Next Article in Journal
A New Collector for Effectively Increasing Recovery in Copper Oxide Ore-Staged Flotation
Next Article in Special Issue
A Novel Method for Improving Low-Temperature Flotation Performance of Nonpolar Oil in the Molybdenite Flotation
Previous Article in Journal
Numerical Simulation Based Targeting of the Magushan Skarn Cu–Mo Deposit, Middle-Lower Yangtze Metallogenic Belt, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comparison of Mineralogical and Thermal Storage Characteristics for Two Types of Stone Coal

1
School of Energy and Power Engineering, Changsha University of Science and Technology, Changsha 410114, China
2
School of Chemical Engineering, North China University of Science and Technology, Tangshan 063009, China
*
Authors to whom correspondence should be addressed.
Minerals 2019, 9(10), 594; https://doi.org/10.3390/min9100594
Submission received: 29 August 2019 / Revised: 18 September 2019 / Accepted: 25 September 2019 / Published: 29 September 2019

Abstract

:
Understanding the mineralogical and chemical characteristics of stone coal is imperative for exploring the thermal storage characteristics of stone coal. Two types stone coal (SCwt and SChc) were sampled from the pond of a stone coal mine in Henan province (China), and the chemical composition, phase composition, and thermal behavior of the stone coal were investigated. Furthermore, the petrography of the stone coal was studied in detail. The mineral phases of the stone coal were quartz, kaolinite, roscoelite, and goethite, as distinguished by reflected light microscopy and further proven by scanning electron microscope-energy dispersive spectrometer (SEM-EDS). The thermal conductivity of SCwt was 0.19 W m−1 K−1, while that of SChc was 0.24 W m−1 K−1. Stearic acid (SA) was blended with SCwt and SChc to prepare SA/SCwt and SA/SChc composites via an impregnation method, respectively. The thermogravimetric (TG) curves show that the loading capacity of SChc, at 17.40%, is higher than that of SCwt (16.63%). The thermal energy storage capacities of SA/SCwt and SA/SChc composites were 29.21 J g−1 and 33.02 J g−1 according to a differential scanning calorimetry (DSC) analysis. Therefore, SChc is a potential candidate for thermal storage applications due to more obvious thermal storage characteristics, including higher thermal conductivity and loading capacity.

1. Introduction

Nowadays, energy and environmental issues are two major challenges of society, and improving energy utilization efficiency and protecting the environment have received widespread attention [1,2,3,4,5,6]. It is obvious that inexpensive energy, which is a prerequisite for sustainable economic growth and development, can be produced from coal [7]. Some trace elements in coal also have significant potential for industrial extraction and utilization [8]. With the increasing market demand for high-performance materials and the shortage of high-grade ores, it becomes more and more necessary and urgent to exploit and utilize low-grade resources for high-performance materials [9]. Recently, utilizing low-grade resources to produce high-performance materials and high purity metals has become a research hotspot [10,11,12,13,14].
Stone coal is a combustible, low heat value, high-rank sedimentary rock mainly derived from early Paleozoic bacteria and algae after saponification and coalification in a marine-influenced environment (e.g., epicontinental sea, lagoon, or bay) [15]. It is mainly distributed in southern China, where coal resources are scarce, and the reserves are extremely rich. The total reserves of stone coal are 61.88 Gt, covering an area of about 30,000 km2 [15]. Because of the simple mining conditions of stone coal and the shortage of coal resources, stone coal resources will be widely used. In recent years, the comprehensive utilization of stone coal resources has developed rapidly [16]. The utilization of stone coal is mainly in four aspects: fuel power (stone coal power generation), building materials, agricultural fertilizer, and refined rare metals [15]. While making full use of the low calorific energy value of stone coal, research on the extraction of vanadium and other associated elements can not only improve the utilization value of stone coal but also provide more urgently needed rare and precious metal materials [16]. Stone coal after the extraction of vanadium also has the potential for producing composite phase change materials. Therefore, the study of the mineralogical characteristics of stone coal is very important. So far, very little information related to the composite phase change materials that use stone coal is available, and only research in vanadium extraction from stone coal has been conducted. Stone coal is now used to extract beneficial elements, mainly small-scale extraction of vanadium pentoxide. Vanadium is an important rare metal that has been widely used in ferrous and non-ferrous alloys to improve its hardness, tensile strength, and fatigue resistance, and vanadium oxide is widely applied [17,18,19]. In China, apart from vanadium titano-magnetite, carbonaceous shale, also called stone coal, is another important vanadium-bearing resource due to the vast amount of deposits [20]. The gross reserve of vanadium in terms of V2O5 in stone coal accounts for more than 87% of the domestic reserve of vanadium [21]. Generally, most vanadium in stone coal exists in terms of isomorphism, where V3+ substitutes for Al3+ in the crystal lattice of aluminosilicates including muscovite, tourmaline, and garnet, because V3+ has a similar electronegativity and ionic radius to Al3+ [22,23]. In addition, small amounts of vanadium in the form of adsorbed state adsorb on the surface of clay minerals such as kaolinite [24]. After the extraction of vanadium, if research on stone coal after extracting vanadium continues, the utilization of stone coal resources will be fully, effectively, and reasonably developed. Due to the complex mineral composition of stone coal, in order to fully exploit and utilize abundant stone coal resources and guide the preparation of composite materials using stone coal as a raw material, it is necessary to thoroughly explore the mineralogical and thermal storage characteristics of stone coal.
In recent years, mineralogical characteristics have been used to study stone coal. Researchers have studied stone coal by various characterization methods so as to explore the nature of minerals, extract vanadium, and prepare materials. To explore the flotation technology of high-carbon stone coal bearing vanadium, Wang et al. [9] studied the mineralogical characteristics of stone coal by XRD, scanning electron microscope-energy dispersive spectrometer (SEM-EDS), and optical micrographs. To optimize the roasting method and conditions to improve the leaching efficiency of vanadium pentoxide (V2O5), Zeng et al. [25] investigated the oxidizing roasting of stone coal in a laboratory fluidized bed reactor by means of X-ray fluorescence (XRF), SEM-EDS, and XRD. Furthermore, Hu et al. [26] explored the separation of V over Fe impurities from V-bearing shale using oxalic acid as an eco-friendly leachant by XRF, XRD, and SEM-EDS. Xie et al. [16] also explored activated carbon (AC), vanadium oxide (V2O5), and V2O5/AC photocatalyst prepared by in situ utilization of stone coal ore by XRD, XRF, Fourier transformation infrared spectroscopy (FTIR), SEM, Brunauer-Emmett-Teller (BET), and thermogravimetric-differential thermal analyzer (TG-DTA). In addition, Gao et al. [27,28,29,30] explored the nature of minerals by calculating surface broken bonds combined with mineralogical features. Therefore, the mineralogical characteristics of stone coal can serve as a guide for subsequent work on stone coal. However, the use of stone coal as a raw material for the preparation of phase change composites also requires research on its thermal storage characteristics.
As a natural mineral, stone coal has the potential to be used to prepare composite thermal storage materials. However, few studied related to stone coal-based composite thermal storage materials have been reported. Some scholars have studied a composite of coal-series kaolin (coal gangue) and fatty acid as form-stable phase change material [31,32]. Stone coal with a lower carbonization degree has similar properties to coal gangue [33], so it is meaningful to study the thermal storage characteristics of stone coal. In this paper, mineralogical and thermal storage studies of stone coal were conducted with the purpose of preparing thermal storage materials by the comparison of the mineralogical and thermal storage characteristics of two types of stone coal.

