Next Article in Journal
Copper Sulfation from Enargite Roasting Using Coal and Fayalite Slag Mixture
Previous Article in Journal
Application of Quantum Chemistry in the Study of Flotation Reagents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insight into Crystalline Structure and Physicochemical Properties of Quartz-Carbon Ore

1
Hunan Key Lab of Mineral Materials and Application, School of Minerals Processing and Bioengineering, Central South University, Changsha 410083, China
2
Jiangxi Province Natural Resources Interests and Reserve Security Center, Department of Natural Resources of Jiangxi Province, Nanchang 330025, China
3
Jiangxi Mineral Resources Guarantee Service Center, Department of Natural Resources of Jiangxi Province, Nanchang 330025, China
4
Engineering Research Center of Nano-Geomaterials of Ministry of Education, China University of Geosciences, Wuhan 430074, China
5
Laboratory of Advanced Mineral Materials, China University of Geosciences, Wuhan 430074, China
6
Faculty of Materials Science and Chemistry, China University of Geosciences, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(12), 1488; https://doi.org/10.3390/min13121488
Submission received: 27 October 2023 / Revised: 16 November 2023 / Accepted: 24 November 2023 / Published: 27 November 2023

Abstract

:
Composites made from carbon and nanominerals show great potential for thermal phase change materials, environmental water treatment, and biomass conversion. In 2019, a micro and nano-quartz-carbon ore was discovered in Feng-cheng City, Jiangxi Province. The study of the structural and physicochemical changes of quartz-carbon ore (QZC) during calcination is essential for the preparation of QZC-based composites and to broaden their application areas. Firstly, the SiO2 crystal structure evolution of QZC during calcination was investigated using in-situ X-ray diffraction (XRD), 29Si magic-angle sample spinning nuclear magnetic resonance (MAS NMR), and Fourier transform infrared FTIR spectroscopy. Then, the changes in carbon during calcination were investigated using Raman spectroscopy, 13C MAS NMR, and X-ray photoelectron spectroscopy (XPS). In addition, changes in the QZC morphology were observed using scanning electron microscopy (SEM) and transmission electron microscopy (TEM) techniques. Finally, the evolution of the physicochemical properties of QZC during calcination was revealed using thermogravimetric (TG), Brunauer–Emmet–Teller (BET), resistivity, thermal conductivity, and zeta potential techniques.

Graphical Abstract

1. Introduction

Over the past decades, nanominerals have attracted considerable scientific interest due to their unique electronic and physical properties [1,2,3]. They have broad application prospects concerning energy storage and conversion [4,5], biological therapy [6], and environmental restoration [7]. Carbon materials have been widely studied and applied for their excellent thermodynamic properties, chemical stability, high specific surface area, high development, and good biocompatibility [8]. In recent years, the synergistic preparation of carbon/nanomineral composites from nanomineral and carbon materials has been widely studied [9]. Zhao et al. prepared carbon-coated aluminosilicate nanosheets with graded porosity using sucrose as a carbon source and kaolinite as a nanomineral [10]. They further loaded stearic acid to develop a composite phase change material with photothermal conversion and thermal energy storage capabilities. Hao et al. prepared graphene oxide/montmorillonite composite aerogels for the removal of planar hydrated copper ions from complex wastewater systems, and the addition of montmorillonite can improve the adsorption performance of graphene oxide-based aerogels and reduce the cost [11]. Binder-free continuous films of mesoporous carbon materials were synthesized on graphite collectors using ordered mesoporous silica films as hard templates; the synthesized films with excellent performance in terms of specific capacitance, rate performance, and electrochemical stability can be used as electrodes in microcapacitors [12]. Composites combined with carbon and nanominerals show great potential for thermal phase change materials, environmental water treatment, and biomass conversion [13,14,15]. However, the preparation of functional carbon/nanomineral composites by the addition of carbon materials to nanominerals often has the disadvantages of complicated and expensive processes. Therefore, natural carbon-containing nanominerals are critical for resolving the aforementioned issues.
In 2019, micro and nano-quartz-carbon ore was discovered in Feng-cheng City, Jiangxi Province [16]. The resource reserves of this deposit are large-scale. Tang et al. studied the process mineralogy of quartz-carbon ore [17]. After the ore mineral processing separation and purification tests, mesoporous silicon, mesoporous carbon, silica, nano-silica powder, and other materials can be obtained, and their industrial value is very high [16]. In addition, little has been reported on the crystal structure and physicochemical properties of QZC. Therefore, the systematic study of the crystal structure, morphology, and surface physicochemical properties of QZC is of great theoretical significance for the realization of the functionalized design of QZC.
Understanding the structural and compositional evolution of QZC during gradual calcination in different atmospheres at high temperatures is essential to facilitate its application in catalysts, adsorbents, and phase change materials for thermal storage. This study investigates the thermal stability of QZC using thermogravimetry/differential scanning calorimetry (TG/DSC). It tracks the structural changes in QZC in real time during thermal treatment via in-situ X-ray diffraction (XRD) and Fourier transform infrared spectrometry (FTIR). Furthermore, the physicochemical properties of minerals at different temperatures were systematically characterized using the Brunauer–Emmet–Teller (BET), laser particle size analyzer, zeta potential, scanning electron microscopy (SEM), and transmission electron microscopy (TEM) techniques, with the goal of providing a theoretical foundation for studying calcined QZC in composite materials, with the aim of laying the groundwork for the full utilization of the QZC and synthetic QZC composites, as well as for expanding the direction of applications.

2. Materials and Methods

2.1. Materials

The ore sample studied in this paper came from a drill hole in Fengcheng, Jiangxi Province, China (Figure S1). It comprised a stable black powder, which could not be ignited, was less dense than water, and was mainly composed of quartz, carbon, and other impurities. For testing and research purposes, the ore sample was ground and mixed, the QZC was sieved through an 80 mesh sieve, and the sieve offsets were dried in an oven at 50 °C for 24 h to obtain QZC. Table S1 shows the sample oxides’ chemical composition and carbon content analyzed by an X-ray fluorescence (XRF) spectrometer and total organic carbon analyzer (TOC).

