Hostname: page-component-848d4c4894-nr4z6 Total loading time: 0 Render date: 2024-05-16T09:38:30.049Z Has data issue: false hasContentIssue false

Thermal Analysis and Thermal Reactions of Smectites: a Review of Methodology, Mechanisms, and Kinetics

Published online by Cambridge University Press:  01 January 2024

Arkadiusz Derkowski*
Affiliation:
Institute of Geological Sciences, Polish Academy of Sciences, Senacka 1, 31-002 Krakow, Poland
Artur Kuligiewicz
Affiliation:
Institute of Geological Sciences, Polish Academy of Sciences, Senacka 1, 31-002 Krakow, Poland
Rights & Permissions [Opens in a new window]

Abstract

Smectites are a group of minerals traditionally analyzed by thermal methods due to their exceptionally large adsorbed-water contents and the presence of OH groups, which makes them unique among all common soil- and rock-forming minerals. The dehydration reaction of smectite is a low-temperature endothermic effect that ends typically below 200°C. Although the removal of bulk interlayer water requires activation energy (Ea) of just above 30 kJ/mol, the removal of the last few H2O molecules attached strongly to interlayer cations requires Ea > 100 kJ/mol. Dehydroxylation is the loss of structural OH groups that proceeds as evolution of H2O molecules out of the smectite structure and occurs in the 300–900°C range. In trioctahedral species, dehydroxylation is combined with recrystallization and proceeds usually at > 700°C. In dioctahedral species, the temperature of dehydroxylation is controlled by the type of octahedral vacancy, having trans-vacant and cis-vacant distinguished by the boundary at ~ 600°C, and by the octahedral cation–OH bond strength, following the order Mg > Al > Fe. The Ea of dehydroxylation correlates linearly with the temperature of maximum dehydroxylation; from > 170 kJ/mol for Cs+-exchanged beidellite and nontronite, through ~ 300 kJ/mol in Mg-rich montmorillonite, to > 500 kJ/mol in trioctahedral saponite. Dehydration and dehydroxylation of smectites can be accompanied by a number of other phenomena, such as dehydrogenation or defluorination. At high temperatures, smectite amorphization and recrystallization occurs. Unless amorphized and/or recrystallized, smectites can undergo rehydration and rehydroxylation, which are opposite reactions to dehydration and dehydroxylation, respectively. This review discusses the details of the above-mentioned thermal reactions of smectites, focusing on thermogravimetric methods, evolved gas analysis, and structural alterations. Factors affecting the accuracy and precision of thermal analysis of smectite are discussed along with examples of best laboratory practices. The paper also provides the most recent description and critical evaluation of smectite reaction kinetics.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium,provided the original work is properly cited.
Copyright
Copyright © 2023 The Authors.

Introduction

Thermal analysis is the oldest instrumental method applied to clays and clay minerals. The first application of thermal analysis to classify and identify smectitic species (Le Chatelier, Reference Le Chatelier1887) happened long before the discovery of X-ray radiation and the identification of the actual smectite structure using X-ray diffractometry (XRD). Right after the description of the smectite structure by Hofman et al. (Reference Hofmann, Endell and Wilm1933), the already-developed differential thermal analysis (DTA) methodology quickly become a major tool in studying smectites, along with their classification and transformations (Grim & Bradley, Reference Grim and Bradley1940; Grim & Rowland, Reference Grim and Rowland1942a, Reference Grim and Rowlandb). Already in the early years of investigation, the nature of the main DTA peaks in a smectite heating curve had been identified correctly: the low-temperature endothermic DTA peak is caused by dehydration and the high-temperature endothermic DTA peak is caused by dehydroxylation (Fig. 1). These two reactions are followed by structure decomposition and recrystallization.

Fig. 1 An example of TGA and DTG curves of samples representing almost-pure dioctahedral (SCa-3 montmorillonite) and trioctahedral (high-charge synthetic saponite; Pelletier et al., Reference Pelletier, Michot, Humbert, Barrès, de la Caillerie and Robert2003) smectites. The dehydration reaction at <200°C is followed by dehydroxylation above ~300°C, which are shown in the corresponding MS signal of H2O (m/z = 18 and 17). Thermal removal of carboxyl groups from organic impurities and the decomposition of carbonate impurities are demonstrated by the CO2 (m/z = 44) signal below 500°C and above 600°C, respectively. The analysis was performed under dry N2 at a flow of 100 mL/min, with a constant heating rate of 20°C/min. Ion currents not to scale. Note: All natural smectite samples shown here and in the following figures come from the Source Clays Repository of The Clay Minerals Society. Unless specified otherwise, all thermal analyses presented in this paper were performed using a TA Discovery TGA (USA) instrument and ThermoStar GDS 320 MS from Pfeiffer (Göttingen, Germany); ~20 mg of loosely packed powder, using a wide 50 µL Platinum HT sample pan without a lid

Thermal Analysis Methodology

Thermal analysis is a ubiquitous and widely used general term that refers to a number of analytical methods monitoring the change in a material’s property with temperature, usually measured over small intervals of time or temperature change. The principles and various methods have been described in numerous reviews and books (Byrn et al., Reference Byrn, Zografi and Chen2017; Gabbott, Reference Gabbott2008; Plante et al., Reference Plante, Fernández and Leifeld2009), including excellent, in-depth reviews devoted specifically to clay minerals (Emmerich, Reference Emmerich and Christidis2010; Stucki et al., Reference Stucki, Bish and Mumpton1990). Therefore, the present paper focuses only on the application of thermal analysis to smectites (also applicable to vermiculites and smectite-interstratified phases) and only to smectites in a form corresponding to their most common natural occurrence in rocks and sediments, i.e. having inorganic interlayer cations. Thermal analysis of smectites with adsorbed organic compounds is a field in and of itself with dedicated reviews (e.g. Xi et al., Reference Xi, Martens, He and Frost2005; Plante et al., Reference Plante, Fernández and Leifeld2009; Yariv et al., Reference Yariv, Borisover and Lapides2011).

Among many thermal analysis methods (see Emmerich, Reference Emmerich and Christidis2010), a few have been used widely for smectites and related clay minerals, and are the most meaningful for their characterization: thermogravimetric analysis (TGA), differential thermal analysis (DTA), and its modern successor – differential scanning calorimetry (DSC). If a sample is analyzed simultaneously using more than one of the techniques (e.g. TGA and DTA), such a mode of analysis is often referred to as Simultaneous Thermal Analysis (STA). The evolved gas analysis (EGA), which allows determination of the composition of gases and vapors evolved from a sample during heating in a TGA or DTA/DSC analyzer, is a form of STA. Among the different means of analyzing gas evolved from clay samples, infrared spectroscopy (IR) and mass spectrometry (MS; also coupled with gas chromatography – GC-MS) are the most commonly used.

Thermogravimetric analysis presents weight change with temperature or time. In most cases this corresponds to the extent of a clay’s weight loss (in mg or weight %) during heating in an inert gas atmosphere, such as N2, Ar, or He. However, under specific circumstances, mass gain can occur during TGA experiments, i.e. when an oxidizing agent or water molecules are present in the sample’s environment. Common practice when reporting TGA results is to present a first derivative of the original mass change curve over time or temperature. Such a derivative is referred to as the DTG curve. Differential thermal analysis returns the temperature difference between the sample and an inert reference material under the same heating conditions. Similar to DTA, DSC measures the heat flow (J K–1 mol–1) needed to maintain the same temperature of the sample and the inert reference material, which enables the heat capacity to be obtained (Giese, Reference Giese, Stucki, Bish and Mumpton1990). All the methods mentioned require some kind of purge gas which, in general, is meant to remove the thermal reaction products. A moisture analyzer is a low-temperature alternative to the TGA instrument, capable of measuring weight change (usually due to dehydration) from room temperature typically up to 200 or 250°C, with no purge gas flow.

Any thermal analysis can be performed in two modes depending on the rate of heating: (1) isothermal – temperature is increased quickly to the desired value and kept constant for the remaining part of the experiment; (2) non-isothermal – temperature varies during the experiment. This mode is represented by, for example, heating or cooling with a constant rate to the final temperature. The most common way of performing thermal analysis is to use a predefined (isothermal or non-isothermal) temperature program. Controlled-rate thermal analysis (CRTA) is a specific case of non-isothermal or quasi-isothermal conditions in modern TGA instruments, where the sample’s reaction rate (e.g. weight loss per time) is set as constant, but the heating rate varies (Ding & Frost, Reference Ding and Frost2002; Rouquerol et al., Reference Rouquerol, Rouquerol and Llewellyn2013). As the temperature controller keeps the reaction rate constant during CRTA, temperature change with time for a particular experiment is not known beforehand. Whereas DTA and DSC usually need the linear heating/cooling mode (with exceptions, see Boller et al., Reference Boller, Jin and Wunderlich1994), TGA can operate in any isothermal or non-isothermal mode, or a combination of both.

Measurement Accuracy and Precision

Thermal-reaction temperatures are commonly reported based on the position of the corresponding DTG or DTA peak. A number of factors affect the position of the peak, which can hamper straightforward comparisons among reaction temperatures reported in various studies (Emmerich, Reference Emmerich and Christidis2010). First, a positive correlation occurs between the DTG or DTA peak temperatures and the heating rate in non-isothermal conditions due to kinetic effects. Second, the peak temperature is dependent on the particular thermal analyzer’s temperature calibration, which should be based on at least two points in the range corresponding to the tested reaction. Curie temperature reference materials are commonly used for TGA instruments, whereas temperature and enthalpy of fusion are used in DTA and DSC instruments (Della Gatta et al., Reference Della Gatta, Richardson, Sarge and Stølen2006). Theoretically, ~1 K (1°C) should be considered as a standard error in the thermal analysis of smectite in well-calibrated instruments. In practice, however, several factors may cause the actual sample temperature to deviate from the measured value even more (see Emmerich, Reference Emmerich and Christidis2010 for comprehensive review). As a result of a poorly crystalline structure, heterogeneous cation distribution, the particularly small size of its crystallites (both planar and crystallite thickness), and their diverse means of agglomeration, smectite samples do not return sharp, well-defined thermal events. These features, along with operator-dependent choices during data mathematical processing, all result in the real accuracy and precision of determining reaction temperature being in the range of a few °C.

Accuracy of the mass-loss measurement can be tested using a homogeneous material with a well-constrained mass loss in a given temperature interval, such as Ca-oxalate monohydrate (Bish & Duffy, Reference Bish, Duffy, Stucki, Bish and Mumpton1990; Plante et al., Reference Plante, Fernández and Leifeld2009). Thermal drift (i.e. the profile of mass or heat flow change with temperature in the absence of any reactive material) should be measured for each thermal analysis instrument and a mode of analysis. Thermal drift can be determined on an empty pan or an unreactive sample with a well-calibrated instrument under conditions corresponding to the experimental sample analysis and the test results should be applied to correct an experimental sample curve. A well-calibrated TGA instrument and method should return the mass-loss precision in the range of 0.1%.

Thermal Reactions of Smectites

Dehydration

Dehydration is the loss of molecular water (H2O) adsorbed on or within clay-mineral particles. Upon drying, H2O molecules are removed sequentially following the strength of their interactions with the smectite surface. The entire physisorbed water condensed in meso- and micropores of particles is removed at room temperature under the flow of dry purge gas. Bulk (bound via H-bonds to other H2O molecules) smectite interlayer water is in equilibrium with the external moisture and, therefore, it is removed along with the physisorbed water albeit at a lower rate (Emmerich et al., Reference Emmerich, Giaudo, Schuhmann, Schnetzer, Kaden and Thissen2018; Ferrage et al., Reference Ferrage, Lanson, Michot and Robert2010; Salles et al., Reference Salles, Douillard, Denoyel, Bildstein, Jullien, Beurroies and Van Damme2009). Therefore, distinction between the bulk interlayer and the external, physisorbed water is hardly possible in thermal analysis as no temperature–time cutoffs are applicable. The removal of physisorbed and bulk interlayer water is completed much below 100°C. Due to the difference in solidification and fusion temperatures of physisorbed water, the DSC-based thermoporometry (freezing conditions, below 0°C) is the best temperature-differentiation method that can be applied to address the distinction between total interlayer and micropore capillary water (Grekov et al., Reference Grekov, Montavon, Robinet and Grambow2019; Salles et al., Reference Salles, Bildstein, Douillard, Jullien, Raynal and Van Damme2010).

In a fully hydrated smectite, some water molecules form a hydration shell around the smectite’s interlayer cations. Once the bulk interlayer water is removed during dehydration, the remaining H2O molecules become organized around the interlayer cations in a fashion and number dependent on the cation’s hydration energy (Salles et al., Reference Salles, Douillard, Bildstein, El Ghazi, Prélot, Zajac and Van Damme2015), expressed as the cation’s hydration enthalpy (Table 1). The larger the absolute value of the cation hydration enthalpy (see Table 1 for an explanation of ΔH o), the lower the probability for the cation to bind directly to the smectite’s siloxane surface, the stronger the interaction between water molecules and the interlayer cation, and, therefore, the greater the energy that is required to remove any remaining strongly bound water (Webb et al., Reference Webb, Stanley and Scheiner1986).

Table 1 The hydration enthalpy (–ΔH o) after Smith (Reference Smith1977) and effective ionic radii for a given coordination number of inorganic cations commonly occurring in smectite interlayers in natural conditions and used in laboratory experiments, after Shannon (Reference Shannon1976)

* After Sidey (Reference Sidey2016)

The cation’s hydration enthalpy is reflected by the pattern of smectite dehydration. Except for Li+-saturated smectites, the interlayer’s monovalent cations produce a single DTG dehydration peak below 100°C, whereas two DTG peaks or a high-temperature shoulder (up to ~200°C) usually correspond to high-enthalpy cations (polyvalent and Li+; Fig. 2; El-Barawy et al., Reference El-Barawy, Girgis and Felix1986). The fewer the H2O molecules around interlayer cations, the greater the activation energy (E a) of dehydration and the higher the temperature needed to remove the remaining H2O. In extreme cases, H2O dissociation leads to the hydrolysis of interlayer cations by OH that can remain in the smectite interlayer, even up to 500°C, which affects the course of dehydroxylation (Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017).