2. Materials and Methods

2.1. Materials

Two kinds of stone coal (weathered stone coal and high-carbon stone coal) were sampled from the pond of a stone coal mine in Henan, China. After being transported to the laboratory, these stone coal samples were air-dried, ground to pass through a 0.074 mm metal sieve, homogenized, and stored in polypropylene plastic sample bags. The weathered stone coal and high-carbon stone coal were labeled as SCwt and SChc, respectively.

2.2. Preparation of the SA/SC Composites

The SA/SC composites were synthesized via vacuum impregnation [34,35], and stearic acid (CH3(CH2)16COOH, SA) with a phase change temperature range of 56–69.9 °C was supplied by Tianjin Hengxing Chemical Reagent Co., Ltd., Tianjin, China. The following is the fabrication process: 10 g SA (an excess) and 5.0 g SC were put into an Erlenmeyer flask, which was concatenated to a suction pump by using a unit for preventing backward suction. A vacuum was created in the Erlenmeyer flask at −0.1 MPa for 5 min, and then its temperature was maintained at 95 °C for 30 min via a constant temperature bath. The vacuum was released and the flask heated to 80 °C for 5 min in an ultrasonic bath before cooling to room temperature. The resulting mixture was filtered at 80 °C to produce the final stearic acid/stone coal (SA/SC) composites. The composites were labeled as SA/SCwt and SA/SChc, respectively.

2.3. Characterization

The elemental composition of raw stone coals (SCwt and SChc) was determined by X-ray fluorescence (XRF, PANalytical B.V., Almelo, The Netherlands) using radiation at an acceleration voltage of 60 kV and a current of 100 mA. X-ray diffraction (XRD, Bruker Corporation, Billerica, MA, USA) was carried out using a D8 ADVANCE analyzer (Cu-Kα) under the following conditions: voltage 40 kV, current 250 mA, scan range from 3 to 80°, and step size of 0.5°. The crystalline phases of samples were identified using the software Jade 6.5 compiled by Materials Data Inc. (MDI, California, CA, USA). The stone coal was embedded in the epoxy resin and polished for the petrography analysis, and the photomicrographs were then obtained using an Axiovert 200 mat photomicroscope; furthermore, a scanning electron microscope (SEM)-energy dispersive spectrometer (EDS, JEOL, Tokyo, Japan) was used on the polished samples under the backscattered electron (BSE, JEOL, Tokyo, Japan) mode by JEOL JSM-6490LV. The samples were pressed onto a conductive adhesive tab mounted on a copper stub and coated with gold before the test, and the corresponding microstructures were investigated using a Zeiss Sigma 500 scanning electron microscope (SEM, Carl Zeiss AG, Oberkochen, Germany) operated at an accelerating voltage. Fourier transformation infrared spectroscopy (FTIR, SHIMADZU, Japan) spectra were recorded using an IR tracer-100AH FTIR spectrometer in the range of 4000~400 cm−1 at 20 kV. A thermogravimetric analysis (TGA, HENVEN HCT-3, Beijing Henven Scientific Instrument Factory, Beijing, China) was conducted at a heating rate of 15 °C·min−1 up to 1100 °C in oxygen atmosphere. The thermal conductivity of two samples pressed into tablets was recorded using a DRX-I-300. Before testing the samples, the samples had to be to be prepared into wafers with a thickness of about 1 mm and a diameter of 12 mm. Differential scanning calorimetry (DSC) analysis of stearic acid and the composites was performed using a TA instruments DSC Q10 (V9.9 Build 303, New Castle, DE, USA) at a heating rate of 5 °C/min under a constant stream of argon at atmospheric pressure. Liquid nitrogen was used to cool the sample during the freezing period.

3. Results

3.1. Elemental Composition

The main components of SCwt were Si, Al, and Fe, while the content of Ca, K, Ba, V, and P was less (Table 1). The main components of SChc were Si, Al, and Fe, while the content of Cr, K, Ba, V, and P was less. The content of V was 8182.42 ppm and 5479.98 ppm for SCwt and SChc, respectively, indicating that the SCwt includes more vanadium. Other valuable elements, such as Cu, Zn, Ni, and Mo, are low in the mineral composition, making little sense if recycled.

3.2. Mineralogical Composition

Figure 1 shows the XRD patterns of the SCwt and SChc, indicating the respective phase composition such as quartz, kaolinite, roscoelite, and goethite [36]. It can clearly be seen from the XRD curve that the phase of the two stone coals is similar. The major phase of SCwt and SChc was quartz. Combined with the results of the XRF, Fe, Ca, and Si in SCwt were slightly more than in SChc. The content of V in SCwt is slightly greater than in SChc, whereas the content of Al and K in SCwt is slightly lower than in SChc. This may be because the goethite in SCwt contains a small amount of vanadium [37]. Vanadium seems to be affiliated mainly with roscoelite.