2.2. Preparation of QZC-A and QZC-N

A series of calcined silica carbonite samples were obtained by heating the QZC to the specified temperatures (from 100 to 1000 °C). QZC calcined under different atmospheres, air, and nitrogen (N2), were labeled as QZC-A and QZC-N, respectively. The heating rate was 5 °C/min, held for 120 min after reaching the target temperature, then cooled naturally. The QZC becomes less malleable after heat treatment. The QZC and stearic acid (SA) were uniformly mixed using solvent evaporation, and the QZC mixture was pressed into small disc-shaped pieces to measure thermal conductivity.

2.3. Characterization

The chemical composition of the QZC was examined using a wavelength dispersive X-ray fluorescence spectrometer (XRF, PANalytical Axios, Amsterdam, the Netherlands). Determination of the total carbon (TC) values of QZC was performed using a total organic carbon analyzer (elementary vario TOC select, Frankfurt, Germany). X-ray diffraction (D8 ADVANCE, Bruker Corp., Saarbrucken, Germany) was used to study the evolution of the crystal structure and phase composition of the QZC samples after treatment with different temperatures and atmospheres. FTIR spectra were obtained by an FTIR spectrometer (BRUKER Tensor II, Karlsruhe, Germany) over a wavelength range of 4000–400 cm−1. Analysis of changes in the local chemical environment and structure of H, C, and Si in the sample was conducted using solid-state NMR Popper NMR (ADVANCE III 400 WB, Saarbrucken, Germany). Sample surface characterization and carbon valence were characterized via X-ray photoelectron spectroscopy (ESCALAB Xi+, Massachusetts, MA, USA) using an Al Kα monochromator X-ray source. Changes in the chemical bonding environment led to important changes in photoelectron energy, which allowed chemical information to be obtained. C1s binding energy was 284.8 eV and was the reference for all measurements. Raman is the most efficient method for examining the structure of carbon without harming the substance since it is the fingerprint spectrum of the substance. The graphitization degree of the QZC was studied using a Raman spectrometer (LabRAM HR Evolution, Paris, France) at a 325 nm excitation wavelength.
The microstructure of the samples was investigated by scanning electron microscopy (SEM, ZEISS SIGMA 300, Oberkochen, Germany) with an energy dispersive (EDS) X-ray spectroscopy and transmission electron microscopy (TEM, JEOL JEM-2100F, Tokyo, Japan). Thermogravimetric and differential scanning calorimetric studies (TG/DSC, NETZSCH STA 449F3, Selb, Germany) were carried out to determine the thermal stability of raw QZC at a heating rate of 10 °C/min from 30 to 1000 °C. The specific surface area and the pore size distributions of the supporting materials were measured according to the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods using a nitrogen adsorption apparatus (ASAP 2020 Plus, Micromeritics, Norcross, GA, USA) at 77 K. The zeta potential of the samples was tested using a zeta potential analyzer (Brookhaven/90Plus PALS, Malvern, PA, USA) in the pH range of 1–12. The pH of the system was adjusted with 0.1 mol/L HCl or NaOH solution, respectively, using KCl as the electrolyte. A thermal conductivity meter (XIATECH TC3100, Xi’an, China) was used to characterize the pattern of changes in the thermal conductivity of the samples during crystal structure modulation and compositional changes.

