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

Only $11.99/month after trial. Cancel anytime.

Surface and Interface Chemistry of Clay Minerals
Surface and Interface Chemistry of Clay Minerals
Surface and Interface Chemistry of Clay Minerals
Ebook848 pages6 hours

Surface and Interface Chemistry of Clay Minerals

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Surface and Interface Chemistry of Clay Minerals, Volume 9, delivers a fundamental understanding of the surface and interface chemistry of clay minerals, thus serving as a valuable resource for researchers active in the fields of materials chemistry and sustainable chemistry. Clay minerals, with surfaces ranging from hydrophilic, to hydrophobic, are widely studied and used as adsorbents. Adsorption can occur at the edges and surfaces of clay mineral layers and particles, and in the interlayer region. This diversity in properties and the possibility to tune the surface properties of clay minerals to match the properties of adsorbed molecules is the basis for study. This book requires a fundamental understanding of the surface and interface chemistry of clay minerals, and of the interaction between adsorbate and adsorbent. It is an essential resource for clay scientists, geologists, chemists, physicists, material scientists, researchers, and students.

  • Presents scientists and engineers with a resource they can rely on for their own research and work involving clay minerals
    • Includes an in-depth look at ion exchange, adsorption of inorganic and organic molecules, including polymers and proteins, and catalysis occurring at the surfaces of clay minerals
    • Includes materials chemistry of clay minerals with chiral clay minerals, optical materials and functional films
    LanguageEnglish
    Release dateNov 5, 2018
    ISBN9780081024331
    Surface and Interface Chemistry of Clay Minerals

    Related to Surface and Interface Chemistry of Clay Minerals

    Titles in the series (7)

    View More

    Related ebooks

    Earth Sciences For You

    View More

    Related articles

    Reviews for Surface and Interface Chemistry of Clay Minerals

    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

      Surface and Interface Chemistry of Clay Minerals - Elsevier Science

      Surface and Interface Chemistry of Clay Minerals

      First Edition

      R. Schoonheydt

      Center for Surface Chemistry and Catalysis, Catholic University of Leuven, Leuven, Belgium

      C.T. Johnston

      Departments of Earth, Atmospheric, and Planetary Sciences and Agronomy, Purdue University, West Lafayette, IN, United States

      F. Bergaya

      Interfaces, Confinement, Matériaux et, Nanostructures (ICMN), Centre National de la, Recherche Scientifique (CNRS), Orléans, France

      Series Editor

      F. Bergaya

      Table of Contents

      Cover image

      Title page

      Copyright

      Dedication

      Contributors

      Preface

      Acknowledgements

      1: Clay minerals and their surfaces

      Abstract

      1.1 TO or 1:1 and TOT or 2:1 clay minerals

      1.2 Structural considerations

      1.3 Isomorphous substitution

      1.4 Consequences of isomorphous substitution

      1.5 Surfaces, surface areas, and surface sites

      1.6 Surface atoms

      1.7 Molecule–molecule and molecule–surface interactions

      2: Determination of surface areas and textural properties of clay minerals

      Abstract

      2.1 Introduction

      2.2 Nonswelling and nonmicroporous clay minerals

      2.3 Microporous clay minerals

      2.4 Swelling clay minerals

      2.5 Concluding remarks

      3: Quantum-chemical modelling of clay mineral surfaces and clay mineral–surface–adsorbate interactions

      Abstract

      3.1 Quantum mechanical description of interatomic interactions

      3.2 Simulations of clay minerals structure

      3.3 Elastic properties of clay minerals

      3.4 Redox processes

      3.5 Interaction of clay minerals with organic compounds

      3.6 Acid–base properties of edge surfaces and cation complexation

      3.7 Outlook

      4: Clay mineral–water interactions

      Abstract

      4.1 Introduction

      4.2 Water interactions with ‘neutral clay mineral surfaces’

      4.3 Water interactions with ‘charged clay mineral surfaces’ (ion–dipole)

      4.4 Molecular probe and reporter group studies of smectite–water interactions

      4.5 Probing the hydrophobic/hydrophilic character of clay mineral surfaces

      4.6 Conclusions

      5: Adsorption of heavy metals including radionuclides

      Abstract

      5.1 Clay mineral adsorption mechanisms and modelling

      5.2 Adsorption of heavy metals and radionuclides on 2:1 clay minerals

      5.3 Adsorption of iron on Mt

      6: From transition metal ion complexes to chiral clay minerals

      Abstract

      6.1 Introduction

      6.2 Stereochemistry of a clay mineral surface

      6.3 Chirality recognition by a clay mineral surface modified with metal complexes

      6.4 Solid-state VCD towards molecular recognition on a clay mineral surface

      6.5 Summary and future development

      Appendix. Basic strategy of applying vibrational circular dichroism (VCD) spectroscopy to solid or film samples

