Next Article in Journal
Constraints on the Genesis of the Shuangwang Gold Deposit in Qinling Orogen, Central China: Evidence from In Situ Trace Element and Sulfur Isotope
Previous Article in Journal
Genesis of the Wuzhutang Granite and Associated W–Sn–Be Mineralization in the Xuebaoding Mining Area, Sichuan Province, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Crystallographic Distortion Mechanism during the Annealing of Kaolinite

1
State Key Laboratory of Mining Response and Disaster Prevention and Control in Deep Coal Mines, Anhui University of Science and Technology, Huainan 232001, China
2
School of Material Science and Engineering, Anhui University of Science and Technology, Huainan 232000, China
*
Authors to whom correspondence should be addressed.
Minerals 2022, 12(8), 994; https://doi.org/10.3390/min12080994
Submission received: 6 July 2022 / Revised: 28 July 2022 / Accepted: 2 August 2022 / Published: 5 August 2022

Abstract

:
The distortion process of kaolinite annealed from 25 °C to 550 °C for different holding times can be characterized using a thermogravimetric/differential scanning calorimeter (TG/DSC) for thermal analysis, X-ray diffraction (XRD) for establishing the crystal structure, the Fourier transform infrared spectrum (FTIR) for identifying the functional groups, and a scanning electron microscope (SEM) for establishing the microstructure. Dehydroxylation is the main reaction during annealing from 25 °C to 550 °C and leads to kaolinite crystal distortion. A stable crystal structure during distortion was obtained by optimizing the bulk phase with quantum chemistry. Then, the crystal structure was studied by using ab initio multiple scattering calculations for X-ray absorption of the fine structure (XAFS). The results of X-ray absorption near the edge structures (XANES) determined that peak shifts and intensity phases slightly increased. The crystal structure distortion of kaolinite during annealing can be explained by the experimental and simulation results. This work provides theoretical support for identifying kaolinite with different degrees of distortion and has the potential for further developments in coal gangue separation.

1. Introduction

With the rapid development of artificial intelligence, compared with traditional coal gangue separation technology, including manual separation, gravity coal preparation, flotation, and other methods, intelligent photoelectric separation has the advantages of nondestructive testing and intelligent remote control; it creates no sewage, and can be adapted to an underground environment, and therefore has been widely researched [1,2]. To remove gangue from raw coal is conducive to improving the efficiency of coal utilization. There are numerous different separation methods, such as flotation, gravity separation, etc.; however, for downhole sorting, photoelectric separation has previously not been possible. In recent years, intelligent photoelectric separation has been widely developed and applied, based on X-ray detection technology and image detection technology. Recognition depends on the different attenuations of X-rays through the coal and gangue, but a deep recognition principle has not yet been found [3,4,5]. A variety of clay minerals are widely mixed in coal seams and are important in the composition of gangue. Among them, kaolinite is used in a large variety of technological applications; it is necessary to separate it from the coal gangue mixture to improve the purity of the coal and facilitate subsequent processing and application. Kaolinite is a form of water-containing aluminosilicate. Its theoretical structure is Al2Si2O5(OH)4; octahedral aluminosilicate with a structure of 1:1 is formed by connecting the octahedral aluminosilicate with the tetrahedral aluminosilicate and stacking along the c axis [6]. In this study, kaolinite is the main object of study.
The existing coal gangue identification equipment is based on the use of the level of gray value to represent the percentage of X-ray attenuation after passing through coal gangue [7,8]. Coal and gangue are distinguished by the difference in gray value. However, it is difficult to distinguish gangue when the gray value difference is not obvious. In order to improve the identification accuracy of coal gangue using intelligent photoelectric separation, based on X-rays for kaolinite and coal, the physical and chemical characteristics of coal and kaolinite have attracted much attention due to their attenuation contribution to X-rays.
The structure of kaolinite has been studied since the 1930s [9]; experimental, theoretical, and combined strategies have been developed to analyze the structure of kaolinite, and several crystal models of kaolinite have been proposed, inspired by changes in the crystal structure of kaolinite caused by heating [10,11,12,13,14,15]. When kaolinite is annealed at 550 °C, as the heating time increases, the hydroxyl of kaolinite is gradually removed and the crystal is transformed into amorphous kaolin [16,17]. In this work, annealing is used to change the crystal structure of kaolinite to obtain kaolinites with different degrees of distortion.
XAFS (X-ray absorption fine-structure spectroscopy) techniques, such as X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS), is a widely-used material characterization technique that is able to obtain information regarding the oxidation state of the atoms, the average coordination number, interatomic distances, and degrees of disorder from the atomic environment to the absorbing atom [18,19,20].
In this study, kaolinite with different crystallinities was obtained for experimentation by using the characteristics of the dehydroxylation of kaolinite at 550 °C and the complete disappearance of the crystal structure [17], identified by thermogravimetric analysis, scanning electron microscopy, Fourier-transform infrared spectrometry, and X-ray diffraction [21,22]. At the same time, material studio software was used to simulate the theoretical kaolinite crystal and the partially deformed kaolinite [23]. The kaolinite crystal structure data was based on mp-541152 Al2Si2H4O9 reported by the Materials Project [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40], and the theoretical XRD was compared with the actual implementation results. The EXAFS signal changes of kaolinite crystal and partially deformed kaolinite crystal were simulated by multiple scattering computation, based on ab initio and ATHENA software [41,42,43]; multiple scattering computation based on ab initio is an implementation of real-space multiple scattering theory (RSMS) and is capable of calculating a wide range of X-ray and electron spectra, while ATHENA is a program capable of supporting common XAS data-processing chores. These results will help to provide new options for separating coal (which is amorphous) and gangue (which is mostly crystalline) from coal gangue.