3. Results and Discussion

3.1. Structural Characterization of Silicon and Carbon in QZC

Changes in the mineral crystal structure and physical phase during heat treatment were tracked in real-time. In-situ XRD with air atmosphere was performed to determine the relevant effects of temperature on the QZC structure and phase. Figure 1a shows the in-situ XRD spectra of the QZC during the heat treatment under the air atmosphere. The silica structure in the QZC is stable and highly crystalline, and there are no major changes in the distinctive peaks during the heat treatment process. After magnified observation of the characteristic peaks, it can be seen that the SiO2 peak shifts towards a large angle after 550 °C, the cell parameters become larger, and the lattice becomes larger, transforming from low-temperature quartz to high-temperature quartz. As an example, the characteristic peak of the (101) crystal surface at 2θ = 26–28 ° (Figure 1b). Meanwhile, the value of d (100) gradually increased from 4.25 nm to a maximum of 4.29 nm at 550 °C and then gradually decreased to 4.28 nm (Figure S2).
In order to further explore the evolution of the local chemical environment of Si atoms in the QZC after heat treatment [18], NMR tests were performed on the QZC, QZC400-N, QZC600-N, and QZC800-N samples. The 29Si spectra of sample QZC show symmetrical signal spikes at −91.90 ppm and −107.94 ppm, corresponding to the typical Q3 and Q4 structures of the Si atoms in the QZC (Figure 1c). The Q4 structure, i.e., Si atoms on [SiO4] tetrahedra attached to four other equivalent Si-O tetrahedra [19], the sharp 29Si NMR signals reflect the highly ordered, crystalline nature of the precursor [20]. As the calcination temperature increased, the NMR chemical shifts corresponding to Q4 of QZC400-N, QZC600-N, and QZC800-N shifted from the original −107.94 ppm to 108.02, 108.01 and 108.15 ppm, respectively. The weakening and disappearance of the resonance peaks of the corresponding Q3 structures of samples QZC600-N and QZC800-N and the broadening of the resonance peaks of Q4 indicate a change in the bond lengths, bond angles, and cation-oxygen bond strengths of the QZC. The lower polymeric silica in Q3 is converted to higher polymeric silica in Q4.
During the heat treatment process, the silica phase and coordination in the QZC undergo changes, but its crystallinity remains stable and almost unchanged. This property makes it suitable for high-temperature stabilization and fire-retardant materials. The relative crystallinity curve from Figure S3 shows that the crystallinity index of the QZC remains relatively stable, with slight variations above and below 100%. However, there is a direct drop in crystallinity to 88.5% at 300 °C, which is attributed to the removal of surface hydroxyl groups (−OH) from the QZC. This removal does not affect the SiO2 crystallinity index. The XRD analysis does not reflect the presence of carbon peaks in the QZC, and the XRD patterns of the samples remain unchanged after heat treatment in both the air and N2 atmosphere (Figure S4). Additionally, the crystallinity index of the QZC at 600 °C remained largely unchanged, indicating that the carbon in the QZC was amorphous. The βcosθ~sinθ fit lines for the heat-treated QZC are almost identical straight lines with very similar slopes (Figure 1d), and the crystal structure is stable and unchanged [21].
The FTIR technique was utilized to investigate the changing pattern of the groups on the surface of the QZC during the heat treatment process. The presence of a large number of (−OH) on the surface of the QZC can be seen in Figure 1e. The characteristic absorption peaks of (−OH) stretching vibrations appear in the range of 4000–3500 cm−1, with two strong absorption peaks at wave numbers 3618.21 cm−1 and 3687.04 cm−1, respectively, and are attributed to the (−OH) within the QZC, formed by the stretching vibrations of the inner (−OH) and the inner surface (−OH), respectively [22]. The inner hydroxyl axis is distributed almost parallel to the layer and, therefore, has a lower frequency. Compared to the QZC, the hydroxyl absorption peaks of QZC500-A at these two sites disappear [23]. A single characteristic absorption peak was observed at 1567.89 cm−1, which corresponds to the stretching vibration of the carbon-carbon bond, as confirmed by NMR spectroscopy [24]. The carbon-carbon absorption peak at QZC600-A disappears due to the reaction of carbon with oxygen at high temperatures. After calcination under a nitrogen atmosphere, the stretching vibration of the carbon-carbon bond in QZC did not disappear, and this is the main difference between the FTIR spectra of QZC-A and QZC-N (Figure 1f). Five characteristic absorption peaks were identified in the range 400–800 cm−1. The double peaks at 440.88 cm−1 and 510.10 cm−1 were Si-O-Si bending vibration absorption peaks, and the single characteristic absorption peak at 691.97 cm−1 was a Si-O-Si stretching vibration absorption peak [25]. The Si-O-Si symmetric stretching vibration frequency of crystalline SiO2 splits and is located as a double peak at 774.26 cm−1 and 792.57 cm−1 [26]. In contrast, the amorphous SiO2 symmetric stretching vibration frequency is a single peak located at around 800 cm−1, so based on the shape of the symmetric stretching vibration absorption peak, the QZC can be judged to be crystalline silica. The single characteristic absorption peak at 911.77 cm−1 is the −OH oscillating vibrational band and disappears at QZC500, corresponding to the presence of two absorption peaks at 3618.21 cm−1 and 3687.04 cm−1. The Si-O-Si anti-symmetric stretching vibration absorption peaks in the silica-oxygen tetrahedra are at 1027.25 cm−1 and 1004.96 cm−1, where QZC500-A shows a blue shift, indicating an enhanced and stable Si-O-Si bond compared to the QZC.
QZC is mainly composed of SiO2 and amorphous carbon. The alterations in silicon’s crystal structure and coordination, as well as the functional groups, are revealed above. The characteristics of carbon matter in physics will then be revealed using Raman, 13C NMR, and XPS. All samples show two broad peaks at 1345 cm−1 and 1580 cm−1, corresponding to the D and G bands, with carbon defects estimated by the relative intensity ratio (ID/IG) [27,28]. The ID/IG values for QZC, QZC600-N, and QZC800-N are 1.02, 1.04, and 1.06, respectively, gradually increasing with increasing charring temperature (Figure 2a), indicating an increase in defect carbon and a decrease or no change in graphite carbon [29,30]. The carbon-carbon double bond absorption peak at QZC800-N starts to weaken (Figure 1f), the carbon-carbon double bond starts to decrease, and the carbon-carbon double bond corresponds to the carbon hybridization mode of sp2 hybridization. Figure 1f corresponds to the results of the data in Figure 2a.
The carbon-carbon double bond in the QZC is represented by a symmetric signal spike chemically shifted at 120.15 ppm in the 13C NMR spectrum (Figure 2b). As the calcination temperature rises, the resonance peaks for QZC400-N, QZC600-N, and QZC800-N shift to 117.57, 116.96, and 108.32 ppm, respectively. The resonance peaks progressively expanded, and the intensity of QZC800-N declined dramatically, indicating that the carbon-carbon double bond was reduced and the C hybridization mode corresponding to the carbon-carbon double bond was sp2, consistent with the Raman results. O1s, C1s, Si2s, and Si2p were all detected in the QZC at different carbonization temperatures with binding energies of 528, 285, 154, and 103 eV, respectively (Figure 2c). To illustrate the structural evolution of elemental carbon further, we compared the high-resolution spectra of C1s of QZC at different calcination temperatures. Four diffraction peaks were present at binding energies of 284.80 eV, 285.20 eV, 285.80 eV, and 289.10 eV for all samples, corresponding to sp2-bound carbon(sp2-C), sp3-bound carbon(sp3-C), C-OH, and −COOH, respectively [13,31,32] (Figure 2d–f). Therefore, the results of the C1s high-resolution spectra show that the relative intensity of the sp3-bound carbon(sp3-C) increases with increasing carbonization temperature at 285.40 eV, with relative intensities of 25,277.76 eV, 26,194.96 eV, and 26,634.67 eV, respectively, indicating an improvement in the crystal structure of the carbon in the QZC. Raman, 13C NMR, and XPS results all indicate that as the calcination temperature increases, the defective carbon content rises, favoring a large number of adsorption sites for the applied material. The increase in defective carbon in QZC by calcination facilitates the preparation of carbon-based defective materials for applications in the direction of electrocatalysis, lithium-sulfur batteries, and energy storage.