3.3. FTIR Analysis

The FTIR spectra of samples are shown in Figure 2. The absorption at 3442 cm−1 and 2359 cm−1 in all samples was attributed to loosely bound water (ν(H–O–H)) [38] and Al–OH [39], and the band at 2924 cm−1 in all samples was crystalline water (free water) [39]. The absorption bands at 1730 cm−1 and 1647 cm−1 were bending vibrations from adsorbed water in roscoelite [39]. The bands at 1456 cm−1 in all samples were attributed to stretching vibrations from CO32− in carbonate impurities [39,40]. The band at 1377 cm−1 in all samples was attributed to symmetric stretching vibrations from adsorbed CO2 [41]. The bands at 1094 cm−1 in SCwt and 1089 cm−1 in SChc were attributed to antisymmetric stretching vibrations from Si–O–Si in roscoelite and quartz [42]. The bands at 692 cm−1 and 798 cm−1 in all samples were attributed to symmetric stretching vibrations from Si–O–Si in quartz [42]. The band at 515 cm−1 in all samples was attributed to Si–O–AlVI bending [43]. The bands at 469 cm−1 in all samples were attributed to Si–O and Si–O–Si bending [39].

3.4. TG-DTA Analysis

The TG and DTA curves of SCwt under the flow of O2 are shown in Figure 3a. According to the TG curve, free water was released in the temperature range of 15–100 °C with a mass loss of 1.28%. Dehydration of lattice water occurred in the range of 100–689 °C with a mass loss of 4.82%. From 689 °C to 879 °C, the weight loss was relatively stable, with a mass loss of 0.23%. From 879 °C to 1140 °C, the weight loss maintained a rising trend with a mass loss of 1.66% due to the removal of the structural hydroxyl groups from aluminosilicate minerals [41]. The TG curve of SCwt has no obvious weightlessness step, while the corresponding DTA curve is a smooth curve without an obvious endothermic peak. Figure 3b shows the TG and DTA curves of SChc under the flow of O2. According to the TG curve, free water was released in the temperature range of 15–100 °C with a mass loss of 2.74%. The lattice water and structural hydroxyls (part) were released in 100–436 °C with a mass loss of 4.00%. From 436 to 600 °C, the weight loss increased seriously, with a mass loss of 22.44% due to the carbon oxidized and the dehydration of lattice water [44]. From 600 to 1130 °C, the weight loss maintained the rising trend, with a mass loss of 3.13% due to the removal of the structural hydroxyl groups from aluminosilicate minerals [41]. The TG curve of SChc has an obvious weightlessness step, and in the DTA curve, an endothermic peak above 500 °C suggests that the endothermic enthalpies of SChc increased. There were no obvious chemical reactions in SCwt under O2 atmosphere at room temperature to 1140 °C. This indicates that the SCwt had good thermal stability.

3.5. Reflected Light Microscopy Images Analysis

The study of petrography is useful for the identification of rocks, minerals, and ores [38,45]. The reflected light microscopy images demonstrated that four types of mineral phases were easily distinguished in two samples (Figure 4). The optical characteristic of roscoelite is that the ⊥ Bxa interferogram can be seen on the (001) plane, and the irregular roscoelite particles were surrounded by quartz. According to the optical characteristic of roscoelite, the dark roscoelite contains Fe impurities or carbonaceous impurities (Figure 4b,c). Goethite was distributed in kaolinite in a wide range of sizes. Kaolinite is mainly distributed along the bedding planes and, to a lesser extent, occurs as disseminated fine particles (Figure 4a,d) [8]. Therefore, the mineralography compositions of these samples were quartz, kaolinite, goethite, and roscoelite. However, the amount and size of roscoelite in SCwt was slightly larger than in SChc, and the brownish-yellow area was also significantly larger than in SChc. These results were in accord with the XRF and XRD results.

3.6. SEM-EDS Analysis

The SEM-EDS analysis of the stone coal samples is presented in Figure 5, and atomic composition contrasts are also identified. Figure 5a,c shows that the particles were well separated from each other and that the shape of these particles is irregular. The stone coal samples presented flaky and pellet structures; the size of the flaky structures was 1–2 μm, while the size of the pellet structures was 0.1–1 μm. It is revealed that flaky structures were interspersed throughout the pellet structures. According to the EDS results (Figure 5b,d), the V element emerged in both SCwt and SChc. These results are in accord with the reflected light microscopy images and XRF results of the stone coal.
Secondary electron (SE) images and EDS elemental maps of SCwt and SChc are shown in Figure 6 and Figure 7, respectively. These indicate that Si (Figure 6b and Figure 7b), O (Figure 6c and Figure 7c), and Al (Figure 6d and Figure 7d) emerged in the concentrated distribution, and Fe (Figure 6e and Figure 7e) and V (Figure 6f and Figure 7f) presented the uniform distribution. Si and O mainly originated from quartz, and Al was from aluminosilicate (kaolinite and roscoelite). Fe and V were geared to goethite and roscoelite, respectively. The elemental mapping diagrams of the two stone coal samples are similar, indicating that their elemental composition and phase distributions are similar. The contents of vanadium minerals are usually very low and are commonly detrital materials of terrigenous origin and, to a lesser extent, from epigenetic igneous hydrothermal solutions [15]. Vanadium and roscoelite have a relatively obvious synchronous distribution. Most of the vanadium in stone coal occurs in ground roscoelite. The presence of V leads to the possibility of having roscoelite [7]. This suggests that roscoelite was formed by a complex reaction involving a change in the oxidation–reduction potential of the groundwater and the presence in the sandstone of sensitive associations of minerals containing vanadium-rich iron and titanium oxides. The reduction spots formed by diffusion as part of the same process [46]. The occurrence of roscoelite in depositional environment indicates input from hydrothermal fluids [15]. The roscoelite is of authigenic origin and was probably derived from the interaction between V derived from penetrating V-rich solutions and kaolinite already present in the coal during an early stage of the diagenetic process [47]. The sediment-source region for the coal was mainly due to Mesozoic intermediate-felsic volcanic rocks [8]. These results were in accord with the XRF results of the stone coal.
The EDS of the SC samples under the backscattered electron (BSE) mode is presented in Figure 8, and atomic composition contrasts (Z-contrast) are also identified. The original EDS data are provided in the Supplementary Materials. Four different contrasts can be seen in Figure 8a and are labeled as A1, A2, A3, and A4, respectively. According to the EDS results (Figure 8b), the A1, A2, A3, and A3 were quartz (dark-gray large grain), kaolinite (dark-gray small flake), roscoelite (light gray), and goethite (bright gray), respectively, in the SCwt. This may be because the primary particles of SChc are too small, and the color between the phases is not much different. But relying on the EDS analysis, quartz, kaolinite, and roscoelite were also discovered in the SChc (Figure 8c,d) and labeled as C1, C2, and C3, respectively. Moreover, quartz, kaolinite, and goethite emerged in Figure 8e, and the corresponding EDS is presented in Figure 8f (E1, E2, and E3). These results were in accord with the reflected light microscopy images and XRF results of the stone coal. In addition, relatively little roscoelite, kaolinite, and goethite were detected according to the BSE and EDS. This is also the reason why the characteristic peaks of kaolinite, goethite, and roscoelite are not obvious in the XRD results. Nevertheless, the optical properties of the minerals in the reflected light microscopy and the results of the EDS data and their elemental composition imply the possible presence of kaolinite, goethite, and roscoelite.