      7: Organic pollutant adsorption on clay minerals

      Abstract

      7.1 Pollutants? Definitions and scope of the chapter

      7.2 Classification of pollutants

      7.3 Pollutant adsorption mechanisms on clay minerals: An overview

      7.4 Tools of the trade: Experimental studies of organic pollutants adsorption

      7.5 Adsorption: A gateway to reactivity

      7.6 Conclusion

      Appendix: A list of pollutant structures

      8: Protein adsorption on clay minerals

      Abstract

      8.1 Introduction

      8.2 General considerations on protein adsorption

      8.3 Methodology to study protein adsorption

      8.4 Parameters that influence protein adsorption

      8.5 An overview of the adsorption of different proteins

      8.6 Conclusion

      9: Clay mineral catalysts

      Abstract

      9.1 Introduction

      9.2 Structural formula of some 2:1 clay minerals

      9.3 Properties of Mt

      9.4 Modified Mt for solid acid catalysis

      9.5 Nanoporous Mt supported metal nanoparticles catalysts

      9.6 Conclusion

      10: From polymers to clay polymer nanocomposites

      Abstract

      10.1 Introduction

      10.2 Clay minerals used in clay mineral polymer nanocomposites

      10.3 Structures and surface properties of clay minerals

      10.4 CPN obtained by cation exchange of a hydrophilic polymer with long alkyl chain/or a cationic monomer

      10.5 CPN obtained by grafting of organophilic polymers with a hydrophilic group

      10.6 CPN obtained by melt intercalation of a pristine clay mineral with hydrophilic or organophilic polymer and surfactant

      10.7 Other strategies of CPN synthesis with organophilic polymers

      10.8 Challenge: Exfoliation of clay minerals in CPN

      10.9 Properties and potential applications of CPN

      10.10 Improved properties of CPN—Some recent examples

      10.11 Conclusions and perspectives

      11: From adsorbed dyes to optical materials

      Abstract

      11.1 Introduction

      11.2 Controlled adsorption of dyes onto clay minerals

      11.3 Flattening of the π-electron system

      11.4 Suppression of vibrational motion

      11.5 Switching of the optical properties of the hybrids

      11.6 Future prospective

      12: Preparation and application of clay mineral films

      Abstract

      12.1 Introduction

      12.2 Casting

      12.3 Spin coating

      12.4 LbL deposition

      12.5 LB and LS methods

      12.6 Other methods

      12.7 Concluding remarks

      Index

      Copyright

      Dedication

      Robert Schoonheydt has only nice memories of clay mineral science with excellent PhD students and postdocs in the lab, with truly international scientific collaboration with universities and laboratories in China, India, Japan, the United States, Hungary, Italy, Romania, Spain, and France, to name the most important ones only. He hopes that clay mineral science continues to develop with the positive spirit that he experienced over the 45 years of his career.

      Contributors

      Preface

      R. Schoonheydt; C.T. Johnston; F. Bergaya

      Clay minerals are the most important components of clay, probably the first material that was exploited and used by mankind several thousand years ago. Since these early times, the areas of application of clays and clay minerals have continuously expanded and continue to do so. These applications are summarized in the table below and are somewhat arbitrarily divided into traditional and more recent application areas.

      Table: Areas of application of clays and clay minerals

      All together, these areas of application represent a multibillion kilogram business every year. This success has been made possible by (1) clay's easy availability, with clay deposits in almost every country of the world and (2) the unique physical and chemical properties of the clay minerals.

      Clay minerals, a group of phyllosilicates, are two-dimensional silicates with sizes < 2 μm. This small size means large surface areas. As a consequence, surface chemistry is at the core of the development of clay mineral applications. Surface chemistry is defined as the interaction between adsorbed molecules or ions and the surface atoms of clay minerals. Any molecule or ion can be adsorbed, from small, inert molecules such as N2 all the way to (bio)polymers. The interaction with the surface might be physical, that is, the adsorbed molecule retains its identity, or chemical, that is, a chemical bond is established between a surface atom and the adsorbed molecule.

      To achieve the optimum combination of adsorbed molecule and the clay mineral surface, the latter can be manipulated. Clay mineral surfaces can be realized which fully cover the hydrophilic-to-hydrophobic range. Layer charge, and thus, chemical composition, exchangeable cations, and ion exchange are the most important for realizing this hydrophilic–hydrophobic range. But, as clay minerals occur in nature, water is the most important adsorbed molecule.

      This book starts with an introductory chapter on clay minerals, their internal and external surfaces, surface atoms, and the forces between adsorbed molecules and surface atoms (Chapter 1). This chapter is followed by a fundamental discussion on surface area determination (Chapter 2); the surface area is the basis for surface chemistry and development of materials. Quantum chemical modelling accompanies the interpretation of experimental data and allows extrapolation beyond the experiment (Chapter 3). As clay minerals are found in nature, they contain water, and the study of adsorbed water is as old as the study of clay minerals themselves (Chapter 4).

      The two following chapters treat inorganic cations. In Chapter 5, the emphasis is on radioactive cations; it forms the basic knowledge for radioactive waste containment. With ion exchange of chiral metal ion complexes, chiral clay minerals can be produced with interesting applications in chromatographic separation and chiral catalysis (Chapter 6).

      The following four chapters deal with the interaction of clay mineral surfaces with organic molecules, which is an extremely important research area. This knowledge is necessary for pollution control and pollution abatement (Chapter 7 and Chapter 8). Organic molecules can be synthesized with clay minerals as catalysts or carriers of catalytic particles (Chapter 9). Organic polymers, including biopolymers, interact strongly with clay mineral surfaces; this leads to clay polymer nanocomposites with improved properties, at least when compared to those of the neat polymers (Chapter 10).

      The final two chapters give a hint about possible new developments: the transformation of clay minerals into optical materials (Chapter 11) and clay mineral films (Chapter 12). In both chapters, the organization of adsorbed molecules and of the individual clay mineral layers in a nanometre thick film has to be precisely controlled so as to generate clay mineral nanocomposites with a range of possible new application areas.

      The editors are convinced that the readers will find knowledge and data on the surface and interface chemistry of clay minerals in this book and inspiration for new research.