Fig. 2 DTG curves in the temperature range corresponding to the dehydration of Na+-, Ca2+-, and La3+-exchanged montmorillonites. The analysis was performed under dry N2 at a flow of 50 mL/min, with a constant heating rate of 5°C/min

NH4 +-exchanged smectites show very low mass loss due to dehydration as the ammonium cation has a low hydration enthalpy (Table 1); however, the smectite undergoes an exothermic deammonization reaction at 300–400°C resulting in weight loss and leaving smectite in H+ form (Mackenzie & Bishui, Reference Mackenzie and Bishui1958; Seiffarth & Kaps, Reference Seiffarth and Kaps2009).

Unless a sample contains an impurity that provides mass loss at low temperature (< 200°C, e.g. gypsum, zeolites, halloysite, and soil organic matter; see the Impurities section below), the quantity of H2O removed from a clay during thermal dehydration corresponds to the smectite content in a sample (both as a discrete mineral and as a fraction of smectite-interstratified phases), and also depends on the hydration enthalpy of the interlayer cation, and – to a minor extent – the smectite layer charge (Środoń & McCarty, Reference Środoń and McCarty2008; Środoń et al., Reference Środoń, Zeelmaekers and Derkowski2009; Webb et al., Reference Webb, Stanley and Scheiner1986). If used for a semi-quantitative estimation of smectite content, a sample needs to be pre-equilibrated at a selected relative humidity (RH) that produces a particular degree of interlayer H2O filling, the latter also being dependent on the interlayer cation type (Berend et al., Reference Berend, Cases, Francois, Uriot, Michot, Masion and Thomas1995; Cases et al., Reference Cases, Bérend, François, Uriot, Michot and Thomas1997). Despite this ambiguity, the measured fraction of mass loss due to dehydration is good enough for a relative comparison among samples tested. For example, pure Ca2+-saturated smectite equilibrated at ~50% RH contains ~16.5 wt.% of H2O (as determined by drying at 200°C; Środoń & McCarty, Reference Środoń and McCarty2008); the corresponding mass loss upon dehydration decreases proportionally to illite interstratification in illite-smectite (Środoń et al., Reference Środoń, Zeelmaekers and Derkowski2009) and smectite mineral content in a sample (Środoń, Reference Środoń2009). When equilibrated at ~50% RH, dehydration would correspond to 6% mass loss in K+-saturated smectite, ~10% in Na+- and Li+-saturated smectite, and almost 20% in Mg2+-saturated smectite (Berend et al., Reference Berend, Cases, Francois, Uriot, Michot, Masion and Thomas1995; Cases et al., Reference Cases, Bérend, François, Uriot, Michot and Thomas1997). The above-mentioned semi-quantitative measurement should not be carried out under any dry gas purge (common in TG instruments) as it triggers dehydration during the sample-weight stabilization under dry gas flow; a moisture analyser is thus preferred (Środoń et al., Reference Środoń, Zeelmaekers and Derkowski2009).

Dehydroxylation

Dehydroxylation is the major thermal reaction of smectites and other 2:1 layer type clay minerals. It occurs between ~300 and 900°C and results in the extraction of a H2O molecule formed by two adjacent OH groups and leaves one residual oxygen atom (O r) in the structure:

(1) 2 n OH - = n H 2 O + n O r 2 -

where n is the number of water molecules eliminated during dehydroxylation.

The actual thermogravimetric manifestation of the reaction is the loss of H2O after its migration through the interlayer and out of the crystallite. The theoretical mass loss upon dehydroxylation is calculated as the molecular weight of H2O per a smectite’s formula unit molecular weight. Dehydroxylation produces an evolved gas pattern of H2O exactly following the DTG curve and TG mass loss, corresponding quantitatively to Eq. 1 (Fig. 1). In Ca2+- or Na+-exchanged Fe-poor smectites mass loss corresponding to dehydroxylation is 4.7–4.9%, whereas in nontronite it can be as low as 4.3%. Exchanging a smectite with a heavy cation such as Cs+ can decrease the theoretical mass loss significantly. The smectite’s mass loss due exclusively to dehydroxylation may not be clearly distinguishable in TGA analysis, and thus, may not correspond to the theoretical value. In numerous dioctahedral smectites, especially those having an interlayer cation of high hydration enthalpy (Ca2+, Mg2+, etc.), the onset of dehydroxylation overlaps with the dehydration of residual, strongly bound water (Bray & Redfern, Reference Bray and Redfern2000; Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017).

The first control on the temperature of dehydroxylation is the octahedral occupancy. In dioctahedral 2:1 layer phyllosilicates the proton from one OH group migrates either along the shared edge of the octahedron or jumps across the octahedral vacancy to join the adjacent OH group (Guggenheim et al., Reference Guggenheim, Chang and Koster van Groos1987; Ogloza & Malhotra, Reference Ogloza and Malhotra1989; Muñoz-Santiburcio et al., 2016). In dioctahedral smectites, dehydroxylation transforms the octahedral structural arrangement into a five-fold coordinated bipyramidal structure and the previously octahedral cations remain in a similar arrangement to the pristine structure (for details see Drits et al., Reference Drits, Besson and Muller1995). The 2:1 layer thus remains at least partially intact. In the case of trioctahedral phyllosilicates, due to the lack of an octahedral vacancy where the residual oxygen could reside, OH groups cannot be removed without a reconstruction of the octahedral sheet; their dehydroxylation is accompanied by recrystallization and occurs at higher temperatures than in dioctahedral species, therefore (Fig. 1). The high-temperature DTG peak (usually >700°C) can serve as a distinction between trioctahedral layers or domains and dioctahedral counterparts in natural smectites. Synthetic saponites, however, showed strong dependence of the dehydroxylation temperature on the octahedral sheet cation composition: Mg > Ni > Co > Zn (Vogels et al., Reference Vogels, Kloprogge, Geus and Beers2005). Vogels et al. (Reference Vogels, Kloprogge, Geus and Beers2005) proposed that the smaller the octahedral cations, the greater the corrugation of the tetrahedral sheet, which results in greater thermal stability.

For dioctahedral smectites, the next factor controlling the temperature of dehydroxylation is octahedral symmetry. Drits et al. (Reference Drits, Besson and Muller1995, Reference Drits, Lindgreen, Salyn, Ylagan and McCarty1998) showed that the cis-vacant (cv) structure has a significantly higher dehydroxylation temperature (>600°C) than the trans-vacant (tv) layer structure (< 600°C; Fig. 3). Those authors assigned that difference to the longer distance between non-shared OH edges in the cv structure than the shared OH edge of the octahedron in the tv coordination, resulting in the less probable and longer H+ jump between the OH groups in the cv coordination. Besides, in contrast to one-step dehydroxylation of tv layers, dehydroxylation of aluminous cv smectite is a two-step process that involves the Al cation migration to cis-sites and results in the pseudo-tv structure of the dehydroxylate (Drits et al., Reference Drits, Besson and Muller1995; Stackhouse et al., Reference Stackhouse, Coveney and Benoit2004), the same as that of the dehydroxylated pristine tv smectite. In tv Mg-Fe-rich smectites (e.g. nontronite), however, some cations migrate to trans-positions during dehydroxylation in order to redistribute the Fe–O bonds, resulting in a cv dehydroxylated structure (Muller et al., 2010a, b Drits & Zviagina, Reference Drits and Zviagina2009).

Fig. 3 DTG curves showing the difference in dehydroxylation of trans-vacant (tv) and cis-vacant (cv) dioctahedral smectites along their dominating octahedral cation-OH composition (after Derkowski et al., Reference Derkowski, Drits and McCarty2012a). Analysis preformed under dry N2 at 50 mL/min flow, with a constant heating rate of 5°C/min

Following Drits et al. (Reference Drits, Lindgreen, Salyn, Ylagan and McCarty1998), Wolters and Emmerich (Reference Wolters and Emmerich2007) and Emmerich et al. (Reference Emmerich, Wolters, Kahr and Lagaly2009) adopted 600°C as the dehydroxylation temperature boundary and suggested using it for the determination of cv and tv compounds in smectites.

A broad temperature range of dehydroxylation is observed for all dioctahedral clay minerals, including smectites. Guggenheim et al. (Reference Guggenheim, Chang and Koster van Groos1987) and Guggenheim (Reference Guggenheim1990) explained this effect as a continuous recombination of bonds between the remaining OH and octahedral cations in a progressively dehydroxylated 2:1 layer. Drits et al. (Reference Drits, Derkowski and McCarty2012a,Reference Drits, McCarty and Derkowskib) and Derkowski et al. (Reference Derkowski, Drits and McCarty2012a) suggested that the reaction model is homogeneous; the 2:1 layers are only either non-dehydroxylated or completely dehydroxylated, but the temperature of maximum dehydroxylation as well as the temperature range of dehydroxylation are controlled by the bond strength between OH groups and octahedral cations, i.e. the octahedral sheet composition (Drits et al., Reference Drits, McCarty and Derkowski2012b) and the crystallite thickness distribution (see also Pérez-Maqueda et al., Reference Pérez-Maqueda, Blanes, Pascual and Pérez-Rodríguez2004, Perez-Rodriguez et al., Reference Perez-Rodriguez, Duran, Sanchez Jimenez, Franquelo, Perejón, Pascual-Cosp and Pérez-Maqueda2010). The latter is not applicable to smectites due to their ultrafine planar size and small crystallite thickness. The relative order of dehydroxylation temperature for tv dioctahedral smectites can be expressed by the following sequence (Derkowski et al., Reference Derkowski, Drits and McCarty2012a), (Fig. 3):

Mg - OH - Mg > Al - OH - Mg > Al - OH - Al > Al - OH - Fe > Fe - OH - Fe

which is supported by the quantum mechanical calculations. The order of dehydroxylation temperature is controlled by the octahedral cation–OH bond strength, thus O…H distance (Molina-Montes et al., Reference Molina-Montes, Timón, Hernández-Laguna and Sainz-Díaz2008; Botella et al., Reference Botella, Timón, Escamilla-Roa, Hernandez-Languna and Sainz-Díaz2004). Although 300°C is applied commonly as the lowest temperature boundary of dehydroxylation (Fig. 3), long isothermal treatment provides energy sufficient to shift the onset of reaction to <300°C in the tv Fe-OH-Fe environment, i.e. nontronite (Środoń & McCarty, Reference Środoń and McCarty2008; Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017).

Heterogeneity of the octahedral sheet’s cation composition contributes to the broad range of smectite dehydroxylation temperature; in smectite that has one cation species dominating the octahedral sheets the temperature range of dehydroxylation is narrower than those observed in smectites with heterogeneous octahedral composition (Drits et al., Reference Drits, Derkowski and McCarty2012a). In trioctahedral smectites the octahedral sheet is rarely fully occupied by cations. Locally occurring vacancies, e.g. in stevensite, produce dioctahedral domains which dehydroxylate at temperatures corresponding to the dioctahedral smectite (i.e. montmorillonite).

The interlayer cations’ hydration enthalpy controls the temperature of final dehydration and, thus, the degree of overlap for the dehydration and dehydroxylation reactions in some dioctahedral smectites (Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017; Fig. 4). Also, the interlayer cation size and its potential for migration into the octahedral sheet play a role in the pathway of smectite dehydroxylation. Early DTA studies demonstrated that the interlayer cation influences the temperature of maximum dehydroxylation (El-Barawy et al., Reference El-Barawy, Girgis and Felix1986; Mackenzie & Bishui, Reference Mackenzie and Bishui1958). Later, Bray and Redfern (Reference Bray and Redfern2000) and Kuligiewicz and Derkowski (Reference Kuligiewicz and Derkowski2017) found differences in the kinetics of the dehydroxylation reaction in a smectite containing monovalent and divalent interlayer cations. Cs+ is the largest inorganic cation that can pillar the interlayer (Table 1), allowing H2O molecules generated during dehydroxylation to diffuse easily, resulting in a decrease in the temperature of dehydroxylation in dioctahedral smectites (Fig. 4). The more beidellitic the smectite structure is, the greater the effect of Cs+-pillaring (Fig. 4: beidellite > beidellitic montmorillonite and nontronite > montmorillonite) because of the repulsion forces between adjacent siloxane surfaces created by Al substitution in the tetrahedral sheet and also the potential for the Cs+ cation to be located above the Al-substituted tetrahedra instead of the ditrigonal cavity (Cuadros, Reference Cuadros2002; Ngouana & Kalinichev, Reference Ngouana and Kalinichev2014). The cation size, however, seems to result in an opposite trend in trioctahedral smectites; the larger the interlayer cation, the higher the temperature of dehydroxylation (Fig. 4). The ability to block efficiently the ditrigonal cavity by a large cation, which suppresses H+ migration into the adjacent OH group, is likely the reason for this effect.

Fig. 4 DTG curves of smectite reference samples saturated with various inorganic cations. The analysis was performed under a 100 mL/min flow of dry N2, with 4 h isothermal heating at the temperatures indicated on the figures, followed by heating to 1000°C at a constant rate of 5°C/min

In montmorillonites, where layer charge is generated in the octahedral sheet, dehydration enables small cations to migrate into the ditrigonal cavity of the tetrahedral sheet and even further into the octahedral vacancy, forming pseudo-trioctahedral domains (Emmerich et al., Reference Emmerich, Madsen and Kahr1999; Hofmann & Klemen, Reference Hofmann and Klemen1950; Komadel et al., Reference Komadel, Madejová and Bujdák2005, and references therein; Skoubris et al., Reference Skoubris, Chryssikos, Christidis and Gionis2013). The Hofmann-Klemen effect affects the pattern of dehydroxylation, forming an additional DTG peak at temperatures corresponding to trioctahedral domains. Not only Li+, commonly used in reducing layer charge, but also other small cations can produce such thermal effects. Koster van Groos and Guggenheim (Reference Koster van Groos and Guggenheim1989) explained differences observed between the dehydroxylation of Mg2+- and Ca2+-exchanged montmorillonite as resulting from the different abilities of Mg2+ and Ca2+ to migrate toward the residual oxygen during the dehydroxylation reaction. The ability to form pseudo-trioctahedral domains must then depend on the kinetics of dehydration and small-cation migration through the ditrigonal cavity. Indeed, as presented in Fig. 5, Li+ migration (Li+ –ΔH o = 519 kJ/mol; Table 1) forms the high-temperature DTG peak weakly dependent on the dehydration temperature prior to dehydroxylation, whereas Mg2+-montmorillonite (Mg2+ –ΔH o = 1921 kJ/mol; Table 1) produces a pseudo-trioctahedral DTG signature at ~900°C, which is proportional to the temperature and time of isothermal dehydration (Fig. 5). Fast dehydroxylation during ramp-heating itself may not be sufficient to quickly and completely dehydrate Mg2+ and fix the cation in the ditrigonal cavity or the octahedral sheet.