3.2. Micromorphological Analysis

The color of the QZC varies throughout calcination, with a dramatic shift from black to pinkish white at 600 °C, attributed to the decomposition of its amorphous carbon and a transition to pure white at 1000 °C (Figure S5). It can be observed that the QZC is in a lamellar structure with tiny irregularly shaped crystals, and the original mineral grain size is in the range of tens of nanometers to about 1 micron in the micro-nano structure (Figure 3 and Figure S6). Silicon and carbon exist together without a distribution pattern (Figure 3 and Figure S7). After calcination at high temperatures, the QZC dehydroxylation appears as slabbing, and the particle size becomes larger (Figure S6). Even when heated to 1000 °C, the quartz-carbon ore retains an excellent lamellar structure, indicating its thermal resilience. Natural QZC crystal particles are made up of micro and nanosheets with tiny diameters, thin thicknesses, and great crystallinity. After heat treatment, the flakes’ lamellar structure stays unaltered. The amorphous carbon and SiO2 crystal structures are visible in Figure 4. The FTIR scan reveals that the heat-treated QZC retains a strong crystal structure, exhibiting its high thermal stability, which is consistent with the XRD results.

3.3. Physico-Chemical Properties of Quartz-Carbon Ore

The above experimental test results show the crystal structure model of QZC and the changes in the physical phase, crystal structure, functional groups, and quartz-carbon during heat treatment to further explore the physicochemical properties of QZC using TG-DSC-DTG, BET, resistivity and thermal conductivity, and zeta potential. They were evaluated for thermal stability under air and N2 atmospheres to reveal more about the carbon content and thermal transformation capabilities of the QZC. The TG-DSC-DTG curve for quartz-carbon ore under air atmosphere (Figure 5a) over the range of 20–1000 °C degrees can be divided into three phases. The first stage was from 20 to 157 °C with a QZC weight loss of 0.18% and a small valley of heat uptake at 81.2 °C, which was mainly attributed to the removal of adsorbed water from the QZC and the release of volatile compounds from the mineral sample; A second stage of 157–426 °C was for the removal of interlayer water (structure water). The third stage was 396–700 °C with a weight loss of 13.39%, indicating the complete decomposition of the fixed carbon in the QZC, which decomposes at a higher temperature in the QZC than in coal because of the silicon-covered carbon structure in the QZC and the protection of the carbon by the silicon. The exothermic peak at this stage is 670 °C, with an exothermic peak at 573.2 °C corresponding to the crystalline transition temperature of quartz for the transition from α to β quartz. The TG-DSC-DTG curve under an N2 atmosphere (Figure 5b) can likewise be divided into three stages. The first stage is from 20–162 °C, where the QZC loses 0.28% of its weight, with a small valley of heat absorption at 81.4 °C, which is mainly attributable to the removal of water adsorbed by QZC. A second stage of 162–420 °C was for the removal of interlayer water (structure water). The third stage is 396–1000 °C with a weight loss of 1.89%, which is mainly the decomposition of organic carbon and may contain the removal of humic acids. A peak in 573.1 °C appears as a transition from α to β quartz [33].
The pore organization structure of the material is related to the shape of the adsorption equilibrium isotherm; according to IUPAC classification, the N2 adsorption/desorption isotherm curves of the calcined samples under both air and N2 atmospheres are of type IV. The hysteresis line corresponds to the mixing of H3 and H4 (Figure S9); there is no obvious saturation adsorption plateau, the pore structure is more unevenly distributed, and the QZC lamellar particles accumulate to form cleavage pores at a lower relative. The adsorption that occurred at lower relative pressures P/P0 was mainly monomolecular layer adsorption followed by multilayer adsorption, with pressures P/P0 larger than 0.9 being sufficient for capillary coalescence to occur and the adsorption isotherm showing a sudden surge. According to the pore size distribution graphs of the samples calculated using the BJH method (suitable for mesopores), the raw samples are concentrated at 3.8 nm (Figure 5c,d), and the disappearance of the peak at 3.8 nm in the pore size distribution curve after calcination of the samples in air atmosphere up to 600 °C, as well as the absence of a hysteresis loop, indicates a collapse of the pore structure. Compared to the air atmosphere calcined samples, for the QZC after calcination under an N2 atmosphere, the isothermal curves and pore size distribution curves did not change much within the temperature range of 400–1000 °C. This indicates that the mesopores in the sample are mainly formed by carbon material and are not related to silica [34,35]. The porous structure influenced by silicon and carbon in the QZC can be loaded with catalysts for catalysis and environmental applications such as air purification, water purification, self-purification, sterilization, and deodorization.
Ordinary SiO2 materials have very low electrical conductivity and are generally regarded as insulators, so they can generally be used to produce high-frequency insulators. The carbon in the QZC is a crystal structure formed by sp2 hybridized carbon atoms in a honeycomb arrangement. In contrast, the crystal structure of [SiO4] tetrahedra and sp2 hybridized carbon atoms in the QZC crystal alters the electron transfer path within the SiO2 crystal in the QZC. It has an impact on the electrical conductivity of the QZC. The resistivity in quartz-carbon ore gradually decreased from 161.14 Ω·cm to 19.39 Ω·cm under 2–20 Mpa (Figure 5e), and after calcination under air atmosphere, the carbon in the crystal of QZC600-A was missing, which considerably blocked the electron transfer path in the crystal, and the resistivity directly increased to 233.23–629.24 MΩ·cm, compared with QZC600-A, the other volatile impurities in the sample QZC1000-A were removed entirely and the resistivity increased slightly, which is expected to become a functional anti-static filler for industrial paints.
The thermal conductivity of the QZC after calcination under an air atmosphere, decreases with the absence of carbon (Figure 5f). Heat treatment allows for the modification of the crystal structure and composition of quartz-carbon ore, which has a significant impact on their thermal conductivity. As a result, the implant’s crystal structure, composition, and particle size gradually decrease from 0.121 W·m−1·K−1 to 0.094 W·m−1·K−1. The decrease in thermal conductivity observed in sample QZC600-A can be attributed to the carbon loss that occurred during the heat treatment of quartz-carbon ore, leading to a modification in the crystal structure of the sample. The crystal structure of the quartz-carbon ore experiences an increase in bond lengths and bond angles in the [SiO4] tetrahedra, as well as the elimination of carbon-carbon double bonds. These changes result in a significant number of thermally excited defects. The presence of these defects causes severe phonon-defect scattering, which hinders the propagation of phonons, shortens their average free range, and ultimately reduces the thermal conductivity of QZC. Compared to QZC600-A, QZC1000-A is completely calcined, with the disappearance of volatile fraction, and the thermal conductivity is slightly reduced from 0.103 W·m−1·K−1 to 0.094 W·m−1·K−1, mainly due to the effect of particle size distribution. The QZC1000-A particle size is larger compared to QZC600-A, and the non-homogeneity is increased, producing secondary pores, which increases the interfacial area of the QZC, enhancing the phonon scattering effect and reducing the thermal conductivity of the QZC [36,37]. Consideration can be given to separating the silicon and carbon in the QZC using some physicochemical methods and using the carbon to make materials with high thermal conductivity, for example, for making heat storage materials, which increase the channels for heat transfer through the increase of defective carbon after heat treatment.
The surface potential of the quartz-carbon ore calcined under different temperatures and atmospheres changes significantly, according to the zeta potential function results (Figure 5g,h). The raw ore and QZC400-A appear to have an isoelectric point in an acidic environment, 4.10 and 3.16, respectively, and the potential values are negative after 600 °C and increase with pH. This indicates that the potential of the QZC is influenced by the carbon material in the sample, with the carbon material carrying a positive point in an acidic environment and the carbon material carrying a negative charge after calcination and decomposition due to the dehydroxylation of the QZC. After calcination, the potential of the QZC shows a negative value and increases, which can be used to treat positively charged pollutants, such as Methylene blue and Rhodamine B.