      July 2018

      Acknowledgements

      The three volume editors (R. Schoonheydt, C. T. Johnston, and F. Bergaya) would like to thank all the authors and their coworkers for their contributions to this book and for delivering chapters with a high scientific level on time. Extensive written and oral discussions between the authors and the editors allowed improved chapters as a result.

      Cliff Johnston would like to express his deep appreciation to both Robert Schoonheydt and Faiza Bergaya for their tireless work on this project. As a junior faculty member, he first met Robert Schoonheydt about 30 years ago. Through his collaboration and sabbatical stays at K.U. Leuven with Robert, he was introduced to many excellent European clay scientists, including Faiza Bergaya in 1995, and to a number of outstanding Japanese clay scientists that included Akihiko Yamagishi and Hisako Sato, all of whom challenged and shaped his thinking about clay science.

      Faiza Bergaya, who is also acting as series editor, would particularly like to thank Amy Shapiro (the acquisitions editor, Earth and Planetary Sciences, Elsevier) and Tasha Frank (the senior editorial project manager, Elsevier) for their patience and their continuous help during the achievement of this book. She thanks Robert Schoonheydt and Cliff Johnston for having accepted the huge editorial task of this Volume 9 and for the friendly relationship established during their close collaboration in these past months. Moreover, she would like to thank Susan Dennis (the acquisitions editor, Physical Sciences Series, Elsevier) for her confidence for several years.

      The author of Chapter 9 is grateful to Dr D. Ramaiah, Director, CSIR-North East Institute of Science and Technology, Jorhat-785006, Assam, India, for his kind support. The author is also grateful to CSIR, New Delhi, for sanctioning the Emeritus Scientist Scheme (No. 21(1029)/16/EMR-II).

      Finally, the support of the Elsevier team is highly appreciated by the three editors.

      1

      Clay minerals and their surfaces

      Robert A. Schoonheydt*; Cliff T. Johnston†; Faïza Bergaya‡    * Centre for Surface Chemistry and Catalysis, Catholic University of Leuven, Leuven, Belgium

      † Departments of Earth, Atmospheric, and Planetary Sciences and Agronomy, Purdue University, West Lafayette, IN, United States

      ‡ Interfaces, Confinement, Matériaux et Nanostructures (ICMN), Centre National de la Recherche Scientifique (CNRS), Orléans, France

      Abstract

      Clay minerals are introduced with their main structural features and unit cell formulae. Isomorphous substitution creates a negative lattice charge, leading to cation exchange, swelling, and intercalation. Edge surfaces and interlayer space are defined together with the surface atoms: O, OH, and exchangeable cations. They are of primary importance to understand the interaction of adsorbed molecules and ions with the surface of clay minerals and for surface–surface interactions. The forces involved in all these interactions are always attractive and induce the organization of the system, which is counteracted by the thermal energy.

      Keywords

      TO and TOT clay minerals; Isomorphous substitution; Cation exchange; Swelling; Intercalation; van der Waals energy; H-bonding; Lennard-Jones potential

      1.1 TO or 1:1 and TOT or 2:1 clay minerals

      Clay minerals belong to the phyllosilicate family (Bailey, 1988b; Brigatti et al., 2013). Each layer of these layered silicates consists of T sheets of corner-sharing Si tetrahedra, [SiO4], and O sheets of edge-sharing Al octahedra, [Al(OH)2O4]. The T and O sheets are connected with a common oxygen atom (Fig. 1.1).

      Fig. 1.1 Association of T and O sheets (example: schematic structure of kaolinite). Red balls ( dark grey in the print version) are oxygen; purple balls ( grey in the print version) are aluminium, and grey balls are hydrogen atoms.

      The possible combinations of T and O sheets among both TO and TOT layers lead to a classification of nine distinguished main groups of clay minerals reported in Table 1.1. The 2D connectivity of the planar layer at the molecular level should be distinguished from the morphology at an upper level of organization (Ras et al., 2007).

      Table 1.1

      The structure of TO (or 1:1) clay minerals layers, as shown in Fig. 1.1, is representative of the kaolinite structure. The chemical structural formula of the unit cell is [Si2Al2O5(OH)4] with two types of structural OH groups. The inner surface OH groups are located on the terminating surface of the Al-octahedral sheet. The H atoms are bonded to O atoms of the O sheet and point towards the basal oxygen atoms of the opposing T sheet (Fig. 1.1) of the next layer with which they engage in H-bonding (Johnston, 2010). The second type of structural OH group is the inner OH. It is the Al–OH group located on the common plane of oxygen atoms joining the O and T sheets (Fig. 1.1). The sum of the formal charges on the atoms in the TO structure equals zero (Si = + 4; Al = + 3; H = + 1, O = − 2).

      The structure of TOT of 2:1 clay minerals layers (Fig. 1.2) is composed of an O sheet (composed of Al³ +, Fe³ + or Mg² + cations), sandwiched between two Si⁴ + T sheets. The chemical formula of the unit cell is [Si8Al4O20(OH)4] or [Si4Al2O10(OH)2] for a half unit cell, and is represented by the clay mineral pyrophyllite. In contrast to the TO structure (e.g., kaolinite), the OH groups of TOT structures (e.g., pyrophyllite) are all inner OH groups. When Al³ + cations occupy the centre of the octahedral cavities in the O sheet, the TOT or TO layer is called dioctahedral. If the O sheet is filled with Mg² + cations, the TOT or TO layer is called trioctahedral, and the chemical formula of the TOT unit cell is [Si8Mg6O20(OH)4]. A representative example of the TOT clay mineral is talc. An example of a trioctahedral TO clay mineral is serpentine, with [Si2Mg3O5(OH)4] as the chemical formula of the unit cell. All these reported unit cell compositions are ideal as they do not carry any permanent structural charge on the basal surfaces (Brigatti et al., 2013).

