Hostname: page-component-7bb8b95d7b-2h6rp Total loading time: 0 Render date: 2024-10-02T14:31:13.119Z Has data issue: false hasContentIssue false

Density functional theory study of the atomic and electronic structures of trans-vacant 1M Al-rich illite

Published online by Cambridge University Press:  01 March 2024

Wei Gao
Affiliation:
State Key Laboratory of Geomechanics and Deep Underground Engineering, China University of Mining and Technology, Beijing, China School of Mechanics and Civil Engineering, China University of Mining and Technology-Beijing, Beijing, China
Jian Zhao*
Affiliation:
State Key Laboratory of Geomechanics and Deep Underground Engineering, China University of Mining and Technology, Beijing, China School of Mechanics and Civil Engineering, China University of Mining and Technology-Beijing, Beijing, China
Man-Chao He
Affiliation:
State Key Laboratory of Geomechanics and Deep Underground Engineering, China University of Mining and Technology, Beijing, China School of Mechanics and Civil Engineering, China University of Mining and Technology-Beijing, Beijing, China
*
Corresponding author: Jian Zhao; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Illite is a common clay mineral that is found in a wide range of geological settings. The good thermal stability and non-swelling properties of illite make it valuable in ceramic materials, paints and coatings, drilling fluids, agriculture and geological studies. To gain a deeper understanding of the physical and chemical properties of illite, in the present paper the atomic and electronic structures of a typical trans-vacant 1M Al-rich illite were constructed and calculated using density functional theory. The calculated indirect band gap of Al-rich illite was 4.99 eV. The electronic analysis revealed that the interactions in the tetrahedral sheet were more stable than those in the octahedral sheet. The substitution of Al atoms noticeably reduced the stability of the tetrahedral sheet in Al-rich illite. Other properties of Al-rich illite, including the density of states, electron population/charge, electronic charge density and bonding interaction, are also discussed and analysed in detail.

Type
Article
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press on behalf of The Mineralogical Society of the United Kingdom and Ireland

Illite is a type of mica-like phyllosilicate clay mineral. Its name was introduced by Grim et al. (Reference Grim, Bray and Bradley1937) and derived from the name of the state of Illinois, USA. Illite is a widespread mineral found in diverse geological environments, primarily originating from the weathering and alteration of other minerals such as micas and feldspars (Righi & Meunier, Reference Righi, Meunier and Velde1995; Shoval, Reference Shoval2023). The good thermal stability (He et al., Reference He, Makovicky and Øsbæck1995) and non-swelling properties (Ruiz Pestana et al., Reference Ruiz Pestana, Kolluri, Head-Gordon and Lammers2017) of illite make it valuable in a wide range of industrial and geological applications, including ceramic materials, paints and coatings, drilling fluids, agriculture and geological studies (Steiger, Reference Steiger1982; Feng et al., Reference Feng, Faiia, WoldeGabriel, Aronson, Poage and Chamberlain1999; Sedmale et al., Reference Sedmale, Randers, Rundans and Seglins2017; Li et al., Reference Li, Ou, Wang, Yan and Zhou2018; Zhang et al., Reference Zhang, Xu, Christidis and Zhou2020; El Halim et al., Reference El Halim, Daoudi, El Ouahabi and Fagel2022).

Existing experimental and theoretical data have confirmed the 2:1-type dioctahedral layered structure of illite, which consists of stacked layers primarily bound by van der Waals forces (Brigatti & Guggenheim, Reference Brigatti and Guggenheim2002). An individual 2:1 basic layer comprises two tetrahedral sheets sandwiching one octahedral sheet to form a T–O–T structure (Bergaya & Lagaly, Reference Bergaya and Lagaly2013). The tetrahedral sheets are composed of Si–O tetrahedrons with some of the Si atoms substituted by Al atoms, whereas the octahedral sheet consists of Al, Mg or Fe cations coordinated with OH groups (Meunier & Velde, Reference Meunier, Velde, Meunier and Velde2004). The interlayer space of illite primarily accommodates K ions, and the interlayer charge generally ranges between 0.60 and 0.85 per O10(OH)2 (Rieder et al., Reference Rieder, Cavazzini, D'Yakonov, Frank-Kamenetskii, Gottardi and Guggenheim1998). The octahedral sheet of illite exhibits two types of structure – cis-vacant and trans-vacant – determined by the configuration of the coordinating OH groups (Drits et al., Reference Drits, McCarty and Zviagina2006; Drits et al., Reference Drits, McCarty and Derkowski2012). In terms of polytypes, it has been observed that the 1M and 2M l polytypes are most frequently encountered in illite species (Zoeller & Brockamp, Reference Zoeller and Brockamp1997; Zviagina & Drits, Reference Zviagina and Drits2019).