2. Materials and Methods

2.1. Materials

In this work, the standard kaolin clay sample Raw-Ka was purchased from Aladdin (Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China) of pharmaceutical grade. The sample has high purity, fine particle size, and high kaolinite content (>95%). The sample was dried at 90 °C for 24 h before the experiment. Anhydrous alcohol (at analytically pure grade) was purchased from Aladdin, Shanghai, China. The chemical composition and particle size analysis (PSD) results are listed in Table A1 and Figure A1, respectively.

2.2. Preparation

The evolution of the kaolinite crystal structure distortion was achieved through thermal treatment. Firstly, in order to find the appropriate temperature for studying the distortion of kaolinite under heat treatment, Raw-Ka was heat-treated in a laboratory in a programmable muffle furnace (tube furnace) in the air for 2 h at different temperatures (400, 500, 550, 600, 700, and 800 °C). Raw-Ka was also heated at the same rate from 25 °C to 550 °C, with different holding times (10, 20, 30, 60, 90, and 120 min, respectively), labeled as Ka-10 min, Ka-20 min, Ka-30 min, Ka-60 min, Ka-90 min, and Ka-120 min, respectively.

2.3. Characterization

The composition of Raw-Ka was investigated using an X-ray fluorescence spectrometer (XRF, Thermo Fisher, Waltham, MA, USA). The PSD was carried out with a SALD-7101 ultraviolet laser nanoparticle-size analyzer (Shimadzu, Kyoto, Japan). The TG/DSC of Raw-Ka was performed using an SDTA851e thermogravimetric (Maitler Toledo, Greifensee, Switzerland), in a temperature range from 25 to 800 °C at 10 °C/min, in the air. The powder XRD data were recorded using a Smartlab SE (Rigaku, Tokyo, Japan, using CuKα radiation). The FTIR absorption spectra were recorded on a NICOLET 380 (Thermo Fisher, Waltham, MA, USA) from 400 to 400 cm−1, using 32 scans with a resolution of 4 cm−1 on a KBr sampling sheet. The SEMs were recorded with a Zeiss Sigma300 (Zeiss, Oberkochen, Germany).

2.4. Simulation of Crystal Structure Distortion

Firstly, an ideal kaolinite crystal structure was constructed, a representative kaolinite crystal structure was formed according to the Materials Project mp-541152 Al2Si2H4O9 (denoted as Sim-Ka). The Materials Project mp-541152 corresponds to a crystal structure file of kaolinite [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40]. The kaolinite lattice constant values are a = 5.213 Å, b = 7.479 Å, c = 9.052 Å, α = 91.787°, β = 89.728°, γ = 104.975°, in the space group P1. With the disordering of the kaolinite, in the process of annealing, hydroxyl groups were removed from the alumina tetrahedron in order of O33H15→O32H14→O31H13→O27H9. After constructing the kaolinite cell model, geometric optimization was carried out. All calculations were based on DFT and were carried out using the CASTEP module [44] in the Material Studio software. Generalized gradient approximation (GGA) with a Perdew–Burke–Ernzerhof (PBE) scheme and a 3 × 2 × 2 Monkhorst-Pack K-point grid was used to achieve geometric optimization [45]. The geometric optimization was set according to the following convergence accuracy: The plane-wave energy cut-off that was used in the calculations was 450.0 eV. The total self-consistent field (SCF) energy change was 1 × 10−6 eV/atom, while the energy between optimization steps was 1 × 10−5 eV/atom. The atomic orbitals involved in the calculation were H1S1, O 2s22p4, Al 3s23p1, and Si 3s23p2, respectively.
Notably, each hydroxyl removal was accompanied by applying geometric optimization. The CSTEP outputs were denoted as Phase 1, Phase 2, Phase 3, and Phase 4.

2.5. XAS Calculation

The absorption coefficient μ, which gives the probability that the X-ray will be absorbed according to the Beer–Lambert Law [46], is mainly involved when discussing X-ray absorption:
I = I 0 e μ t
where I0 is the X-ray intensity incident on a sample, t is the sample thickness, and I is the intensity transmitted through the sample.
At most X-ray energies, the absorption coefficient μ is a function of energy, sample density ρ, the atomic number Z, atomic mass A, and the X-ray energy, E, which is roughly that reported in [47].
μ ρ Z 4 A E 3
When the incident X-ray has an energy equal to that of the binding energy of a core-level electron, there is a sharp rise in absorption, with an absorption edge corresponding to the core level to the continuum. The energy dependence of the X-ray absorption coefficient μ(E) in transmission mode is that reported in [48].
μ E = ln I I 0
The EXAFS fine-structure function χ(E) as a change to μ(E) and μ0(E), as reported in [48]:
χ E = μ E μ 0 E Δ μ 0 E
where μ(E) is the measured absorption coefficient and μ0(E) is a smooth background function, representing the absorption of an isolated atom.
EXAFS is best understood in terms of the wave behavior of the photoelectron that is created in the absorption process. It is common to convert the X-ray energy to k [24], the wave number of the photoelectron, which is defined as:
k = 2 m ( E E 0 ) ħ 2
where E0 is the absorption edge energy, m is the electron mass, and ħ is Planck’s constant. The theoretical expression of EXAFS [49] can be obtained by introducing k to the equation:
x k = j N j f j k e 2 k 2 σ j 2 k R j 2 sin 2 k R j + δ j k
Under the conditions of known scattering amplitude f k and phase displacement δ k via numerical analysis, the number of neighboring atoms is N, the distance to the neighboring atom is R, and the disorder in the neighboring distance is σ 2 .
The theoretical structure and simulation analysis of the intermediate state of dehydroxylation and the structure change of kaolinite are carried out using Materials Studio and ATHENA. The EXAFS change caused by the increase of kaolinite disorder has been studied elsewhere [41,42,43].
The CASTEP module in the Materials Studio software was used to optimize the structure. The kaolinite model of mp-541152 was used to optimize the crystal model by using the PBE exchange-correlation function in the generalized gradient approximation (GGA) [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,42].
The XAFS at the Si4 sites was calculated by multiple scattering computations based on ab initio, with Si as the central absorption atoms. The coordination number, coordination distance, and species information of the absorption atoms were given. The first coordination shells of Si4 at the K edge were calculated and analyzed.
The self-consistent field (SCF) and the full multiple scattering (fms) were calculated for the unit cell. A Hedin–Lundqvist-type exchange-correlation potential was used.