4. Conclusions

In this study, the material composition, morphology, crystal structure, and physicochemical properties of quartz-carbon ore were first studied. The quartz-carbon ore consists of well-crystallized SiO2 and amorphous carbon; SiO2 in the QZC contains hydroxyl groups and has Q4 and Q3 structures, and its amorphous carbon is connected by carbon-carbon double bonds and contains hydroxyl and carboxyl groups. The interlocking pattern between SiO2 and amorphous carbon appears through Si and C. Notably, the Si-C interlocking structure of the ore was proposed for the first time in conjunction with the formation process of the QZC. Next, the evolution of the crystal structure, morphology, and physical and chemical properties were systematically investigated by heat treatment of the ore under air and nitrogen atmospheres. The SiO2 in the QZC is well crystalline, and the stable crystal structure after heat treatment was proven by the Si atomic coordination changes, and the transformation from lower polymeric Si in Q3 to higher polymeric Si in Q4 happens after 600 °C. The four characterization results from the XRD, 13C NMR, Raman, and XPS correspond and demonstrate a reduction in carbon-carbon double bonds, an increase in defective carbon, and an increase in the relative strength of sp3 hybridized carbon. TG, BET, resistivity, thermal conductivity, and zeta tests on samples of the ore after heat treatment under air and N2 atmospheres subsequently concluded that the collapse of mesopores, the strength of resistivity, the high thermal conductivity, and the change in potential are all closely related to the carbon in the ore. Our results systematically confirm the structure evolution of the QZC after heat treatment under different atmospheres, providing a theoretical basis for the preparation of functional QZC composites and a direction for high-value applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min13121488/s1, Figure S1: Drill core sample of quartz-carbon ore; Figure S2: Corresponding d(100) values and peak positions; Figure S3: relative crystallinity of QZC; Figure S4: XRD curves of pristine QZC and calcined samples at different temperatures (a) under air and (b) under N2 atmospheres; Figure S5: whiteness change in the QZC during heat treatment in air; Figure S6: particle size distribution of pristine QZC and calcined samples at different temperatures (a) under air and (b) under N2 atmospheres; Figure S7: SEM images of the (a) QZC, (b) QZC400-N, (c) QZC500-N, (d) QZC600-N, (e) QZC800-N, (f) QZC1000-N; the mapping of (g) QZC and (h) QZC1000-N samples; TEM images of the (a) QZC; Figure S8: TEM and HRTEM images of (b) QZC, (c) QZC400-N, (d) QZC600-N, (e) QZC800-N, (f) QZC1000-N; Figure S9: N2 adsorption/desorption isotherm curves of pristine QZC and calcined samples at different temperatures (a) under air and (b) under N2 atmospheres; QZC polarizing microscope features; Table S1: chemical composition of pristine QZC by XRF and TOC; References [38,39,40,41,42].

Author Contributions

Methodology, X.L. and X.Z.; validation, X.L., J.W. and Y.T.; formal analysis, X.W.; investigation, X.L.; resources, X.L. and X.W.; data curation, X.L. and J.Z.; writing—original draft preparation, X.L.; writing—review and editing, H.Y., J.L. and Q.W.; conceptualization, H.Y.; Supervision, H.Y.; funding acquisition, H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program of China (2022YFC2904804), the CUG Scholar Scientific Research Funds at China University of Geosciences (Wuhan) (2019152), the Fundamental Research Funds for the Central Universities at China University of Geosciences (Wuhan), and the National Science Fund for Distinguished Young Scholars (51225403).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Phuekphong, A.F.; Imwiset, K.J.; Ogawa, M. Designing Nanoarchitecture for Environmental Remediation Based on the Clay Minerals as Building Block. J. Hazard. Mater. 2020, 399, 122888. [Google Scholar] [CrossRef] [PubMed]
  2. Zhou, Y.; LaChance, A.M.; Smith, A.T.; Cheng, H.; Liu, Q.; Sun, L. Strategic Design of Clay-Based Multifunctional Materials: From Natural Minerals to Nanostructured Membranes. Adv. Funct. Mater. 2019, 29, 1807611. [Google Scholar] [CrossRef]
  3. Li, Y.; Dong, S.; Qiao, J.; Liang, S.; Wu, X.; Wang, M.; Zhao, H.; Liu, W. Impact of Nanominerals on the Migration and Distribution of Cadmium on Soil Aggregates. J. Clean. Prod. 2020, 262, 121355. [Google Scholar] [CrossRef]
  4. Gao, D.C.; Sun, Y.; Fong, A.M.; Gu, X. Mineral-Based Form-Stable Phase Change Materials for Thermal Energy Storage: A State-of-the Art Review. Energy Storage Mater. 2022, 46, 100–128. [Google Scholar] [CrossRef]
  5. Zhao, X.; Tang, Y.; Xie, W.; Li, D.; Zuo, X.; Yang, H. 3D Hierarchical Porous Expanded Perlite-Based Composite Phase-Change Material with Superior Latent Heat Storage Capability for Thermal Management. Constr. Build. Mater. 2023, 362, 129768. [Google Scholar] [CrossRef]
  6. Liao, J.; Qian, Y.; Sun, Z.; Wang, J.; Zhang, Q.; Zheng, Q.; Wei, S.; Liu, N.; Yang, H. In Vitro Binding and Release Mechanisms of Doxorubicin from Nanoclays. J. Phys. Chem. Lett. 2022, 13, 8429–8435. [Google Scholar] [CrossRef]
  7. Yu, M.; Hua, Y.; Sarwar, M.T.; Yang, H. Nanoscale Interactions of Humic Acid and Minerals Reveal Mechanisms of Carbon Protection in Soil. Environ. Sci. Technol. 2023, 57, 286–296. [Google Scholar] [CrossRef]
  8. Egbedina, A.O.; Bolade, O.P.; Ewuzie, U.; Lima, E.C. Emerging Trends in the Application of Carbon-Based Materials: A Review. J. Environ. Chem. Eng. 2022, 10, 107260. [Google Scholar] [CrossRef]
  9. Huang, H.; Huang, D.; Chen, S.; Wang, G.; Chen, Y.; Tao, J.; Chen, H.; Gao, L. Removing Antibiotic Resistance Genes under Heavy Metal Stress with Carbon-Based Materials and Clay Minerals: By Sorption Alone? Chem. Eng. J. 2022, 446, 137121. [Google Scholar] [CrossRef]
  10. Zhao, X.; Tang, Y.; Zhang, X.; Zuo, X.; Yang, H. Development of Carbon-Coated Aluminosilicate Nanolayers Composite Shape-Stabilized Phase Change Materials with Enhanced Photo-Thermal Conversion and Thermal Storage. Appl. Clay Sci. 2022, 229, 106678. [Google Scholar] [CrossRef]
  11. Hao, X.; Yang, S.; Tao, E.; Li, Y. High Efficiency and Selective Removal of Cu(Ⅱ) via Regulating the Pore Size of Graphene Oxide/Montmorillonite Composite Aerogel. J. Hazard. Mater. 2022, 424, 127680. [Google Scholar] [CrossRef] [PubMed]
  12. Leyva-García, S.; Lozano-Castelló, D.; Morallón, E.; Cazorla-Amorós, D. Silica-Templated Ordered Mesoporous Carbon Thin Films as Electrodes for Micro-Capacitors. J. Mater. Chem. A 2016, 4, 4570–4579. [Google Scholar] [CrossRef]
  13. Tang, Y.; Zhao, X.; Li, D.; Zuo, X.; Tang, A.; Yang, H. Nano-Porous Carbon-Enabled Composite Phase Change Materials with High Photo-Thermal Conversion Performance for Multi-Function Coating. Sol. Energy Mater. Sol. Cells 2022, 248, 112025. [Google Scholar] [CrossRef]
  14. Wei, J.; Tu, C.; Yuan, G.; Bi, D.; Xiao, L.; Theng, B.K.G.; Wang, H.; Ok, Y.S. Carbon-Coated Montmorillonite Nanocomposite for the Removal of Chromium(VI) from Aqueous Solutions. J. Hazard. Mater. 2019, 368, 541–549. [Google Scholar] [CrossRef]
  15. Tao, Y.; Dai, Y.; Zhang, Z.; Geng, M.; Liu, F.; Na, H.; Zhu, J. Formation of Hydroxyl-Rich Carbon Layer Coated Silica Microspheres and Its Application to Enhance Hydrolysis of Cellulose to Sugar. Carbon 2023, 202, 276–285. [Google Scholar] [CrossRef]
  16. Wang, X.; Hu, Z.; Xiao, Y.; Du, G.; Wang, P. Discovery of a natural micro nano silicon carbon deposit deposited by planting silica. Geol. Rev. 2021, 67, 1829–1837. [Google Scholar] [CrossRef]
  17. Tang, S.; Shuai, H.; Zhao, R.; Du, G.; Wang, X.; Wang, J. Process Mineralogy of Micro/Nano Silicon-Carbon Ore Obtained from Jiangxi, China. Minerals 2022, 12, 700. [Google Scholar] [CrossRef]
  18. Venetos, M.C.; Dwaraknath, S.; Persson, K.A. Effective Local Geometry Descriptor For 29Si NMR Q4 Anisotropy. J. Phys. Chem. C 2021, 125, 19481–19488. [Google Scholar] [CrossRef]
  19. Bradford, H.; Ryder, A.; Henderson, J.; Titman, J.J. Structure of Ancient Glass by 29Si Magic Angle Spinning NMR Spectroscopy. Chem.—Eur. J. 2018, 24, 7474–7479. [Google Scholar] [CrossRef] [PubMed]
  20. Dorn, R.W.; Marro, E.A.; Hanrahan, M.P.; Klausen, R.S.; Rossini, A.J. Investigating the Microstructure of Poly (Cyclosilane) by 29Si Solid-State NMR Spectroscopy and DFT Calculations. Chem. Mater. 2019, 31, 9168–9178. [Google Scholar] [CrossRef]
  21. Cui, J.; Cheng, F.; Lin, J.; Yang, J.; Jiang, K.; Wen, Z.; Sun, J. High Surface Area C/SiO2 Composites from Rice Husks as a High-Performance Anode for Lithium Ion Batteries. Powder Technol. 2017, 311, 1–8. [Google Scholar] [CrossRef]
  22. Huang, H.; Liu, J.; Liu, H.; Evrendilek, F.; Buyukada, M. Pyrolysis of Water Hyacinth Biomass Parts: Bioenergy, Gas Emissions, and by-Products Using TG-FTIR and Py-GC/MS Analyses. Energy Convers. Manag. 2020, 207, 112552. [Google Scholar] [CrossRef]
  23. Fernandez, R.; Martirena, F.; Scrivener, K.L. The Origin of the Pozzolanic Activity of Calcined Clay Minerals: A Comparison between Kaolinite, Illite and Montmorillonite. Cem. Concr. Res. 2011, 41, 113–122. [Google Scholar] [CrossRef]
  24. Mermoux, M.; Rousseau, A. FTIR and 13C NMR study of graphite oxide. Carbon 1991, 29, 469–474. [Google Scholar] [CrossRef]
  25. Farias, T.M.B.; Gennari, R.F.; Chubaci, J.F.D.; Watanabe, S. FTIR Spectra and TL Properties of Quartz Annealed at High Temperatures. Phys. Procedia 2009, 2, 493–496. [Google Scholar] [CrossRef]
  26. Ellerbrock, R.; Stein, M.; Schaller, J. Comparing Amorphous Silica, Short-Range-Ordered Silicates and Silicic Acid Species by FTIR. Sci. Rep. 2022, 12, 11708. [Google Scholar] [CrossRef] [PubMed]
  27. Gong, Y.; Li, D.; Luo, C.; Fu, Q.; Pan, C. Highly Porous Graphitic Biomass Carbon as Advanced Electrode Materials for Supercapacitors. Green Chem. 2017, 19, 4132–4140. [Google Scholar] [CrossRef]
  28. Lin, Q.; Zhang, J.; Kong, D.; Cao, T.; Zhang, S.W.; Chen, X.; Tao, Y.; Lv, W.; Kang, F.; Yang, Q.H. Deactivating Defects in Graphenes with Al2O3 Nanoclusters to Produce Long-Life and High-Rate Sodium-Ion Batteries. Adv. Energy Mater. 2019, 9, 1803078. [Google Scholar] [CrossRef]
  29. Tang, C.; Zhang, Q. Nanocarbon for Oxygen Reduction Electrocatalysis: Dopants, Edges, and Defects. Adv. Mater. 2017, 29, 1604103. [Google Scholar] [CrossRef]
  30. Yang, Y.; Tian, C.; Wang, J.; Sun, L.; Shi, K.; Zhou, W.; Fu, H. Facile Synthesis of Novel 3D Nanoflower-like CuxO/Multilayer Graphene Composites for Room Temperature NOx Gas Sensor Application. Nanoscale 2014, 6, 7369–7378. [Google Scholar] [CrossRef]
  31. Atinafu, D.G.; Dong, W.; Wang, C.; Wang, G. Synthesis of Porous Carbon from Cotton Using an Mg(OH)2 Template for Form-Stabilized Phase Change Materials with High Encapsulation Capacity, Transition Enthalpy and Reliability. J. Mater. Chem. A 2018, 6, 8969–8977. [Google Scholar] [CrossRef]
  32. Campos-Roldán, C.A.; Ramos-Sánchez, G.; Gonzalez-Huerta, R.G.; Vargas García, J.R.; Balbuena, P.B.; Alonso-Vante, N. Influence of Sp3-Sp2 Carbon Nanodomains on Metal/Support Interaction, Catalyst Durability, and Catalytic Activity for the Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2016, 8, 23260–23269. [Google Scholar] [CrossRef] [PubMed]
  33. De vos Burchart, E.; van Bekkum, H.; van de Graaf, B. Molecular Mechanics Study on the α-Quartz/β-Quartz Transition. J. Chem. Soc. Faraday Trans. 1992, 88, 1161–1164. [Google Scholar] [CrossRef]
  34. de Mendonça, F.G.; da Cunha, I.T.; Soares, R.R.; Tristão, J.C.; Lago, R.M. Tuning the Surface Properties of Biochar by Thermal Treatment. Bioresour. Technol. 2017, 246, 28–33. [Google Scholar] [CrossRef] [PubMed]
  35. Xiao, F.; Bedane, A.H.; Zhao, J.X.; Mann, M.D.; Pignatello, J.J. Thermal Air Oxidation Changes Surface and Adsorptive Properties of Black Carbon (Char/Biochar). Sci. Total Environ. 2018, 618, 276–283. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, B.; Guo, Y.; Khvesyuk, V.I.; Barinov, A.A.; Wang, M. Heat Conduction of Multilayer Nanostructures with Consideration of Coherent and Incoherent Phonon Transport. Nano Res. 2022, 15, 9492–9497. [Google Scholar] [CrossRef]
  37. Luckyanova, M.N.; Mendoza, J.; Lu, H.; Song, B.; Huang, S.; Zhou, J.; Li, M.; Dong, Y.; Zhou, H.; Garlow, J.; et al. Phonon Lo-calization in Heat Conduction. Sci. Adv. 2018, 4, 9460. [Google Scholar] [CrossRef]
  38. Ohlberge, S.M.; Strickler, D.W. Determination of Percent Crystallinity of Partly Devitrified Glass by X-ray Diffraction. J. Am. Ceram. Soc. 1962, 45, 170–171. [Google Scholar] [CrossRef]
  39. Radwan, O.A.; Humphrey, J.D.; Al-Ramadan, K.A. Quartz Crystallinity Index of Arabian Sands and Sandstones. Earth Sp. Sci. 2021, 8, e2020EA001582. [Google Scholar] [CrossRef]
  40. Marinoni, N.; Broekmans, M.A.T.M. Microstructure of Selected Aggregate Quartz by XRD, and a Critical Review of the Crystallinity Index. Cem. Concr. Res. 2013, 54, 215–225. [Google Scholar] [CrossRef]
  41. Aly, K.A.; Khalil, N.M.; Algamal, Y.; Saleem, Q.M.A. Lattice Strain Estimation for CoAl2O4 Nano Particles Using Williamson-Hall Analysis. J. Alloys Compd. 2016, 676, 606–612. [Google Scholar] [CrossRef]
  42. Vargas-Becerril, N.; Reyes-Gasga, J.; García-García, R. Evaluation of Crystalline Indexes Obtained through Infrared Spectroscopy and X-ray Diffraction in Thermally Treated Human Tooth Samples. Mater. Sci. Eng. C 2019, 97, 644–649. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) In-situ XRD patterns, (b) in-situ XRD peak at around 26.5°, (c) 29Si MAS NMR spectra of the QZC, and calcined samples at different temperatures, (d) βcosθ~sinθ curves; FTIR spectra of the QZC calcined at different temperatures (e) under air and (f) under N2 atmospheres.
Figure 1. (a) In-situ XRD patterns, (b) in-situ XRD peak at around 26.5°, (c) 29Si MAS NMR spectra of the QZC, and calcined samples at different temperatures, (d) βcosθ~sinθ curves; FTIR spectra of the QZC calcined at different temperatures (e) under air and (f) under N2 atmospheres.
Minerals 13 01488 g001
Figure 2. (a) Raman spectra, (b) 13C MAS NMR spectra of the QZC and calcined samples at different temperatures, (c) XPS survey results of the QZC; high-resolution spectra of C1s in the XPS analysis of (d) QZC, (e) QZC 600-N, and (f) QZC 800-N.
Figure 2. (a) Raman spectra, (b) 13C MAS NMR spectra of the QZC and calcined samples at different temperatures, (c) XPS survey results of the QZC; high-resolution spectra of C1s in the XPS analysis of (d) QZC, (e) QZC 600-N, and (f) QZC 800-N.
Minerals 13 01488 g002
Figure 3. SEM images of the (a) QZC, (b) QZC400-A, (c) QZC500-A, (d) QZC600-A, (e) QZC800-A, (f) QZC1000-A; the mapping of (g) QZC and (h) QZC 1000-A samples.
Figure 3. SEM images of the (a) QZC, (b) QZC400-A, (c) QZC500-A, (d) QZC600-A, (e) QZC800-A, (f) QZC1000-A; the mapping of (g) QZC and (h) QZC 1000-A samples.
Minerals 13 01488 g003
Figure 4. TEM images of the (a) QZC; TEM and HRTEM images of (b) QZC, (c) QZC400-A, (d) QZC600-A, (e) QZC800-A, (f) QZC1000-A.
Figure 4. TEM images of the (a) QZC; TEM and HRTEM images of (b) QZC, (c) QZC400-A, (d) QZC600-A, (e) QZC800-A, (f) QZC1000-A.
Minerals 13 01488 g004
Figure 5. TG-DSC curves of pristine QZC (a) under air and (b) under N2 atmospheres; BJH pore size distribution curves of pristine QZC and calcined samples at different temperatures (c) under air and (d) under N2 atmospheres; (e) resistivity of the sample; (f) thermal conductivities of QZC, QZC600-A, and QZC1000-A; zeta potential of pristine QZC and calcined samples at different temperatures (g) under air and (h) under N2 atmospheres.
Figure 5. TG-DSC curves of pristine QZC (a) under air and (b) under N2 atmospheres; BJH pore size distribution curves of pristine QZC and calcined samples at different temperatures (c) under air and (d) under N2 atmospheres; (e) resistivity of the sample; (f) thermal conductivities of QZC, QZC600-A, and QZC1000-A; zeta potential of pristine QZC and calcined samples at different temperatures (g) under air and (h) under N2 atmospheres.
Minerals 13 01488 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, X.; Zhao, X.; Wang, X.; Tang, Y.; Liao, J.; Wu, Q.; Wang, J.; Zhang, J.; Yang, H. Insight into Crystalline Structure and Physicochemical Properties of Quartz-Carbon Ore. Minerals 2023, 13, 1488. https://doi.org/10.3390/min13121488

AMA Style

Liu X, Zhao X, Wang X, Tang Y, Liao J, Wu Q, Wang J, Zhang J, Yang H. Insight into Crystalline Structure and Physicochemical Properties of Quartz-Carbon Ore. Minerals. 2023; 13(12):1488. https://doi.org/10.3390/min13121488

Chicago/Turabian Style

Liu, Xi, Xiaoguang Zhao, Xianguang Wang, Yili Tang, Juan Liao, Qianwen Wu, Jie Wang, Jun Zhang, and Huaming Yang. 2023. "Insight into Crystalline Structure and Physicochemical Properties of Quartz-Carbon Ore" Minerals 13, no. 12: 1488. https://doi.org/10.3390/min13121488

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