Due to the complexity of the chemical components, disordered stacking, cation substitutions, cation configurations and the small crystalline domain size of illite, measuring its atomic structure and crystal properties accurately through experiments remains challenging. With advancements in theoretical research, computational chemistry calculations based on density functional theory (DFT) have been proven to be effective and reliable for investigating clay minerals at the molecular level (Bridgeman, Reference Bridgeman1996; Beermann & Brockamp, Reference Beermann and Brockamp2005; Mercier et al., Reference Mercier, Le Page and Desgreniers2010; He et al., Reference He, Zhao and Fang2012; Teich-McGoldrick et al., Reference Teich-McGoldrick, Greathouse and Cygan2012; Scholtzová & Tunega, Reference Scholtzová and Tunega2019; Yuan et al., Reference Yuan, Wang, He, Fang and Huang2022; Zhao et al., Reference Zhao, Wang, Luan, Cao and He2023). A series of studies has investigated the structure of illite using DFT. Stixrude & Peacor (Reference Stixrude and Peacor2002) discussed two avaliable competing models of illite smectite systems based on a first-principles study. The results revealed that differences in energy and structure between the two models can be understood in terms of local charge balance. Militzer et al. (Reference Militzer, Wenk, Stackhouse and Stixrude2011) calculated the elastic constants of muscovite illite–smectite using DFT theory, and the differences indicated that there were variations in their crystal sturctures. Geatches & Wilcox (Reference Geatches and Wilcox2014) modelled a variety of interlayer-deficient dioctahedral mica models of the 1M illite series based on crystallographic data and average formulae with DFT, and they analyzed the differences between cis-vacant and trans-vacant models. Escamilla-Roa et al. (Reference Escamilla-Roa, Nieto and Sainz-Díaz2016) investigated the stability of the hydronium cation in the structure of illite. The results revealed that the hydronium cation remains in the interlayer of illite at high temperatures. Sánchez-Coronilla et al. (Reference Sánchez-Coronilla, Martín, Fernández-de-Cordova, Santos and Toledo2019) investigated the inclusion of Fe, Cu and Zn in Illite. After analysing the stability and electronic effect of the inclusion system, illite presented good adsorption characteristics for Fe, Cu and Zn in the (1,0,0) site.

In the present paper, a DFT study of a typical trans-vacant 1M Al-rich illite without octahedral substitution is reported by calculating its atomic and electronic structure, density of states (DOS), electron population/charge, electronic charge density and bonding interaction analysis. The current DFT results will contribute to a better understanding of the chemical, physical and mechanical properties of illite from a microscopic perspective.

Method of calculation

The periodic DFT calculations were conducted with the Vienna ab initio simulation package (VASP; Monkhorst & Pack, Reference Monkhorst and Pack1976; Kresse & Furthmüller, Reference Kresse and Furthmüller1996). The interaction between the ions and valence electrons is described by the full-potential frozen-core all-electron projector augmented wave (PAW) method proposed by Blöchl (Reference Blöchl1994) and by Kresse & Joubert (Reference Kresse and Joubert1999). Plane-wave pseudopotentials and periodic boundary conditions were applied when solving the Kohn–Sham DFT equations. The electron exchange potential and correlation energy were calculated using the Perdew–Burke–Ernzerhof (PBE) form of the generalized gradient approximation (GGA; Perdew et al., Reference Perdew, Burke and Ernzerhof1996). In the present paper, a (3 × 4 × 3) Monkhorst–Pack k-point was utilized for Brillouin-zone integrations. The corresponding plane-wave basis cutoff energy was set at 450 eV. Geometry optimization was terminated when the Hellman–Feynman forces acting on each atom fell below 0.01 eV Å–1, and the self-consistency iterations concluded once the total energies converged to within 10–4 eV. The valence electron configurations of the elements in Al-rich illite were determined to be H 1s 1, O 2s 22p 4, Al 3s 23p 1, Si 3s 23p 2 and K 3s 23p 64s 1. The accuracy of the calculations for weak interactions was ensured by employing the DFT-D3 dispersion correction method (Moellmann & Grimme, Reference Moellmann and Grimme2014).

The Mulliken population/charge calculations and crystal orbital Hamilton population (COHP) bonding analyses were conducted using the LOBSTER package (Hughbanks & Hoffmann, Reference Hughbanks and Hoffmann1983; Dronskowski & Bloechl, Reference Dronskowski and Bloechl1993; Maintz et al., Reference Maintz, Deringer, Tchougréeff and Dronskowski2016; Ertural et al., Reference Ertural, Steinberg and Dronskowski2019). Basis sets given by Koga and Maintz that fitted to the atomic VASP GGA-PBE wavefunctions were employed during the projection (Koga et al., Reference Koga, Kanayama, Watanabe, Imai and Thakkar2000).

Results and discussion

The system of KxAl2(Si4–xAlx)O10(OH)2 was used to determine the chemical formula of Al-rich illite. The values of x ranged from 0.60 to 0.85, which conformed to the required layer charge and lattice substitution ratio. Based on the chemical composition analysis of Al-rich illite samples (Drits et al., Reference Drits, Zviagina, McCarty and Salyn2010), in the present paper x = 0.75, for the ideal chemical formula of K0.75Al2(Si3.25Al0.75)O10(OH)2, was selected to represent a type of Al-rich illite. The initial cell configuration of the Al-rich illite was derived from the illite cells constructed by Drits et al. (Reference Drits, Zviagina, McCarty and Salyn2010) and consisted of a trans-vacant dioctahedral 1M polytype structure. K ions were introduced in the interlayer to balance the cell charge. Subsequently, Si atoms in the Si–O tetrahedrons were substituted with Al atoms to afford the K3Al8(Si13Al3)O40(OH)8 chemical formula. The configurations of the three substituted tetrahedral Al atoms were analysed comprehensively through calculations as presented in Table 1. To provide a clearer and more intuitive demonstration of the configurations of the three substituted Al atoms, a selection of representative configurations was chosen, and these are presented (arranged in order of energy) in Fig. 1. Lower energy corresponds to a more stable structure. The results revealed that the local aggregation of Al–O tetrahedrons result in reduced stability of the Al-rich illite cell, which was related to the partial local balance between the interlayer charge and the charge-deficient tetrahedral sheets (Stixrude & Peacor, Reference Stixrude and Peacor2002).

Table 1. The calculated total energy of the Al-rich illite with different configurations of the three substituted tetrahedral Al atoms.

Figure 1. A selection of representative configurations of 13 Si atoms and three substituted Al atoms (shown as tetrahedra) in the Al-rich illite unit cell. (a–l) Representative configurations of Al–O tetrahedra and Si–O tetrahedra arranged in order of energy. Yellow tetrahedra = Al–O tetrahedra; pink tetrahedra = Si–O tetrahedra.

The final optimized crystal structure model of Al-rich illite (with the most stable configuration of substituted tetrahedral Al atoms) is presented in Fig. 2a. The calculated structural parameters were: 2a = 10.454 Å, b = 9.021 Å, c = 10.316 Å, α = 89.93°, β = 101.99° and γ = 89.92°, which were in good agreement with the existing experimental data (Eberl et al., Reference Eberl, Srodon, Lee, Nadeau and Northrop1987; Drits & McCarty, Reference Drits and McCarty2007), as shown in Table 2. The major bond lengths of Al-rich illite are presented in Table 3. Considering the difference in symmetries and positions, as indicated in Fig. 2b, the Al atoms were divided into two kinds: Ala was an Al atom in an octahedral sheet; and Alb was a substituted Al atom in a tetrahedral sheet. The O atoms were divided into five kinds: Oa was an O atom of the hydroxyl in an octahedral sheet; Ob was an inner O atom shared by a Si–O tetrahedron and an Al–O octahedron; Oc was an inner O atom shared by an Al–O tetrahedron and an Al–O octahedron; Od was an O atom in a tetrahedral sheet connected between Si and Alb; and Oe was an O atom in a tetrahedral sheet connected between Si atoms. The results revealed that the length of the bonds in a octahedral sheet (Ala–Oa, Ala–Ob and Ala–Oc) were all longer than those of the bonds in a tetrahedral sheet (Si–Ob, Alb–Oc, Si–Oe, Alb–Od and Si–Od), which implied that the bonding interactions in an octahedral sheet were weaker than those in a tetrahedral sheet. The lengths of the bonds in a Si–O tetrahedron (Si–Oe, Si–Ob and Si–Od) were all shorter than those of the bonds in an Al–O tetrahedron (Alb–Oc and Alb–Od), which indicated that the substitution of Al atoms reduced the connection strength of a tetrahedral sheet.

Figure 2. Crystal structures of the Al-rich illite K3Al8(Si13Al3)O40(OH)8. (a) Unit cell of Al-rich illite. (b) Characteristic atoms and bonds in Al-rich illite (back view). White spheres = hydrogen; red spheres = oxygen; yellow spheres = aluminium; pink spheres = silicon; purple spheres = potassium.

Table 2. Calculated and experimental lattice parameters of the Al-rich illite.

Table 3. Calculated bond lengths in the optimized structure of the Al-rich illite.

Figure 3 displays the band structure of Al-rich illite near the Fermi level. The high-symmetry Brillouin-zone points were G(0,0,0), F(0,0.5,0), Q(0,0.5,0.5) and Z(0,0,0.5). The conduction-band minimum of Al-rich illite is at the G point and the valence-band maximum is at the F point. Thus, Al-rich illite was calculated to have an indirect band gap with a gap width of 4.99 eV. This characteristic illustrated that Al-rich illite can be treated as an insulator.

Figure 3. Calculated band structure for the Al-rich illite. The red and blue lines represent the bottom of the conduction band and the top of the valence band, respectively.

To further investigate the electronic properties of Al-rich illite, the electronic total DOS (TDOS) and electronic partial DOS (PDOS) for H, O, Al, Si and K atoms were calculated, and these are presented in Fig. 4. Five types of O atom and two types of Al atom were individually plotted, taking into account their symmetries and positions. The Fermi energy was set at 0. The TDOS within the energy range of –35 to 0 eV exhibited four prominent peaks, labelled as 1–4 in Fig. 4a,a′. It is clear that peaks 1 and 3 were solely contributed by the K s and p states. Peak 2 was mainly from the O 2s state, with a partial contribution of Si and Al 3s/3p and H 1s states. By contrast, peak 4 was almost entirely contributed by the 2p state of O atoms. After integrating the band structures in Fig. 3, it became evident that the top of the valence band is primarily composed of the 2p state of O atoms, whereas the bottom of the conduction band includes the 3s/3p states of Si and Al atoms. Five different types of O atom exhibited similar PDOS curves, which are attributed to the high ionicity of the oxygen, resulting in charge transfer from adjacent H, Al and Si atoms. As shown in Fig. 4b,c,g, the overlap between the Ob 2s state and H 1s state split into two peaks due to the hybridization with Ala 3s/3p states in the octahedral sheet. Similarly, a particularly narrow peak of the O 1s state arose in the PDOS plot of Oc (Fig. 4c), which contributed peak 2′ of the TDOS plot. In the tetrahedral sheet, as shown in Fig. 4d,e,h–k, the substituted Alb atom presented a different hybrid mode with adjacent O atoms compared to the Si atom. It is clear that the degree of hybridization between the Si 3p state and the adjacent O 2p state was stronger than that between the Alb 3p state and adjacent the O 2p state.

Figure 4. The calculated electronic TDOS and electronic PDOS for the H, O, Al, Si and K atoms of the Al-rich illite. (a & á) TDOS of the Al-rich illite. (b) PDOS of the H atom in the hydroxyl. (c) PDOS of the Ala atom in the octahedron. (d) PDOS of the substituted Alb atom in the tetrahedron. (e) PDOS of the Si atom in the tetrahedron. (f) PDOS of the interlayer K atom. (g) PDOS of the Oa atom in the hydroxyl. (h) PDOS of the inner Ob atom. (i) PDOS of the inner Oc atom. (j) PDOS of the Od atom connected between the Si and Alb atoms. (k) PDOS of the Oe atom connected between Si atoms.

The atomic Mulliken population/charge analysis and the COHP bonding analysis for characteristic atoms and bonds in Al-rich illite were implemented using the LOBSTER package, and these are summarized in Tables 4 & 5. The Mulliken charge populations provide a comprehensive quantification of the charge transfers occurring within the crystals. The results revealed that five types of O atom gained 0.96–1.16 e, while the adjacent K lost 1.01 e, H lost 0.42 e, Si lost 2.18 e, Alb lost 1.73 e and Ala lost 1.62 e. In the tetrahedral sheet, the Si atom lost more electrons than the substituted Alb atom, which indicated a stronger charge transfer between the Si atom and adjacent O atoms. Moreover, the residual electrons in the Al 3s/3p and Si 3s/3p states revealed the covalent character of the Si–O and Al–O bonds in the tetrahedrons and octahedrons. COHP can reveal the nature of covalent interactions between bonding atoms, whereas the integrated COHP (ICOHP) values up to the Fermi-level yields are the bond contributions to the band-structure energy, which provides clues as to the bond strength. The calculated ICOHP values of the bonds Ala–Oa, Ala–Ob, Si–Ob, Alb–Oc, Si–Oe, Alb–Od and Si–Od were –4.17, –3.96, –7.63, –5.47, –7.63, –5.83 and –8.36 eV, respectively, indicating that the bonding interaction strength of these bonds followed the order Si–Od > Si–Ob > Si–Oe > Alb–Od > Alb–Oc > Ala–Oa > Ala–Ob. The results revealed that, in the tetrahedral sheet, the Si–O bonds were significantly stronger than the Al–O bonds, suggesting that the substitution of Alb weakened the stability of the silica tetrahedra. However, the Si–O and Al–O bonds in the tetrahedral sheet were all stronger than the Al–O bond in the octahedral sheet, indicating that the tetrahedral sheet was more stable than the octahedral sheet, even in the presence of substitution.

Table 4. Atomic Mulliken population/charge analysis and COHP bonding analysis for characteristic atoms and bonds in the Al-rich illite.

Table 5. Atomic Mulliken population analysis with specific orbitals for characteristic atoms of the Al-rich illite.

To gain a greater understanding of the bonding properties and charge distributions of the atoms in the tetrahedral sheet of Al-rich illite, the 3D and 2D electronic charge density contour plots (Fig. 5) were sampled within the plane containing the Si, Alb and Od atoms, which were positioned at the centre of the diagram. To better observe the charge density variation between atoms, regions with an electronic charge density exceeding 1 e Å–3 were coloured white. The results revealed that the charge density around the Od atom was substantial, indicating the high electronegativity of the O atom. The overlap of charge density within the Si–Od and Alb–Od bond regions indicated covalent bonding properties. The dense contour lines within the Si–Od bond region implied a stronger covalent bonding interaction compared to the Alb–Od bond, which is consistent with the conclusions presented above. Furthermore, the sparse contour lines around the Alb atom revealed an ionic bonding characteristic of the Alb–Od bond. This mixed ionic/covalent bonding property has also been discovered in other clay minerals such as kaolinite and montmorillonite.

Figure 5. The sampled (a) 3D and (b) 2D electronic charge density contour plots of the Al-rich illite.

Further COHP analyses including main orbital-pair contributions were performed to investigate the bonding interactions of Si–Od, Alb–Od and Ala–Ob, as shown in Fig. 6. A comparison of the Si–Od bond (ICOHP = –8.36 eV) and Alb–Od bond (ICOHP = –5.83 eV) in the tetrahedral sheet (Fig. 6a,b) revealed that all of the interactions (3s–2s, 3s–2p, 3p–2s and 3p–2p) of the Si–Od bond were enhanced compared to the Alb–Od bond. Particularly noteworthy was the 3p–2p orbital pair (ICOHP = –2.94 eV), which reached a similar characteristic strength level as the 3p–2s orbital pair (ICOHP = –2.95 eV). The results revealed a high degree of hybridization between the orbitals of the Si and Od atoms, especially the Si 3p and Od 2s/2p states. Some 3p/3s–2s antibonding states were found in the energy range from –10 to –3 eV for the Si–Od bond, which somewhat attenuated the interactions of the 3p–2s and 3s–2p bonding characteristics. Furthermore, a comparison of Ala–Ob bond (ICOHP = –3.96 eV) interaction strengths and the main orbital-pair contribution is also presented in Fig. 6c. It was observed that the Ala–Ob bond in octahedra presented weaker bonding interactions in all orbital pairs compared to the Alb–Od and Si–Od bonds in tetrahedra, which resulted in lower structural stability of the octahedral sheet compared to the tetrahedral sheet.

Figure 6. The COHP analyses including the main orbital-pair contributions for three types of bonds of the Al-rich illite: (a) the Si–Od bond in the tetrahedral sheet; (b) the Alb–Od bond in the tetrahedral sheet; and (c) the Ala–Ob bond in the octahedral sheet.

Conclusion

In the present study, the atomic structure of a typical trans-vacant 1M Al-rich illite was constructed and systematically investigated using the DFT method. The calculated structural parameters were in good agreement with the available experimental values. The calculated indirect band gap of Al-rich illite was 4.99 eV, and as a result it can be treated as an insulator. The bond-length analyses indicated that the interactions in the tetrahedral sheet were more stable than those in the octahedral sheet. The ICOHP values of the Si–O and Al–O bonds in the tetrahedral sheet were –8.36 and –5.83 eV, respectively, which indicated that the substitutions of Al atoms reduced the stability of the tetrahedral sheet. The electronic charge density implied a mixed ionic/covalent bonding property in Al-rich illite. The present findings serve as a valuable reference for use in experimental and theoretical mineralogical investigations into illite.

Financial support

This work was supported by the National Natural Science Foundation of China (grant number 41702317), the Key Laboratory of Geotechnical and Underground Engineering of the Ministry of Education of Tongji University (KLE-TJGE-B2006) and the Fundamental Research Funds for the Central Universities (2023ZKPYSB01).

Conflicts of interest

The authors declare none.

References

Beermann, T. & Brockamp, O. (2005) Structure analysis of montmorillonite crystallites by convergent-beam electron diffraction. Clay Minerals, 40, 113.CrossRefGoogle Scholar
Bergaya, F. & Lagaly, G. (2013) Handbook of Clay Science, 2nd edition. Elsevier, Amsterdam, The Netherlands, 1752 pp.Google Scholar
Blöchl, P.E. (1994) Projector augmented-wave method. Physical Review B, 50, 1795317979.CrossRefGoogle ScholarPubMed
Bridgeman, C.H. (1996) Ab-initio total energy study of uncharged 2:1 clays and their interaction with water. Molecular Physics, 89, 879888.CrossRefGoogle Scholar
Brigatti, M.F. & Guggenheim, S. (2002) Mica crystal chemistry and the influence of pressure, temperature, and solid solution on atomistic models. Reviews in Mineralogy and Geochemistry, 46, 197.CrossRefGoogle Scholar
Drits, V.A. & McCarty, D.K. (2007) The nature of structure-bonded H2O in illite and leucophyllite from dehydration and dehydroxylation experiments. Clays and Clay Minerals, 55, 4558.CrossRefGoogle Scholar
Drits, V.A., McCarty, D.K. & Zviagina, B.B. (2006) Crystal-chemical factors responsible for the distribution of octahedral cations over trans- and cis-sites in dioctahedral 2:1 layer silicates. Clays and Clay Minerals, 54, 131152.CrossRefGoogle Scholar
Drits, V.A., Zviagina, B.B., McCarty, D.K. & Salyn, A.L. (2010) Factors responsible for crystal-chemical variations in the solid solutions from illite to aluminoceladonite and from glauconite to celadonite. American Mineralogist, 95, 348361.CrossRefGoogle Scholar
Drits, V.A., McCarty, D.K. & Derkowski, A. (2012) Mixed-layered structure formation during trans-vacant Al-rich illite partial dehydroxylation. American Mineralogist, 97, 19221938.CrossRefGoogle Scholar
Dronskowski, R. & Bloechl, P.E. (1993) Crystal orbital hamilton populations (COHP): energy-resolved visualization of chemical bonding in solids based on density-functional calculations. Journal of Physical Chemistry, 97, 86178624.CrossRefGoogle Scholar
Eberl, D.D., Srodon, J., Lee, M., Nadeau, P.H. & Northrop, H.R. (1987) Sericite from the Silverton Caldera, Colorado; correlation among structure, composition, origin, and particle thickness. American Mineralogist, 72, 914934.Google Scholar
El Halim, M., Daoudi, L., El Ouahabi, M. & Fagel, N. (2022) Characterization of clays from the Fez area (northern Morocco) for potential uses in the ceramics industry. Clay Minerals, 57, 139149.CrossRefGoogle Scholar
Ertural, C., Steinberg, S. & Dronskowski, R. (2019) Development of a robust tool to extract Mulliken and Löwdin charges from plane waves and its application to solid-state materials. RSC Advances, 9, 2982129830.CrossRefGoogle ScholarPubMed
Escamilla-Roa, E., Nieto, F. & Sainz-Díaz, C.I. (2016) Stability of the hydronium cation in the structure of illite. Clays and Clay Minerals, 64, 413424.CrossRefGoogle Scholar
Feng, X., Faiia, A.M., WoldeGabriel, G., Aronson, J.L., Poage, M.A. & Chamberlain, C.P. (1999) Oxygen isotope studies of illite/smectite and clinoptilolite from Yucca Mountain: implications for paleohydrologic conditions. Earth and Planetary Science Letters, 171, 95106.CrossRefGoogle Scholar
Geatches, D.L. & Wilcox, J. (2014) Ab initio investigations of dioctahedral interlayer-deficient mica: modelling 1M polymorphs of illite found within gas shale. European Journal of Mineralogy, 26, 127144.CrossRefGoogle Scholar
Grim, R.E., Bray, R.H. & Bradley, W.F. (1937) The mica in argillaceous sediments. American Mineralogist, 22, 813829.Google Scholar
He, C., Makovicky, E. & Øsbæck, B. (1995) Thermal stability and pozzolanic activity of calcined illite. Applied Clay Science, 9, 337354.CrossRefGoogle Scholar
He, M.-C., Zhao, J. & Fang, Z.-J. (2012) First-principles study of atomic and electronic structures of kaolinite in soft rock. Chinese Physics B, 21, 039101.CrossRefGoogle Scholar
Hughbanks, T. & Hoffmann, R. (1983) Chains of trans-edge-sharing molybdenum octahedra: metal–metal bonding in extended systems. Journal of the American Chemical Society, 105, 35283537.CrossRefGoogle Scholar
Koga, T., Kanayama, K., Watanabe, T., Imai, T. & Thakkar, A.J. (2000) Analytical Hartree–Fock wave functions for the atoms Cs to Lr. Theoretical Chemistry Accounts, 104, 411413.CrossRefGoogle Scholar
Kresse, G. & Furthmüller, J. (1996) Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Physical Review B, 54, 11169.CrossRefGoogle ScholarPubMed
Kresse, G. & Joubert, D. (1999) From ultrasoft pseudopotentials to the projector augmented-wave method. Physical Review B, 59, 1758.CrossRefGoogle Scholar