3. Results and Discussion

3.1. TG/DSC Measurements

The distortion of kaolinite is highly correlated with heat treatment. The results obtained by the thermogravimetric/differential scanning calorimeter (TG/DSC) for Raw-Ka was presented in Figure 1. The sample has a strong internal thermal effect in the range of 450 °C to 600 °C. The endothermic peak at 500 °C corresponds to the dehydroxylation of kaolinite, which is consistent with the descriptions found in previous studies. In addition, 17.53% of weight loss occurred throughout the heating process, which corresponds to the loss of the hydroxyl weight of kaolinite during heat treatment [16].

3.2. FTIR

Dehydroxylation is the main change of kaolinite during annealing; it plays a pivotal role in the distortion of kaolinite and the corresponding structural characteristics of the samples. The FTIR absorption spectra of Raw-Ka, Ka-10 min, Ka-30 min, and Ka-120 min is shown in Figure 2a.
Details of FTIR spectra is shown if Figure 2b. The prominent peaks of the stretching vibration phase of the Al-OH groups appeared at 3695, 3654, and 3620 cm−1 in the FTIR spectra, respectively, corresponding to the inner surface hydroxyls, the out-of-plane stretching vibrations, and the inner hydroxyl between the tetrahedral and octahedral. Several bands within 1100 cm−1 to 1000 cm−1 corresponded to Si-O in the in-plane stretching modes. For inner Al-Al-OH bonds, a peak was observed at 916 cm−1. The weak bends at 792, 755, and 698 cm−1 were assigned for Si-O bending. In the lower-frequency region, the Si-O-Al, Si-O-Si, and Si-O bending peaks appeared at 538, 470, and 428 cm−1, respectively [50,51]. With the increase in heat-treatment time, the reflection intensity has obviously been reduced in Ka-10 min, Ka-20 min, and Ka-120 min. In terms of annealing holding for 120 min, almost all the peaks of Al-OH groups were absent, while the FTIR spectrum degraded into a relatively flat line, which consisted of three smooth peaks at 1064, 809, and 462 cm−1. The peak at 1064 cm−1 was for tetrahedral sheet vibration; the peak at 809 cm−1 was for free silica or quartz; a new peak appeared at 451 cm−1 for structural distortion [11,22,52]. The results show the dehydroxylation phenomenon, which leads to the distortion of the kaolinite.

3.3. Microstructural Characterization

Figure 3 shows images of the original kaolinite (Raw-Ka) and annealed kaolinite samples taken with a Zeiss Sigma300 (Zeiss, Oberkochen, Germany). Pictures show several random particles of the four samples, scanned by a scanning electron microscope, with the typical agglomeration features of clay aggregates and clear platelet morphology. Aggregates of hexagonal spoon kaolinite slabs can be clearly observed, dominated by face-to-face and edge-to-edge contacts.
For Raw-Ka, Ka-10 min, the nano-scale internal structure of the deposition layer can be identified. The Ka microphotograph (Figure 3a) shows that the samples have an obvious layered structure and that these layers are stacked approximately directly. The Ka-10 min photograph (Figure 3b) shows that the kaolinite at 550 °C for 10 min still has an obviously layered structure, but it also shows that this reflects a certain degree of deformation. As the holding time increased from 30 min to 120 min, the layered structure of kaolinite from Figure 3c,d fractured and deformed due to dehydroxylation and crystal structure collapse, which was consistent with the results obtained by XRD.

3.4. Crystal Structure

The crystal structures of Mp-541152 Al2Si2H4O9 [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40] as Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4 are shown in Figure 4. Among them, Phase 1, Phase 2, Phase 3, and Phase 4 are stable structures that may exist in the annealing distortion process of geometrically optimized kaolinite and are shown in Figure 4b–e. With the dehydroxylation and distortion of kaolinite, the Al-O octahedral structure degrades into a relatively unstable Al-O tetrahedral structure. At the same time, the distortion of the Al-O groups leads to a series of movement and rotation transformations of Si-O groups, but the Si-O tetrahedron structure remains intact, on the whole. With the transformation of the Al-O octahedron into an Al-O tetrahedron, the structure of the whole aged rock tends to be unstable. With the continuous annealing process, kaolinite eventually loses its crystal structure and becomes amorphous metakaolin. In this study, based on the results of TG/DSC and FTIR XRD patterns, combined with the results of other studies, we used software to simulate the possible change process of crystal structure in the distortion process of kaolinite [11,17]. The distortion of kaolinite in the process of Sim-Ka → Phase 4 change was mainly studied.