3.7. The Thermal Conductivity of Stone Coal

Thermal conductivity is a fundamental factor for composite thermal storage material. The thermal conductivity of SCwt is 0.19 W m−1 k−1, while the thermal conductivity of SChc is 0.24 W m−1 k−1. The thermal conductivity of SChc is slightly greater due to the sample having a high carbon content. It indicates that the SChc is more suitable for use as a composite thermal storage material. According to the lattice theory, it is known that stone coal contains a large amount of silicate, and the Si–O bond is the most powerful in the composition, so the crystal vibration of the grid is the most efficient in this direction. According to the composition analysis of stone coal, the content of stone coal is complicated and the Si–O bond in the unit cell increases, which will increase the vibration coupling between the cells, and the phonon collision will increase, thus improving the thermal resistance. According to the composition analysis, the metal ion composition in stone coal is also very different. However, the metal ion generally does not participate in the phonon conduction, but the vibration of the Si–O bond skeleton also partially excites the metal ion vibration, which also increases the difficulty of phonon propagation through the Si–O skeleton, thereby reducing the thermal conductivities of stone coal [48]. The aforementioned results also provide the way for enhancing the thermal conductivity of stone coal, such as adding the carbon materials.

3.8. Thermal Stability and Loading Capacity

As shown in Figure 9a, the curves of SCwt and SChc under N2 are similar to those under O2, but are not as weightless as under O2. This was due to the release of free water and dehydration of lattice water, not an oxidizing reaction (Figure 9b). Only one single decomposition process existed in the tested temperature range of SA/SCwt, which was due to the degradation of SA. The TG curve of SA/SChc has two obvious weightlessness steps. An endothermic peak above 200 °C suggested decomposition of pure SA, and another endothermic peak in the curve of SA/SChc above 500 °C was due to the dehydration of lattice water. As can be seen in Figure 9a, the maximum mass fractions of SA in SA/SCwt and SA/SChc were 16.63% and 17.40%, respectively. Thus, the loading capacity of SChc for phase change materials is higher than that of SCwt. The complete decomposition temperatures (Tmax) for SA/SCwt and SA/SChc were in the range of 320–340 °C, indicating that the stone coal-based composite phase change materials (PCMs) had good thermal stability under 300 °C. Moreover, the TG curves and DTA curves of SCwt, SChc, SA/SCwt, and SA/SChc under the flow of N2 are shown in Figure 9.

3.9. Thermal Storage Behavior of Composite Phase Change Materials

It can be concluded that the thermal storage behavior of the composites was very close to that of the pure stearic acid (Figure 10), indicating no chemical reaction between stearic acid and stone coal during the preparation of the composites. The phase change properties were presented in Table 2. The SA had a melting phase change temperature (Tm) of 52.91 °C in the heating curve and a freezing phase change temperature (Tf) of 53.10 °C in the cooling curve (Figure 10). DSC curves show that the heating curves of the two composites exhibit a peak; the phase change of SA/SCwt was at 52.18 °C (Figure 10), while the phase change of SA/SChc was at 52.77 °C (Figure 10). Furthermore, two transitions (a liquid–solid phase transition of SA/SCwt at 50.15 °C and a liquid–solid phase transition of SA/SChc at 53.24 °C) appeared in the cooling curve. The latent heats of melting (ΔHm) and freezing (ΔHf), therefore, were calculated to be 190.2 J g−1 and 191.5 J g−1, respectively, for stearic acid. However, the latent heats of melting and freezing were 29.21 J g−1 and 30.55 J g−1 for SA/SCwt, and 33.02 J g−1 and 33.17 J g−1 for SA/SChc, respectively. These values are lower than their theoretic values (Table 2).
The decrease in the latent heats of the composites is not only attributable to the lower fraction of stearic acid within the composites but could also be related to the crystallinity of the stearic acid in the composites. Crystallization of the stearic acid in the composites may be inhibited by interactions between the stearic acid and the substrate, which would also reduce the latent heats of the composites. The crystallinity of stearic acid (Fc) and efficient energy (Eef) in the composite were calculated by [49,50,51,52,53]:
F c = Δ H c o m p o s i t e Δ H P C M β × 100 % ,
E e f = Δ H P C M × F c ,
where ΔHcomposite and ΔHPCM are the latent heats of the composites and pure stearic acid, respectively, and β represents the loading of stearic acid in the composites. The crystallinity of stearic acid in the SA/SChc (99.78%) was clearly higher than in the SA/SCwt (92.35%). The stearic acid in the SA/SChc had a higher Eef (189.78 J g−1) than the SA/SCwt (175.64 J g−1). In this experiment, the loading of stearic acid in SA/SCwt is higher than that in SA/SChc, but the latent heat of phase change in SA/SChc is higher than SA/SCwt, and the crystallinity of stearic acid in SA/SChc is also higher than in SA/SCwt. This is because the mineral composition in stone coal is complex, and the adsorption of components in SCwt may affect the stearic acid molecules, reducing the regularity of stearic acid molecules and resulting in a decrease in crystallinity and a decrease in enthalpy.