      Fig. 1.2 Representation of a TOT layer with an O sheet sandwiched between two T sheets: (example: structure of dioctahedral montmorillonite showing exchangeable cations and interlayer water). Red balls ( grey in the print version) are oxygen; gold balls ( light grey in the print version) are silicon, green balls ( light grey in the print version) are magnesium, purple balls ( grey in the print version) are aluminium, and grey balls are hydrogen atoms.

      1.2 Structural considerations

      The Si tetrahedra in the tetrahedral sheet each have three corner-shared oxygen atoms. The fourth or apical oxygen atom points towards the octahedral sheet. The chemical representation of the tetrahedral sheet is [(Si2O5)² −]n. The arrangement of these dimers in one sheet leads to the hexagonal arrangement of the Si tetrahedra, as shown in Fig. 1.3.

      Fig. 1.3 [001] projection of T and O sheets (example: schematic structure of kaolinite). Silicon tetrahedrons shown in gold ( light grey in the print version) (corner shared), and aluminium octahedrons shown in purple ( dark grey in the print version) (edge shared).

      O bond length is on average 0.161 nm, leading to the ‘ideal’ unit cell dimensions given in Table 1.2.

      Table 1.2

      Similarly, the Al octahedra form a sheet by sharing their edges (Fig. 1.3), and the cation positions form regular hexagons. The unit cell dimensions are given in Table 1.2. Such a sheet is essentially the same as the Al(OH)3 sheet of the gibbsite lattice. In the case of Al³ + octahedra, two-thirds of the cationic positions are occupied. The TOT and TO clay minerals with Al in the octahedral sheet are called dioctahedral. For divalent cations such as Mg² +, the octahedral sheet is identical to the Mg(OH)2 sheet of brucite, where all of the cationic positions are filled. The corresponding TOT and TO clay minerals are called trioctahedral. The unit cell dimensions of the tetrahedral and octahedral sheets are close (Table 1.2). They can be linked by sharing the apical oxygens of the Si tetrahedra. There is, however, a small mismatch which is compensated by rotation of the Si tetrahedra. The hexagonal array is distorted from an ideal hexagonal structure to a ditrigonal structure where three O atoms move inwards, three O atoms move outwards, and the Si–O–Si angle takes a value between 180 and 120 degrees, 180 degree being for the ideal hexagonal array (Fig. 1.4) (Bailey, 1988a).

      Fig. 1.4 Hexagonal array of Si-tetrahedrons showing ditrigonal symmetry.

      Other factors influencing this angle and deformations in octahedral and tetrahedral sheets result from isomorphous substitution and defects.

      1.3 Isomorphous substitution

      Heterovalent isomorphous substitution (referred to as isomorphous substitution) is the substitution of a cation in the lattice by a cation of lower charge and about equal size. Examples are Si⁴ + by Al³ + in tetrahedral sheets, Al³ + by Mg² + in octahedral sheets of dioctahedral layers, and Mg² + by Li+ in octahedral sheets of trioctahedral layers (Schoonheydt and Johnston, 2013). Fe is invariably present in nature and in the structure of clay minerals. Fe³ + can substitute for Si⁴ + in tetrahedral sheets; Fe² + and Fe³ + can substitute for Al³ + in octahedral sheets. Ionic radii of the cations and O atoms, the ratios of the radii of the cations, rc to that of the oxygens, ro (rc/ro) and the preferred coordination environment are given in Table 1.3 (Huheey et al., 1997).

      Table 1.3

      From Huheey, J.E., Keiter, E.A., Keiter, R.L., 1997. Inorganic Chemistry. Principles of Structure and Reactivity. Prentice Hall, New York, pp. 737.

      The sheets comprising the structures of clay minerals can be considered as sheets of closely packed O atoms. The cations merely fill the holes between the O atoms. This is illustrated in Fig. 1.5 for the Al octahedron and Si tetrahedra.

      Fig. 1.5 Packed oxygen atoms of (1) a tetrahedral sheet ( left ) shown in space filled ( left side ) and polyhedral representation. Similar depiction of the (2) octahedral sheet shown on the right in both space filled and polyhedral representation. The polyhedron shown in green represents isomorphic substitution of Mg for Al.

      When the atoms are considered as hard spheres one obtains 2ro/(2rc + 2ro) = cos 45 degree = 0.707 or rc/ro = 0.414 in the case of the octahedron. For tetrahedral coordination, one obtains the radius ratio rc/ro = 0.225. Cationic radii with rc/ro between 0.414 and 0.225 prefer tetrahedral coordination and substitute preferentially in the tetrahedral sheet according to Pauling's 1st rule. If rc/ro is larger than 0.414, octahedral coordination is preferred or substitution in the octahedral sheets. This is given in the second column of Table 1.3. In reality, the ions are not hard spheres, and the rc/ro ratios do not strictly obey the rule but must be close. The most common TOT clay minerals with isomorphous substitution are reported in Table 1.4 with their unit cell formula and their names.

      Table 1.4

      IV and VI refer, respectively, to the tetrahedral and octahedral sheets.