3.5. Powder XRD

The distortion of kaolinite due to the elimination of structure hydroxyl groups at 550–800 °C [17] corresponds with the obvious changes in the XRD reflections between Ka-550 °C and Ka-800 °C (Figure A2). The results of Figure A2. show that 550 °C is suitable for preparing kaolinite with different degrees of distortion.
To study the relationship between the structural distortion of kaolinite and X-ray absorption, kaolinite was annealed at 550 °C for different holding times. Figure 5 shows the XRD patterns of the sample at Raw-Ka, Ka-10 min, Ka-20 min, Ka-30 min, Ka-60 min, and Ka-120 min. With the annealing process, the diffraction peak of the sample gradually weakened and finally became flat, indicating the collapsing behavior of the crystal texture, especially of the (001) reflection at 2θ = 12° and (002) reflection at 2θ = 25°. These peaks are gradually decreasing with the increase in time, which corresponds to the poorly ordered kaolinite. The presence of poorly ordered kaolinite implies that annealing this kaolin will transform it into highly reactive metakaolin. Combined with previous studies, it can be concluded that the disappearance of the diffraction peak of kaolinite corresponds to the dehydroxylation of kaolinite into amorphous kaolin during heat treatment [53].
After a heating duration of 120 min, the peak of (001) diffraction cannot be distinguished from the background intensity, which indicates that the crystal structure of kaolinite is completely distorted and that most of the kaolinite in the sample is transformed into amorphous metakaolin [53]. It was proved that with the increase in holding time, the kaolinite crystal degenerated and finally lost its crystal structure, evolving into amorphous metakaolin.
Figure 6a shows the XRD simulation patterns of Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4. Material Studio was used to simulate the XRD pattern and the XRD pattern with the removal of partial hydroxyl and the changes of atomic position that denoted Phase 1, Phase 2, Phase 3, and Phase 4. With the removal of the partial hydroxyl and the change of Si atom position, it can be observed that the diffraction peak changes significantly. The θm moves from 12.25° to 11.79°, and the peak intensity in the range of 2θ = 20–30° is no longer consistent with the intensity shown in PDF 78–2110, while certain peaks with a smaller intensity have appeared. The details are shown in Figure 6b. The spectrum shows an amorphous structure with crystalline peaks of kaolinite, suggesting incomplete dehydroxylation compared with the spectrum shown in Figure 5 and in previous studies [11,17].

3.6. XAFS Calculation

To compare the different structures of kaolinite, the spectrums of the simulated kaolinite structure are calculated by multiple scattering computations, based on ab initio [19,54,55,56]. The multiple scattering calculation is based on the simulated samples including Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4, which correspond to the incomplete dehydroxylation of kaolinite.
Figure 7 presents the simulated XAFS spectroscopy for the samples at different reaction stages including Si K-edge XANES absorption, the EXAFS functions k2 χ(k), and the radial structural function (RSF) of samples with the calculated theoretical multiple scattering spectra.
The peak features shown in Figure 7a correspond to the simulated XANES spectra at the Si K-edge of Sim-Ka and samples at different reaction stages. The calculated main characteristic pick of Si of the Sim-Ka K-edge (absorption peak A) is E = 1847.5 eV, due to the permitted transition of Si 1 s electrons to an antibonding 3p-like state, and is related to tetrahedral coordination, which is one of the apparent characteristics of SiO2 and is consistent with the report [12,20].
For the normalized X-ray absorption spectra of the Si K-edge, the position of the characteristic peak of Si changed from 1846.7 eV to 1845.1 eV.
In addition, in all cases, the post-edge resonances that are present are similar to those reported in other articles for kaolinite. The resonances at 1860.7 eV (absorption peak B) are the result of multiple scattering processes within the (SiO4)-tetrahedron. Slight differences are observed between Sim-Ka and Phase 1, Phase 2, and Phase 3, while mutation occurred during Phase 1 to Phase 2, and during Phase 3 to Phase 4, which corresponds to the distortion of the kaolinite crystal structure, while the structural change of (SiO4)-tetrahedron resulted in the change in X-ray absorption.
The X-ray absorption spectrum was converted into wave vector space using the formula k = 2 π 2 m E E 0 , where E is the incident X-ray energy, E0 is the energy of the absorption edge, m is the electron mass, and h is the Planck constant. In order to enhance the signal strength, the data is weighted by k2. Figure 7b shows the al K-varying k2 χ(k) of the sample. It can be found that with the distortion of the kaolinite crystal structure, the curve deviates from its original position, indicating the change in the structure at the selected atomic site.
Figure 7c presents the different Fourier transforms of the EXAFS oscillations for the studied samples, measured at the Si K-edge, with the k-range of the Hanning window from 3 to around 15. The simulated spectrum possesses most of the characteristics found in the experimental measurements, which shows that the calculated structure is consistent with the X-ray absorption technique. The peaks in the graph correspond to the Si/O coordination shell at the first nearest-neighbor position, around the Si atoms. For the radial structural function of Si (IV), Figure 7c shows that the atomic distance between Si-O is maintained at 1.32 Å (without phase correction), but with the distortion of the crystal structure, the coordination number of Si increases significantly, which may be due to the difference in the multiple scattering calculation results caused by the position change of the silicon atom in the center of the (SiO4)-tetrahedron structure and the distance change between the silicon atom and the surrounding oxygen atoms. Corresponding to Figure 8, the height of the curve at 1.32 Å has a sudden change, which may correspond to the obvious change in the position of the Si atom. However, the change in the length of the Si-O bond is small, while the influence on the position of the peak is not reflected.

4. Conclusions

A thermal treatment was utilized for regulating the distortion of kaolin to obtain kaolinite with different degrees of distortion.
The process when the kaolinite crystal structure disappears and transforms into metakaolin during annealing was studied and characterized. The Materials Studio software was used to simulate the dehydroxylation process of two groups of kaolinite during annealing and the model was optimized. The XAFS results of the model were calculated by multiple scattering. The XAFS result for Si shows that the structural changes in the dehydroxylation process of kaolinite will lead to significant changes in the X-ray absorption coefficients of Si, especially the X-ray absorption spectra of Si, which have a certain regularity. With the progression of dehydroxylation, the characteristic peak position of the Si atom and (SiO4)-tetrahedron shifts to a low-energy position, and the intensity increases abruptly. The results were Fourier-transformed into R-space, and it was observed that with the dehydroxylation, the k-space data of Si obviously deviated from the initial value. In particular, the phenomena in R-space corresponded to changes in the spatial structure around silicon atoms, which is caused by the lack of hydroxyl. The distortion of (SiO4)-tetrahedron to a minimum-energy structure resulted in changes in the position of the Si atom and in the length of the Si-O bond.
This work is devoted to exploring the structural characteristics and the identification mechanism of kaolinite with different distortions, which may be helpful to identify kaolinite in coal gangue. The crystallographic structure of kaolinite is different from that of amorphous coal in its natural state, and this may make it easier for scientists to identify whether there is a crystal structure. However, it is necessary to study the low crystallinity and excessive impurities of kaolinite further in actual coal gangue.

