easy to make, especially if they do not tend to polymerize. A few of them such as Zr and Ti were early industrially used by the Schott company for coating applications (Dislich 1983) and are thus quite affordable. Moreover, alkoxides can also be mixed with much cheaper metal salts provided that a purification step is included in the procedure.

    The sol-gel synthesis of ceramics will never be able to compete for the mass production of some large-scale materials such as window glass for which the conventional processes can rely on much cheaper raw materials. Sol-gel processing becomes, however, much more interesting for highly advanced ceramics.

    But sol-gel processing is not the only chemical route that leads to better ceramics. Another procedure includes the use of organometallic compounds in which an organic group is directly bonded to a metal without any oxygen atom intermediate. The chemistry of organometallics essentially concerns the synthesis of complex polymers, especially related to carbides and nitrides. But when this chemistry is carried out in a liquid medium, it can still be considered as a sol-gel process. Precipitation and coprecipitation techniques are also used and are sometimes even considered as a side branch of sol-gel processing. The chemical reactions concerned in this case are the same as those occurring in sol-gel synthesis. They often lead to the production of colloidal particles, but they can also be re-dispersed into a stable sol.

    Two other techniques are the thermal decomposition of precursors in the vapor phase (Mazdiyasni et al. 1965) and hydrothermal processing (Sapieszko and Matijevic 1980). The first one is also known as chemical vapor decomposition (CVD). Some CVD protocols use quite sophisticated nucleation and growth procedures such as heating by laser and, as in sol-gel processing, very pure products can be obtained depending on the CVD precursor used, which can be an alkoxide. The hydrothermal process is carried out in solutions placed in an autoclave and at higher temperatures than those required in sol-gel processing. This technique also produces crystals with a size of the order of a micrometer or larger. Hydrothermal synthesis is not within the scope of this book, although its wet chemistry aspect is essentially similar to that of sol-gel processing.

    1.5 Organization of the Book

    This book aims at offering a synthetic view of the various aspects and interest of the materials produced by sol-gel processing. These aspects are presented according to the scheme of sol-gel process illustrated in Fig. 1.1. The liquid medium chemistry is first discussed in Chap. 2 for metal salt precursors, in Chap. 3 for alkoxides, and in Chap. 4 for non-oxides (chalcogenide, fluorides, carbides and nitrides, simple organic gels, and carbon gels). Chapter 5 is devoted to the nucleation and growth of nanoparticles from these solutions and Chap. 6 describes their behavior, dispersion or coagulation, in a sol. The phenomenon of gelation is described in Chap. 7 and the gels themselves and their properties are the subject of Chap. 8 regarding wet gels (including their drying), and Chap. 9 regarding the dry gels (Chap. 5). Presentation of two recent families of sol-gel-derived materials is given in Chap. 10 for hybrid organic-inorganic materials and Chap. 11 regarding surfactant-templated materials. The phase transformations of gels at higher temperature and their sintering are the subjects of Chaps. 12 and 13, respectively. These chapters were only marginally modified in comparison with their equivalent from the first edition of this book. On the other hand, different applications were significantly developed and are presented in Chap. 14.

    Overall, from molecules in an aqueous medium to porous ceramics undertaking sintering, a constant phenomenon is a competition between the formation of dense structures and of open ones. This book is a journey through the various stages of this competition.

    References

    J. Alemán, A.V. Chadwick, J. He, M. Hess, K. Horie, R.G. Jones, P. Kratochvíl, I. Meisel, I. Mita, G. Moad, S. Penczek, R.F.T. Stepto, Pure Appl. Chem. 79, 1801–1829 (2007)Crossref

    C.J. Brinker, G.W. Scherer, Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing, vol 1990 (Academic Press, New York, 1990), p. 908

    A. Corma, Chem. Rev. 97, 2373–2420 (1997)Crossref

    A. Cossa, Il Nuovo Cimento 3(3), 228–230 (1870)

    B.V. Derjaguin, L.D. Landau, Acta Physicochim. 14, 633–662 (1941)

    H. Dislich, J. Non-Cryst. Solids 57(57), 371–388 (1983)Crossref

    M. Ebelmen, Ann. Chim. Phys. 16, 129–166 (1846)

    M. Ebelmen, C. R. Acad. Sci. Paris 25, 854–856 (1847)

    M. Faraday, Philos. Trans. R. Soc. Lond. 147, 145–181 (1857)

    P.J. Flory, J. Am. Chem. Soc. 63(63), 3083–3100 (1941)Crossref

    J.P. Flory, Principles of Polymer Chemistry (Cornell University Press, Ithaca, New York, 1953)

    P.J. Flory, Disc. Faraday Soc. 57(54), 7–8 (1974)Crossref

    R.W. Ford, Drying, Institute of Ceramics, Textbook Series (MacLaren and Sons, London, 1964)

    H. Freundlich, Colloid and Capillary Chemistry (Dutton Ed., New York, 1923)

    W.H. Gitzen, Alumina as a Ceramic Material (The American Ceramic Society, Columbus, OH, 1970)

    T. Graham, Philos. Trans. R. Soc. Lond. 151, 183–224 (1861)

    G.H. Haertling, C.E. Land, Ferroelectrics 3, 269–280 (1972)Crossref

    J.M. Hammersley, Proc. Cambridge Phil. Soc. 53, 642–645 (1957)Crossref

    P.C. Hiemenz, Principles of Colloid and Surface Chemistry (Marcel Dekker, New York, 1977). Also: Hiemenz P.C. and Rajagopalan R. 3rd Edition Taylor and Francis (1997)

    R.K. Iler, The Chemistry of Silica (Wiley, New York, 1979)

    S.S. Kistler, Nature 127, 741 (1931)Crossref

    S.S. Kistler, J. Phys. Chem. 36, 52–64 (1932)Crossref

    J. Livage, J. Lemerle, Annu. Rev. Mater. Sci. 12, 103–122 (1982)Crossref

    B.B. Mandelbrot, Fractals: Form, Chances and Dimensions (Freeman, San Francisco, 1977)