Li, H., Ou, J., Wang, X., Yan, Z. & Zhou, Y. (2018) Immobilization of soil cadmium using combined amendments of illite/smectite clay with bone chars. Environmental Science and Pollution Research, 25, 2072320731.CrossRefGoogle ScholarPubMed
Maintz, S., Deringer, V.L., Tchougréeff, A.L. & Dronskowski, R. (2016) LOBSTER: a tool to extract chemical bonding from plane-wave based DFT. Journal of Computational Chemistry, 37, 10301035.CrossRefGoogle ScholarPubMed
Mercier, P.H.J., Le Page, Y. & Desgreniers, S. (2010) Kaolin polytypes revisited ab initio at 10 GPa. American Mineralogist, 95, 11171120.CrossRefGoogle Scholar
Meunier, A. & Velde, B. (2004) The mineralogy of illite – what is illite? Pp. 362 in: Illite: Origins, Evolution and Metamorphism (Meunier, A. & Velde, B., editors). Springer, Berlin, Germany.CrossRefGoogle Scholar
Militzer, B., Wenk, H.-R., Stackhouse, S. & Stixrude, L. (2011) First-principles calculation of the elastic moduli of sheet silicates and their application to shale anisotropy. American Mineralogist, 96, 125137.CrossRefGoogle Scholar
Moellmann, J. & Grimme, S. (2014) DFT-D3 study of some molecular crystals. Journal of Physical Chemistry C, 118, 76157621.CrossRefGoogle Scholar
Monkhorst, H.J. & Pack, J.D. (1976) Special points for Brillouin-zone integrations. Physical Review B, 13, 51885192.CrossRefGoogle Scholar
Perdew, J.P., Burke, K. & Ernzerhof, M. (1996) Generalized gradient approximation made simple. Physical Review Letters, 77, 38653868.CrossRefGoogle ScholarPubMed
Rieder, M., Cavazzini, G., D'Yakonov, Y.S., Frank-Kamenetskii, V.A., Gottardi, G., Guggenheim, S. et al. (1998) Nomenclature of the micas. Clays and Clay Minerals, 46, 586595.CrossRefGoogle Scholar
Righi, D. & Meunier, A. (1995) Origin of clays by rock weathering and soil formation. Pp. 43161 in: Origin and Mineralogy of Clays: Clays and the Environment (Velde, B., editor). Springer, Berlin, Germany.CrossRefGoogle Scholar
Ruiz Pestana, L., Kolluri, K., Head-Gordon, T. & Lammers, L.N. (2017) Direct exchange mechanism for interlayer ions in non-swelling clays. Environmental Science & Technology, 51, 393400.CrossRefGoogle ScholarPubMed
Sánchez-Coronilla, A., Martín, E.I., Fernández-de-Cordova, F.J., Santos, F.J. & Toledo, J.H. (2019) A theoretical study on the inclusion of Fe, Cu, and Zn in illite clays. Journal of Nanomaterials, 2019, 4546350.CrossRefGoogle Scholar
Scholtzová, E. & Tunega, D. (2019) Density functional theory study of the stability of the tetrabutylphosphonium and tetrabutylammonium montmorillonites. Clay Minerals, 54, 4148.CrossRefGoogle Scholar
Sedmale, G., Randers, M., Rundans, M. & Seglins, V. (2017) Application of differently treated illite and illite clay samples for the development of ceramics. Applied Clay Science, 146, 397403.CrossRefGoogle Scholar
Shoval, S. (2023) Characterization of the clayey sediments in the environment of exposed mudflats on the western dead sea shore. Clay Minerals. https://doi.org/10.1180/clm.2023.33.CrossRefGoogle Scholar
Steiger, R.P. (1982) Fundamentals and use of potassium/polymer drilling fluids to minimize drilling and completion problems associated with hydratable clays. Journal of Petroleum Technology, 34, 16611670.CrossRefGoogle Scholar
Stixrude, L. & Peacor, D.R. (2002) First-principles study of illite–smectite and implications for clay mineral systems. Nature, 420, 165168.CrossRefGoogle ScholarPubMed
Teich-McGoldrick, S.L., Greathouse, J.A. & Cygan, R.T. (2012) Molecular dynamics simulations of structural and mechanical properties of muscovite: pressure and temperature effects. Journal of Physical Chemistry C, 116, 1509915107.CrossRefGoogle Scholar
Yuan, R., Wang, W.-M., He, Y., Fang, Y. & Huang, X.-L. (2022) Molecular dynamics modelling of Na-montmorillonite subjected to uniaxial compression and unidirectional shearing. Clay Minerals, 57, 241252.CrossRefGoogle Scholar
Zhang, J.R., Xu, M.D., Christidis, G.E. & Zhou, C.H. (2020) Clay minerals in drilling fluids: functions and challenges. Clay Minerals, 55, 111.CrossRefGoogle Scholar
Zhao, J., Wang, Y.-F., Luan, Z.-L., Cao, Y. & He, M.-C. (2023) Theoretical analysis of the effect of doping with Na(I), K(I), Mg(II), Ca(II) and Fe(II) on the electronic and mechanical properties of pyrophyllite. Clay Minerals, 58, 113120.CrossRefGoogle Scholar
Zoeller, M. & Brockamp, O. (1997) 1M- and 2M 1-illites: different minerals and not poly types. European Journal of Mineralogy, 9, 821828.CrossRefGoogle Scholar
Zviagina, B.B. & Drits, V.A. (2019) Structural factors affecting the crystal-chemical variability in Al-rich K-dioctahedral 2M 1 micas. Clay Minerals, 54, 169179.CrossRefGoogle Scholar
Figure 0

Table 1. The calculated total energy of the Al-rich illite with different configurations of the three substituted tetrahedral Al atoms.

Figure 1

Figure 1. A selection of representative configurations of 13 Si atoms and three substituted Al atoms (shown as tetrahedra) in the Al-rich illite unit cell. (a–l) Representative configurations of Al–O tetrahedra and Si–O tetrahedra arranged in order of energy. Yellow tetrahedra = Al–O tetrahedra; pink tetrahedra = Si–O tetrahedra.

Figure 2

Figure 2. Crystal structures of the Al-rich illite K3Al8(Si13Al3)O40(OH)8. (a) Unit cell of Al-rich illite. (b) Characteristic atoms and bonds in Al-rich illite (back view). White spheres = hydrogen; red spheres = oxygen; yellow spheres = aluminium; pink spheres = silicon; purple spheres = potassium.

Figure 3

Table 2. Calculated and experimental lattice parameters of the Al-rich illite.

Figure 4

Table 3. Calculated bond lengths in the optimized structure of the Al-rich illite.

Figure 5

Figure 3. Calculated band structure for the Al-rich illite. The red and blue lines represent the bottom of the conduction band and the top of the valence band, respectively.

Figure 6

Figure 4. The calculated electronic TDOS and electronic PDOS for the H, O, Al, Si and K atoms of the Al-rich illite. (a & á) TDOS of the Al-rich illite. (b) PDOS of the H atom in the hydroxyl. (c) PDOS of the Ala atom in the octahedron. (d) PDOS of the substituted Alb atom in the tetrahedron. (e) PDOS of the Si atom in the tetrahedron. (f) PDOS of the interlayer K atom. (g) PDOS of the Oa atom in the hydroxyl. (h) PDOS of the inner Ob atom. (i) PDOS of the inner Oc atom. (j) PDOS of the Od atom connected between the Si and Alb atoms. (k) PDOS of the Oe atom connected between Si atoms.

Figure 7

Table 4. Atomic Mulliken population/charge analysis and COHP bonding analysis for characteristic atoms and bonds in the Al-rich illite.

Figure 8

Table 5. Atomic Mulliken population analysis with specific orbitals for characteristic atoms of the Al-rich illite.

Figure 9

Figure 5. The sampled (a) 3D and (b) 2D electronic charge density contour plots of the Al-rich illite.

Figure 10

Figure 6. The COHP analyses including the main orbital-pair contributions for three types of bonds of the Al-rich illite: (a) the Si–Od bond in the tetrahedral sheet; (b) the Alb–Od bond in the tetrahedral sheet; and (c) the Ala–Ob bond in the octahedral sheet.