Author Contributions

Conceptualization, Q.Z., J.X. and J.Z.; methodology, Q.Z., J.X. and W.Z. (Wei Zhou); software, Q.Z. and J.X.; validation, Q.Z. and J.Z.; formal analysis, L.L. and J.Y.; investigation, Q.Z.; resources, J.X., W.Z. (Wei Zhou) and J.Z.; data curation, W.Z. (Wenliang Zhu); writing—original draft preparation, Q.Z.; writing—review and editing, Q.Z. and J.X.; visualization, Q.Z. and J.X.; supervision, J.X. and J.Z.; project administration, J.Z.; funding acquisition, Q.Z., J.X., W.Z. (Wei Zhou) and J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Complete Set of Solid Waste Large-Scale Utilization Technology and Integrated Demonstration in Large Coal Electrochemistry Base (2019YFC1904304), the Anhui Provincial Natural Science Foundation Project (2108085ME160), the Deep Coal Mining Response and Disaster Prevention and Control State Key Laboratory Open Fund (SKLMRDPC19KF11), and the Anhui University of Science and Technology Graduate Innovation Fund Project in 2021 (2021CX1010).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used to support the findings of the study are available from the first author upon request.

Acknowledgments

The authors would also like to express their sincere appreciation to the reviewers for their constructive comments on the revision and improvement of the manuscript. All the support is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Particle size analysis of the Raw-Ka sample.
Figure A1. Particle size analysis of the Raw-Ka sample.
Minerals 12 00994 g0a1
Figure A2. XRD patterns of Raw-Ka, Ka-400 °C, Ka-500 °C, Ka-550 °C, Ka-600 °C, Ka-700 °C and Ka-800 °C.
Figure A2. XRD patterns of Raw-Ka, Ka-400 °C, Ka-500 °C, Ka-550 °C, Ka-600 °C, Ka-700 °C and Ka-800 °C.
Minerals 12 00994 g0a2
Table A1. XRF analysis of the Raw-Ka sample.
Table A1. XRF analysis of the Raw-Ka sample.
OxideContent (%)Standard Error
SiO254.820.25
Al2O343.160.25
TiO20.6540.033
Fe2O30.480.024
Na2O0.2250.011
MgO0.1640.008
P2O50.1620.008
K2O0.1530.008
SO30.05660.0028
ZnO0.04150.0021
CaO0.03050.0015
V2O50.01470.0008
Cl0.01380.0009
Cr2O30.00580.0005
CeO20.00560.002
Nd2O30.0030.0011
Re2O70.00250.0012
Ga2O30.00220.0003
MnO0.00190.0004
Sc2O30.00170.0004
ZrO20.00150.0003