      Often isomorphous substitution occurs simultaneously in the tetrahedral and octahedral sheets. The extent of isomorphous substitution in the tetrahedral sheet is represented by x, and that in the octahedral sheet by y. The total amount of isomorphous substitution is then x + y. The consequence of isomorphous substitution is the presence of excess negative charge: the sum of the negative charges in the unit cell is larger than the sum of the positive charges. The excess negative charge in the structure is compensated by cations, indicated by M+ in Table 1.4. These cations do not belong to the structure and are exchangeable by other cations. They are called exchangeable cations. For the smectites (or swelling clay minerals) of Table 1.4, the amount of negative charges per unit cell is x + y = 0.4–1.2. Higher charges occur for vermiculites (x + y = 1.2–1.8) and micas (x + y = 2–4).

      Layer charge density is the average charge per unit surface. It is 0.4/nm²–1.3/nm² for smectites. This is an important quantity, as it determines the amount and molecular arrangement of organic exchangeable cations, especially in the case of bulky cations such as long chain alkylammoniums and cationic dyes. Thus, if the size of the monovalent exchangeable organic cation exceeds the average surface area per charge in the structure, the exchangeable organic cations must adopt special configurations in order to compensate the negative lattice charge sites (Laird et al., 1989). These configurations might include bilayers, dimers, trimers, and higher aggregates, and even crystalline compaction. Layer charge has two important consequences that has 3 important consequences that are discussed in Section 1.4.

      1.4 Consequences of isomorphous substitution

      1.4.1 Cation exchange

      Cation exchange is the substitution of exchangeable cations originally present, B, by other types of cations in solution (A) (Bergaya et al., 2013). The reaction can formally be written as

         (1.1)

      where CM means clay mineral.

      This equilibrium reaction is characterized by a thermodynamic equilibrium constant, Kc given by

         (1.2)

      where () are activities.

      Activities are expressed as the product of the concentration and activity coefficients. In solution, the activity of the aqueous species, (Am +) can be represented by (Am +)n = [Am +]nγMn where [Am +] is the concentration of A in moles dm− 3 and γMn is the single ion activity coefficient. There are no simple expressions for activities of ions on solid surfaces. Gaines and Thomas (1953) assumed that the activity of the cation on the solid phase was represented by its equivalent fractions, while Vanselow used molar fractions of the ions on the solid (Sposito, 1981a,b):

         (1.3)

         (1.4)

         (1.5)

      If m = n, the Gaines–Thomas and Vanselow selectivity coefficients are identical.

      1.4.2 Cation exchange capacity

      Cation exchange capacity (CEC) is the amount of a cation that can be exchanged by another cation on the surface of a clay mineral. It is expressed in cmol(+)/kg, which is numerically equivalent to meq/100 g, where mol(+) represents moles of electrical charge. Theoretically, it can be obtained from the amount of isomorphous substitution, as shown by the structural formulae of the unit cells in Table 1.4. For instance, for beidellite, with Na+ as the exchangeable cation, the CEC is 0.27 mmol/g if x = 0.2 and 0.81 mmol/g if x = 0.6. This is called the pH-independent CEC.

      At the edges of the clay mineral layers (Figs 1.1 and 1.2), the O atoms are coordinatively unsaturated and pick up protons from the aqueous environment, thus forming surface OH groups. They can pick up a proton or lose their proton, depending on the pH conditions. In acidic conditions, the edges are positively charged; at alkaline pH, they are negatively charged. In between, there is a pH at which the sum of the positive charges equals the sum of the positive charges. This is called the point of zero charge. The OH groups at the edges are bonded to Si, Al, or Mg and the point of zero charge depends on the chemical composition of the clay mineral layers, particle size, presence of impurities, and method of determination of the point of zero charge. Values reported in the literature can vary over the range pH 3–7 (Lagaly, 2006).

      The protons at the edges are exchangeable, and charge resulting from these edge sites is pH-dependent and increases with increasing pH. This contribution to the CEC is called the pH-dependent CEC. Factors influencing this pH-dependent CEC are particle size, degree of swelling (see Section 1.4.3), and type of cation. The sum of the pH-independent and pH-dependent exchange capacities is the total CEC.

      The CEC is dependent on the type of exchangeable cation for two reasons:

      (1)Not all the exchange sites are available or accessible. This is because the dilute clay mineral dispersion is composed of single layers, particles, and aggregates of particles. In these particles, not all exchange sites might be accessible.

      (2)The charge density of the particles might be different from particle to particle. Exchange sites are accessible dependent on this charge density. If it is too high, swelling is limited, and cations might not be able to exchange on these high charge density particles.

      The exchange of bulky cations, such as metal ion complexes, organic cations, and cationic polymers, including cationic proteins, renders the surface hydrophobic, thus creating an environment for adsorption of these cations in excess of the CEC. This means that the bulky cations with their associated anions prefer to be located in the hydrophobic environment of the clay mineral particles and not in the hydrophilic water environment of the solution. The amount of adsorbed cations exceeds the CEC.

      1.4.3 Intercalation and swelling

      Swelling is the spontaneous increase of the interlayer distance of clay mineral particles, due to the adsorption of molecules in the interlayer space. As a consequence, the c dimension of the unit cell increases, while the a and b dimensions remain constant. It is observed in the XRD spectrum by a shift of the d001 line to lower 2θ values. The process is governed by (i) the charge density of the clay mineral layers; (ii) the type of exchangeable cation in the interlayer space; and (iii) the type of adsorbing molecule or ion.

      If the charge density of the clay mineral layers is zero or close to zero, there are no exchangeable cations in the interlayer space, and the clay mineral does not swell. Such is the case of kaolinite, pyrophyllite, and talc. The charge density of smectites, σ = 0.4–1.2e per unit cell, is ideal for swelling in water with Li+ and Na+ as the exchangeable cations in the interlayer space. The interlayer distance can increase to infinity, or the smectite particles are disassembled into single layers, diffusing randomly in the liquid water. If the charge density is above 1.2e per unit cell, swelling in water is limited to one or two layers of water molecules in the interlayer space. This is the case of vermiculites. When the charge density is 2 or higher, as for micas, swelling does not occur.