    E. Matijevic, Acc. Chem. Res. 14, 22–29 (1981)Crossref

    E. Matijevic, in Monodisperse Colloids (Preparation, Properties and Applications), and Interactions in Mixed Colloidal Systems (Heterocoagulation, Adhesion and Microflotation). Conference Presented at the Université de Bordeaux I, France (9–10 June 1987)

    K.S. Mazdiyasni, Ceram. Int. 8, 42–56 (1982)Crossref

    K.S. Mazdiyasni, C.T. Lynch, J.S. Smith II, J. Am. Ceram. Soc. 48, 372–375 (1965)Crossref

    T.P. Niesen, M.R. DeGuire, J. Electroceram. 6, 169–207 (2001)Crossref

    Y. Ohya, Aqueous Precursors, in Handbook of Sol-Gel Science and Technology, ed. by L. Klein, M. Aparicio, A. Jitianu, (Springer, New York, 2016). (Chapters 2–10)

    R. Roy, J. Am. Ceram. Soc. 39, 145 (1956)Crossref

    C. Sanchez, F. Ribot, New J. Chem. 18, 1007–1047 (1994)

    R.S. Sapieszko, E. Matijevic, Corrosion 36, 522–530 (1980)Crossref

    D.L. Segal, J. Non-Cryst. Solids 63, 183–191 (1984)Crossref

    D. Seyferth, G.H. Wiseman, in Ultrastructure Processing of Ceramics, Classes and Composites, ed. by L. L. Hench, D. R. Ulrich, (Wiley, New York, 1984), pp. 265–271

    A. Stock, K. Somieski, Silicon hydrides. X. Ber. Dtsch. Chem. Ges. 54, 740–758 (1921)Crossref

    N. Tohge, G.S. Moore, J.D. Mackenzie, J. Non-Cryst. Solids 63, 95–103 (1984)Crossref

    E.J.W. Verwey, Rec. Trav. Chim. 60, 625–633 (1941)Crossref

    E.J.W. Verwey, J.T.G. Overbeek, Theory of the Stability of Lyophobic Colloids (Elsevier, Amsterdam, 1948)

    T.A. Wheat, J. Can. Ceram. Soc. 46, 11–18 (1977)

    B.J.J. Zelinski, D.R. Uhlmann, J. Phys. Chem. Solids 45, 1069–1090 (1984)Crossref

    © Springer Nature Switzerland AG 2020

    A. C. PierreIntroduction to Sol-Gel Processinghttps://doi.org/10.1007/978-3-030-38144-8_2

    2. The Sol-Gel Chemistry of Oxides from Metal Salts

    Alain C. Pierre¹  

    (1)

    allée des écureuils, Université Claude Bernard-Lyon 1, ROCHETAILLEE SUR SAONE, France

    Alain C. Pierre

    Keywords

    Metal saltsHydrolysisCondensationPartial chargeComplexationPechini

    2.1 Introduction

    If inorganic sols and gels can be obtained by various methods, they are often directly synthesized from chemical reactants dissolved in a liquid medium. A chemical reactant which contains the cation M present in the final inorganic sol or gel is called a chemical precursor . Its chemical transformations are complex and involve a competition at the molecular level between the reactions responsible for the formation of open structures and those leading to dense solids. These same reactions are also responsible for the controlled dispersion of dense colloidal particles in a sol or their agglomeration into a gel or a precipitate.

    In sol-gel processing, many types of precursors can be used, provided that a solvent is available to dissolve them. In practice, two main groups are distinguished: the metallic salts and the alkoxides. The general formula of a metallic salt is MmXn where M is the metal, X an anionic group, and m and n stoichiometric coefficients. These precursors often have an ionic structure. They comprise chlorides such as aluminum chloride A1C13, but also sulfates, oxysulfates, nitrates, and organic salts such as acetates, citrates, and lactates (Ohya 2016). In this chapter, the focus is on the chemical reactions occurring in liquid medium for this type of precursors.

    The solids synthesized by sol-gel are mainly oxides. Hence water is usually present as a major reactant in order to bind oxygen atoms to the cations. Moreover it is also often the main solvent. Hence, the electronic properties of the water molecule are important to summarize, as they are largely linked to the chemical transformations of the precursors. At last a number of other reactants are important: acids, bases, solvents other than water, and various complexing agents. All these aspects and the main physical chemistry theory, the partial charge theory, which permits to some extent to predict or explain the molecular complexes which are formed, are reviewed.

    2.2 Solvents

    It is necessary to consider water separately from the other solvents.

    2.2.1 Water

    The water molecule is, in the Lewis representation, V-shaped (Fig. 2.1). The oxygen atom is surrounded by four electron pairs; which comprise one covalent bond with each of the two hydrogen atoms, plus two unshared electron pairs. According to the valence shell electron pair repulsion (VSEPR) model, the oxygen therefore occupies the center position of a tetrahedron and the two hydrogen atoms occupy two of the tetrahedron corners. The angle $$ \theta =\hat{\mathrm{HOH}} $$ is 104.5° in the gaseous state, and varies from 118° to 120° in the liquid state.

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig1_HTML.png

    Fig. 2.1

    Lewis representation of a water molecule

    Such a configuration provides a permanent polarization to the water molecule, with a dipole moment defined as

    $$ \overrightarrow{\mu}=\left(\delta q\right)\kern0.24em \overrightarrow{d} $$

    (2.1)

    In the physical representation, the distance vector $$ \overrightarrow{d} $$ originates at the center of the negative charges (−δq) on the unshared electron pair side of the oxygen atom, and ends at the center of the positive charges (+δq) of the H atoms. The magnitude of this dipole moment is |μ| = 1.85 D (1D = 1 Debye = 3.336 × 10−30 C m) (Atkins 1994).

    The water molecule belongs to the symmetry group C2v (Chabanel and Gressier 1991). The following theoretical calculations of the molecular orbitals of H2O can be made by linear combination of the 2s and 2p atomic orbitals of the oxygen atom and the 1s orbitals of the two hydrogen atoms. The electrons of the water molecule are therefore placed on the four molecular orbitals which have the following wave functions and energies:

    $$ {\varPsi}_{\mathrm{w}}\left(2{\mathrm{a}}_1\right)=0.85{\varPsi}_0\left(2\mathrm{s}\right)+0.13{\varPsi}_0\left(2{\mathrm{p}}_z\right)+0.81\left({\varPsi}_{\mathrm{HA}}\left(1\mathrm{s}\right)+{\varPsi}_{\mathrm{HB}}\left(1\mathrm{s}\right)\right)\kern2em E=-36\;\mathrm{eV} $$

    (2.2)

    $$ {\varPsi}_{\mathrm{w}}\left(1{\mathrm{b}}_2\right)=0.4{\varPsi}_0\left(2{\mathrm{p}}_y\right)+0.78\;\left({\varPsi}_{\mathrm{HA}}\left(1\mathrm{s}\right)-{\varPsi}_{\mathrm{HB}}\left(1\mathrm{s}\right)\right)\kern2em E=-19\;\mathrm{eV} $$

    (2.3)

    $$ {\varPsi}_{\mathrm{w}}\left(3{\mathrm{a}}_1\right)=0.46{\varPsi}_0\left(2\mathrm{s}\right)-0.83{\varPsi}_0\left(2{\mathrm{p}}_z\right)-0.33\left({\varPsi}_{\mathrm{HA}}\left(1\mathrm{s}\right)+{\varPsi}_{\mathrm{HB}}\left(1\mathrm{s}\right)\right)\kern2em E=-14\;\mathrm{eV} $$

    (2.4)

    $$ {\varPsi}_{\mathrm{w}}\left(1{\mathrm{b}}_1\right)={\varPsi}_0\left(2{\mathrm{p}}_x\right)\kern2em E=-12\;\mathrm{eV} $$

    (2.5)

    In these equations Ψw represents the molecular orbitals of a water molecule, while Ψ0 designates atomic orbitals of the oxygen atom, and ΨHA and ΨHB the respective atomic orbitals of the two hydrogen atoms A and B. The corresponding energy diagram, as illustrated in Fig. 2.2 with a shape of the electron presence probability clouds, demonstrates that the 3a1 molecular orbital of water is largely delocalized outside the water molecule. This orbital is responsible for the Lewis base character of water. The molecule is therefore able to donate a pair of electrons to a ligand group and build a σ bond with it. On the other hand, the 1b1 molecular orbital is strictly nonbinding, and hence has a very weak π donor character.

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig2_HTML.png

    Fig. 2.2

    Energy diagram and shape of the electron presence probability clouds for the molecular orbitals of water (adapted from Jorgensen and Salem (1973))

    It is possible to estimate the partial electrical charge δq (thereafter simply termed δ) carried by each atom from the molecular orbital wave functions. Since the wave function ΨM of a molecule AB is obtained by linear combination of ΨA and ΨB

    $$ {\varPsi}_{\mathrm{M}}=a{\varPsi}_{\mathrm{A}}+b{\varPsi}_{\mathrm{B}} $$

    (2.6)

    The electronic charge density is obtained by integrating on the whole volume

    $$ {\varPsi}_{\mathrm{M}}^2={a}^2{\varPsi}_{\mathrm{A}}^2+{b}^2{\varPsi}_{\mathrm{B}}^2+2 ab{\varPsi}_{\mathrm{A}}{\varPsi}_{\mathrm{B}} $$

    (2.7)

    The a² coefficient gives the electron charge density contributed by the A atom only, and the b² coefficient by the B atom only, and the overlap integral of 2abΨAΨB is attributed in equal proportion to the A and B atoms. The difference between the total charge density of the A atom in the molecule AB and in the isolated state gives an estimation of the partial charge δ(A) carried by the A atom. For the water molecule, this estimation gives a negative partial charge for the oxygen atom of δ(O) = −0.4 and a positive partial charge δ(H) = +0.2 for each hydrogen atom.

    Since its molecules are polarized, water is an excellent liquid medium in which to dissolve ionic solutes. This property is expressed by the following coulomb electrostatic force F between two electrical charges q and q′:

    $$ F=\frac{1}{4\pi {\upepsilon}_{\mathrm{r}}\kern0.24em {\varepsilon}_0}\frac{q{q}^{\prime }}{r^2} $$

    (2.8)

    in which ε0 = 8.8542 × 10−12 F m−1 is the dielectric permittivity of vacuum and εr the relative dielectric constant which has no dimension. Water at 25 °C has a relative dielectric constant of εr = 78.4; consequently it largely attenuates the coulomb interaction between two electrical charges.

    As illustrated in Fig. 2.3, the polar structure of water molecules drives the formation of hydrogen bonds between oxygen and hydrogen atoms of different molecules. In this type of bond, a hydrogen atom fluctuates by tunnel effect between two minimum energy positions only separated by a small energy barrier, of the order of 20–40 kJ mol−1. Such bonds can also be formed with fluoride or nitrogen atoms instead of oxygen.

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig3_HTML.png

    Fig. 2.3

    Hydrogen bonding between two water molecules

    A direct consequence of these bonds is that, according to molecular dynamic calculations (Stillinger 1980; Jolivet et al. 1994), very few water molecules can be isolated in the pure liquid state. Most water molecules are bound by up to four hydrogen bonds to neighboring molecules. Those random groups have an average lifetime of the order of 10−10 s. Water molecules also auto-dissociate according to the equilibrium reaction:

    $$ 2{\mathrm{H}}_2\mathrm{O}\leftrightarrow \kern0.6em {\mathrm{H}}_3{\mathrm{O}}^{+}+{\mathrm{H}\mathrm{O}}^{-} $$

    (2.9)

    with equilibrium constant

    $$ {K}_{\mathrm{w}}=\left[{\mathrm{H}}_3{\mathrm{O}}^{+}\right]\kern0.36em \left[{\mathrm{H}\mathrm{O}}^{-}\right]={10}^{-14}\kern0.24em \mathrm{at}\kern0.24em {25}^{\circ}\mathrm{C} $$

    (2.10)

    Hence water is a protic solvent. Furthermore, H3O+ and HO− ions also associate themselves with other water molecules by hydrogen bonds, so that they actually exist in liquid water as [H9O4]+ and [H7O4]−. The latter anion is the strongest possible base in aqueous solutions. As a consequence, O²− practically does not exist in water and, when a solid oxide is dissolved, it immediately undergoes an acid-base protonation reaction.

    2.2.2 Nonaqueous Solvents

    Apart for the molten salts, nonaqueous solvents are often polar with a molecular structure characterized by both a permanent dipole moment μ and a relative dielectric constant εr. εr not only depends on a permanent dipole moment μ, but also on the polarizability α of the molecule, which is itself defined according to the relation

    $$ {\mu}^{\ast }=\alpha E $$

    (2.11)

    μ∗ defines the induced dipole moment which adds to the permanent dipole moment when the molecule is submitted to an electric field E.

    A high relative dielectric constant (εr > 40) is often due to the existence of a permanent dipole moment. Such molecules have good ionizing properties and can therefore dissolve other polar solute. On the other hand when the solvent’s relative dielectric constant is low (εr < 20), it has a weak ionizing property and can only dissolve less polar solute. A list of frequently used solvents with their relative dielectric constant and dipole moment is gathered in Table 2.1.

    Table 2.1

    List of some solvents with their dielectric properties

    Adapted from Lagowski (1976)

    Solvents are classified as protic when they can exchange a proton, and as aprotic when they cannot do so. They can also be classified as acidic, in the Brønsted sense when they are able to donate a proton or in the Lewis sense when they are able to accept a pair of electrons. Similarly, a base, according to Brønsted, is able to accept a proton, and according to Lewis to donate a pair of electrons. A solvent is amphoteric when it can behave both as a base and an acid. Amphoteric solvents include:

    Mineral acids (HCN, HX, HNO3, H2SO4, H2S)

    Carboxylic acids R-COOH

    Water, the first alcohols (CH3OH, C2H5OH, …), and phenol C6H5OH

    Ammonia NH3 and amines (RNH2, RR′NH)

    Amides (R-CO-NH2, R-CO-NHR′)

    Organic solvents are frequently used in sol-gel processing as they permit to control the reaction of alkoxide precursors with water, and hence to direct with more flexibility the sol-gel product structure.

    2.3 Basis of Cation Transformations in Solution

    2.3.1 The Partial Charge Model

    Sol-gel precursors undergo chemical reactions both with water and with the other species present in the solution. One of the most efficient models used to predict those reactions is the partial charge model (PCM) which has been recently elaborated by Henry and Livage (Livage et al. 1988; Henry et al. 1990) after a principle developed by Sanderson (1971). It is based on the electrical interactions between the partial electric charges, δ, carried by each atom and molecule. Since the chemical potential of the electrons in an atom i or a molecule C depends on the partial electric charges δ(i) or δ(C) carried either by i or by C, and since the electronegativity χ(i) of i or χ(C) of C is directly related to this chemical potential, this model can also be expressed in terms of the particles’ electronegativity (Parr et al. 1978).

    The derivative of the function E = f(ne) which associates the total energy E of an isolated atom to the number of electrons ne of this atom (Fig. 2.4) describes the influence of the electric charge carried by an atom on its electronegativity. The first ionization energy, I1, of an isolated atom is also a function of the first and second partial derivatives of f(ne) with respect to ne (Parr et al. 1978):

    $$ {I}_1=E\left(z-1\right)-E(z)\approx -\frac{\partial E}{\partial {n}_{\mathrm{e}}}+\frac{1}{2}\frac{\partial^2E}{\partial\;{n}_{\mathrm{e}}^2} $$

    (2.12)

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig4_HTML.png

    Fig. 2.4

    First ionization and affinity energies of an atom (adapted from Chermette and Lissilour (1985))

    The electron affinity of this atom is

    $$ A=E(z)-E\left(z+1\right)\approx -\frac{\partial E}{\partial {n}_{\mathrm{e}}}-\frac{1}{2}\frac{\partial^2E}{\partial\;{n}_{\mathrm{e}}^2} $$

    (2.13)

    The electronegativity of an isolated atom is therefore defined as

    $$ {\chi}^{\mathrm{a}}=\frac{1}{2}\left({I}_1+A\right)=-\frac{\partial E}{\partial\;{n}_{\mathrm{e}}} $$

    (2.14)

    Since its hardness is defined as

    $$ {\eta}^{\mathrm{a}}=\frac{1}{2}\left({I}_1-A\right)=\frac{1}{2}\frac{\partial^2\;E}{\partial\;{n}_{\mathrm{e}}^2} $$

    (2.15)

    and the chemical potential of the electrons in this atom is

    $$ {\mu}_{\mathrm{e}}^{\mathrm{a}}=\frac{\partial E}{\partial\;{n}_{\mathrm{e}}} $$

    (2.16)

    we have

    $$ {\chi}^{\mathrm{a}}=-{\mu}_{\mathrm{e}}^{\mathrm{a}} $$

    (2.17)

    For any chemical transformation, equilibrium is achieved when all phases have the same chemical potential. In the same manner, the electrons of a molecule transfer from atom to atom until they all reach the same chemical potential or electronegativity. Each atom thus gains an electronic charge (δne). The definition of their hardness then leads to the following equation:

    $$ {\mu}_{\mathrm{e}}\left(\mathrm{i}\right)={\mu}_{\mathrm{e}}^{\mathrm{a}}\left(\mathrm{i}\right)-{\eta}_{\mathrm{i}}^{\mathrm{a}}\left(\delta {n}_{\mathrm{e}}\right) $$

    (2.18)

    Since the electronegativity of an atom χ(i) is the opposite of the chemical potential of the electrons μe(i) this is equivalent to

    $$ \chi \left(\mathrm{i}\right)={\chi}_{\mathrm{i}}^{\mathrm{a}}+{\eta}_{\mathrm{i}}^{\mathrm{a}}\left(\delta {n}_{\mathrm{e}}\right) $$

    (2.19)

    Electric charge transfer keeps proceeding until all atoms in all species reach an equilibrium where they all have the same electronegativity. Statistically, in order to reach such an equilibrium an ion such as H+ may leave a complex whenever its partial charge reaches the value δ(H) = +1, an anion X− if its partial charge reaches the value δ(X) = −1, or a molecule HX if its total partial charge reaches the value δ(HX) = 0. In this manner, the partial charge model provides the statistical thermodynamical basis needed to understand the possible evolution of a complex in solution.

    Several electronegativity scales exist. The absolute electronegativity $$ {\chi}_{\mathrm{i}}^{\mathrm{a}} $$ and hardness $$ {\eta}_{\mathrm{i}}^{\mathrm{a}} $$ concern an isolated atom i. The Mulliken electronegativity $$ {\chi}_{\mathrm{i}}^{\mathrm{M}} $$ and hardness $$ {\eta}_{\mathrm{i}}^{\mathrm{M}} $$ take into account the electronic state of an atom in its average valence structure. At last, the Pauling electronegativity $$ {\chi}_{\mathrm{i}}^{\mathrm{P}} $$ and hardness $$ {\eta}_{\mathrm{i}}^{\mathrm{P}} $$ involve the average structure configuration in which the atoms are engaged. It is therefore related to the binding enthalpies ΔHij between atoms i and j, ΔHii between two atoms i, and ΔHjj between two atoms j according to the relation

    $$ {\chi}_{\mathrm{i}}^{\mathrm{P}}-{\chi}_{\mathrm{j}}^{\mathrm{P}}={\left[\Delta {\mathrm{H}}_{\mathrm{i}\mathrm{j}}-\frac{1}{2}\left(\Delta {\mathrm{H}}_{\mathrm{i}\mathrm{i}}+\Delta {\mathrm{H}}_{\mathrm{j}\mathrm{j}}\right)\right]}^{1/2} $$

    (2.20)

    The Pauling $$ {\chi}_{\mathrm{i}}^{\mathrm{P}} $$ and Mulliken $$ {\chi}_{\mathrm{i}}^{\mathrm{M}} $$ electronegativities are linked by the following equation:

    $$ {\upchi}_{\mathrm{i}}^{\mathrm{M}}=\frac{\chi_{\mathrm{i}}^{\mathrm{P}}}{0.335}+0.615 $$

    (2.21)

    The electronegativity usually taken as the reference in the partial charge model is the Allred and Rochow’s electronegativity $$ {\chi}_{\mathrm{i}}^0 $$ and the corresponding hardness $$ {\eta}_{\mathrm{i}}^0 $$ , because it takes into consideration both the valence state and the shape of an atom X in its average polarization state. These characteristics are reported in Table 2.2. Furthermore, in this scale the hardness and electronegativity are linked by the following relationship:

    $$ {\eta}_{\mathrm{i}}^0=1.36\kern0.24em \sqrt{\chi_{\mathrm{i}}^0} $$

    (2.22)

    Table 2.2

    Allred-Rochow electronegativities

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Tab2_HTML.png

    After Jolivet et al. (1994)

    Let us consider in an aqueous medium and at a given pH a complex C such as [M(OH)y(H2O)N−y](zy)+. According to the partial charge model at equilibrium, the electronegativities are such that

    $$ \chi \left(\mathrm{C}\right)=\chi \left({\mathrm{H}}^{+}\right)=\chi \left({\mathrm{H}}_2\mathrm{O}\right) $$

    (2.23)

    This makes it possible to predict statistically which complex of a precursor M will exist in a solution, simply by calculating the electronegativities of all the possible complexes.

    Furthermore, for a simple atom i with a partial charge δi

    $$ \chi \left(\mathrm{i}\right)={\chi}_{\mathrm{i}}^0+{\eta}_{\mathrm{i}}^0{\delta}_{\mathrm{i}} $$

    (2.24)

    where $$ {\chi}_{\mathrm{i}}^0 $$ is the Allred-Rochow electronegativity and $$ {\eta}_{\mathrm{i}}^0 $$ the Allred-Rochow hardness. Hence the partial charge of atom i is

    $$ {\delta}_{\mathrm{i}}=\frac{\chi -{\chi}_{\mathrm{i}}^0}{\eta_{\mathrm{i}}^0}=\frac{\chi -{\chi}_{\mathrm{i}}^0}{1.36\kern0.24em \sqrt{\chi_{\mathrm{i}}^0}} $$

    (2.25)

    In the case of a complex molecule Cz+ composed of several elements, we must consider its total charge z, or formal charge, which is defined as

    $$ z={\sum}_{\mathrm{i}}{\delta}_{\mathrm{i}} $$

    (2.26)

    Since at equilibrium all atoms in C reach the same electronegativity, χ(i) = χ(C) and hence

    $$ \chi \left(\mathrm{C}\right)=\frac{\sum_{\mathrm{i}}\sqrt{\chi_{\mathrm{i}}^0}+1.36z}{\sum_{\mathrm{i}}\frac{1}{\sqrt{\chi_{\mathrm{i}}^0}}} $$

    (2.27)

    For water at pH = 7, δ(H) = +0.2 and δ(O) = −0.4 (see Sect. 2.2).

    So that for water

    $$ z={\sum}_{\mathrm{i}}{\delta}_{\mathrm{i}}=\delta \left(\mathrm{O}\right)+2\delta \left(\mathrm{H}\right)=0 $$

    (2.28)

    $$ \chi \left({\mathrm{H}}_2\mathrm{O}\right)=\frac{2\sqrt{\chi_{\mathrm{H}}^0}+\sqrt{\chi_{\mathrm{O}}^0}}{\frac{2}{\sqrt{\chi_{\mathrm{H}}^0}}+\frac{1}{\sqrt{\chi_{\mathrm{O}}^0}}}=2.491 $$

    (2.29)

    For an acidic or basic solution , pH ≠ 7 and the water molecules carry a partial charge δ(H2O) ≠ 0 either positive or negative. This partial charge represents the average charge of the H+ and OH− ions as if they were evenly shared between all H2O molecules. As mentioned in Sect. 2.2, this partial charge is actually a result of the fast transformation of complex groups of water molecules bound to each other by hydrogen bonds. The Nernst equation then provides the expression of the chemical potential of a proton:

    $$ \mu \left({\mathrm{H}}^{+}\right)={\mu}^0\left({\mathrm{H}}^{+}\right)-0.06\kern0.24em \mathrm{pH} $$

    (2.30)

    Finally, if considering that μ(H+) is proportional to μ(H+) and that χ(H+) takes the values 2.491 and 2.631 at pH = 7 and pH = 0, respectively, where H+ is statistically present in [H7O3]+ and [H9O4]+ species, the following formula can statistically be written as

    $$ \chi \left({\mathrm{H}}^{+}\right)=\chi \left({\mathrm{H}}_2\mathrm{O}\right)=2.631-0.02\;\mathrm{pH} $$

    (2.31)

    2.3.2 Transformation Mechanisms of Complexes

    Various atomic or molecular groups called ligands can bind to a complex C or a cation M, either directly or by substituting to another ligand. The transformation mechanism depends on the partial charge of the different atoms in all complex species. Those ligands with a negative partial charge are nucleophilic, and those with a positive charge are electrophilic. Similarly, in a substitution reaction, the new ligand with the highest partial negative charge, Y, is the nucleophile while the group in the metal complex with the highest positive charge, X, is the leaving group.

    Direct addition of a new ligand to a complex C occurs when the coordination number of the cation in the complex is not fully satisfied. The mechanism then involved is a nucleophilic addition symbolized by AN which may be rather complex. Substitution of a ligand by another one occurs when the coordination number of the metal is already full. In this case the reaction is expressed as an exchange of a Lewis base by another one in order to form a Lewis acid. For instance in the following example, Cl− substitutes H2O:

    $$ {\mathrm{Cl}}^{-}+{\left[\mathrm{Co}{\left({\mathrm{OH}}_2\right)}_6\right]}^{2+}\to {\left[\mathrm{Co}\mathrm{Cl}{\left({\mathrm{OH}}_2\right)}_5\right]}^{+}+{\mathrm{H}}_2\mathrm{O} $$

    (2.32)

    Three different substitution mechanisms exist: the dissociative, the associative, and the interchange mechanisms. In the first step of the dissociative substitution mechanism, enough thermal energy must be available in order to break the bond between the leaving group, X, and the complex. This X ligand then becomes labile, not very stable. Consequently, an intermediary complex in which the metal M has a reduced coordination number is formed and it can eventually be observed by analytical techniques such as NMR (Fig. 2.5a). In a second step the entering group, Y, completes again the normal coordination number of the cation M. Since the rate constant of such a reaction does not depend on the concentration of the entering Y ligand, this dissociative mechanism is a unimolecular nucleophilic substitution, designated by SN1. This is for instance the case of the substitution of H2O by a ligand L such as ammonia (NH3) or pyridine in [Ni(OH2)6]²+.

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig5_HTML.png

    Fig. 2.5

    Substitution mechanisms of a ligand Y for A ligand X. (a) Dissociative SN1 mechanism; (b) associative SN2 mechanism; (c) interchange SN2 mechanism. After McMurry (1995)

    Water is formed in the first step:

    $$ {\left[\mathrm{Ni}{\left({\mathrm{OH}}_2\right)}_6\right]}^{2+}\to {\left[\mathrm{Ni}{\left({\mathrm{OH}}_2\right)}_5\right]}^{2+}+{\mathrm{H}}_2\mathrm{O} $$

    (2.33)

    while L enters in the second step:

    $$ {\left[\mathrm{Ni}{\left({\mathrm{OH}}_2\right)}_5\right]}^{2+}+\mathrm{L}\to {\left[\mathrm{Ni}{\left({\mathrm{OH}}_2\right)}_5\mathrm{L}\right]}^{2+} $$

    (2.34)

    In the associative substitution mechanism, the entering group Y binds to the complex before departure of the other ligand. Therefore, in this first step, an intermediary complex is formed in which the cation M has an increased coordination number (Fig. 2.5b). This intermediate may also be observable by analytical techniques. The X leaving group separates from the complex only in a second step. As the rate constant of this mechanism depends on the concentration of both the entering and the leaving groups, it is a bimolecular nucleophilic substitution, termed SN2. Some examples are given further on.

    Finally, in the interchange substitution mechanism no intermediary complex can be observed, in which the metal M has either an increased or a decreased coordination number. The reaction thus proceeds in only one step and as in a SN2 substitution (Fig. 2.5c).

    Various factors can influence the reaction mechanisms leading to the exchange of ligands. Steric strain on the reaction center, for instance, inhibits associative reactions in favor of the dissociative ones. Chelating ligands also have an important effect as they are polydentate. Hence they can bind to a metal atom by several bonds and become difficult to substitute. The acetylacetonato ion [CH3COCHCOCH3]− is, for example, bidentate; if L is the chelating ligand, it can form a ring (LML) with the metal. The distance between the two binding points is the bite distance (Fig. 2.6). Table 2.3 gives a list of a few important chelating ligands.

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig6_HTML.png

    Fig. 2.6

    Binding of a chelating ligand with a metal complex. An example of a bidentate chelating ligand: the acetylacetonato ion. After Shriver et al. (1994)

    Table 2.3

    Typical chelating ligands (2: bidentate, 3: tridentate, 4: tetradentate, 6: hexadentate). The coordinating atoms are in parenthesis

    Adapted from Shriver et al. (1994)

    2.4 Hydrolysis of Cations in Solution

    In sol-gel processing, when metal salts are used as the cation precursors, they are often dissolved in an aqueous medium. A metal salt MX usually dissociates into ions which are dispersed in the solution, and the anions’ negative charge Xz− is balanced by the positive charge of the metal cation Mz+. In this example, the cation and the anion then have the same absolute formal charge z. These anions are sometimes considered as impurities, in which case they must be eliminated in order to produce, for example, pure oxide ceramics. However, they can also be invaluable in channeling the chemical transformations within the solution. In any case, all ions first solvate with water molecules, a reaction due to the polar nature of water.

    2.4.1 Ion Solvation in Water

    When in solution in water, the positive charge z+ of a cation attracts the negative partial charge pole, that is, the oxygen atom, of H2O molecules (δO) < 0). Overall, a cation is therefore entrapped by a number N of water molecules which constitute the first solvation shell (Fig. 2.7). This shell is tightly bonded to the metal cation Mz+ so that the chemical formula of the complex formed by the solvated cation is [M(H2O)N]z+. The number N is fixed for a given type of metal; its value often ranges from 4 to 8 and is frequently equal to 6, such as in [A1(H2O)6]³+. A water molecule also solvates the proton and Sect. 2.2 reported that the most frequent value of N for H+ is 4, which corresponds to the complex [H(H2O)4]+ also written as [H9O4]+.

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Fig7_HTML.png

    Fig. 2.7

    Solvation of (a) a cation and (b) an anion

    A second shell of water molecules surrounds this first solvation shell. Its oxygen atoms are turned towards the hydrogen atoms of the first shell. However this second shell is much less rigid than the first one and it does not have to be taken into consideration in the chemical evolution of the precursor.

    Anions are also solvated by water. In this case these molecules are turned the other way around with the hydrogen atoms oriented towards the anion (Fig. 2.7). Nevertheless, the solvation of anions is not as important as for cations. In fact, only the solvated cations need to be taken into consideration to explain the hydrolysis of metal salt precursors, a complex chemical transformation leading to the formation of oxides.

    2.4.2 Hydrolysis of Cations in Aqueous Media

    Hydrolysis is the deprotonation of a solvated metal cation. It consists of the loss of a proton by one or several of the water molecules which surround the metal M in the first solvation shell (Baes Jr. and Mesmer 1976). As a consequence, the aquo ligand molecule, H2O which is bonded to the metal center, is either transformed to a hydroxo ligand, OH− whenever only one proton leaves, or to an oxo ligand, O²− whenever two protons leave (Schmidt et al. 1986).

    2.4.2.1 The Formation of Hydroxo Ligands

    A hydroxo ligand is formed when the solvated metal is an acid and when water therefore acts as a Lewis base (Baes Jr. and Mesmer 1976). This corresponds to the following reaction:

    $$ {\displaystyle \begin{array}{c}{\left[\mathrm{M}{\left({\mathrm{O}\mathrm{H}}_2\right)}_N\right]}^{z+}+{\mathrm{H}}_2\mathrm{O}\leftrightarrow {\left[\mathrm{M}\left(\mathrm{OH}\right){\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-1}\right]}^{\left(z-1\right)}+{\mathrm{H}}_3{\mathrm{O}}^{+}\operatorname{}{K}_{11}\\ {}\mathrm{acid}\kern0.48em +\mathrm{Lewisbase}\kern0.48em \leftrightarrow \kern0.48em \mathrm{conjugated}\kern0.17em \mathrm{base}\kern0.6em +\kern0.6em \mathrm{conjugated}\kern0.17em \mathrm{acid}\end{array}} $$

    (2.35)

    K11 is the equilibrium constant of the first deprotonation reaction of a complex involving only one metal atom. This complex can undergo other successive deprotonations, globally described by the following reaction for h consecutive loss of protons:

    $$ {\left[\mathrm{M}{\left({\mathrm{O}\mathrm{H}}_2\right)}_N\right]}^{z+}+h{\mathrm{H}}_2\mathrm{O}\leftrightarrow {\left[\mathrm{M}{\left(\mathrm{OH}\right)}_h{\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-h}\right]}^{\left(z-h\right)}+h{\mathrm{H}}_3{\mathrm{O}}^{+} $$

    (2.36)

    In [M(OH2)N] all the ligands are water molecules; it is therefore the most acidic form of the metal complex. On the contrary, [M(OH)h(OH2)Nh](zh)+ is a more basic form of the metal complex. It is also an aquo-hydroxo complex since it contains both aquo (H2O) and hydroxo (OH−) ligands.

    If the metal has an acidic oxide , the following equivalent deprotonation reaction (2.35) explains the formation of hydroxo ligands by addition of a base to the solution:

    $$ {\left[\mathrm{M}{\left({\mathrm{OH}}_2\right)}_N\right]}^{z+}+{\mathrm{OH}}^{-}\leftrightarrow {\left[\mathrm{M}\left(\mathrm{OH}\right){\left({\mathrm{OH}}_2\right)}_{N-1}\right]}^{\left(z-1\right)+}+{\mathrm{H}}_2\mathrm{O} $$

    (2.37)

    The latter reaction proceeds according to the mechanism in reaction (2.38) (Livage et al. 1988; Jolivet et al. 1994):

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Equ38_HTML.png

    (2.38)

    In this mechanism, a free OH− nucleophilic anion attacks one of the hydrogen atoms of one of the water molecules in the first solvation shell of the metal M. Since this hydrogen carries a positive partial charge (δ(H) > 0), an electron charge transfer occurs between the incoming OH− ion and the original metal complex. Consequently, the partial charge, δ(H2O), of the H2O group composed of the incoming OH− ion and the attacked H atom increases until it becomes null, δ(H2O) = 0. When this state is reached, an independent water molecule leaves the metal complex. Such deprotonation reactions occur as long as, for an H2O group ligand, δ(O)free water < δ(O)complex < 0. In this case, the deprotonation reaction can be written (Jolivet et al. 1994) as

    $$ \left[\mathrm{M}{\left(\mathrm{OH}\right)}_z{\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-z}\right]+{\mathrm{H}}_2\mathrm{O}\leftrightarrow {\left[\mathrm{M}{\left(\mathrm{OH}\right)}_{z+1}{\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-z-1}\right]}^{-}+{\mathrm{H}}_3{\mathrm{O}}^{+} $$

    (2.39)

    Actually, as indicated before, the H+ ion is also solvated by water.

    While the above metals which have acidic oxides form hydroxo ligands, those which have basic oxides are characterized by the formation of oxo ligands, O²−. For such metals, a hydroxo ligand can still be produced with an acid by attack of a free H+ ion on the nucleophilic oxygen of an oxo ligand (Baes Jr. and Mesmer 1976; Jolivet et al. 1994). In the case of water, the reaction is the following one:

    $$ {\displaystyle \begin{array}{c}{\left[\mathrm{M}\mathrm{O}{\left({\mathrm{OH}}_2\right)}_{N\cdot 1}\right]}^{\left(z-2\right)+}+{\mathrm{H}}_2\mathrm{O}\leftrightarrow {\left[\mathrm{M}\left(\mathrm{OH}\right){\left({\mathrm{OH}}_2\right)}_{N-1}\right]}^{\left(z-1\right)+}+{\mathrm{OH}}^{-}\\ {}{\mathrm{H}}^{+}+{\mathrm{OH}}^{-}\end{array}} $$

    (2.40)

    The complex reacts with an acid in a similar reaction:

    $$ {\left[\mathrm{M}\mathrm{O}{\left({\mathrm{O}\mathrm{H}}_2\right)}_{N\cdot 1}\right]}^{\left(z-2\right)+}+{\mathrm{H}}_3{\mathrm{O}}^{+}\leftrightarrow {\left[\mathrm{M}\left(\mathrm{OH}\right){\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-1}\right]}^{\left(z-1\right)+}+{\mathrm{H}}_2\mathrm{O} $$

    (2.41)

    The corresponding mechanism is described in reaction (2.42):

    ../images/94851_2_En_2_Chapter/94851_2_En_2_Equ42_HTML.png

    (2.42)

    Such reactions keep proceeding as long as 0 < δ(H)complex < δ′(H)water (Jolivet et al. 1994).

    2.4.2.2 Formation of Oxo Ligands

    As mentioned previously, an oxo ligand is an O²− anion bonded to a metal M within a complex. It is formed by the deprotonation of a hydroxo ligand according to the following acid-base reaction (Livage et al. 1988):

    $$ {\displaystyle \begin{array}{c}{\left[\mathrm{M}\left(\mathrm{OH}\right){\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-1}\right]}^{\left(z-1\right)+}+{\mathrm{H}}_2\mathrm{O}\leftrightarrow {\left[\mathrm{M}\mathrm{O}{\left({\mathrm{O}\mathrm{H}}_2\right)}_{N-1}\right]}^{\left(z-2\right)+}+{\mathrm{H}}_3{\mathrm{O}}^{+}\\ {}\mathrm{Acid}+\mathrm{Lewis}\kern0.5em \mathrm{base}\leftrightarrow \mathrm{conjugated}\kern0.5em \mathrm{base}+\mathrm{conjugated}\kern0.5em \mathrm{acid}\end{array}} $$

    (2.43)

    The product obtained, [MO(OH2)N−1](z−2)+, is an aquo-oxo complex since it contains both water and oxo ligands. Nevertheless, oxo-hydroxo and oxo-hydroxo-aquo complexes, with general chemical formula [MOx(OH)y(OH2)Nxy](zy−2x)+, also exist. An example is the vanadium complex [VO(OH)2(OH2)3]+ (Jolivet et al. 1994).

    2.4.2.3 Application of the Partial Charge Model to the Hydrolysis of Cations

    According to the partial charge model , the electronegativity of any complex can be calculated from its formal charge and from the Allred and Rochow electronegativity of each atom present in the complex. The electronegativity of an aquo-hydroxo complex C = [M(OH)h(OH2)N.h](zh)+ is for instance (Jolivet et al. 1994)

    $$ \chi \left(\mathrm{C}\right)=\frac{\sqrt{\chi_{\mathrm{M}}^0}+N\sqrt{\chi_{\mathrm{O}}^0}+\left(2N-h\right)\sqrt{\chi_{\mathrm{H}}^0}+1.36\left(z-h\right)}{\frac{1}{\sqrt{\chi_{\mathrm{M}}^0}}+\frac{N}{\sqrt{\chi_{\mathrm{O}}^0}}+\frac{2N-h}{\sqrt{\chi_{\mathrm{H}}^0}}} $$

    (2.44)

    If the electronegativity of this complex, χ(C), is equal to the electronegativity of water, then

    $$ \chi \left({\mathrm{H}}^{+}\right)=\chi \left({\mathrm{H}}_2\mathrm{O}\right)=2.631-0.02\;\mathrm{pH}=\chi \left(\mathrm{C}\right) $$

    (2.45)

    which permits to derive the equation (Jolivet et al. 1994)

    $$ h=\left(\frac{1}{1+0.014\mathrm{pH}}\right)\left(1.36z-N\left(0.236-0.038\mathrm{pH}\right)-\frac{2.621-0.02\mathrm{pH}-{\chi}_{\mathrm{M}}^0}{\sqrt{\chi_{\mathrm{M}}^0}}\right) $$

    (2.46)

    For instance

    $$ \mathrm{at}\;\mathrm{pH}=0;h=1.36z-0.24N-\frac{2.621-{\chi}_{\mathrm{M}}^0}{\sqrt{\chi_{\mathrm{M}}^0}} $$

    (2.47)

    $$ \mathrm{at}\;\mathrm{pH}=14;\kern0.36em h=1.14z+0.25N-\frac{0.836\left(2.341-{\chi}_{\mathrm{M}}^0\Big)\right.}{\sqrt{\chi_{\mathrm{M}}^0}} $$

    (2.48)

    Equation (2.46) can be used to estimate the number h of hydroxo ligands present in an aquo-hydroxo complex formed with a metal M. Likewise, other corresponding expressions exist for all types of complexes involving oxo ligands. The most acidic and the most basic forms of some metal cations as calculated by the partial charge model are reported in Table 2.4.

    Table 2.4

    Most acidic and most basic complexes for some metal with formula [MONH2N−h](z−h)+, together with the experimental and calculated value of h. The number h of hydroxo ligands is estimated by the partial charge model

    Enjoying the preview?
    Page 1 of 1