Fig. 5 The effect of cation migration into the octahedral sheet. DTG curves of Li+-exchanged and Mg2+-exchanged montmorillonites preheated isothermally under a flow of 100 mL/min of dry N2 for 4 h at various temperatures (as stated in the legend) followed by heating to 1000°C at a constant rate of 5°C/min. The DTG curves of the SCa-2 sample saturated with other cations and preheated isothermally at 300°C are given as reference

Dehydrogenation

Dehydrogenation proceeds during thermal treatment of Fe(II)-bearing phyllosilicates, resulting in the removal of one hydrogen atom while simultaneously oxidizing one Fe(II) cation that is octahedrally coordinated with the OH group (Farmer et al. Reference Farmer, Russell, McHardy, Newman, Ahlrichs and Rimsaite1971; MacKenzie & Berezowski, Reference MacKenzie and Berezowski1984; Rancourt et al., Reference Rancourt, Mercier, Cherniak, Desgreniers, Kodama, Robert and Murad2001; Sanz et al., Reference Sanz, Gonzalez-Carreno and Gancedo1983; Wicks & Ramik, Reference Wicks, Ramik, Stucki, Bish and Mumpton1990).

In vacuum and in an inert gas atmosphere, the dehydrogenation reaction proceeds as follows:

(2) 2 Fe 2 + + OH - 2 Fe 3 + + O r 2 - + H 2

resulting in the formation of H• radicals that evolve from a sample as H2 gas, which can be recorded in an EGA mass spectroscopy signal as m/z = 2 (Lempart et al., Reference Lempart, Derkowski, Strączek and Kapusta2020). The reaction proceeding via Eq. 2 produces negligible mass loss, 1/9 of that produced by dehydroxylation (Eq. 1). Dehydrogenation usually overlaps with the temperature range corresponding to dehydroxylation, which has led to numerous incorrect interpretations of the reaction mechanisms and products (Heller-Kallai & Rozenson, Reference Heller-Kallai and Rozenson1980). Dehydrogenation removes the hydroxyl H, reducing the amount of OH groups available for dehydroxylation, which results in a weight loss that is much lower than the theoretical value for dehydroxylation. In contrast to what has been reported in the literature, Fe(II)-free phyllosilicates do not undergo dehydrogenation and do not produce H2 under normal pressure (Heide & Földvari, Reference Heide and Földvari2006 cf. Lempart et al., Reference Lempart, Derkowski, Strączek and Kapusta2020).

Under O2-bearing gas (e.g. air), the reaction follows the same pathway as in Eq. 2, with H• radicals combined with ambient oxygen to produce H2O.

(3) 2 Fe 2 + + OH - + 1 / 2 O 2 2 Fe 3 + + O r 2 - + H 2 O

The presence of oxygen enhances dehydrogenation, shifting the reaction to lower temperatures. The MS curve of m/z = 18 of evolved gas in the reaction from Eq. 3 is excessive to the DTG curve, because H2O is not produced within the heated crystallite and thus does not produce the corresponding weight loss (Lempart et al., Reference Lempart, Derkowski, Strączek and Kapusta2020).

In contrast to Eq. 3, only a specific portion of structural Fe(II) can oxidize under O2 (e.g. air) following the reaction:

(4) 2 Fe 2 + + 1 2 O 2 2 Fe 3 + + O r 2 -

because the structural Fe(II) in a 2:1 layer is not readily available for binding with ambient oxygen and the reaction presented in Eq. 3 is preferred unless VIFe(II)/OH ratio > 1. Such a reaction as in Eq. 4 is possible during or after the high-temperature recrystallization when Fe(II) is exposed to ambient O2 (Lempart et al., Reference Lempart, Derkowski, Luberda-Durnaś, Skiba and Błachowski2018).

Because of structural constraints, Fe(II) in natural smectites occurs in numerous trioctahedral species (Kohyama et al., Reference Kohyama, Shimoda and Sudo1973; Stucki, Reference Stucki, Stucki, Goodman and Schwertmann1988); therefore, these phases are the most prone to dehydrogenation. In nature, dehydrogenation may be responsible for the depletion of OH groups in so-called oxy-smectites (Dainyak et al., Reference Dainyak, Dritz, Kudryavtzev, Simanovitch and Slonimskaya1981). Dehydrogenation following Eq. 2 or Eq. 3 is expected upon heating Fe-reduced nontronites (Heller-Kallai, Reference Heller-Kallai1997a; Lear & Stucki, Reference Lear and Stucki1985).

In F-bearing smectites, usually trioctahedral and synthetic smectites (Christidis et al., Reference Christidis, Aldana, Chryssikos, Gionis, Kalo, Stöter, Breu and Robert2018), defluorination occurs as F• radical loss above 500°C and HF loss above 850°C (Heller-Kallai Reference Heller-Kallai1997b; Wicks & Ramik, Reference Wicks, Ramik, Stucki, Bish and Mumpton1990).

Smectite Structural Transformation upon Heating

Amorphization and recrystallization are the reactions that occur during thermal destabilization of the dehydroxylation and dehydrogenation products (Grim & Rowland, Reference Grim and Rowland1942a,Reference Grim and Rowlandb; Meyers & Speyer, Reference Meyers and Speyer2003). At temperatures higher than dehydroxylation, dioctahedral smectites undergo amorphization and then recrystallize into a topotaxic set of mullite, spinels, quartz or cristobalite, and pyroxenes (Bradley & Grim, Reference Bradley and Grim1951; Brett et al., Reference Brett, MacKenzie and Sharp1970). The degree and temperature of amorphization and transformation depends on the octahedral-sheet composition. Beidellite dehydroxylate is the most stable high-temperature smectite phase, the irreversible amorphization of which has not been observed below 900°C (Derkowski et al., Reference Derkowski, Drits and McCarty2012a; Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017; Kawano & Tomita, Reference Kawano and Tomita1991b), whereas recrystallization occurs above 1200°C (Bradley & Grim, Reference Bradley and Grim1951). Sequentially heated Mg-rich dioctahedral smectites show an onset of amorphization at the end of dehydroxylation, whereas maximum amorphization is reached by ~800°C (Garg & Skibsted, Reference Garg and Skibsted2014;  Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017), which provides their pozzolanic properties, but irreversibly removes their adsorption potential. In smectites rich in redox-active metals, i.e. nontronite, the temperature of transformation and the composition of transformation products depend heavily on the type of ambient gas (MacKenzie & Rogers, Reference MacKenzie and Rogers1977).

In trioctahedral smectites, transformation proceeds along with dehydroxylation; after initial amorphization, a pyroxene-like structure and an olivine-like structure are formed (Vogels et al., Reference Vogels, Kloprogge, Geus and Beers2005).

In contrast to dehydroxylation and dehydration, phase transformation is manifested in the DTA or DSC curves and not in TGA and TGA-MS, because it is not associated with a mass change or evolution of volatiles. However, note that at high temperature (>800°C), especially for isothermal experiments, some alkali cations can be removed due to vaporization, which may be responsible for occasionally observed minor and slow weight loss over a broad temperature range.

Rehydration and Rehydroxylation

Rehydration is the adsorption of H2O molecules by a dried material. Rehydroxylation is the reconstruction of the structural OH groups, i.e. reversal of the reaction in Eq. 1, in a dehydroxylated smectite structure that has not undergone phase transition (amorphization or recrystallization). Whereas rehydroxylation in general must be preceded by H2O diffusion through the dehydroxylated structure, advanced rehydroxylation requires specifically the diffusion within the interlayer in order to access all potential reaction sites available in the 2:1 layer. Both rehydration and rehydroxylation can be observed in situ as mass gain once wet gas is delivered directly to the TGA chamber (Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017). When exposed to air, dried smectite having divalent cations in the interlayer starts resorbing H2O from ambient moisture within seconds (Fig. 6; Kanik et al., Reference Kanik, Longstaffe, Kuligiewicz and Derkowski2022). For this reason, in practice, smectite material for rehydration and rehydroxylation tests also needs to be pre-dehydrated or pre-dehydroxylated in situ. The rate and extent of rehydration in pristine, dried (i.e. dried below the temperature of dehydroxylation) smectite depends heavily on the pre-drying conditions and, thus, the extent of interlayer cation fixation, and the temperature and RH during rehydration (Fig. 6; Derkowski et al., Reference Derkowski, Drits and McCarty2012b, Gailhanou et al., Reference Gailhanou, Vieillard, Blanc, Lassin, Denoyel, Bloch, De Weireld, Gaboreau, Fialips, Madé and Giffaut2017). Structurally, the extent of rehydration is controlled by the bonding between the siloxane surface and the interlayer cation; the greater the layer charge and the lower the cation hydration enthalpy, the stronger the bonding. K+-exchanged high-layer charge smectites would not rehydrate extensively keeping their interlayer closed, whereas Ca2+-exchanged smectite is prone to fast rehydration (Derkowski et al., Reference Derkowski, Drits and McCarty2012b; Kawano & Tomita, Reference Kawano and Tomita1991a, Reference Kawano and Tomitab). Once a small interlayer cation, like Li+ is dehydrated and migrates through the ditrigonal cavity in montmorillonite (Hofmann-Klemen effect; Fig. 5), however, the interlayer is fixed permanently, prohibiting rehydration (Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017; Komadel et al., Reference Komadel, Hrobáriková and Koppelhuber-Bitschnau2002). A similar effect occurs even if the interlayer cations do not migrate into the 2:1 layer: if layer charge is located in the octahedral sheet, dehydrated interlayer cations are bonded strongly to the 2:1 layer, resulting in rehydration slower than if the charge came from tetrahedral cation substitution (Kawano & Tomita, Reference Kawano and Tomita1991a,Reference Kawano and Tomitab).

Fig. 6 Rehydration of a smectite reference sample (SWy-3), saturated with cations of contrasting hydration enthalpy, at various temperatures, after prior in situ drying for 4 h at 200°C. Mass gain is shown as fractional mass gain at a constant H2O content of 9 g/kg air (not constant RH), normalized to the sample mass after drying at 200°C

Rehydroxylation occurs mostly in the intact dehydroxylated phase that preserves the layer structure and can undergo the reverse of the Eq. 1 reaction (Derkowski et al., Reference Derkowski, Drits and McCarty2012a). Once smectite crystallites reach the high-temperature amorphization or recrystallization, the structural integrity is lost. Therefore, dehydroxylated trioctahedral smectite, Mg-rich montmorillonite, and nontronite do not have great potential for rehydroxylation, whereas beidellite is the most prone to rehydroxylation because its layer structure remains intact at high temperatures (Derkowski et al., Reference Derkowski, Drits and McCarty2012a).

Because dehydroxylation of originally tv and cv aluminum smectites produces a similar tv phase, when rehydroxylated it resembles a tv structure (Drits et al., Reference Drits, Besson and Muller1995; Muller et al., Reference Muller, Drits, Plançon and Besson2000a, Reference Muller, Drits, Plançon and Robertb). Indeed, when analyzed with thermal methods, such a rehydroxylated phase dehydroxylates at the temperature range typical for tv beidellite (<500°C; irrespective of the temperature of dehydroxylation of its pristine structure (Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017; Fig. 7). A tv Fe,Mg-rich smectite that adopts a cv character after dehydroxylation can probably return to tv when rehydroxylated (Drits & Zviagina, Reference Drits and Zviagina2009; Muller et al., Reference Muller, Drits, Plançon and Robert2000b). The most intense rehydroxylation occurs under H2O vapor at temperatures of 250–350°C, even at extremely low H2O activity. At such conditions, up to tens of % of the dehydroxylated dioctahedral smectite structure can be rehydroxylated within an hour of reaction (Fig. 7; Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017). The mass gain observed during isothermal treatment of the prior dehydroxylated phase has been proven to correspond to true rehydroxylation by XRD and infrared spectroscopy (Kawano & Tomita, Reference Kawano and Tomita1989, Reference Kawano and Tomita1991b; Muller et al., Reference Muller, Drits, Plançon and Robert2000b; Derkowski et al., Reference Derkowski, Drits and McCarty2012a; Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017). At reaction temperatures providing intense rehydroxylation, the interlayer’s H2O diffusion does not result in permanent rehydration (Derkowski et al., Reference Derkowski, Drits and McCarty2012b).

Fig. 7 An example of an in-situ experiment of an isothermal rehydroxylation reaction observed with TGA (after Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017) on a Cs+-exchanged beidellite (SBId-1) and montmorillonite (SWy-2). Note the initial dehydroxylation to 800°C under dry gas and switching to H2O-rich gas for isothermal rehydroxylation, then secondary dehydroxylation to 1000°C under dry gas. The mass gain under wet gas corresponds to the rebuilding of OH groups

Rehydroxylation is site-specific and is influenced by several factors. If the layer charge is located in the tetrahedral sheet, the opening of the interlayer space during rehydroxylation is relatively easy due to large electrostatic repulsion between adjacent layers. Interlayer cation type also affects the rehydroxylation behavior of a smectite (Kawano & Tomita, Reference Kawano and Tomita1989, Reference Kawano and Tomita1991b; Emmerich et al., Reference Emmerich, Plötze and Kahr2001; Derkowski et al., Reference Derkowski, Drits and McCarty2012a; Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017). Interlayer cations with relatively large radii (>0.13 nm in the eight-fold coordination; Shannon, Reference Shannon1976) act as pillars, which leads to a relatively large interlayer space of the dehydroxylate and enhances its rehydroxylation potential. Cations with large hydration potential also increase H2O migration in the interlayer region, which is the prerequisite for further H2O migration into the octahedral sheet and reformation of OH groups (Derkowski et al., Reference Derkowski, Drits and McCarty2012a,Reference Derkowski, Drits and McCartyb). If an anhydrous cation has a small ionic radius (< 0.08 nm, Li+, Mg2+), it can migrate into a pseudohexagonal cavity or into a vacant position of the former octahedral sheet and physically block access to the residual oxygen, which results in a reduced ability to rehydroxylate. Such cation migration results in irreversible collapse of the interlayer space (Emmerich, Reference Emmerich2000; Emmerich et al., Reference Emmerich, Plötze and Kahr2001; Derkowski et al., Reference Derkowski, Drits and McCarty2012a). In practice, rehydroxylation in smectites is controlled by the fraction of AlAlOH environment in the octahedral sheet, with no Al substitution in the neighboring tetrahedral sites (“pyrophyllite-like” domains), resulting in the lack of charge-compensating interlayer cations blocking the ditrigonal cavity for H2O migration. Another factor is the valence of the exchangeable cation – the greater the valence, the fewer cations are present in the interlayer space and the lower the chance of blocking the pathway of potential rehydroxylation (Derkowski et al., Reference Derkowski, Drits and McCarty2012a).