4. Conclusions

The mineralogical and thermal storage characteristics of two types of stone coal (SCwt and SChc) were investigated by XRF, XRD, FTIR, TGA-DTA, reflected light microscopy, thermal conductivity, and DSC. The content of V2O5 was 1.46 wt.% and 0.9778 wt.% for SCwt and SChc, respectively. The main mineral phase in the two samples was quartz, kaolinite, roscoelite, and goethite. The results of combining reflected light microscopy images and scanning electron microscopy of the stone coal show that most of the roscoelite in the stone coal is dispersed around the quartz. The thermal conductivity of SCwt and SChc was 0.19 W m−1 k−1 and 0.24 W m−1 k−1, respectively. The SA/SCwt and SA/SChc composites were prepared by impregnating SA into SCwt and SChc, respectively. The loading capacities of SCwt and SChc for SA were 16.63% and 17.40%, respectively. The latent heat value of SA/SChc (33.02 J g−1) was higher than that of SA/SCwt (29.21 J g−1). The composite PCMs had good thermal stability under 300 °C. Therefore, the SChc is a potential candidate for thermal storage applications due to its higher thermal conductivity and loading capacity. The results can provide a base for preparing thermal storage materials from stone coal.

Supplementary Materials

The following are available online at https://www.mdpi.com/2075-163X/9/10/594/s1, Figure S1: The original EDS data (a) and BSE (b) of A1 in Figure 8; Figure S2: The original EDS data (a) and BSE (b) of A2 in Figure 8; Figure S3: The original EDS data (a) and BSE (b) of A3 in Figure 8; Figure S4: The original EDS data (a) and BSE (b) of A4 in Figure 8; Figure S5: The original EDS data (a) and BSE (b) of C1 in Figure 8; Figure S6: The original EDS data (a) and BSE (b) of C2 in Figure 8; Figure S7: The original EDS data (a) and BSE (b) of C3 in Figure 8; Figure S8: The original EDS data (a) and BSE (b) of E1 in Figure 8; Figure S9: The original EDS data (a) and BSE (b) of E2 in Figure 8; Figure S10: The original EDS data (a) and BSE (b) of E3 in Figure 8.

Author Contributions

Conceptualization, C.L.; methodology, C.L., W.L., and Z.H.; validation, C.L.; formal analysis, C.L.; investigation, C.L. and H.M.; data curation, H.M., B.X., B.Z., X.Z., and M.W.; writing—original draft preparation, H.M.; writing—review and editing, C.L., W.L., and J.C.; supervision, C.L.; project administration, C.L.; funding acquisition, C.L.

Funding

This work was supported by the National Natural Science Foundation of China (51874047, 51504041); the Training Program for Excellent Young Innovators of Changsha (kq1802007); the Fund for University Young Core Instructors of Hunan Province; the Outstanding Youth Project of Hunan Provincial Department of Education (18B148); and the Hunan Province 2011 Collaborative Innovation Center of Clean Energy and Smart Grid.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, Z.; Li, M.; Li, Y.; Wang, L.; Zhu, J.; Meng, W.; Li, C.; Zhou, H.; Dai, L. Electrospun nitrogen-doped carbon nanofiber as negative electrode for vanadium redox flow battery. Appl. Surf. Sci. 2019, 469, 423–430. [Google Scholar] [CrossRef]
  2. Yang, S.; Zhang, M.; Wu, X.; Wu, X.; Zeng, F.; Li, Y.; Duan, S.; Fan, D.; Yang, Y.; Wu, X. The excellent electrochemical performances of ZnMn2O4/Mn2O3: The composite cathode material for potential aqueous zinc ion batteries. J. Electroanal. Chem. 2019, 832, 69–74. [Google Scholar] [CrossRef]
  3. Cui, J.; Wu, X.; Yang, S.; Li, C.; Tang, F.; Chen, J.; Chen, Y.; Xiang, Y.; Wu, X.; He, Z. Cryptomelane-Type KMn8O16 as Potential Cathode Material—for Aqueous Zinc Ion Battery. Front. Chem. 2018, 6, 352. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, Z.; Li, Y.; Zhu, J.; Li, B.; Li, C.; Wang, L.; Meng, W.; He, Z.; Dai, L. Synthesis and performance of a graphene decorated NaTi2(PO4)3/C anode for aqueous lithium-ion batteries. J. Alloys Compd. 2019, 791, 176–183. [Google Scholar] [CrossRef]
  5. Yi, Z.; Huang, J.; Cen, C.; Chen, X.; Zhou, Z.; Tang, Y.; Wang, B.; Yi, Y.; Wang, J.; Wu, P. Nanoribbon-ring cross perfect metamaterial graphene multi-band absorber in THz range and the sensing application. Results Phys. 2019, 14, 102367. [Google Scholar] [CrossRef]
  6. He, X.; Sun, Z.; Zou, Q.; Wu, L.; Jiang, J. Electrochemical Behavior of Co(II) Reduction for Preparing Nanocrystalline Co Catalyst for Hydrogen Evolution Reaction from 1-ethyl-3-methylimidazolium Bisulfate and Ethylene Glycol System. J. Electrochem. Soc. 2019, 166, D57–D64. [Google Scholar] [CrossRef]
  7. Papanicolaou, C.; Kotis, T.; Foscolos, A.; Goodarzi, F. Coals of Greece: A review of properties, uses and future perspectives. Int. J. Coal Geol. 2004, 58, 147–169. [Google Scholar] [CrossRef]
  8. Dai, S.; Guo, W.; Nechaev, V.P.; French, D.; Ward, C.R.; Spiro, B.F.; Finkelman, R.B. Modes of occurrence and origin of mineral matter in the Palaeogene coal (No. 19-2) from the Hunchun Coalfield, Jilin Province, China. Int. J. Coal Geol. 2018, 189, 94–110. [Google Scholar] [CrossRef]
  9. Wang, L.; Sun, W.; Zhang, Q.-P. Recovery of vanadium and carbon from low-grade stone coal by flotation. Trans. Nonferrous Met. Soc. China 2015, 25, 3767–3773. [Google Scholar] [CrossRef]
  10. Fang, S.; Xu, L.; Wu, H.; Tian, J.; Lu, Z.; Sun, W.; Hu, Y. Adsorption of Pb(II)/benzohydroxamic acid collector complexes for ilmenite flotation. Miner. Eng. 2018, 126, 16–23. [Google Scholar] [CrossRef]
  11. Shu, K.; Xu, L.; Wu, H.; Fang, S.; Wang, Z.; Xu, Y.; Zhang, Z. Effects of ultrasonic pre-treatment on the flotation of ilmenite and collector adsorption. Miner. Eng. 2019, 137, 124–132. [Google Scholar] [CrossRef]
  12. Huang, Z.; Cheng, C.; Liu, Z.; Zeng, H.; Feng, B.; Zhong, H.; Luo, W.; Hu, Y.; Guo, Z.; He, G.; et al. Utilization of a new Gemini surfactant as the collector for the reverse froth flotation of phosphate ore in sustainable production of phosphate fertilizer. J. Clean. Prod. 2019, 221, 108–112. [Google Scholar] [CrossRef]
  13. Wu, H.; Tian, J.; Xu, L.; Fang, S.; Zhang, Z.; Chi, R. Flotation and adsorption of a new mixed anionic/cationic collector in the spodumene-feldspar system. Miner. Eng. 2018, 127, 42–47. [Google Scholar] [CrossRef]
  14. Fang, S.; Xu, L.; Wu, H.; Shu, K.; Xu, Y.; Zhang, Z.; Chi, R.; Sun, W. Comparative studies of flotation and adsorption of Pb(II)/benzohydroxamic acid collector complexes on ilmenite and titanaugite. Powder Technol. 2019, 345, 35–42. [Google Scholar] [CrossRef]
  15. Dai, S.; Zheng, X.; Wang, X.; Finkelman, R.B.; Jiang, Y.; Ren, D.; Yan, X.; Zhou, Y. Stone coal in China: A review. Int. Geol. Rev. 2018, 60, 736–753. [Google Scholar] [CrossRef]
  16. Xie, T.; Liao, T.; Liu, C.; Xu, L.; Yang, J.; Zhu, Q.; Wang, J.; Zhang, X. Synthesis of multifunctional photocatalyst vanadium oxide/activated carbon via in situ utilization of stone coal ore. Ceram. Int. 2019, 45, 4934–4944. [Google Scholar] [CrossRef]
  17. Wang, S.; Liu, G.; Hu, P.; Zhou, Y.; Ke, Y.; Li, C.; Chen, J.; Cao, T.; Long, Y. Largely lowered transition temperature of a VO2/carbon hybrid phase change material with high thermal emissivity switching ability and near infrared regulations. Adv. Mater. Interfaces 2018. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Cen, C.; Liang, C.; Yi, Z.; Chen, X.; Li, M.; Zhou, Z.; Tang, Y.; Yi, Y.; Zhang, G. Dual-band switchable terahertz metamaterial absorber based on metal nanostructure. Results Phys. 2019, 14, 102422. [Google Scholar] [CrossRef]
  19. Lv, Y.; Zhang, L.; Cheng, G.; Wang, P.; Zhang, T.; Li, C.; Jiang, Y.; He, Z.; Dai, L.; Wang, L. Preparation of carbon nanosheet by molten salt route and it’s application in catalyzing VO2+/VO2+ redox reaction. J. Electrochem. Soc. 2019, 166, A953–A959. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Bao, S.; Liu, T.; Chen, T.; Huang, J. The technology of extracting vanadium from stone coal in China: History, current status and future prospects. Hydrometallurgy 2011, 109, 116–124. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Zhang, Y.; Bao, S.; Chen, T.; Liu, X. Effect of Stone Coal Chemical Composition on Sintering Behavior during Roasting. Ind. Eng. Chem. Res. 2013, 53, 157–163. [Google Scholar] [CrossRef]
  22. He, D.; Feng, Q.; Zhang, G.; Ou, L.; Lu, Y. An environmentally-friendly technology of vanadium extraction from stone coal. Miner. Eng. 2007, 20, 1184–1186. [Google Scholar] [CrossRef]
  23. Wang, T.; Xu, L.; Liu, C.; Zhang, Z. Calcified roasting-acid leaching process of vanadium from low-grade vanadium-containing stone coal. Chin. J. Geochem. 2014, 33, 163–167. [Google Scholar] [CrossRef]
  24. Bian, Y.; Zhang, Y.M.; Bao, S.; Zhao, Y. Analytical Method for Vanadium Occurrence State in Stone Coal and Corresponding Chemical Explanation. Min. Metall. Eng. 2013, 33, 62–70. [Google Scholar] [CrossRef]
  25. Zeng, X.; Wang, F.; Zhang, H.; Cui, L.; Yu, J.; Xu, G. Extraction of vanadium from stone coal by roasting in a fluidized bed reactor. Fuel 2015, 142, 180–188. [Google Scholar] [CrossRef]
  26. Hu, P.; Zhang, Y.; Huang, J.; Liu, T.; Yuan, Y.; Xue, N. Eco-Friendly Leaching and Separation of Vanadium over Iron Impurity from Vanadium-Bearing Shale Using Oxalic Acid as a Leachant. ACS Sustain. Chem. Eng. 2017, 6, 1900–1908. [Google Scholar] [CrossRef]
  27. Gao, Z.; Fan, R.; Ralston, J.; Sun, W.; Hu, Y. Surface broken bonds: An efficient way to assess the surface behaviour of fluorite. Miner. Eng. 2019, 130, 15–23. [Google Scholar] [CrossRef]
  28. Gao, Z.; Xie, L.; Cui, X.; Hu, Y.; Sun, W.; Zeng, H. Probing Anisotropic Surface Properties and Surface Forces of Fluorite Crystals. Langmuir 2018, 34, 2511–2521. [Google Scholar] [CrossRef]
  29. Li, C.; Gao, Z. Tune surface physicochemical property of fluorite particles by regulating the exposure degree of crystal surfaces. Min. Eng. 2018, 128, 123–132. [Google Scholar] [CrossRef]
  30. Tian, M.; Gao, Z.; Sun, W.; Han, H.; Sun, L.; Hu, Y. Activation role of lead ions in benzohydroxamic acid flotation of oxide minerals: New perspective and new practice. J. Colloid Interface Sci. 2018, 529, 150–160. [Google Scholar] [CrossRef]
  31. Liu, S.; Yang, H. Composite of Coal-Series Kaolinite and Capric–Lauric Acid as Form-Stable Phase-Change Material. Energy Technol. 2015, 3, 77–83. [Google Scholar] [CrossRef]
  32. Liu, S.; Yang, H. Stearic acid hybridizing coal–series kaolin composite phase change material for thermal energy storage. Appl. Clay Sci. 2014, 101, 277–281. [Google Scholar] [CrossRef]
  33. Lin, H.; Li, G.; Dong, Y.; Li, J. Effect of pH on the release of heavy metals from stone coal waste rocks. Int. J. Min. Process. 2017, 165, 1–7. [Google Scholar] [CrossRef]
  34. Li, C.; Xie, B.; Chen, J.; Chen, Z.; Sun, X.; Gibb, S.W. H2O2-microwave treated graphite stabilized stearic acid as a composite phase change material for thermal energy storage. RSC Adv. 2017, 7, 52486–52495. [Google Scholar] [CrossRef]
  35. Li, C.; Xie, B.; Chen, J. Graphene-decorated silica stabilized stearic acid as a thermal energy storage material. RSC Adv. 2017, 7, 30142–30151. [Google Scholar] [CrossRef] [Green Version]
  36. Sun, W.; Wang, L.; Cao, X.; Liu, R.; Xu, L.; Han, H. Flotation technology and adsorption mechanism of vanadium extraction from stone coal. Chin. J. Nonferrous Metals 2012, 22, 2069–2074. [Google Scholar]
  37. Li, M.; Liu, P.; Wang, F.; Cui, R. Study on Process Mineralogy of a Stone Coal Vanadium Ore in Hubei. Conserv. Util. Miner. Resour. 2013, 3, 41–46. [Google Scholar] [CrossRef]
  38. Li, C.; Ouyang, J.; Yang, H. Novel sensible thermal storage material from natural minerals. Phys. Chem. Miner. 2013, 40, 681–689. [Google Scholar] [CrossRef]
  39. Liu, J.; Zhang, Y.; Huang, J.; Liu, T.; Yuan, Y.; Huang, X. Influence of Mechanical Activation on Mineral Properties and Process of Acid Leaching from Stone Coal. Chin. J. Rare Met. 2014, 38. [Google Scholar] [CrossRef]
  40. Fleet, M.E.; Liu, X. Carbonate apatite type A synthesized at high pressure: New space group (P3) and orientation of channel carbonate ion. J. Solid State Chem. 2003, 174, 412–417. [Google Scholar] [CrossRef]
  41. Li, C.; Zeng, L.; Fu, H.; Chen, J.; He, J.; He, Z. Mineralogical and chemical characteristics of the lead-zinc tailing and contaminated soil from the mine tailing pond in Hunan Province (China). Physicochem. Probl. Miner. Process. 2017, 53, 1133–1147. [Google Scholar] [CrossRef]
  42. Lin, H.; Wang, X.; Dong, Y.; Xu, X.; Zhang, Y. Bioleaching of Vanadium-Bearing Stone Coal by Heterotrophic Bacteria. Chin. J. Rare Met. 2017, 41. [Google Scholar] [CrossRef]
  43. Li, C.; Fu, L.; Ouyang, J.; Tang, A.; Yang, H. Kaolinite stabilized paraffin composite phase change materials for thermal energy storage. Appl. Clay Sci. 2015, 115, 212–220. [Google Scholar] [CrossRef]
  44. Wu, H.; Wei, C.; Fan, G.; Li, M.; Deng, Z.; Ge, H. Occurrence of Vanadium in High Carbon Stone Coal and Priority of Preparation Process. J. Kunming Univ. Sci. Technol. (Sci. Technol.) 2008, 33. [Google Scholar] [CrossRef]
  45. Shutong, X.; Wen, S.; Yican, L.; Laili, J.; Shouyuan, J.; Okay, A.I.; Sengör, A.M.C. Diamond from the Dabie Shan Metamorphic Rocks and Its Implication for Tectonic Setting. Science 1992, 256, 80–82. [Google Scholar] [CrossRef]
  46. van Panhuys-Sigler, M.; Trewin, N.H.; Still, J. Roscoelite associated with reduction spots in Devonian red beds, Gamrie Bay, Banffshire. Scott. J. Geol. 1996, 32, 127–132. [Google Scholar] [CrossRef]
  47. Dai, S.; Xie, P.; Jia, S.; Ward, C.R.; Hower, J.C.; Yan, X.; French, D. Enrichment of U-Re-V-Cr-Se and rare earth elements in the Late Permian coals of the Moxinpo Coalfield, Chongqing, China: Genetic implications from geochemical and mineralogical data. Ore Geol. Rev. 2017, 80, 1–17. [Google Scholar] [CrossRef]
  48. Yang, G.; Yuan, Y.; Shen, Y.; Qu, Y. Analysis on thermal conductivity of cretaceous sandstone associated with microstructure. Coal Eng. 2015, 47, 82–85. [Google Scholar] [CrossRef]
  49. Li, C.; Xie, B.; Chen, D.; Chen, J.; Li, W.; Chen, Z.; Gibb, S.W.; Long, Y. Ultrathin graphite sheets stabilized stearic acid as a composite phase change material for thermal energy storage. Energy 2019, 166, 246–255. [Google Scholar] [CrossRef]
  50. Li, C.; Xie, B.; Chen, J.; He, Z.; Chen, Z.; Long, Y. Emerging mineral-coupled composite phase change materials for thermal energy storage. Energy Convers. Manag. 2019, 183, 633–644. [Google Scholar] [CrossRef]
  51. Li, C.; Xie, B.; He, Z.; Chen, J.; Long, Y. 3D structure fungi-derived carbon stabilized stearic acid as a composite phase change material for thermal energy storage. Renew. Energy 2019, 140, 862–873. [Google Scholar] [CrossRef]
  52. Li, C.; Zhang, B.; Xie, B.; Zhao, X.; Chen, J.; Chen, Z.; Long, Y. Stearic acid/expanded graphite as a composite phase change thermal energy storage material for tankless solar water heater. Sustain. Cities Soc. 2019, 44, 458–464. [Google Scholar] [CrossRef]
  53. Li, C.; Wang, M.; Xie, B.; Ma, H.; Chen, J. Enhanced properties of diatomite-based composite phase change materials for thermal energy storage. Renew. Energy 2020, 147, 265–274. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the stone coal (a), partial enlargement of the XRD of SCwt (b) and partial enlargement of the XRD of SChc (c).
Figure 1. XRD patterns of the stone coal (a), partial enlargement of the XRD of SCwt (b) and partial enlargement of the XRD of SChc (c).
Minerals 09 00594 g001
Figure 2. Fourier transformation infrared spectroscopy (FTIR) spectra of stone coal (SC).
Figure 2. Fourier transformation infrared spectroscopy (FTIR) spectra of stone coal (SC).
Minerals 09 00594 g002
Figure 3. Thermogravimetric analysis (TGA) (weight loss curve and differential thermal analysis (DTA) curves of (a) SCwt and (b) SChc in oxygen atmosphere.
Figure 3. Thermogravimetric analysis (TGA) (weight loss curve and differential thermal analysis (DTA) curves of (a) SCwt and (b) SChc in oxygen atmosphere.
Minerals 09 00594 g003
Figure 4. Reflected light microscopy images of SCwt (a,b) and SChc (c,d).
Figure 4. Reflected light microscopy images of SCwt (a,b) and SChc (c,d).
Minerals 09 00594 g004
Figure 5. Scanning electron microscope and energy dispersive X-ray spectroscopy of SCwt (a,b) and SChc (c,d).
Figure 5. Scanning electron microscope and energy dispersive X-ray spectroscopy of SCwt (a,b) and SChc (c,d).
Minerals 09 00594 g005
Figure 6. Secondary electron (SE) image and energy dispersive X-ray spectroscopic (EDS) elemental maps of SCwt. (a) SE image, (b) Si, (c) O, (d) Al, (e) Fe, (f) V.
Figure 6. Secondary electron (SE) image and energy dispersive X-ray spectroscopic (EDS) elemental maps of SCwt. (a) SE image, (b) Si, (c) O, (d) Al, (e) Fe, (f) V.
Minerals 09 00594 g006
Figure 7. Secondary electron (SE) image and energy dispersive X-ray spectroscopic (EDS) elemental maps of SChc. (a) SE image, (b) Si, (c) O, (d) Al, (e) Fe, (f) V.
Figure 7. Secondary electron (SE) image and energy dispersive X-ray spectroscopic (EDS) elemental maps of SChc. (a) SE image, (b) Si, (c) O, (d) Al, (e) Fe, (f) V.
Minerals 09 00594 g007
Figure 8. Backscattered electron z-contrast (BSE-Z) images and energy dispersive X-ray spectroscopy (EDS) of quartz, kaolinite, goethite, and roscoelite in SCwt (a,b) and SChc (cf). (A1: quartz; A2: kaolinite; A3: roscoelite; A4: goethite; C1: quartz; C2: kaolinite; C3: roscoelite; E1: quartz; E2: kaolinite; E3: goethite).
Figure 8. Backscattered electron z-contrast (BSE-Z) images and energy dispersive X-ray spectroscopy (EDS) of quartz, kaolinite, goethite, and roscoelite in SCwt (a,b) and SChc (cf). (A1: quartz; A2: kaolinite; A3: roscoelite; A4: goethite; C1: quartz; C2: kaolinite; C3: roscoelite; E1: quartz; E2: kaolinite; E3: goethite).
Minerals 09 00594 g008
Figure 9. Thermogravimetric (TG) (a)—differential scanning calorimetry (DSC) (b) of SCwt, SChc, SA/SCwt, and SA/SChc.
Figure 9. Thermogravimetric (TG) (a)—differential scanning calorimetry (DSC) (b) of SCwt, SChc, SA/SCwt, and SA/SChc.
Minerals 09 00594 g009
Figure 10. DSC curves of pure stearic acid (SA), SA/SCwt, and SA/SChc.
Figure 10. DSC curves of pure stearic acid (SA), SA/SCwt, and SA/SChc.
Minerals 09 00594 g010
Table 1. X-ray fluorescence (XRF) results of samples (ppm).
Table 1. X-ray fluorescence (XRF) results of samples (ppm).
NaMgAlSiPSO3ClKCa
SCwt764.19564641,876.47308,186.674366.2018,70018,0008546.8120,642.86
SChc487.452588.445,837.53243,231.82841.96567010,652.812068.57
TiVCrCuZnAsSeRbSr
SCwt13628182.42841.586161348.15143.9435.5927.43725.15
SChc1468.85479.989713.74393.695.49237.12033.83445.08
YZrMoBaPbCoMnFeNi
SCwt126.0263.6346.6730,623.53176.3715.73123.9438,745330.4
SChc57.546.6142.6720,451.5041.7711.8179.7221,886.9815.77
Table 2. Thermal properties of pure SA and SA/SC composites.
Table 2. Thermal properties of pure SA and SA/SC composites.
SampleLoadage (β, %)Melting Temperature (Tm, °C)Freezing Temperature (Tf, °C)Latent Heat of Melting (ΔHm, J g−1)Latent Heat of Freezing (ΔHf, J g−1)Theoretic Values of ΔHmHth, J g−1)Crystallinity of SA (Fc, %)Efficient Energy per Unit Mass of SA (Eef, J g−1)
SA10052.9153.10190.2191.5-100-
SA/SCwt16.6352.1850.1529.2130.5531.6392.35175.64
SA/SChc17.4052.8153.2433.0233.1733.0999.78189.78

Share and Cite

MDPI and ACS Style

Li, C.; Ma, H.; Xie, B.; Zhang, B.; Zhao, X.; Wang, M.; He, Z.; Li, W.; Chen, J. A Comparison of Mineralogical and Thermal Storage Characteristics for Two Types of Stone Coal. Minerals 2019, 9, 594. https://doi.org/10.3390/min9100594

AMA Style

Li C, Ma H, Xie B, Zhang B, Zhao X, Wang M, He Z, Li W, Chen J. A Comparison of Mineralogical and Thermal Storage Characteristics for Two Types of Stone Coal. Minerals. 2019; 9(10):594. https://doi.org/10.3390/min9100594

Chicago/Turabian Style

Li, Chuanchang, Huan Ma, Baoshan Xie, Bo Zhang, Xinbo Zhao, Mengfan Wang, Zhangxing He, Wei Li, and Jian Chen. 2019. "A Comparison of Mineralogical and Thermal Storage Characteristics for Two Types of Stone Coal" Minerals 9, no. 10: 594. https://doi.org/10.3390/min9100594

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