References

  1. Neubert, K.; Wotruba, H. Investigations on the Detectability of Rare-Earth Minerals Using Dual-Energy X-ray Transmission Sorting. J. Sustain. Metall. 2017, 3, 3–12. [Google Scholar] [CrossRef]
  2. Osipov, S.; Libin, E.; Chakhlov, S.; Osipov, O.; Shtein, A. Parameter Identification Method for Dual-Energy X-ray Imaging. NDTE Int. 2015, 76, 38–42. [Google Scholar] [CrossRef]
  3. Pan, Y.; Bi, Y.; Zhang, C.; Yu, C.; Li, Z.; Chen, X. Feeding Material Identification for a Crusher Based on Deep Learning for Status Monitoring and Fault Diagnosis. Minerals 2022, 12, 380. [Google Scholar] [CrossRef]
  4. Jin, J.; Yuan, S.; Lv, Z.; Sun, Q. Development of Backfill Concrete Including Coal Gangue and Metakaolin and Prediction of Compressive Strength by Extreme Learning Machine. Minerals 2022, 12, 330. [Google Scholar] [CrossRef]
  5. Sun, A.; Jia, W.; Hei, D.; Yang, Y.; Cheng, C.; Li, J.; Wang, Z.; Tang, Y. Application of Concave Point Matching Algorithm in Segmenting Overlapping Coal Particles in X-ray Images. Miner. Eng. 2021, 171, 107096. [Google Scholar] [CrossRef]
  6. Brigatti, M.F.; Galán, E.; Theng, B.K.G. Chapter 2—Structure and mineralogy of clay minerals. In Developments in Clay Science; Bergaya, F., Lagaly, G., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; Volume 5, pp. 21–81. ISBN 1572-4352. [Google Scholar]
  7. Dou, D.; Wu, W.; Yang, J.; Zhang, Y. Classification of Coal and Gangue under Multiple Surface Conditions via Machine Vision and Relief-SVM. Powder Technol. 2019, 356, 1024–1028. [Google Scholar] [CrossRef]
  8. Dou, D.; Zhou, D.; Yang, J.; Zhang, Y. Coal and Gangue Recognition under Four Operating Conditions by Using Image Analysis and Relief-SVM. Int. J. Coal Prep. Util. 2020, 40, 473–482. [Google Scholar] [CrossRef]
  9. Pauling, L. The Structure of the Chlorites. Proc. Natl. Acad. Sci. USA 1930, 16, 578–582. [Google Scholar] [CrossRef] [Green Version]
  10. Chen, J.; Min, F.; Liu, L.; Cai, C. Systematic Exploration of the Interactions between Fe-Doped Kaolinite and Coal Based on DFT Calculations. Fuel 2020, 266, 117082. [Google Scholar] [CrossRef]
  11. Li, J.; Zuo, X.; Zhao, X.; Ouyang, J.; Yang, H. Insight into the Effect of Crystallographic Structure on Thermal Conductivity of Kaolinite Nanoclay. Appl. Clay Sci. 2019, 173, 12–18. [Google Scholar] [CrossRef]
  12. Richard, D.; Martínez, J.M.; Mizrahi, M.; Andrini, L.; Rendtorff, N.M. Assessment of Structural Order Indices in Kaolinites: A Multi-Technique Study Including EXAFS. J. Electron. Spectrosc. Relat. Phenom. 2022, 254, 147128. [Google Scholar] [CrossRef]
  13. Richard, D.; Rendtorff, N.M. First Principles Study of Structural Properties and Electric Field Gradients in Kaolinite. Appl. Clay Sci. 2019, 169, 67–73. [Google Scholar] [CrossRef]
  14. Zhang, S.; Liu, Q.; Yang, Y.; Wang, D.; He, J.; Sun, L. Preparation, Morphology, and Structure of Kaolinites with Various Aspect Ratios. Appl. Clay Sci. 2017, 147, 117–122. [Google Scholar] [CrossRef]
  15. Hu, X.; Zhang, M.; Xu, W.; Tu, M.; Yin, Z.; Zhang, X. A Dynamic Coupled Elastoplastic Damage Model for Rock-like Materials Considering Tension-Compression Damage and Pressure-Dependent Behavior. Minerals 2022, 12, 851. [Google Scholar] [CrossRef]
  16. Shvarzman, A.; Kovler, K.; Grader, G.S.; Shter, G.E. The Effect of Dehydroxylation/Amorphization Degree on Pozzolanic Activity of Kaolinite. Cem. Concr. Res. 2003, 33, 405–416. [Google Scholar] [CrossRef]
  17. Wan, Q.; Rao, F.; Song, S. Reexamining Calcination of Kaolinite for the Synthesis of Metakaolin Geopolymers-Roles of Dehydroxylation and Recrystallization. J. Non-Cryst. Solids 2017, 460, 74–80. [Google Scholar] [CrossRef]
  18. Andrini, L.; Gauna, M.R.; Conconi, M.S.; Suarez, G.; Requejo, F.G.; Aglietti, E.F.; Rendtorff, N.M. Extended and Local Structural Description of a Kaolinitic Clay, Its Fired Ceramics and Intermediates: An XRD and XANES Analysis. Appl. Clay Sci. 2016, 124–125, 39–45. [Google Scholar] [CrossRef]
  19. Kelly, S.D.; Hesterberg, D.; Ravel, B. Analysis of soils and minerals using x-ray absorption spectroscopy. In Methods of Soil Analysis Part 5—Mineralogical Methods; 2008; Volume 5, pp. 387–464. [Google Scholar]
  20. Mathew, K.; Zheng, C.; Winston, D.; Chen, C.; Dozier, A.; Rehr, J.J.; Ong, S.P.; Persson, K.A. High-Throughput Computational X-ray Absorption Spectroscopy. Sci. Data 2018, 5, 180151. [Google Scholar] [CrossRef] [PubMed]
  21. Danner, T.; Norden, G.; Justnes, H. Characterisation of Calcined Raw Clays Suitable as Supplementary Cementitious Materials. Appl. Clay Sci. 2018, 162, 391–402. [Google Scholar] [CrossRef]
  22. Drits, V.A.; Sakharov, B.A.; Dorzhieva, O.V.; Zviagina, B.B.; Lindgreen, H. Determination of the Phase Composition of Partially Dehydroxylated Kaolinites by Modelling Their X-ray Diffraction Patterns. Clay Miner. 2019, 54, 309–322. [Google Scholar] [CrossRef]
  23. Andrini, L.; Moreira Toja, R.; Conconi, M.S.; Requejo, F.G.; Rendtorff, N.M. Halloysite Nanotube and Its Firing Products: Structural Characterization of Halloysite, Metahalloysite, Spinel Type Silicoaluminate and Mullite. J. Electron. Spectrosc. Relat. Phenom. 2019, 234, 19–26. [Google Scholar] [CrossRef]
  24. Newville, M. Fundamentals of XAFS. Rev. Mineral. Geochem. 2014, 78, 33–74. [Google Scholar] [CrossRef]