Experimental Factors Affecting the Results of Thermal Analysis of Smectites

Carrier Gas Composition and Pressure

Smectite thermal analysis is usually performed under dry gas purge in order to remove the products of dehydroxylation or dehydration. The presence of water vapor in the ambient gas changes the reaction pattern, shifting the reaction to higher temperatures following the H2O content in the purge gas (Emmerich et al. Reference Emmerich, Steudel and Merz2017; Koster van Groos & Guggenheim, Reference Koster van Groos and Guggenheim1987, Reference Koster van Groos and Guggenheim1989; Levy & Hurst, Reference Levy and Hurst1993). Smectite dehydration and dehydroxylation reactions under high H2O activity and high pressure were the subject of dedicated experiments enabling the prediction of smectite behavior under conditions relevant to geologic systems (Koster van Groos & Guggenheim, Reference Koster van Groos, Guggenheim, Stucki, Bish and Mumpton1990, and references therein). Vacuum conditions decrease the smectite reaction temperatures with respect to ≥1 bar conditions (Wicks & Ramik, Reference Wicks, Ramik, Stucki, Bish and Mumpton1990).

An analysis of redox-sensitive minerals requires use of a virtually oxygen-free purge gas (Bottom, Reference Bottom and Gabbott2008, Bish & Duffy, Reference Bish, Duffy, Stucki, Bish and Mumpton1990). The presence of oxygen in the purge gas can be tested in such cases using CuC2O4 (Lempart et al., Reference Lempart, Derkowski, Luberda-Durnaś, Skiba and Błachowski2018). The purge gas is considered virtually oxygen-free if no mass gain is detected during the experiment with pure CuC2O4.

Sample Form and Packing

With some exceptions, smectite samples are typically analyzed in a powder form. Before a thermal experiment, powdered samples should be homogenized and preferably sifted through a 100 or 200 μm mesh sieve, which allows for the reduction of the size of smectite aggregates present in the sample. In the case of montmorillonite that has a temperature of amorphization close to that of dehydroxylation, the presence of large aggregates leads to the formation of macroscopic bubbles and hollow spheres, especially for samples with exchangeable cations of high hydration enthalpy, such as Ca2+, Mg2+, or Li+ (Fig. 8; Emmerich, Reference Emmerich and Christidis2010; Emmerich et al., Reference Emmerich, Steudel and Merz2017). The spheres are probably produced by slow heat transfer inward and slow diffusion of H2O outward from the aggregates. Interestingly, the formation of such spheres has virtually no effect on the DTG curve or mass loss during a TG experiment (Fig. 8).

Fig. 8 The texture (upper) and DTG pattern (lower) of unsifted and sifted Ca2+-exchanged montmorillonite sample (SCa-3) after heating to >1000°C at a constant rate of 20°C/min (DTG pattern) and 500°C/min under a flow of 50 mL/min dry N2 gas

The way powder is loaded into a holder can also affect the reaction temperature (Emmerich, Reference Emmerich and Christidis2010); a pressed sample delays the diffusion of the reaction volatile products out of the holder due to decreased permeability. Although the effect is not greater than a few °C, and thus is within the temperature error (Schilling, Reference Schilling1990), use of a loose powder or exactly the same mode of pressing for all samples being compared is recommended. Also, the experiments involving surface reaction with ambient gas (i.e. dehydrogenation under oxidizing conditions) may be strongly affected by the sample permeability to gases.

The maximum reaction temperature correlates positively with the sample mass due to the heat capacity and the diffusion rate of volatile reaction products (Schilling, Reference Schilling1990; Wolters & Emmerich, Reference Wolters and Emmerich2007). However, as found by Derkowski and Kuligiewicz (unpublished data), up to 20% relative of smectite sample weight difference has a negligible effect on the reaction temperature and remains within the analytical error; 20% would thus be the maximum difference in the sample weights recommended for relative comparison. The greater the sample weight, the lower the relative influence of total measurement errors (weighing imprecision, thermal drift) that are usually expressed in absolute quantities, i.e. are independent of the sample size. However, upon heating, a smectite sample can change its bulk volume (due to recrystallization, amorphization) or electrostatic properties (due to dehydration, cation fixation), resulting in the holder overflow and the drop of a sample portion off the holder, which is detrimental to the analysis. Fortunately, a material drop is usually abrupt enough to be distinguishable from a slower thermal reaction of smectites, allowing the results to be rejected from interpretation. Such a sample drop is typical for Li+-montmorillonites at the temperature of dehydroxylation, when a sample is blown off the holder by purge gas (Fig. 9). Using a holder with a lid may prevent that problem (e.g. Emmerich et al., Reference Emmerich, Steudel and Merz2017; Wolters & Emmerich, Reference Wolters and Emmerich2007), but this would also induce a delay in removal of volatiles from the sample. The delayed removal of the reaction products by the gas purge results in momentary production of a water vapor atmosphere around the sample.

Fig. 9 TGA patterns of Li+- and Ca2+-exchanged montmorillonite (SWy-2) heated isothermally for 4 h at temperatures as indicated in the legend followed by ramp-heating at 10°C/min under a flow of 50 mL/min of dry N2. Note the abrupt mass loss at almost the same temperature in Li+-exchanged samples, originating from the sample dropping off the holder

Factors Affecting Evolved Gas Analysis (EGA)

Isothermal reactions are usually not recorded in EGA. The EGA requires a high heating rate and high carrier gas flow, unless a vacuum system is used (Wicks & Ramik, Reference Wicks, Ramik, Stucki, Bish and Mumpton1990 cf. Lempart et al., Reference Lempart, Derkowski, Strączek and Kapusta2020). High heating rates (and possibly large sample masses) are usually needed to keep the evolved gas in a concentration range that is easily determined by the applied spectrometric method.

In the MS signal, H2O is represented by four spectral masses, the dominant m/z = 18, and the minor m/z = 17 (OH+), whereas, the H3O+ (m/z = 19) and O+ (m/z = 16) can be detected in orders of magnitude lower intensity (Meija et al., Reference Meija, Mester and D'Ulivo2006; Emmerich, Reference Emmerich and Christidis2010). With pure dehydration and dehydroxylation reactions, the m/z 17 follows precisely m/z 18, and both signals overlap the DTG curve (Fig. 1). The NH3 evolution from NH4 +-smectites would produce excess m/z 17 whereas dehydrogenation under O2 (or air) would produce an additional H2O signal distinguishable as an excess to the DTG curve (Heller-Kallai, Reference Heller-Kallai1997b, Fig. 5 in Lempart et al., Reference Lempart, Derkowski, Strączek and Kapusta2020; Fig. 1). Dehydrogenation that produces H2 (m/z = 2) is difficult to detect and requires a fast sample-heating rate (≥20°C/min; Lempart et al., Reference Lempart, Derkowski, Strączek and Kapusta2020) in MS or employing a GC–MS system.

Halogen elements present in trioctahedral smectites substituting OH groups (mostly F, also Cl, B) evolve from a smectite sample as hydrogen halides or radicals forming gaseous diatomic molecules (Heller-Kallai, Reference Heller-Kallai1997b; Wicks & Ramik, Reference Wicks, Ramik, Stucki, Bish and Mumpton1990). Strong reactivity of H-halide molecules with metals makes them react with any metal-based elements of the analyzer setup, however, and so ceramic pans and skimmer TGA-MS interfaces are suggested.

Impurities

The mass loss recorded with TGA in a pure smectite should correspond to the cation-dependent dehydration and the theoretical mass loss corresponding to dehydroxylation (Emmerich, Reference Emmerich and Christidis2010; Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017). In nature, smectite never occurs as a pure mineral, even in bentonite beds. Also, even a thorough chemical treatment aimed at removing impurities and grain-size fractionation (Jackson, Reference Jackson1969), usually leaves some impurity in the studied smectite fraction. Some impurities are not reactive in the temperature range corresponding to major smectite reactions, e.g. the presence of quartz and feldspar only reduces the theoretical mass loss due to dehydroxylation. Other impurities, however, have their own thermal reactions that can strongly affect the thermal analysis pattern of smectite samples (Emmerich, Reference Emmerich and Christidis2010). The presence of discrete or interstratified kaolinite frequently coexisting with aluminous smectites (such as beidellites), increases the sample’s mass loss in the dehydroxylation range, as kaolinite has three times more OH groups per formula unit to form H2O than smectite. The above-mentioned impurities can be detected by XRD and for kaolinite also by infrared spectroscopy.

During smectite cation exchange with divalent cations, atmospheric CO2 dissolved in the exchange solution binds with the cations forming carbonate impurities. In TGA tests they provide a significant mass loss corresponding to 44–52% of the carbonate content depending on the type of carbonate; the same mass loss would come from naturally occurring carbonate minerals present in a clayey sample (Emmerich & Smykatz-Kloss, Reference Emmerich and Smykatz-Kloss2002; Emmerich, Reference Emmerich and Christidis2010), if they were not carefully removed during the chemical pretreatment (Jackson, Reference Jackson1969). Therefore, even minute quantities of carbonate minerals, untraceable using other methods (e.g. XRD), can affect the smectite mass-loss determination. Thermal decarbonatization resulting in CO2 evolved usually at temperatures of 700–900°C can be observed with EGA, however, e.g. by monitoring m/z = 44 in the TGA-MS setup (Fig. 1), in order to apply appropriate mass-loss corrections.

Sedimentary or soil organic matter (SOM) can be bound strongly to smectite surfaces and is difficult to remove even by extensive chemical treatment (Mikutta et al., Reference Mikutta, Kleber, Kaiser and Jahn2005; Szymański et al., Reference Szymański, Drewnik, Stolarczyk, Musielok, Gus-Stolarczyk and Skiba2022). Thermal decomposition of recalcitrant SOM under neutral gas results in various evolved molecules, usually below 500°C, identifiable by EGA-MS as a series of various signals with m/z > 50 (Emmerich, Reference Emmerich and Christidis2010), along with those belonging to hydrocarbons (methane with m/z = 16), and CO2 (m/z = 44; Fig. 1). The SOM-evolved H2O (m/z = 18) and NH3 (m/z = 17) overlap with the smectite-derived H2O signal. Under neutral gas, only a portion of SOM decomposes, leaving charcoal as a high-temperature product. If run under oxidizing gas (i.e. air), SOM oxidizes completely, usually below 400°C, resulting in evolved H2O and CO2 identifiable by EGA, and the excess mass loss equal the SOM content. The difference in mass loss and evolved gas composition (especially CO2) between the experiments under neutral and oxidizing gas can reveal the SOM presence. However, structural Fe(II) present in the smectite or associated phases can decrease mass loss under oxidizing conditions, canceling the SOM effect, leaving CO2 (m/z = 44) as the most reliable proxy.

Hydrous sulfate minerals (e.g. gypsum) are the first to be removed during smectite sample pretreatment. However, if present in a sample even in minute amounts, their mass loss upon dehydration at 100–200°C (for example, 21% in gypsum) increases smectite dehydration. Desulfurization occurring at >700°C – in the range of trioctahedral smectite dehydroxylation – results in a loss of almost 50% of gypsum mass, which can be detected easily using EGA-MS (m/z: 64, 48, 32; Emmerich, Reference Emmerich and Christidis2010; Dietel et al., Reference Dietel, Warr, Bertmer, Steudel, Grathoff and Emmerich2017). If analyzed under oxidizing conditions, trace sulfides, e.g. pyrite, reacting with O2 also release SO2 albeit at lower temperatures (<500°C; Dietel et al., Reference Dietel, Warr, Bertmer, Steudel, Grathoff and Emmerich2017). When reacting with H2O coming from a smectite’s dehydroxylation at 500–800°C, sulfides produce H2S molecules (Levy & White, Reference Levy and White1988).

Kinetic Calculations of Smectite Thermal Reactions

Theoretical Background

As in any other reaction, de- and re-hydration and de- and re-hydroxylation can be tracked using common kinetic approaches. The comprehensive review of methodology applicable in solid-state kinetics studies of thermally stimulated processes can be found elsewhere (e.g. Khawam & Flanagan, Reference Khawam and Flanagan2006a,Reference Khawam and Flanaganb; Vyazovkin et al., Reference Vyazovkin, Burnham, Criado, Pérez-Maqueda, Popescu and Sbirrazzuoli2011, Reference Vyazovkin, Burnham, Favergeon, Koga, Moukhina, Pérez-Maqueda and Sbirrazzuoli2020). A brief introduction of the key methodological points is presented below, with special emphasis on methods applicable to the thermal reactions of smectite.