      In general terms, swelling occurs because the attractive forces between the exchangeable cations and the adsorbing molecules overcome the attractive forces between the negative charge density of the clay mineral layers and the exchangeable cations. These attractive forces are influenced by the charge density of the clay mineral particles, by the size and shape of these particles, and by the charge of the exchangeable cations.

      There are two limiting cases of intercalation of molecules in the interlayer space.

      (1)Random intercalation: there is no preference of the incoming molecules for any of the clay mineral particles in the sample, and within one particle, the molecules are randomly distributed in the interlayer space.

      (2)Interstratification: the interlayer spaces are filled up successively. Interlayer spaces might be filled up randomly or in organised fashion (Fig. 1.6).

      Fig. 1.6 Possible organization of molecules ( black balls ) in the interlayer space.

      Random intercalation is only possible if all clay mineral particles in the sample are identical in terms of charge density, number of layers per particle, size, and shapes of the particles. Such an ideal situation does not occur. In the case of interstratification, the incoming molecules favour the particles that offer the largest attractive forces towards the molecules. Once one molecule has entered the interlayer space, it has opened an easy pathway for subsequent molecules, until at some loading, the molecule–molecule repulsion prohibits the entry of additional molecules. In reality, intercalation is a mix of random intercalation and interstratification.

      XRD experiments give some hints as to what is happening. The d001 line reflects the distance between two successive clay mineral layers in the clay mineral particles of the sample. This distance is the sum of the height of the clay mineral layer or 0.96 nm in the case of smectites, and the thickness of the interlayer space. Thus, as the thickness of the interlayer space increases with adsorption of molecules in the interlayer space, so will the d001 line shift to smaller values. In the case of random intercalation, the d001 line shifts gradually as more molecules are intercalated. In the case of interstratification, different d001 lines may be observed dependent of the degree and the order of interstratification.

      1.5 Surfaces, surface areas, and surface sites

      In a clay mineral sample, there are single layers, particles, and aggregates of particles. A single layer has a planar surface and an edge surface. A particle has an external surface and the interlayer space. An aggregate has an external surface, interlayer surface, and pores (Fig. 1.7).

      Fig. 1.7 Definition of planar surfaces, edge surfaces and of interlayer space From Bergaya, F., Theng, B.K.G., Lagaly, G., 2006. Handbook of Clay Science. Elsevier, Amsterdam.

      There are then two types of porosity in a clay mineral sample: the interlayer space and the pores between the particles in the aggregates. The interlayer space is available for adsorption if the molecules or pillars keeping the clay mineral layers apart have free space between them for adsorption of incoming molecules or can be replaced/substituted by the incoming molecules. The interlayer space is usually microporous; that is, the free space between the layers is less than 2 nm. The pores in aggregates are interparticle pores. They can be mesoporous (2 nm < d < 50 nm) or macroporous (d > 50 nm).

      The surface area of a clay mineral is the surface which is occupied by the probe molecule used to determine the surface area. This is usually N2. One N2 molecule occupies 0.16 nm² (i.e., cross-sectional area). If one can determine the number of adsorbed N2 molecules per unit mass the surface area can be obtained. Several situations can be envisaged. (1) The interlayer space is filled up with molecules, usually water. In this case, N2 cannot substitute for water in the interlayer space, and one measures only the external surface area of the clay mineral sample. (2) The interlayer space is filled with pillars or bulky molecules, and the distance between the molecules or pillars is larger than the size of the N2 molecule. Upon adsorption, the N2 molecules occupy the external surface area and the free interlayer surface area not occupied by the pillars/bulky molecules.

      If the goal is to measure the interlayer surface area, a probe molecule is necessary which is strongly/preferentially adsorbed in the interlayer space and forms a closely packed monolayer. From the amount adsorbed and the size of the molecule, one calculates the interlayer surface area (see Chapter 2). A popular molecule with which to do so is ethylene glycol.

      1.6 Surface atoms

      The overwhelming majority of the surface atoms of clay minerals are oxygen atoms (Figs 1.1 and 1.2). The O atoms of the Si tetrahedral sheets of smectites and kaolinite are coordinatively saturated. They are attractive towards adsorbing molecules by dispersion or nonpolar interactions and through electrostatic interaction for clay minerals with isomorphous substitution, because they carry a negative charge. The Al octahedral sheet of a kaolinite layer is terminated with OH groups (see Fig. 1.1). These OH groups are polar and capable of H-bonding between layers. These bonds keep the kaolinite layers together by H-bonding with the O atoms of the Si tetrahedral sheet of the opposing layer in the particle. Consequently, these OH groups are not accessible for adsorption of molecules in the interlayer region of kaolinite. Adsorption is only possible with strongly polar molecules that are capable of breaking up the H-bonding between successive layers of kaolinite in a particle. Examples of these types of polar molecules include hydrazine, dimethyl sulfoxide, potassium acetate, and urea.

      The O atoms at the edges of clay mineral particles are coordinatively unsaturated and carry a negative charge. The interaction of these O atoms with an aqueous environment is pH dependent:

      The OH groups are polar and interact with adsorbing molecules by polar–polar, polar–apolar, and dispersion interactions. They may also adsorb cations by a surface complexation reaction:

         (1.6)

      In the interlayer space of smectites, the O atoms of the Si tetrahedral sheets and the exchangeable cations are the main adsorption sites (Fig. 1.2). Polar molecules such as water complete the coordination sphere of the exchangeable cations and have H-bonding interactions with the O atoms of the Si tetrahedral sheets and with each other, thus forming one or two layers hydrates. Traditionally, smectites have been considered to be hydrophilic materials. However, if the exchangeable cations are exchanged with organic cations such as alkylammonium cations ([H4 − xN-(R)x]+, x = 1–4; R = CnH2n + 1), the surface is rendered hydrophobic. In this case, the positively charged headgroup is adsorbed on the siloxane surface to effectively neutralize the negative charges of the smectite layers. The alkyl chains can adopt different configurations, depending on the length of the chain and the charge density of the clay mineral layers. Hydrophilic and hydrophobic clay minerals are capable of adsorbing different types of molecules: polar molecules in the first case, apolar molecules in the second case. Intermediate situations can be created by exchanging bulky inorganic cationic complexes, cationic dyes, and cationic polymers in the interlayer space of smectites. These cations do not leave much space for adsorption in the interlayer space and create an amphiphilic environment.

      The adsorption phenomena discussed so far are examples of physical adsorption (or physisorption) phenomena. Upon adsorption, the adsorbing molecules and the surface atoms retain their identity, and no chemical bond is established. Chemisorption occurs when a chemical bond is established between the surface atoms and the adsorbing molecules. Both the surface and the molecule must carry reactive groups. Examples are:

         (1.7)

         (1.8)

      A comparison between physisorption and chemisorption is given in Table 1.5. Physisorption is a weak interaction between the adsorbed molecules and the surface and leads to the formation of adsorbed mono- and multilayers. Such an ordering phenomenon is counteracted by temperature. Thus, the amount of physisorbed molecules decreases with increasing temperature. Physisorption and chemisorption are exothermic processes. A continuous scale of enthalpy of adsorption can be established from physisorption to chemisorption. However, it is difficult and dangerous to distinguish physisorption from chemisorption on the basis of enthalpy changes alone accompanying these phenomena. Other criteria must be taken into account, such as reversibility and temperature dependence. For example, a molecule might physisorb at low temperatures and chemisorb at high temperatures. Above all, the chemical bond between the surface and the molecule must be clearly established to conclude that the adsorption process is chemisorption.

      Table 1.5

      1.7 Molecule–molecule and molecule–surface interactions

      Physisorption or physicochemical interactions between molecules and surfaces involve molecule–molecule interactions and molecule–surface interactions. In the case of chemisorption, the molecule–surface interactions result in the formation of a chemical bond between the surface atoms and the adsorbed molecules. Additional physical interactions between the chemisorbed molecule and the surface atoms and intermolecular interactions between chemisorbed molecules may occur too, but they are weak when compared with the molecule–surface chemical bond formation and are usually neglected. In this chapter, three types of interactions are discussed: (1) molecule–molecule, (2) molecule–surface, and (3) surface-surface.

      1.7.1 Molecule–molecule interactions

      1.7.1.1 Ion-ion interactions

      If the molecules are ions, they will undergo electrostatic attraction and repulsion according to Coulomb's law:

         (1.9)

      With Qi and Qj representing the point charges of the ions and rij is the distance between the ions i and j. ɛ is the dielectric constant of the medium in which the ions reside, and ɛ0 is the vacuum permittivity. The use of ɛ implies that this medium is considered a continuum. The dielectric constant of the medium is required, because Coulomb's energy is due to a long-range electrostatic interaction, which decays with distance as r− 1. For all charges Qi and Qj in the medium, the total Coulombic interaction energy is

         (1.10)

      Charge–dipole interaction is the electrostatic interaction between a point charge Qi and a permanent point dipole as shown in Fig. 1.8. The electrostatic interaction energy is the sum of the Coulombic interactions between the point charge Qi and the charged qj and − qj on the permanent dipole:

         (1.11)

      Fig. 1.8 Ion–dipole interactions From Israelachvili, J.N., 1989. Intermolecular and Surface Forces. Academic Press, London.

      This expression can be rearranged into

         (1.12)

      In this case, the negative pole of the dipole is attracted towards the point charge; the positive pole is repelled. The result is that the dipole is aligned with the point charge as shown in Fig. 1.8. Such an orientation is counteracted by increasing temperature as the thermal energy induces a random orientation of the dipoles in the system. Eq. (1.12) then becomes

         (1.13)

      The charge–dipole interaction energy decreases with distance as r⁴. It is a short range interaction between neighbouring charges and dipoles. With short range interactions, the dielectric constant of the medium is usually set equal to 1. If one wants to express this interaction per mole of dipoles or per mole of charges, the interaction energy of Eq. (1.13) has to be multiplied by Avogadro's number Na.

      1.7.1.2 Dipole–dipole interactions

      The interaction between two fixed dipoles is the sum of the electrostatic attractions between the poles with opposite charges and the electrostatic repulsion between poles of like charges. The result is an alignment or organization of the dipoles as shown in Fig. 1.9. The maximum attractive interaction is obtained for θ1 = θ2 = 0: Edd = − μμ2/4πɛ0r³.

      Fig. 1.9 Possible arrangement of dipoles.

      For ϑ1 = 90 degree and ϑ2 = 0 as well as for ϑ1 = 0 and ϑ2 = 90 degree Edd = 0. For ϑ1 = ϑ2 = 90 degree, Edd is repulsive: μ1μ2/4πɛ0r³. If freely rotating dipoles are taken into account together with the thermal energy, the dipole–dipole interaction energy becomes

         (1.14)

      1/r⁶ indicates that Edd is a short range interaction: only neighbouring dipoles interact significantly. In this case, the dielectric constant is usually set equal to 1.

      1.7.1.3 Charge–nonpolar interaction

      A charge + Q interacts with a neighbouring nonpolar molecule as it attracts the electrons and repels the positively charged nuclei. Mainly, the electrons in the outer molecular orbitals or bonding molecular orbitals make a significant contribution to this interaction. The result is that a small dipole moment is induced in the nonpolar molecule, μi. The induced dipole moment is proportional to the electric field E of the charge + Q: μi = α E. The proportionality constant α is the polarizability of the molecule. With the induced dipole moment in Cm and E in volt, the dimensions of the polarizability are m³. The charge–nonpolar interaction energy is then

         (1.15)

      Also, dipoles will undergo a shift of their bonding electrons under the influence of an electric field of a charge Q. This means that a dipole moment will be induced, and that the total dipole moment is μT = μ + μi. In general, μi ≪ μ, and for dipolar molecules, the induced dipole moment is usually neglected.

      1.7.1.4 Dipolar–nonpolar interaction

      A nonpolar molecule in the immediate neighbourhood of a dipolar molecule is polarized in the same way as in the vicinity of a charge Q. The positive pole of the dipole attracts the bonding electrons and repels the nuclei. The negative pole does the reverse. For a fixed dipole, the interaction energy is given by

         (1.16)

      Averaged over all angles, that is, taking the thermal energy into account

         (1.17)

      1.7.1.5 Nonpolar–nonpolar interactions

      Two neighbouring nonpolar molecules attract each other. This is explained as follows. In a nonpolar molecule, the time-averaged distributions of negative charges (electrons) and positive charges (nuclei) coincide, and there is no dipole moment. However, at any time t, the distributions of the positive and negative charges do not coincide, and an instantaneous dipole moment is generated. The instantaneous dipoles of neighbouring molecules interact and are at the basis of the nonpolar–nonpolar interaction energy, also called dispersion energy of van der Waals energy. Empirical formulae for these interactions are:

      For two identical spherical molecules:

         (1.18)

      For two different spherical molecules:

         (1.19)

      where I represents the ionization potential of the molecules. One remarks that the dispersive energy is attractive, short range, and independent of temperature.

      Table 1.5 summarizes the expressions of the different types of intermolecular interaction energies. They have been calculated for a charge Q = 1.6 × 10− 19C, a dipole moment of 6.17 × 10− 30Cm or 1.85D (D = Debye) T = 298 K, r = 0.4 × 10− 9 nm (4 Å), and α = 2.886 × 10− 40 m³. The ionization energy is that of methane 12.6 eV or 1215 kJ/mol, and the dispersion energy is that of the methane–methane interactions. These energies have to be compared with those of the thermal energy, kT, which is 2.5 kJ/mol for T = 298 K. The following remarks have to be made.

      1.All the intermolecular interactions are attractive and tend to induce organization in the molecular systems in which they are active. Thermal energy, kT, induces disorder and counteracts the intermolecular interactions. At room temperature, kT equals 4.21 × 10− 21 J or 2.5 kJ/mol. This is larger than Edd, Edn, and Enn, and disorder is prevalent. Intermolecular interactions with ions are larger than the thermal energy, and order is to some extent present in these systems at short distance; that is, the ions organize the molecules in their immediate environment. In the case of water, the ions acquire a hydration shell. At 10 K, kT equals 1.38 × 10− 22 J or 73.8 J/mol. Organization will be prevalent in all molecular systems.

      2.Intermolecular interactions are additive. A molecule with dipole moment μ1 in the immediate environment of an ion with charge Q or another molecule with dipole moment μ2 undergoes small electron displacements, which give rise to an induced dipole moment μi. All molecules, whether they have a dipole moment or not, have temporary displacements of nuclei and electrons, giving rise to temporary dipole moments and leading to dispersive interactions. Thus, interacting molecules with a dipole moment have three types of intermolecular interactions: Edd, Edn, and Enn, the former being the largest (Table 1.6).

      Table 1.6

      Examples are calculated with the following parameters:

      Q = 1.6 × 10− 19 C (electron charge); r = 0.4 nm; μ = 1.85D or 6.17 × 10− 30 Cm;

      α = 2.88 × 10− 40 m³; 4πϵ0 = 1.11 × 10− 10; T = 298°K; I1 = I2 = 12.5 eV = 2 × 10− 18 J;

      k = 1.38 × 10− 23 JK−1; ɛ = 78.

      3.As the intermolecular interactions are attractive, they bring the molecules towards each other. This cannot continue indefinitely, because at short distance, electron–electron repulsion increases sharply, and at some characteristic distance, a minimum in energy is obtained at the equilibrium distance between the molecules. The attractive intermolecular interactions and the electron–electron repulsion are combined in the Lennard-Jones potential:

         (1.20)

      Its graphical representation is shown in Fig. 1.10. ELJ = 0 if r = σ, and the minimum ELJ = E0 is obtained for r = 2¹/⁶σ = 1.12σ.

      Fig. 1.10 The Lennard-Jones potential showing energy as a function of distance.

      1.7.1.6 H-bonding: X–H-----Y

      H-bonding is the—mainly—electrostatic interaction between a H atom bonded to an electronegative atom X such as F, O, and N, and the electronegative atom Y of the same molecule, a neighbouring molecule or a surface. Typical H-bonding interaction energies and electronegativities are given in Table 1.7.

      Table 1.7

      Enjoying the preview?
      Page 1 of 1