  25. de Jong, M.; Chen, W.; Geerlings, H.; Asta, M.; Persson, K.A. A Database to Enable Discovery and Design of Piezoelectric Materials. Sci. Data 2015, 2, 150053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wang, A.; Kingsbury, R.; McDermott, M.; Horton, M.; Jain, A.; Ong, S.P.; Dwaraknath, S.; Persson, K. A Framework for Quantifying Uncertainty in DFT Energy Corrections. Nature 2021, 11, 15496. [Google Scholar] [CrossRef] [PubMed]
  27. Munro, J.M.; Latimer, K.; Horton, M.K.; Dwaraknath, S.; Persson, K.A. An Improved Symmetry-Based Approach to Reciprocal Space Path Selection in Band Structure Calculations. NPJ Comput. Mater. 2020, 6, 112. [Google Scholar] [CrossRef]
  28. de Jong, M.; Chen, W.; Angsten, T.; Jain, A.; Notestine, R.; Gamst, A.; Sluiter, M.; Ande, C.K.; van der Zwaag, S.; Plata, J.J.; et al. Charting the Complete Elastic Properties of Inorganic Crystalline Compounds. Sci. Data 2015, 2, 150009. [Google Scholar] [CrossRef] [Green Version]
  29. Jain, A.; Ong, S.P.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. Commentary: The Materials Project: A Materials Genome Approach to Accelerating Materials Innovation. APL Mater. 2013, 1, 011002. [Google Scholar] [CrossRef] [Green Version]
  30. Ding, H.; Dwaraknath, S.S.; Garten, L.; Ndione, P.; Ginley, D.; Persson, K.A. Computational Approach for Epitaxial Polymorph Stabilization through Substrate Selection. ACS Appl. Mater. Interfaces 2016, 8, 13086–13093. [Google Scholar] [CrossRef] [Green Version]
  31. Patel, A.M.; Nørskov, J.K.; Persson, K.A.; Montoya, J.H. Efficient Pourbaix Diagrams of Many-Element Compounds. Phys. Chem. Chem. Phys. 2019, 21, 25323–25327. [Google Scholar] [CrossRef] [Green Version]
  32. Singh, A.K.; Zhou, L.; Shinde, A.; Suram, S.K.; Montoya, J.H.; Winston, D.; Gregoire, J.M.; Persson, K.A. Electrochemical Stability of Metastable Materials. Chem. Mater. 2017, 29, 10159–10167. [Google Scholar] [CrossRef] [Green Version]
  33. Latimer, K.; Dwaraknath, S.; Mathew, K.; Winston, D.; Persson, K.A. Evaluation of Thermodynamic Equations of State across Chemistry and Structure in the Materials Project. NPJ Comput. Mater. 2018, 4, 40. [Google Scholar] [CrossRef]
  34. Jain, A.; Hautier, G.; Ong, S.P.; Moore, C.J.; Fischer, C.C.; Persson, K.A.; Ceder, G. Formation Enthalpies by Mixing GGA and GGA +U Calculations. Phys. Rev. B 2011, 84, 045115. [Google Scholar] [CrossRef] [Green Version]
  35. Zheng, H.; Li, X.-G.; Tran, R.; Chen, C.; Horton, M.; Winston, D.; Persson, K.A.; Ong, S.P. Grain Boundary Properties of Elemental Metals. Acta Mater. 2020, 186, 40–49. [Google Scholar] [CrossRef] [Green Version]
  36. Horton, M.K.; Montoya, J.H.; Liu, M.; Persson, K.A. High-Throughput Prediction of the Ground-State Collinear Magnetic Order of Inorganic Materials Using Density Functional Theory. NPJ Comput. Mater. 2019, 5, 64. [Google Scholar] [CrossRef] [Green Version]
  37. Petousis, I.; Mrdjenovich, D.; Ballouz, E.; Liu, M.; Winston, D.; Chen, W.; Graf, T.; Schladt, T.D.; Persson, K.A.; Prinz, F.B. High-Throughput Screening of Inorganic Compounds for the Discovery of Novel Dielectric and Optical Materials. Sci. Data 2017, 4, 160134. [Google Scholar] [CrossRef] [PubMed]
  38. Persson, K.A.; Waldwick, B.; Lazic, P.; Ceder, G. Prediction of Solid-Aqueous Equilibria: Scheme to Combine First-Principles Calculations of Solids with Experimental Aqueous States. Phys. Rev. B 2012, 85, 235438. [Google Scholar] [CrossRef] [Green Version]
  39. Tran, R.; Xu, Z.; Radhakrishnan, B.; Winston, D.; Sun, W.; Persson, K.A.; Ong, S.P. Surface Energies of Elemental Crystals. Sci. Data 2016, 3, 160080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Aykol, M.; Dwaraknath, S.S.; Sun, W.; Persson, K.A. Thermodynamic Limit for Synthesis of Metastable Inorganic Materials. Sci. Adv. 2018, 4, eaaq0148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data Analysis for X-ray Absorption Spectroscopy Using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef] [Green Version]
  42. Ravel, B.; Newville, M.; Kas, J.; Rehr, J. The Effect of Self-Consistent Potentials on EXAFS Analysis. J. Synchrotron Radiat. 2017, 24, 1173–1179. [Google Scholar] [CrossRef] [PubMed]
  43. Rehr, J.J.; Kas, J.J.; Vila, F.D.; Prange, M.P.; Jorissen, K. Parameter-Free Calculations of X-ray Spectra with FEFF9. Phys. Chem. Chem. Phys. 2010, 12, 5503–5513. [Google Scholar] [CrossRef] [PubMed]
  44. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.; Refson, K.; Payne, M.C. First Principles Methods Using CASTEP. Z. Für Krist. Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef] [Green Version]
  45. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Swinehart, D.F. The Beer-Lambert Law. J. Chem. Educ. 1962, 39, 333. [Google Scholar] [CrossRef]
  47. Koningsberger, D.C.; Mojet, B.L.; Van Dorssen, G.E.; Ramaker, D.E. XAFS Spectroscopy; Fundamental Principles and Data Analysis. Top. Catal. 2000, 10, 143–155. [Google Scholar] [CrossRef]
  48. Gaur, A.; Shrivastava, B.D.; Nigam, H.L. X-ray Absorption Fine Structure (XAFS) Spectroscopy—A Review. Proc. Indian Natl. Sci. Acad. 2013, 79, 921–966. [Google Scholar]
  49. Crozier, E.D. A Review of the Current Status of XAFS Spectroscopy. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1997, 133, 134–144. [Google Scholar] [CrossRef]