The majority of methods in solid-state kinetic analysis of thermally stimulated reactions are based on the following equation:

(5) d α d t = k T f ( α )

which states that the reaction rate (dα/dt) is dependent on the absolute temperature (in K) through the reaction rate constant (k(T)) and on the extent of conversion (α) through the reaction mechanism (f(α)). The extent of conversion is the dimensionless parameter describing progress of the reaction and is scaled between 0 and 1, where 0 corresponds to the start and 1 to the completion of the reaction. The extent of conversion must correspond to a change of some measurable property of a material that represents the reaction progress. In TGA practice, this parameter is mass change and α is expressed conveniently as the ratio of mass change at a given time during the reaction to the mass change corresponding to the completion of the reaction e.g. to full dehydroxylation. Calculating the reaction progress from DTA and DSC data requires prior integration of the peak over time or temperature. The reaction mechanism in Eq. 5 is a function describing the progress of a reaction based on the molecular reaction mechanism and/or reactant geometry and can take various mathematical forms (Khawam & Flanagan, Reference Khawam and Flanagan2006a,Reference Khawam and Flanaganb; Vyazovkin et al., Reference Vyazovkin, Burnham, Criado, Pérez-Maqueda, Popescu and Sbirrazzuoli2011). The dependence of the reaction rate constant on temperature is most commonly expressed with the Arrhenius equation:

(6) k T = A exp - E a R T

where A is the pre-exponential factor, E a is the activation energy, and R is the gas constant.

Substituting Eq. 6 into Eq. 5 gives:

(7) d α d t = A exp - E a R T f α

or in the logarithmic form:

(8) ln d α d t = ln A + ln f α - E a R T

Finding the triplet f(α), E a, and A allows for the full kinetic description of the reaction studied. However, many kinetic studies present only an estimation of E a as the most relevant reaction-specific parameter. Depending on the mode in which the thermal analysis experiment was performed, E a can be calculated using isothermal or non-isothermal data, the latter most commonly coming from linear heating experiments (Fig. 10). In any case, using at least three temperature programs is strongly advised (Vyazovkin et al., Reference Vyazovkin, Burnham, Criado, Pérez-Maqueda, Popescu and Sbirrazzuoli2011; Figs. 10, 11b), and at least five points are required to detect any non-Arrhenius behavior of the reaction, which manifests itself by non-linearity of Eqs. 6 and 8 (Vyazovkin et al., Reference Vyazovkin, Burnham, Favergeon, Koga, Moukhina, Pérez-Maqueda and Sbirrazzuoli2020).

Fig. 10 The examples of linear, ramp-heating of a Na+-beidellite sample to 1000°C under a 50 mL/min flow of dry N2 and at various heating rates. The rates are indicated at the DTG peak in °C/min. The shift between the DTGmax peaks can be used to determine the activation energy of dehydroxylation

Fig. 11 A, C–F Activation energy (E a) of the dehydroxylation of various dioctahedral smectites, calculated using the 0.01 α-increments and the Friedman method, and B an example of E a determination using the Kissinger method and the DTG peak temperature (Fig. 10). Note the similarity of the E a values and trends in the dehydroxylation of two Ca2+-exchanged Mg-rich montmorillonites (E, F). For each TGA run, the total mass loss upon dehydroxylation was divided into 100 portions of equal mass difference (0.01α). For each 0.01α, the ln(dα/dt) was plotted against the corresponding 1/T using four or five data points from the TGA curves recorded with a different heating rate (B). E E a is then calculated from the slope of ln(dα/dt) vs 1/T plot (Eq. 8). The correlation coefficient, R2, was computed for each of 0.01α. All E a values shown (A, C–F) are reported for R2 > 0.997 (arbitrary cut-off),

Two general types of kinetic calculations are used for thermal analysis data. The model-fitting approach is based on fitting one of the reaction models available in the literature to Eq. 5 or its integral form and finding the best-fitting model, simply by comparing its linearity with experimental data. The reaction mechanism may or may not include adjustable parameters, which can later be interpreted in terms of a rate-controlling model (Bray & Redfern, Reference Bray and Redfern2000; Giese, Reference Giese, Stucki, Bish and Mumpton1990; Vyazovkin et al., Reference Vyazovkin, Burnham, Criado, Pérez-Maqueda, Popescu and Sbirrazzuoli2011). In smectites, the isothermal model-fitting approach has been proved useful for determining the kinetics of dehydration (Bray & Redfern, Reference Bray and Redfern1999), dehydroxylation (Bray & Redfern, Reference Bray and Redfern2000), and rehydroxylation (Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017).

The second type of calculations are those using the isoconversional methods (Criado et al., Reference Criado, Sánchez-Jiménez and Pérez-Maqueda2008, Perez-Rodriguez et al., Reference Perez-Rodriguez, Duran, Sanchez Jimenez, Franquelo, Perejón, Pascual-Cosp and Pérez-Maqueda2010; Vyazovkin, Reference Vyazovkin2015, Vyazovkin, Reference Vyazovkin2021), in which E a and A are calculated without any assumptions of a particular kinetic model or the reaction order. The basic principle behind the isoconversional methods is that the value of f(α) in Eq. 5 is independent of T. In other words, the value of f(α) is constant for a given α, irrespective of the temperature program. In its simplest form for non-isothermal conditions (the Friedman method), calculating isoconversional E a involves plotting the values of ln(dα/dt) against 1/T for a given α. The isoconversional approach enables changes in E a to be detected during the reaction progress with no prior assumptions regarding the reaction mechanism. Such changes can, for example, reflect the changes in the bond strength with the reaction progress (Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017).

Another calculation method based on the isoconversional principle and representing its simplified version is the method of Kissinger (Reference Kissinger1957). It is applicable to linear heating conditions and is based on the shift of the reaction peak when changing the heating rate, thus directly applicable to DTA, DSC, and DTG data (Fig. 10). This method employs the fact that for the maximum reaction rate (DTG peak), d2α/dt 2 = 0. Taking this into account, the Kissinger equation can be derived from Eq. 7:

(9) ln β T 2 = ln - A R E a d f ( α ) d α - E a R 1 T

where β is the linear heating rate (dT/dt). Values of all other parameters (T, f(α), and α) in Eq. 9 refer to values at the maximum reaction rate. In practice, the left side of Eq. 9 is plotted against 1/T, which should produce a straight line, the slope of which can be used to calculate E a (i.e.Vyazovkin, Reference Vyazovkin, Burnham, Criado, Pérez-Maqueda, Popescu and Sbirrazzuoli2011; Fig. 11b).

Note that some studies apply kinetics analysis to data obtained from a single temperature program only (i.e. only one isothermal heating temperature or only one linear heating rate), which is especially common in the older literature. While this approach was justifiable because of hardware and computational limitations in those times, it is now generally considered unreliable (Vyazovkin et al., Reference Vyazovkin, Burnham, Criado, Pérez-Maqueda, Popescu and Sbirrazzuoli2011).

Kinetics of Smectite Thermal Reactions

Dehydration

Despite the large number of studies of smectite–water interactions, kinetics studies of smectite dehydration employing thermal analysis methods do not present a consistent picture. Besides the dependence of dehydration energy on the hydration enthalpy of the interlayer cations (Webb et al., Reference Webb, Stanley and Scheiner1986), it has been demonstrated that a change occurs in the reaction mechanism along the course of dehydration (Koster van Groos & Guggenhem, Reference Koster van Groos, Guggenheim, Stucki, Bish and Mumpton1990; Bray & Redfern, Reference Bray and Redfern1999; Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017; Poinsignon et al., Reference Poinsignon, Yvon and Mercier1982; Zabat & Van Damme, Reference Zabat and Van Damme2000). As already mentioned, physisorbed water in the micro- and mesopores and bulk interlayer water represent the ‘weakly-bound water’ (WBW). WBW is removed preferentially during dehydration because of its weaker interactions with the smectite surface. Change in the reaction mechanism during dehydration and the associated change in E a (Table 2) correspond to the completion of weakly bound water removal and the onset of strongly bound water removal. The E a of dehydration of WBW is in the range 35–79 kJ/mol for smectites having monovalent and divalent interlayer cations commonly found in nature (Ca2+, Mg2+, K+, Na+, Li+; Table 2). As deduced from Table 2, E a of WBW dehydration in general does not depend on the exchangeable cation type or chemistry of the 2:1 layer. Once weakly bound water is removed, removal of the sequential portions of strongly bound water requires increasingly higher energy, as is manifested by the continuous increase in E a with reaction progress, up to 190 kJ/mol (Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017). This was interpreted as the result of bond reconfiguration upon removal of water molecules from hydration spheres of interlayer cations, in agreement with theoretical calculations (Pavlov et al., Reference Pavlov, Siegbahn and Sandström1998).

Table 2 Comparison of the E a of dehydration of weakly bound water (WBW) and strongly bound water (SBW) from studies employing various methods

*Zabat and Van Damme (Reference Zabat and Van Damme2000) assumed the first-order reaction mechanism for dehydration and presented a new computation method, using principles similar to those of the Kissinger method

Dehydroxylation

In the majority of dioctahedral smectites studied, E a of dehydroxylation does not vary much with the reaction progress (Fig. 11A,C,D). If the initial (α < 0.2) and final (α > 0.8) portions of the data, corresponding also to lower R2 coefficients because of an overlap with other thermal effects are excluded, the E a values remain similar over a broad range of α for beidellite SBId-1, cv beidellitic montmorillonite SWy-1, and ferruginous smectite SWa-1 (an approximation for nontronite) (Fig. 11). In Mg-rich montmorillonites that have charge generated exclusively in the octahedral sheet, however, E a increases significantly with α, especially in the samples saturated with divalent cations (Fig. 11E,F). The increase of E a in montmorillonite remains in line with the interpretation of Koster van Groos and Guggenheim (Reference Koster van Groos and Guggenheim1989) who suggested the increase in the residual OH bond strength upon sequential removal of OH groups in heterogeneous structure.

The kinetics of smectite dehydroxylation is controlled by the same structural factors as the temperature of dehydroxylation (Figs. 1 and 3). Activation energy of dehydroxylation as computed from linear heating experiments with isoconversional methods correlates linearly with the temperature of maximum dehydroxylation, based on the DTG peak (Fig. 12, Table 3). The observed relationship implies that the rate-controlling mechanism of dehydroxylation is the same in all smectites regardless of their crystal-chemical composition.

Fig. 12 The E a of dehydroxylation of smectite reference samples presenting values averaged over α = 0.2–0.8 calculated using the Friedman method (upper, cf. Figure 11) and E a calculated using the conventional Kissinger method based on DTG peak temperature (lower, cf. Fig. 10), plotted against the maximum of dehydroxylation (DTGmax; Table 3). The E a error in the Kissinger method is ~5%. Filled shapes denote the Cs+-exchanged form; note their lower position with respect to other cations

Table 3 E a of dehydroxylation calculated as using the 0.01 α-increments with values averaged over α = 0.2–0.6 calculated using the Friedman method (Fig. 11) and E a calculated using the conventional Kissinger method based on DTG peak temperature (Fig. 10). The maximum of dehydroxylation (DTGmax) was determined using the 5°C/min constant rate TG measurement

*estimated E a error in the Kissinger method is ~ 5% relative

Another factor influencing the kinetics of smectite dehydroxylation is the interlayer cation type. The application of the model-fitting approach to isothermal data allowed demonstration that the interlayer cation type can modify the dominant reaction model, which provides the best description of the process. Bray and Redfern (Reference Bray and Redfern2000) suggested that dehydroxylation in smectite saturated with divalent cations (Ca2+ or Mg2+) is diffusion-controlled, while dehydroxylation in smectites saturated with monovalent cations (Na+ or K+) follows the first-order reaction kinetics. The influence of the interlayer cation on the dehydroxylation E a value is less clear (Bray & Redfern, Reference Bray and Redfern2000; Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017) but it seems that Mg2+-exchanged smectites have significantly lower E a values than Ca2+- or Na+-exchanged counterparts (Bray & Redfern, Reference Bray and Redfern2000; Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2017). A speculative explanation may be that Mg2+, which is capable of retaining dissociated H2O (Mg-OH) up to temperatures of dehydroxylation, acts as a flux and lowers the E a of dehydroxylation.

All of the above-mentioned results and interpretations come from studies using simple (isothermal or linear heating rate) thermal experiments. An alternative approach based on multiple heating–cooling cycles was employed by Drits et al. (Reference Drits, Derkowski and McCarty2012a) to determine the kinetics of partial dehydroxylation in dioctahedral smectites and other 2:1 layer minerals. The protocol involves heating the samples with a constant heating rate and subsequent cooling by increasing each following cycle’s maximum temperature in small increments (i.e. 25°C). If a temperature increment between cycles is sufficiently small, the increase of α within one and the same cycle is negligible; therefore, α can be considered as constant for the cycle, meeting the requirement of the isoconversional kinetic approach. Within the cycle, the majority of the ln(dα/dt) over 1/T returns a linear fit, which Drits et al. (Reference Drits, Derkowski and McCarty2012a) interpreted as the zeroth-order kinetics of partial dehydroxylation. Such a heating–cooling cycle approach apparently allows tracking of E a in different stages of the reaction or portions of sample reacted at subsequently higher temperatures. However, for a portion of the sample reacting with constant f(α), the Arrhenius relationship represents the population that has already been treated thermally in the preceding cycle, hence, not corresponding to the actual rate-controlling mechanism in full reaction (Muñoz-Santiburcio et al., Reference Muñoz-Santiburcio, Kosa, Hernández-Laguna, Sainz-Díaz and Parrinello2012; Ogloza & Malhotra, Reference Ogloza and Malhotra1989). The difficulty in explaining the concave shape of E a evolution over α (Fig. 13) and the maximum E a being much lower than that for full reaction (Figs. 11 and 12) implies the limited use of the heating–cooling protocol and the need for specific interpretation of the kinetic data.

Fig. 13 An example of the calculation of E a of smectite dehydroxylation for Ca2+-saturated SCa-3 montmorillonite based on a heating–cooling cycles TGA experiment (Drits et al., Reference Drits, Derkowski and McCarty2012a). The maximum temperature (in °C) of each consecutive heating cycle is given in italics. Compare E a values with Fig. 11e

Rehydration and Rehydroxylation

The kinetics of smectite rehydration is methodologically challenging to study because the maximum amount of water which can be adsorbed by the smectite’s structure is dependent on temperature; for a fixed H2O molar content in the carrier gas, the lower the reaction temperature, the greater the rehydration (Fig. 6), but for constant RH the degree of rehydration is likely to be similar in spite of different temperatures (Vieillard et al., Reference Vieillard, Gailhanou, Lassin, Blanc, Bloch, Gaboreau, Fialips and Made2019). This renders non-isothermal experiments inapplicable. The rehydration process can be most conveniently tracked using the TGA experimental setup, allowing in situ dehydration and rehydration (Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017). The mechanism of rehydration at temperatures <100°C cannot be described well by a commonly used solid-state reaction model (such as those presented in Table 1 of Khawam & Flanagan, Reference Khawam and Flanagan2006a). Models developed specifically for adsorption kinetics, such as a pseudo n th order model, usually provide much better description of the process (see Largitte & Pasquier, Reference Largitte and Pasquier2016, for a comprehensive review of models used in the adsorption kinetics). Rehydration kinetics are influenced by the exchangeable cation type; cations with low hydration energy (such as Cs+) adsorb quickly the small amount of water which they are capable of holding. For smectite saturated with cations with high hydration energy (such as Mg2+), the adsorption process continues over longer timescales (Fig. 6).

Isothermal rehydroxylation experiments under H2O vapor at temperatures above those typical for rehydration have shown that kinetics of smectite rehydroxylation can be described by the time-to-the-quarter (TTTQ) model expressed as relative mass gain equal to the temperature-dependent rate constant multiplied by the 4th root of reaction time (Hamilton & Hall, Reference Hamilton and Hall2012; Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017). Explained as corresponding to restricted, single-file diffusion in fired clay minerals, the TTTQ model has been applied to clay-based traditional ceramics and used to predict the time elapsed since firing, hence for archeological dating (Hare et al., Reference Hare, Kärger, Moinester and Piasetzky2016; Wilson et al., Reference Wilson, Carter, Hall, Hoff, Ince, Savage, McKay and Betts2009). Both the root coefficient of the TTTQ (Barret, Reference Barrett2017) and the overall validity of TTTQ model for kinetic calculations have been criticized strongly (Kuligiewicz & Derkowski, Reference Kuligiewicz and Derkowski2021).

Studies of E a of smectite rehydroxylation are still rare and the E a values calculated based on the TTTQ model remain within an unexpectedly broad range, between 19 and 173 kJ/mol for experiments performed at 110–400°C (Derkowski & Kuligiewicz, Reference Derkowski and Kuligiewicz2017; Stevenson & Gurnick, Reference Stevenson and Gurnick2016). However, Kuligiewicz and Derkowski (Reference Kuligiewicz and Derkowski2021) recently proved that the TTTQ model describes well the progress of various reactions governed by unexpectedly different mechanisms. E a obtained with the TTTQ model for those reactions was grossly incorrect despite an apparent agreement with simulated data, making the TTTQ model unreliable. Therefore, although the TTTQ model fits well the smectite isothermal rehydroxylation in some cases and has served for occasional successful dating of clay ceramics, its link to kinetics formalism and the mechanism of rehydroxylation are still unclear and a need exists to develop an alternative approach.

Acknowledgements

The authors thank Katja Emmerich, Georgios Chryssikos, Marek Szczerba, Małgorzata Lempart-Drozd, and three anonymous reviewers for their valuable comments. Nadine Kanik kindly provided corrections of the English language. The work was financed by the National Science Centre, Poland; grant OPUS-13 UMO-2017/25/B/ST10/01675.

Declarations

Conflict of Interest

The authors declare that they have no conflict of interest.

Footnotes

Associate Editor: George Christidis

References

Barrett, G. T. (2017). Rehydroxylation (RHX) dating: Trials on post-medieval brick using a component based approach. Journal of Archaeological Science: Reports, 16, 489502.Google Scholar
Berend, I., Cases, J. M., Francois, M., Uriot, J. P., Michot, L., Masion, A., & Thomas, F. (1995). Mechanism of adsorption and desorption of water vapor by homoionic montmorillonites; 2, The Li+, Na+, K+, Rb+ and Cs+-exchanged forms. Clays and Clay Minerals, 43(3), 324336.CrossRefGoogle Scholar
Bish, D. L. & Duffy, C. J. (1990). Thermogravimetric Analysis of Minerals. In Stucki, J.W., Bish, D., & Mumpton, F. A. (Eds), CMS Workshop Lectures, Vol. 3. Thermal analysis in clay science. The Clay Minerals Society.https://doi.org/10.1346/CMS-WLS-3.4Google Scholar
Boller, A., Jin, Y., & Wunderlich, B. (1994). Heat capacity measurement by modulated DSC at constant temperature. Journal of Thermal Analysis and Calorimetry, 42(2–3), 307330.CrossRefGoogle Scholar
Botella, V., Timón, V., Escamilla-Roa, E., Hernandez-Languna, A., & Sainz-Díaz, C. I. (2004). Hydrogen bonding and vibrational properties of hydroxy groups in the crystal lattice of dioctahedral clay minerals by means of first principles calculations. Physics and Chemistry of Minerals, 31(8), 475486.CrossRefGoogle Scholar
Bottom, R. (2008). Thermogravimetric Analysis. In Gabbott, P. (Ed.), Principles and applications of thermal analysis. John Wiley & Sons.Google Scholar
Bradley, W. F., & Grim, R. E. (1951). High temperature thermal effects of clay and related materials. American Mineralogist, 36(3–4), 182201.Google Scholar
Bray, H. J., & Redfern, S. A. T. (1999). Kinetics of dehydration of Ca-montmorillonite. Physics and Chemistry of Minerals, 26(7), 591600.CrossRefGoogle Scholar
Bray, H. J., & Redfern, S. A. T. (2000). Influence of counterion species on the dehydroxylation of Ca2+-, Mg2+-, Na+-and K+-exchanged Wyoming montmorillonite. Mineralogical Magazine, 64(2), 337346.CrossRefGoogle Scholar
Brett, N. H., MacKenzie, K. J. D., & Sharp, J. H. (1970). The thermal decomposition of hydrous layer silicates and their related hydroxides. Quarterly Reviews, Chemical Society, 24(2), 185207.CrossRefGoogle Scholar
Byrn, S. R., Zografi, G., & Chen, X. S. (2017). Differential scanning calorimetry and thermogravimetric analysis. Solid State Properties of Pharmaceutical Materials (pp. 124142). Wiley.CrossRefGoogle Scholar
Cases, J. M., Bérend, I., François, M., Uriot, L. P., Michot, L. J., & Thomas, F. (1997). Mechanism of adsorption and desorption of water vapor by homoionic montmorillonite: 3. The Mg2+, Ca2+, Sr2+ and Ba2+ exchanged forms. Clays and Clay Minerals, 45(1), 822.CrossRefGoogle Scholar
Christidis, G. E., Aldana, C., Chryssikos, G. D., Gionis, V., Kalo, H., Stöter, M., Breu, J., & Robert, J. L. (2018). The nature of laponite: pure hectorite or a mixture of different trioctahedral phases? Minerals, 8(8), 314.CrossRefGoogle Scholar
Criado, J., Sánchez-Jiménez, P., & Pérez-Maqueda, L. (2008). Critical study of the isoconversional methods of kinetic analysis. Journal of Thermal Analysis and Calorimetry, 92(1), 199203.CrossRefGoogle Scholar
Cuadros, J. (2002). Structural insights from the study of Csexchanged smectites submitted to wetting-and-drying cycles. Clay Minerals, 37(3), 473486.CrossRefGoogle Scholar
Dainyak, L. G., Dritz, V. A., Kudryavtzev, D. I., Simanovitch, I. M., & Slonimskaya, M. V. (1981). Novaya mineral'naya raznovidnost' trioktaedricheskech smektitov iz effuzivnych basaltov Tungusskoi sineklizy (New mineral variant of trioctahedral smectites from effusive basalts of Tungusskaya sineclise). Lithology and Mineral Resources, 6, 123129. (in Russian).Google Scholar
Della Gatta, G., Richardson, M. J., Sarge, S. M., & Stølen, S. (2006). Standards, calibration, and guidelines in microcalorimetry. Part 2. Calibration standards for differential scanning calorimetry (IUPAC Technical Report). Pure and Applied Chemistry, 78(7), 14551476.CrossRefGoogle Scholar
Derkowski, A., & Kuligiewicz, A. (2017). Rehydroxylation in smectites and other clay minerals observed in-situ with a modified thermogravimetric system. Applied Clay Science, 136, 219229.CrossRefGoogle Scholar
Derkowski, A., Drits, V. A., & McCarty, D. K. (2012). Nature of rehydroxylation in dioctahedral 2: 1 layer clay minerals. American Mineralogist, 97, 610629.CrossRefGoogle Scholar
Derkowski, A., Drits, V. A., & McCarty, D. K. (2012). Rehydration in a dehydrated-dehydroxylated smectite in environment of low water vapor content. American Mineralogist, 97, 110127.CrossRefGoogle Scholar
Dietel, J., Warr, L. N., Bertmer, M., Steudel, A., Grathoff, G. H., & Emmerich, K. (2017). The importance of specific surface area in the geopolymerization of heated illitic clay. Applied Clay Science, 139, 99107.CrossRefGoogle Scholar
Ding, Z., & Frost, R. L. (2002). Controlled rate thermal analysis of nontronite. Thermochimica Acta, 389(1–2), 185193.CrossRefGoogle Scholar
Drits, V. A., & Zviagina, B. (2009). Trans-vacant and cisvacant 2: 1 layer silicates: Structural features, identification, and occurrence. Clays and Clay Minerals, 57(4), 405415.CrossRefGoogle Scholar
Drits, V. A., Besson, G., & Muller, F. (1995). An improved model for structural transformation of heat-treated aluminous dioctahedral 2: 1 layer silicates. Clays and Clay minerals, 43(6), 718731.CrossRefGoogle Scholar
Drits, V. A., Lindgreen, H., Salyn, A. L., Ylagan, R., & McCarty, D. K. (1998). Semiquantitative determination of trans-vacant and cis-vacant 2: 1 layers in illites and illite-smectites by thermal analysis and X-ray diffraction. American Mineralogist, 83(12), 11881198.CrossRefGoogle Scholar
Drits, V. A., Derkowski, A., & McCarty, D. K. (2012). Kinetics of partial dehydroxylation in dioctahedral 2: 1 layer clay minerals. American Mineralogist, 97, 930950.CrossRefGoogle Scholar
Drits, V. A., McCarty, D. K., & Derkowski, A. (2012). Mixed-layered structure formation during trans-vacant Al-rich illite dehydroxylation. American Mineralogist, 97, 19221938.CrossRefGoogle Scholar
El-Barawy, K. A., Girgis, B. S., & Felix, N. S. (1986). Thermal treatment of some pure smectites. Thermochimica Acta, 98, 181189.CrossRefGoogle Scholar
Emmerich, K. (2000). Spontaneous rehydroxylation of a dehydroxylated cis-vacant montmorillonite. Clays and Clay Minerals, 48(3), 405408.CrossRefGoogle Scholar
Emmerich, K. (2010). Thermal Analysis in the Characterization and Processing of Industrial Minerals. In Christidis, G. E. (Ed.), Advances in the Characterization of Industrial Minerals. European Mineralogical Union.Google Scholar
Emmerich, K., & Smykatz-Kloss, W. (2002). Exothermic effects in soils during thermal analysis. Clay Minerals, 37(4), 575582.CrossRefGoogle Scholar
Emmerich, K., Madsen, F. T., & Kahr, G. (1999). Dehydroxylation behavior of heat-treated and steam-treated homoionic cis-vacant montmorillonites. Clays and Clay Minerals, 47(5), 591604.CrossRefGoogle Scholar
Emmerich, K., Plötze, M., & Kahr, G. (2001). Reversible collapse and Mg2+ release of de-and rehydroxylated homoionic cis-vacant montmorillonites. Applied Clay Science, 19(1–6), 143154.CrossRefGoogle Scholar
Emmerich, K., Wolters, F., Kahr, G., & Lagaly, G. (2009). Clay profiling: the classification of montmorillonites. Clays and Clay Minerals, 57(1), 104114.CrossRefGoogle Scholar
Emmerich, K., Steudel, A., & Merz, D. (2017). Dehydroxylation of dioctahedral smectites in water vapor atmosphere. Applied Clay Science, 137, 15.CrossRefGoogle Scholar
Emmerich, K., Giaudo, N., Schuhmann, R., Schnetzer, F., Kaden, H., & Thissen, P. (2018). On the Prediction of Water Contents in Na-Saturated Dioctahedral Smectites. Journal of Physical Chemistry C, 122(13), 74847493.CrossRefGoogle Scholar
Farmer, V. C., Russell, J. D., McHardy, W. J., Newman, A. C. D., Ahlrichs, J. L., & Rimsaite, J. Y. H. (1971). Evidence for loss of protons and octahedral iron from oxidized biotites and vermiculites. Mineralogical Magazine, 38(294), 121137.CrossRefGoogle Scholar
Ferrage, E., Lanson, B., Michot, J. L., & Robert, J.-L. (2010). Hydration properties and interlayer organization of water and ions in synthetic Na-smectite with tetrahedral layer charge. Part 1. Results from X-ray difraction profile modeling. Journal of Physical Chemistry C, 114, 45154526.CrossRefGoogle Scholar
Gabbott, P. (2008). Principles and applications of thermal analysis. John Wiley & Sons.CrossRefGoogle Scholar
Gailhanou, H., Vieillard, P., Blanc, P., Lassin, A., Denoyel, R., Bloch, E., De Weireld, G., Gaboreau, S., Fialips, C. I., Madé, B., & Giffaut, E. (2017). Methodology for determining the thermodynamic properties of smectite hydration. Applied Geochemistry, 82, 146163.CrossRefGoogle Scholar
Garg, N., & Skibsted, J. (2014). Thermal activation of a pure montmorillonite clay and its reactivity in cementitious systems. The Journal of Physical Chemistry C, 118(21), 1146411477.CrossRefGoogle Scholar
Giese, R. F. (1990). Precision Scanning Calorimetry of Clay Minerals and Their Intercalates: In Stucki, J. W., Bish, D., & Mumpton, F. A. (Eds), CMS Workshop Lectures, Vol. 3. Thermal analysis in clay science. The Clay Minerals Society. https://doi.org/10.1346/CMS-WLS-3.2Google Scholar
Girgis, B. S., El-Barawy, K. A., & Felix, N. S. (1987). Dehydration kinetics of some smectites: a thermogravimetric study. Thermochimica Acta, 111, 919.CrossRefGoogle Scholar
Grekov, D., Montavon, G., Robinet, J. C., & Grambow, B. (2019). Smectite fraction assessment in complex natural clay rocks from interlayer water content determined by thermogravimetric and thermoporometry analysis. Journal of Colloid and Interface Science, 555, 157165.CrossRefGoogle ScholarPubMed
Grim, R. E., & Bradley, W. F. (1940). Investigation of the effect of heat on the clay minerals illite and montmorillonite. Journal of the American Ceramic Society, 23(8), 242248.CrossRefGoogle Scholar
Grim, R. E., & Rowland, R. A. (1942). Differential thermal analyses of clay minerals and other hydrous materials. Part 1. American Mineralogist, 27, 746761.Google Scholar
Grim, R. E., & Rowland, R. A. (1942). Differential thermal analyses of clay minerals and other hydrous materials. Part 2. American Mineralogist, 27, 901–818.Google Scholar
Guggenheim, S. (1990). The dynamics of thermal decomposition in aluminous dioctahedral 2: 1 layer silicates: A crystal chemical model. Sciences Géologiques, Bulletins et Mémoires, 86(1), 99107.Google Scholar
Guggenheim, S., Chang, Y. H., & Koster van Groos, A. F. (1987). Muscovite dehydroxylation; high-temperature studies. American Mineralogist, 72(5–6), 537550.Google Scholar
Hamilton, A., & Hall, C. (2012). A review of rehydroxylation in fired-clay ceramics. Journal of the American Ceramic Society, 95(9), 26732678.CrossRefGoogle Scholar
Hare, V. J., Kärger, J., Moinester, M., & Piasetzky, E. (2016). Testing the (time)1/4 diffusion law of rehydroxylation in fired clays: evidence for single-file diffusion in porous media? Diffusion fundamentals Organization, 25(5), 111.Google Scholar
Heide, K., & Földvari, M. (2006). High temperature mass spectrometric gas-release studies of kaolinite Al2[Si2O5(OH)4] decomposition. Thermochimica Acta, 446(1–2), 106112.CrossRefGoogle Scholar
Heller-Kallai, L. (1997). Reduction and reoxidation of nontronite: the data reassessed. Clays and Clay Minerals, 45(3), 476479.CrossRefGoogle Scholar
Heller-Kallai, L. (1997). The nature of clay volatiles and condensates and the effect on their environment: A review. Journal of Thermal Analysis and Calorimetry, 50(1–2), 145156.CrossRefGoogle Scholar
Heller-Kallai, L., & Rozenson, L. (1980). Dehydroxylation of dioctahedral phyllosilicates. Clays and Clay Minerals, 28(5), 355368.CrossRefGoogle Scholar
Hofmann, U., & Klemen, R. (1950). Verlust der austauschfähigkeit von lithiumionen an bentonit durch erhitzung. Zeitschrift für Anorganische Chemie, 262(1–5), 9599.CrossRefGoogle Scholar
Hofmann, U., Endell, K., & Wilm, D. (1933). Kristallstruktur und quellung von montmorillonit. Zeitschrift für Kristallographie-Crystalline Materials, 86(1–6), 340348.CrossRefGoogle Scholar
Jackson, M. L. (1969). Soil Chemical Analysis: Advanced Course. Edition 2 pp. 895 pp. UW-Madison.Google Scholar
Kanik, N. J., Longstaffe, F. J., Kuligiewicz, A., & Derkowski, A. (2022). Systematics of smectite hydrogen-isotope composition: structural hydrogen versus adsorbed water. Applied Clay Science, 216, 106338.CrossRefGoogle Scholar
Kawano, M., & Tomita, K. (1989). X-ray studies of rehydration behaviors for montmorillonite. Clay Science, 7(5), 277287.Google Scholar
Kawano, M., & Tomita, K. (1991). Dehydration and rehydration of saponite and vermiculite. Clays and Clay Minerals, 39(2), 174183.Google Scholar
Kawano, M., & Tomita, K. (1991). X-ray powder diffraction studies on the rehydration properties of beidellite. Clays and Clay Minerals, 39(1), 7783.CrossRefGoogle Scholar
Khawam, A., & Flanagan, D. R. (2006). Basics and Applications of Solid-State Kinetics: A Pharmaceutical Perspective. Journal of Pharmaceutical Sciences, 95(3), 472498.CrossRefGoogle Scholar
Khawam, A., & Flanagan, D. R. (2006). Solid-State Kinetic Models: Basics and Mathematical Fundamentals. The Journal of Physical Chemistry B, 110(35), 1731517328.CrossRefGoogle ScholarPubMed
Kissinger, H. E. (1957). Reaction kinetics in differential thermal analysis. Analytical Chemistry, 29, 17021706.CrossRefGoogle Scholar
Kohyama, N., Shimoda, S., & Sudo, T. (1973). Iron-rich saponite (ferrous and ferric forms). Clays and Clay Minerals, 21(4), 229237.CrossRefGoogle Scholar
Komadel, P., Hrobáriková, J., & Koppelhuber-Bitschnau, B. (2002). Hydration of reduced-charge montmorillonite. Clay Minerals, 37(3), 543550.CrossRefGoogle Scholar
Komadel, P., Madejová, J., & Bujdák, J. (2005). Preparation and properties of reduced-charge smectites–a review. Clays and Clay Minerals, 53(4), 313334.CrossRefGoogle Scholar
Koster van Groos, A. F. & Guggenheim, S. (1990). High pressure Differential Thermal Analysis of Clays. In Stucki, J. W., Bish, D., & Mumpton, F. A. (Eds.), CMS Workshop Lectures, Vol. 3. Thermal Analysis in Clay Science. The Clay Minerals Society. https://doi.org/10.1346/CMS-WLS-3.3Google Scholar
Koster van Groos, A. F., & Guggenheim, S. (1987). Dehydration of a Ca-and a Mg-exchanged montmorillonite (SWy-1) at elevated pressures. American Mineralogist, 72(3–4), 292298.Google Scholar
Koster van Groos, A. F., & Guggenheim, S. (1989). Dehydroxylation of Ca-and Mg-exchanged montmorillonite. American Mineralogist, 74(5–6), 627636.Google Scholar
Kuligiewicz, A., & Derkowski, A. (2017). Tightly bound water in smectites. American Mineralogist, 102, 10731090.CrossRefGoogle Scholar
Kuligiewicz, A., & Derkowski, A. (2021). Rehydroxylation of fired clays: is the time to the quarter (TTTQ) model correct? Journal of Archaeological Science, 125, 105301.CrossRefGoogle Scholar
Kuligiewicz, A., Derkowski, A., Szczerba, M., Gionis, V., & Chryssikos, G. D. (2015). Revisiting the infrared spectrum of the water–smectite interface. Clays and Clay Minerals, 63, 1529.CrossRefGoogle Scholar
Largitte, L., & Pasquier, R. (2016). A review of the kinetics adsorption models and their application to the adsorption of lead by an activated carbon. Chemical Engineering Research and Design, 109, 495504.CrossRefGoogle Scholar
Le Chatelier, M. H. (1887). De l'action de la chaleur sur les argiles. Bulletin de la Société française de Minéralogie, 10(5–6), 204211.CrossRefGoogle Scholar
Lear, P. R., & Stucki, J. W. (1985). Role of structural hydrogen in the reduction and reoxidation of iron in nontronite. Clays and Clay Minerals, 33(6), 539545.CrossRefGoogle Scholar
Lempart, M., Derkowski, A., Luberda-Durnaś, K., Skiba, M., & Błachowski, A. (2018). Dehydrogenation and dehydroxylation as drivers of the thermal decomposition of Fe-chlorites. American Mineralogist, 103(11), 18371850.CrossRefGoogle Scholar
Lempart, M., Derkowski, A., Strączek, T., & Kapusta, C. (2020). Systematics of H2 and H2O evolved from chlorites during oxidative dehydrogenation. American Mineralogist, 105, 932944.CrossRefGoogle Scholar
Levy, J. H., & Hurst, H. J. (1993). Kinetics of dehydroxylation, in nitrogen and water vapour, of kaolinite and smectite from Australian Tertiary oil shales. Fuel, 72(6), 873877.CrossRefGoogle Scholar
Levy, J. H., & White, T. J. (1988). The reaction of pyrite with water vapour. Fuel, 67(10), 13361339.CrossRefGoogle Scholar
MacKenzie, K. J. D., & Berezowski, R. M. (1984). Thermal and Mössbauer studies of iron-containing hydrous silicates. V. Berthierine. Thermochimica Acta, 74(1–3), 291312.CrossRefGoogle Scholar
MacKenzie, K. J. D., & Rogers, D. E. (1977). Thermal and Mössbauer studies of iron-containing hydrous silicates: I. nontronite. Thermochimica Acta, 18(2), 177196.CrossRefGoogle Scholar
Mackenzie, R. C., & Bishui, B. M. (1958). The montmorillonite differential thermal curve. II. Effect of exchangeable cations on the dehydroxylation of normal montmorillonite. Clay Minerals Bulletin, 3(20), 276286.CrossRefGoogle Scholar
Meija, J., Mester, Z., & D'Ulivo, A. (2006). Mass spectrometric separation and quantitation of overlapping isotopologues H2O/HOD/D2O and H2Se/HDSe/D2Se mixtures. Journal of the American Society for Mass Spectrometry, 17(7), 10281036.CrossRefGoogle ScholarPubMed
Meyers, K. S., & Speyer, R. F. (2003). Thermal analysis of clays. In Handbook of Thermal Analysis and Calorimetry (2nd ed.). Elsevier Science BV, pp. 261306Google Scholar
Mikutta, R., Kleber, M., Kaiser, K., & Jahn, R. (2005). Organic matter removal from soils using hydrogen peroxide, sodium hypochlorite, and disodium peroxodisulfate. Soil Science Society of America Journal, 69(1), 120135.CrossRefGoogle Scholar
Molina-Montes, E., Timón, V., Hernández-Laguna, A., & Sainz-Díaz, C. I. (2008). Dehydroxylation mechanisms in Al3+/Fe3+ dioctahedral phyllosilicates by quantum mechanical methods with cluster models. Geochimica et Cosmochimica Acta, 72(16), 39293938.CrossRefGoogle Scholar
Muller, F., Drits, V. A., Plançon, A., & Besson, G. (2000). Dehydroxylation of Fe3+, Mg-rich dioctahedral micas: (I) structural transformation. Clay Minerals, 35(3), 491504.CrossRefGoogle Scholar
Muller, F., Drits, V. A., Plançon, A., & Robert, J. P. (2000). Structural transformation of 2: 1 dioctahedral layer silicates during dehydroxylation-rehydroxylation reactions. Clays and Clay Minerals, 48, 572585.CrossRefGoogle Scholar
Muñoz-Santiburcio, D., Kosa, M., Hernández-Laguna, A., Sainz-Díaz, C. I., & Parrinello, M. (2012). Ab initio molecular dynamics study of the dehydroxylation reaction in a smectite model. The Journal of Physical Chemistry C, 116(22), 1220312211.CrossRefGoogle Scholar
Ngouana, B. F., & Kalinichev, A. G. (2014). Structural arrangements of isomorphic substitutions in smectites: Molecular simulation of the swelling properties, interlayer structure, and dynamics of hydrated Cs-montmorillonite revisited with new clay models. The Journal of Physical Chemistry C, 118(24), 1275812773.CrossRefGoogle Scholar
Ogloza, A. A., & Malhotra, V. M. (1989). Dehydroxylation induced structural transformations in montmorillonite: an isothermal FTIR study. Physics and Chemistry of Minerals, 16(4), 378385.CrossRefGoogle Scholar
Pavlov, M., Siegbahn, P. E., & Sandström, M. (1998). Hydration of beryllium, magnesium, calcium, and zinc ions using density functional theory. The Journal of Physical Chemistry A, 102(1), 219228.CrossRefGoogle Scholar
Pelletier, M., Michot, L. J., Humbert, B., Barrès, O., de la Caillerie, J. B. D., & Robert, J.-L. (2003). Influence of layer charge on the hydroxyl stretching of trioctahedral clay minerals: A vibrational study of synthetic Na- and K-saponites. American Mineralogist, 88, 18011808.CrossRefGoogle Scholar
Pérez-Maqueda, L. A., Blanes, J. M., Pascual, J., & Pérez-Rodríguez, J. L. (2004). The influence of sonication on the thermal behavior of muscovite and biotite. Journal of the European Ceramic Society, 24(9), 27932801.CrossRefGoogle Scholar
Perez-Rodriguez, J. L., Duran, A., Sanchez Jimenez, P. E., Franquelo, M. L., Perejón, A., Pascual-Cosp, J., & Pérez-Maqueda, L. A. (2010). Study of the dehydroxylationrehydroxylation of pyrophyllite. Journal of the American Ceramic Society, 93(8), 23922398.CrossRefGoogle Scholar
Plante, A. F., Fernández, J. M., & Leifeld, J. (2009). Application of thermal analysis techniques in soil science. Geoderma, 153(1–2), 110.CrossRefGoogle Scholar
Poinsignon, C., Yvon, J., & Mercier, R. (1982). Dehydration energy of the exchangeable cations in montmorillonite—A DTA study. Israel Journal of Chemistry, 22(3), 253255.CrossRefGoogle Scholar
Rancourt, D. G., Mercier, P. H. J., Cherniak, D. J., Desgreniers, S., Kodama, H., Robert, J. L., & Murad, E. (2001). Mechanisms and crystal chemistry of oxidation in annite: resolving the hydrogen-loss and vacancy reactions. Clays and Clay Minerals, 49(6), 455491.CrossRefGoogle Scholar
Rouquerol, F., Rouquerol, J., & Llewellyn, P. (2013). Thermal analysis. In: Developments in Clay Science (Vol. 5, pp. 361379). Elsevier.Google Scholar
Salles, F., Douillard, J. M., Denoyel, R., Bildstein, O., Jullien, M., Beurroies, I., & Van Damme, H. (2009). Hydration sequence of swelling clays: Evolutions of specific surface area and hydration energy. Journal of Colloid and Interface Science, 333(2), 510522.CrossRefGoogle ScholarPubMed
Salles, F., Bildstein, O., Douillard, J. M., Jullien, M., Raynal, J., & Van Damme, H. (2010). On the cation dependence of interlamellar and interparticular water and swelling in smectite clays. Langmuir, 26(7), 50285037.CrossRefGoogle ScholarPubMed
Salles, F., Douillard, J. M., Bildstein, O., El Ghazi, S., Prélot, B., Zajac, J., & Van Damme, H. (2015). Diffusion of interlayer cations in swelling clays as a function of water content: Case of montmorillonites saturated with alkali cations. The Journal of Physical Chemistry C, 119(19), 1037010378.CrossRefGoogle Scholar
Sanz, J., Gonzalez-Carreno, T., & Gancedo, R. (1983). On dehydroxylation mechanisms of a biotite in vacuo and in oxygen. Physics and Chemistry of Minerals, 9(1), 1418.CrossRefGoogle Scholar
Schilling, M. R. (1990). Effects of sample size and packing in the thermogravimetric analysis of calcium montmorillonite STx-1. Clays and Clay Minerals, 38(5), 556558.CrossRefGoogle Scholar
Seiffarth, T., & Kaps, C. (2009). Structural characterization of (Cu2+, Na+)- and (Cu2+, NH4+)-exchanged bentonites upon thermal treatment. Clays and Clay Minerals, 57(1), 4045.CrossRefGoogle Scholar
Shannon, R. D. (1976). Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica A, 32(5), 751767.CrossRefGoogle Scholar
Sidey, V. (2016). On the effective ionic radii of ammonium. Acta Crystallographica B., B72, 626633.CrossRefGoogle Scholar
Skoubris, E. N., Chryssikos, G. D., Christidis, G. E., & Gionis, V. (2013). Structural characterization of reduced-charge montmorillonites. Evidence based on FTIR spectroscopy, thermal behavior, and layer-charge systematics. Clays and Clay Minerals, 61(2), 8397.CrossRefGoogle Scholar
Smith, D. W. (1977). Ionic hydration enthalpies. Journal of Chemical Education, 54(9), 540.CrossRefGoogle Scholar
Środoń, J. (2009). Quantification of illite and smectite and their layer charges in sandstones and shales from shallow burial depth. Clay Minerals, 44(4), 421434.CrossRefGoogle Scholar
Środoń, J., & McCarty, D. K. (2008). Surface area and layer charge of smectite from CEC and EGME/H2O-retention measurements. Clays and Clay Minerals, 56(2), 155174.CrossRefGoogle Scholar
Środoń, J., Zeelmaekers, E., & Derkowski, A. (2009). The charge of component layers of illite-smectite in bentonites and the nature of end-member illite. Clays and Clay Minerals, 57(5), 649671.CrossRefGoogle Scholar
Stackhouse, S., Coveney, P. V., & Benoit, D. M. (2004). Density-functional-theory-based study of the dehydroxylation behavior of aluminous dioctahedral 2: 1 layer-type clay minerals. Journal of Physical Chemistry B., 108, 96859694.CrossRefGoogle Scholar
Stevenson, C. M., & Gurnick, M. (2016). Structural collapse in kaolinite, montmorillonite and illite clay and its role in the ceramic rehydroxylation dating of low-fired earthenware. Journal of Archaeological Science, 69, 5463.CrossRefGoogle Scholar
Stucki, J. W. (1988). Structural iron in smectites. In: Stucki, J. W., Goodman, B. A., & Schwertmann, U. (Eds.), Iron in Soils and Clay Minerals (Chap. 17, pp. 625664). NATO ASI Series, vol 217. Dordrecht: Springer. https://doi.org/10.1007/978-94-009-4007-9CrossRefGoogle Scholar
Stucki, J.W., Bish, D., & Mumpton, F.A. (1990). Thermal Analysis in Clay Science. The Clay Minerals Society CMS Workshop Lectures (vol. 3)CrossRefGoogle Scholar
Szymański, W., Drewnik, M., Stolarczyk, M., Musielok, Ł., Gus-Stolarczyk, M., & Skiba, M. (2022). Occurrence and stability of organic intercalation in clay minerals from permafrost-affected soils in the High Arctic – A case study from Spitsbergen (Svalbard). Geoderma, 408, 115591.CrossRefGoogle Scholar
Vieillard, P., Gailhanou, H., Lassin, A., Blanc, P., Bloch, E., Gaboreau, S., Fialips, C. I., & Made, B. (2019). A predictive model of thermodynamic entities of hydration for smectites: Application to the formation properties of smectites. Applied Geochemistry, 110, 104423.CrossRefGoogle Scholar
Vogels, R. J. M. J., Kloprogge, J. T., Geus, J. W., & Beers, A. W. F. (2005). Synthesis and characterization of saponite clays: Part 2. Thermal Stability. American Mineralogist, 90(5–6), 945953.CrossRefGoogle Scholar
Vyazovkin, S. (2015). Isoconversional kinetics of thermally stimulated processes. Springer.CrossRefGoogle Scholar
Vyazovkin, S. (2021). Determining Preexponential Factor in Model-Free Kinetic Methods: How and Why? Molecules, 26(11), 3077.CrossRefGoogle ScholarPubMed
Vyazovkin, S., Burnham, A. K., Criado, J. M., Pérez-Maqueda, L. A., Popescu, C., & Sbirrazzuoli, N. (2011). ICTAC Kinetics Committee recommendations for performing kinetic computations on thermal analysis data. Thermochimica acta, 520(1–2), 119.CrossRefGoogle Scholar
Vyazovkin, S., Burnham, A. K., Favergeon, L., Koga, N., Moukhina, E., Pérez-Maqueda, L. A., & Sbirrazzuoli, N. (2020). ICTAC Kinetics Committee recommendations for analysis of multi-step kinetics. Thermochimica Acta, 689, 178597.CrossRefGoogle Scholar
Webb, S. W., Stanley, D. A., & Scheiner, B. J. (1986). Thermal analysis of ion-exchanged montmorillonite. Particulate Science and Technology, 4(2), 131–42.CrossRefGoogle Scholar
Wicks, F. J. & Ramik, R. A. (1990). Vacuum Thermogravimetric Analysis and Evolved Gas Analysis by Mass Spectrometry. In Stucki, J. W., Bish, D., & Mumpton, F. A. (Eds.), CMS Workshop Lectures, Vol. 3. Thermal Analysis in Clay Science. The Clay Minerals Society. https://doi.org/10.1346/CMS-WLS-3.5Google Scholar
Wilson, M. A., Carter, M. A., Hall, C., Hoff, W. D., Ince, C., Savage, S. D., McKay, B., & Betts, I. M. (2009). Dating fired-clay ceramics using long-term power law rehydroxylation kinetics. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 465(2108), 24072415.CrossRefGoogle Scholar
Wolters, F., & Emmerich, K. (2007). Thermal reactions of smectites—Relation of dehydroxylation temperature to octahedral structure. Thermochimica Acta, 462(1–2), 8088.CrossRefGoogle Scholar
Xi, Y., Martens, W., He, H., & Frost, R. L. (2005). Thermogravimetric analysis of organoclays intercalated with the surfactant octadecyltrimethylammonium bromide. Journal of Thermal Analysis and Calorimetry, 81(1), 9197.CrossRefGoogle Scholar
Yariv, S., Borisover, M., & Lapides, I. (2011). Few introducing comments on the thermal analysis of organoclays. Journal of Thermal Analysis and Calorimetry, 105(3), 897906.CrossRefGoogle Scholar
Zabat, M., & Van Damme, H. (2000). Evaluation of the energy barrier for dehydration of homoionic (Li, Na, Cs, Mg, Ca, Ba, Alx(OH)y z+ and La)-montmorillonite by a differentiation method. Clay Minerals, 35(2), 357363.CrossRefGoogle Scholar
Figure 0

Fig. 1 An example of TGA and DTG curves of samples representing almost-pure dioctahedral (SCa-3 montmorillonite) and trioctahedral (high-charge synthetic saponite; Pelletier et al., 2003) smectites. The dehydration reaction at <200°C is followed by dehydroxylation above ~300°C, which are shown in the corresponding MS signal of H2O (m/z = 18 and 17). Thermal removal of carboxyl groups from organic impurities and the decomposition of carbonate impurities are demonstrated by the CO2 (m/z = 44) signal below 500°C and above 600°C, respectively. The analysis was performed under dry N2 at a flow of 100 mL/min, with a constant heating rate of 20°C/min. Ion currents not to scale. Note: All natural smectite samples shown here and in the following figures come from the Source Clays Repository of The Clay Minerals Society. Unless specified otherwise, all thermal analyses presented in this paper were performed using a TA Discovery TGA (USA) instrument and ThermoStar GDS 320 MS from Pfeiffer (Göttingen, Germany); ~20 mg of loosely packed powder, using a wide 50 µL Platinum HT sample pan without a lid

Figure 1

Table 1 The hydration enthalpy (–ΔHo) after Smith (1977) and effective ionic radii for a given coordination number of inorganic cations commonly occurring in smectite interlayers in natural conditions and used in laboratory experiments, after Shannon (1976)

Figure 2

Fig. 2 DTG curves in the temperature range corresponding to the dehydration of Na+-, Ca2+-, and La3+-exchanged montmorillonites. The analysis was performed under dry N2 at a flow of 50 mL/min, with a constant heating rate of 5°C/min

Figure 3

Fig. 3 DTG curves showing the difference in dehydroxylation of trans-vacant (tv) and cis-vacant (cv) dioctahedral smectites along their dominating octahedral cation-OH composition (after Derkowski et al., 2012a). Analysis preformed under dry N2 at 50 mL/min flow, with a constant heating rate of 5°C/min

Figure 4

Fig. 4 DTG curves of smectite reference samples saturated with various inorganic cations. The analysis was performed under a 100 mL/min flow of dry N2, with 4 h isothermal heating at the temperatures indicated on the figures, followed by heating to 1000°C at a constant rate of 5°C/min

Figure 5

Fig. 5 The effect of cation migration into the octahedral sheet. DTG curves of Li+-exchanged and Mg2+-exchanged montmorillonites preheated isothermally under a flow of 100 mL/min of dry N2 for 4 h at various temperatures (as stated in the legend) followed by heating to 1000°C at a constant rate of 5°C/min. The DTG curves of the SCa-2 sample saturated with other cations and preheated isothermally at 300°C are given as reference

Figure 6

Fig. 6 Rehydration of a smectite reference sample (SWy-3), saturated with cations of contrasting hydration enthalpy, at various temperatures, after prior in situ drying for 4 h at 200°C. Mass gain is shown as fractional mass gain at a constant H2O content of 9 g/kg air (not constant RH), normalized to the sample mass after drying at 200°C

Figure 7

Fig. 7 An example of an in-situ experiment of an isothermal rehydroxylation reaction observed with TGA (after Derkowski & Kuligiewicz, 2017) on a Cs+-exchanged beidellite (SBId-1) and montmorillonite (SWy-2). Note the initial dehydroxylation to 800°C under dry gas and switching to H2O-rich gas for isothermal rehydroxylation, then secondary dehydroxylation to 1000°C under dry gas. The mass gain under wet gas corresponds to the rebuilding of OH groups

Figure 8

Fig. 8 The texture (upper) and DTG pattern (lower) of unsifted and sifted Ca2+-exchanged montmorillonite sample (SCa-3) after heating to >1000°C at a constant rate of 20°C/min (DTG pattern) and 500°C/min under a flow of 50 mL/min dry N2 gas

Figure 9

Fig. 9 TGA patterns of Li+- and Ca2+-exchanged montmorillonite (SWy-2) heated isothermally for 4 h at temperatures as indicated in the legend followed by ramp-heating at 10°C/min under a flow of 50 mL/min of dry N2. Note the abrupt mass loss at almost the same temperature in Li+-exchanged samples, originating from the sample dropping off the holder

Figure 10

Fig. 10 The examples of linear, ramp-heating of a Na+-beidellite sample to 1000°C under a 50 mL/min flow of dry N2 and at various heating rates. The rates are indicated at the DTG peak in °C/min. The shift between the DTGmax peaks can be used to determine the activation energy of dehydroxylation

Figure 11

Fig. 11 A, C–F Activation energy (Ea) of the dehydroxylation of various dioctahedral smectites, calculated using the 0.01 α-increments and the Friedman method, and B an example of Ea determination using the Kissinger method and the DTG peak temperature (Fig. 10). Note the similarity of the Ea values and trends in the dehydroxylation of two Ca2+-exchanged Mg-rich montmorillonites (E, F). For each TGA run, the total mass loss upon dehydroxylation was divided into 100 portions of equal mass difference (0.01α). For each 0.01α, the ln(dα/dt) was plotted against the corresponding 1/T using four or five data points from the TGA curves recorded with a different heating rate (B). E Ea is then calculated from the slope of ln(dα/dt) vs 1/T plot (Eq. 8). The correlation coefficient, R2, was computed for each of 0.01α. All Ea values shown (A, C–F) are reported for R2 > 0.997 (arbitrary cut-off),

Figure 12

Table 2 Comparison of the Ea of dehydration of weakly bound water (WBW) and strongly bound water (SBW) from studies employing various methods

Figure 13

Fig. 12 The Ea of dehydroxylation of smectite reference samples presenting values averaged over α = 0.2–0.8 calculated using the Friedman method (upper, cf. Figure 11) and Ea calculated using the conventional Kissinger method based on DTG peak temperature (lower, cf. Fig. 10), plotted against the maximum of dehydroxylation (DTGmax; Table 3). The Ea error in the Kissinger method is ~5%. Filled shapes denote the Cs+-exchanged form; note their lower position with respect to other cations

Figure 14

Table 3 Ea of dehydroxylation calculated as using the 0.01 α-increments with values averaged over α = 0.2–0.6 calculated using the Friedman method (Fig. 11) and Ea calculated using the conventional Kissinger method based on DTG peak temperature (Fig. 10). The maximum of dehydroxylation (DTGmax) was determined using the 5°C/min constant rate TG measurement

Figure 15

Fig. 13 An example of the calculation of Ea of smectite dehydroxylation for Ca2+-saturated SCa-3 montmorillonite based on a heating–cooling cycles TGA experiment (Drits et al., 2012a). The maximum temperature (in °C) of each consecutive heating cycle is given in italics. Compare Ea values with Fig. 11e