  50. Kristó, J.; Frost, R.L.; Felinger, A.; Mink, J. FTIR Spectroscopic Study of Intercalated Kaolinite. J. Mol. Struct. 1997, 410–411, 119–122. [Google Scholar] [CrossRef]
  51. Sadri, S.; Johnson, B.B.; Ruyter-Hooley, M.; Angove, M.J. The Adsorption of Nortriptyline on Montmorillonite, Kaolinite and Gibbsite. Appl. Clay Sci. 2018, 165, 64–70. [Google Scholar] [CrossRef]
  52. Drits, V.A.; Zviagina, B.B.; Sakharov, B.A.; Dorzhieva, O.V.; Savichev, A.T. New Insight into the Relationships between Structural and FTIR Spectroscopic Features of Kaolinites. Clays Clay Miner. 2021, 69, 366–388. [Google Scholar] [CrossRef]
  53. Shafiq, N.; Nuruddin, M.F.; Khan, S.U.; Ayub, T. Calcined Kaolin as Cement Replacing Material and Its Use in High Strength Concrete. Constr. Build. Mater. 2015, 81, 313–323. [Google Scholar] [CrossRef]
  54. Borst, A.M.; Smith, M.P.; Finch, A.A.; Estrade, G.; Villanova-de-Benavent, C.; Nason, P.; Marquis, E.; Horsburgh, N.J.; Goodenough, K.M.; Xu, C. Adsorption of Rare Earth Elements in Regolith-Hosted Clay Deposits. Nat. Commun. 2020, 11, 1–15. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, G.; Zhao, H.; Li, X.; Xia, S. Theoretical Investigation of the Chloride Effect on Aqueous Hg (II) Adsorption on the Kaolinite (001) Surface. Appl. Clay Sci. 2021, 210, 106120. [Google Scholar] [CrossRef]
  56. Park, S.-M.; Lee, J.; Jeon, E.-K.; Kang, S.; Alam, M.S.; Tsang, D.C.; Alessi, D.S.; Baek, K. Adsorption Characteristics of Cesium on the Clay Minerals: Structural Change under Wetting and Drying Condition. Geoderma 2019, 340, 49–54. [Google Scholar] [CrossRef]
Figure 1. TG/DSC in the kaolinite−metakaolin transition region.
Figure 1. TG/DSC in the kaolinite−metakaolin transition region.
Minerals 12 00994 g001
Figure 2. (a) FTIR spectra of Raw−Ka, Ka−10 min, Ka−30 min, Ka−120 min; (b) local amplification of FTIR spectra of Raw−Ka, Ka−10 min, Ka−30 min, Ka−120 min in range of wavenumbers = 400−1700 cm−1 and 3400−4000 cm−1.
Figure 2. (a) FTIR spectra of Raw−Ka, Ka−10 min, Ka−30 min, Ka−120 min; (b) local amplification of FTIR spectra of Raw−Ka, Ka−10 min, Ka−30 min, Ka−120 min in range of wavenumbers = 400−1700 cm−1 and 3400−4000 cm−1.
Minerals 12 00994 g002
Figure 3. SEM images of the studied kaolin (a) Raw-Ka, (b) Ka-10 min, (c) Ka-30 min, (d) Ka-120 min.
Figure 3. SEM images of the studied kaolin (a) Raw-Ka, (b) Ka-10 min, (c) Ka-30 min, (d) Ka-120 min.
Minerals 12 00994 g003
Figure 4. Crystal structure of (a) Sim-Ka, (b) Phase 1, (c) Phase 2, (d) Phase 3 and (e) Phase 4.
Figure 4. Crystal structure of (a) Sim-Ka, (b) Phase 1, (c) Phase 2, (d) Phase 3 and (e) Phase 4.
Minerals 12 00994 g004
Figure 5. (a) XRD patterns of kaolin annealed for different holding times; (b) local amplification of kaolin annealed for different holding time, in the range of θ = 10–30°.
Figure 5. (a) XRD patterns of kaolin annealed for different holding times; (b) local amplification of kaolin annealed for different holding time, in the range of θ = 10–30°.
Minerals 12 00994 g005
Figure 6. (a) The simulated XRD patterns of Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4; (b) partial amplification of simulated XRD patterns of Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4 in the range of θ = 18–26°.
Figure 6. (a) The simulated XRD patterns of Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4; (b) partial amplification of simulated XRD patterns of Sim-Ka, Phase 1, Phase 2, Phase 3, and Phase 4 in the range of θ = 18–26°.
Minerals 12 00994 g006
Figure 7. (a) The normalized X−ray absorption spectra for the Si K−edge for Sim−Ka, Phase 1, Phase 2, Phase 3, and Phase 4; (b) the EXAFS spectrum k2 χ(k) as a function of the wave vector, k, for the Si K−edge for Sim−Ka, Phase 1, Phase 2, Phase 3, and Phase 4; (c) the radial structural function for Sim−Ka, Phase 1, and Phase 2 at the Si K−edge.
Figure 7. (a) The normalized X−ray absorption spectra for the Si K−edge for Sim−Ka, Phase 1, Phase 2, Phase 3, and Phase 4; (b) the EXAFS spectrum k2 χ(k) as a function of the wave vector, k, for the Si K−edge for Sim−Ka, Phase 1, Phase 2, Phase 3, and Phase 4; (c) the radial structural function for Sim−Ka, Phase 1, and Phase 2 at the Si K−edge.
Minerals 12 00994 g007
Figure 8. The microstructure evolution of the Si−O units in kaolinite: (a)Sim−Ka; (b) Phase 1; (c) Phase 2; (d) Phase 3; (e) Phase 4.
Figure 8. The microstructure evolution of the Si−O units in kaolinite: (a)Sim−Ka; (b) Phase 1; (c) Phase 2; (d) Phase 3; (e) Phase 4.
Minerals 12 00994 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zeng, Q.; Xie, J.; Zhou, W.; Zhu, J.; Liu, L.; Yin, J.; Zhu, W. Study of the Crystallographic Distortion Mechanism during the Annealing of Kaolinite. Minerals 2022, 12, 994. https://doi.org/10.3390/min12080994

AMA Style

Zeng Q, Xie J, Zhou W, Zhu J, Liu L, Yin J, Zhu W. Study of the Crystallographic Distortion Mechanism during the Annealing of Kaolinite. Minerals. 2022; 12(8):994. https://doi.org/10.3390/min12080994

Chicago/Turabian Style

Zeng, Qiuyu, Jun Xie, Wei Zhou, Jinbo Zhu, Liangliang Liu, Jianqiang Yin, and Wenliang Zhu. 2022. "Study of the Crystallographic Distortion Mechanism during the Annealing of Kaolinite" Minerals 12, no. 8: 994. https://doi.org/10.3390/min12080994

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop