Next Article in Journal
Cell-Based Meat and Firms’ Environmental Strategies: New Rationales as per Available Literature
Previous Article in Journal
Harnessing the Potential of Storytelling and Mobile Technology in Intangible Cultural Heritage: A Case Study in Early Childhood Education in Sustainability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sustainable Ceramic Materials Manufactured from Ceramic Formulations Containing Quartzite and Scheelite Tailings

by
Jucielle Veras Fernandes
1,
Danyelle Garcia Guedes
1,
Fabiana Pereira da Costa
1,
Alisson Mendes Rodrigues
1,2,*,
Gelmires de Araújo Neves
1,2,
Romualdo Rodrigues Menezes
1,2 and
Lisiane Navarro de Lima Santana
1,2
1
Programa de Pós-Graduação em Ciência e Engenharia de Materiais (PPG-CEMat), Universidade Federal de Campina Grande, Av. Aprígio Veloso-882, Bodocongó, Campina Grande-PB 58429-900, Brazil
2
Unidade Acadêmica de Engenharia de Materiais, Centro de Ciências e Tecnologia, Universidade Federal de Campina Grande, Av. Aprígio Veloso-882, Bodocongó, Campina Grande-PB 58429-900, Brazil
*
Author to whom correspondence should be addressed.
Sustainability 2020, 12(22), 9417; https://doi.org/10.3390/su12229417
Submission received: 8 October 2020 / Revised: 25 October 2020 / Accepted: 26 October 2020 / Published: 12 November 2020
(This article belongs to the Section Sustainable Materials)

Abstract

:
In this study, we develop ceramic formulations based on quartzite and scheelite tailings collected from mining companies in the northeast of Brazil (Rio Grande do Norte State). New ceramic samples (27 wt% of kaolin, 29 wt% of plastic clay, 11 wt% of quartzite tailing, and 0–8 wt% scheelite tailing) were uniaxially pressed in two steps (20 MPa and 50 Mpa for 20 s); dried at 110 °C for 24 h; and sintered at 1150 °C, 1200 °C, and 1250 °C. The main mineralogical phases (mullite, quartz, calcite, and anorthite) of the sintered samples were identified using X-ray diffraction (XRD). After evaluation of the physical-mechanical properties (water absorption, linear shrinkage, apparent porosity, and flexural strength), it was observed that the incorporation of scheelite tailing by up to 8 wt% did not significantly alter the properties of samples sintered at all temperatures. Our results indicate that the new ceramics formulations developed have strong potentials in manufacturing sustainable materials such as ceramic tiles and porcelain stoneware.

1. Introduction

It is indisputable that mining activities are a significant portion of the Gross Domestic Product of several countries (such as Norway, the United States, the United Kingdom, Australia, and Brazil) and provide raw materials for industries that leverage the scientific and technological development of our society. Unfortunately, this activity can also be incredibly harmful to the environment since many tailings are generated [1,2]. Annually, it is estimated that these activities generate about 25 billion tons of solid residues, of which approximately 14 billion tons are rejects that are often improperly disposed of [3,4]. Generally, the mine tailings and water originated in the process are stored together in tanks or are improperly stored in massive residues piles near the mining companies, generating environmental impacts and damage to human health [5,6,7]. The costs associated with processing and logistics for the proper disposal of these mine tailings tend to be obstacles for mining companies. This has attracted attention from researchers interested in reusing this residue to manufacture sustainable materials [8,9,10,11].
Scheelite is a non-ferrous metallic mineral of the tungstate class (calcium tungstate-CaWO4 (CaO at 19.4% and WO3 80.6%)) that is exploited to obtain tungsten, which is widely used in the electrical, automotive, and aerospace industries, among others. Brazil has the largest scheelite deposits (Brejuí mine) in South America and is located in Rio Grande do Norte State, (Figure 1). Of all materials extracted from this mine, only 0.8% corresponds to scheelite, and 99.2% is mine tailings, which is deposited in piles or deposition basins [2,12].
Scheelite tailings have a composition similar to several traditional ceramic raw materials used to manufacture bricks, tiles, mortars, concrete, and porcelain materials, containing calcium oxide (CaO), silicon oxide (SiO2), aluminum oxide (Al2O3), and fluxing oxides [2,12,13]. Therefore, tailings generated from scheelite mining have strong potential to be used as an alternative raw material to substitution for the conventional one [14,15]. Also, due to the high production volume in the ceramic industry, it can absorb a large amount of mine tailings [16]. Several studies have addressed the ceramic industry’s viability in absorbing different mine tailings to manufacture sustainable materials. Among these studies, it is worth mentioning that red bauxite wastes from the production of alumina have been used to produce ceramic bricks [17,18,19,20] and tiles [21,22]. The granite processing waste was used in ceramic bricks [23,24], tiles [24,25], mortars [26], and ceramic membranes [27]. Gneiss rock, slate, and limestone waste were used to produce ecological bricks [28] and tiles [29,30]. However, few studies on the development of mortars [12], concrete formulations [13], and ceramic bricks [2] addressed the recycling of scheelite waste. Studies on the use of scheelite tailing for the development of triaxial ceramic bodies are scarce in the literature, despite the high amount of scheelite waste generated each year.
Quartzite has been extensively investigated as an alternative raw material to obtain mortars, porous ceramic materials, and materials in the sanitary ware industry [31,32,33,34]. The quartzite tailing contains appreciable contents of silicon oxide (SiO2), aluminum oxide (Al2O3), calcium oxide (CaO), and alkaline oxides (K2O and Na2O), which makes it a good substitute for high-value-added raw materials that are used in the ceramic industry [34,35].
In this context, the incorporation of scheelite and quartzite tailings in the ceramic industry stands out as an alternative to minimize the environmental impact caused by its inadequate disposal since it reduces the amount of residues deposited and the extraction of conventional raw materials, which are each increasingly scarce and expensive. However, the application of this type of mine tailings to the production of ceramic tiles still need to be thoroughly investigated.
Given the situations described above, the quartzite and scheelite tailings were incorporated in ceramic formulations with the objective of utilizing these potential raw materials to manufacture ceramic tile and porcelain stoneware. First, the kaolin, plastic clay, and quartzite and scheelite tailings were characterized using X-ray fluorescence, XRD, particle size distributions (laser diffraction), and thermal analytical techniques (TG and DTA). Uniaxially pressed samples (50 mm × 15 mm × 5 mm) were sintered at 1150 °C, 1200 °C, and 1250 °C for 40 min. Subsequently, their main mineralogical phases were identified, and physical-mechanical properties (linear shrinkage, water absorption, bulk density, and flexural strength module) were evaluated.

2. Materials and Methods

2.1. Raw Materials

The scheelite tailing (SR) was collected from existing mine piles of the Mineração Tomaz Salustiano S.A. company, located in the Currais Novos City, Rio Grande do Norte State (Brazil). The quartzite tailing came from the Tecquímica company, located in the Várzea City, Paraíba State (Brazil). The Kaolin (K) was supplied from Rocha Minérios, located in Juazeirinho City, Paraíba State (Brazil); plastic clay (PC) and feldspar (F) were supplied from the Armil Mineração do Nordeste, located in Parelhas City, Rio Grande do Norte State (Brazil).

2.2. Samples Preparation and Sintering Treatments

The samples investigated in this research were prepared from 27wt% of kaolin, 29wt% of plastic clay, 11wt% of quartzite tailing, and different contents of scheelite tailing. The nominal compositions of the ceramic formulations (wt%) and their respective nomenclatures are summarized in Table 1.
Before the sintering step, all raw materials were sieved (74 µm), wet-mixed (7% moisture content by wt.), and uniaxially pressed in two stages (Servitech, model CT-335, Tubarão, Brazil) in a rectangular mold (50 mm × 15 mm × 5 mm). In the first stage, called pre-pressing, a uniaxially pressing was applied at 13.5 Mpa for 10 s. The samples were then divided into two groups, where the first group was pressed at 20 Mpa and the second at 50 Mpa. In both, the pressing time was equal to 20 s. Finally, the pressed samples were dried at 110 °C for 24 h and sintered in a conventional electric oven (Flyever Equipment FE 50 RP, São Carlos, Brazil) at temperatures of 1150 °C, 1200 °C, and 1250 °C. The sintering protocol consisted of heating (30 °C·min−1) the samples from room temperature until the temperature of interest (1150 °C, 1200 °C, and 1250 °C) and performing an isothermal treatment there for 40 min. After sintering treatment, the oven was turned off, and the samples were cooled to room temperature.

2.3. Characterizations of the Raw Materials and Samples after Sintering Treatment

The chemical composition of the raw materials was measured with the aid of X-ray fluorescence (Shimadzu, model EDX 720, Kyoto, Japan) [36,37,38]. The mineralogical phases of the raw materials and the sintered samples were identified by X-ray diffraction (Shimadzu XRD 6000 model, Kyoto, Japan), with Cu X-ray wavelength = 40 Kv/30 mA), 2θ range from 5°–60°, an angular step of 0.02° step, counting time of ½ s, and JCPDS database [39,40,41,42]. The particle size distribution was determined by laser diffraction (Cilas, model 1064LD, Orléans, France) [36,37]. The thermal behaviors were evaluated by thermogravimetry (TG)/derivative thermogravimetry (DTG) (DTG Shimadzu, model TA 60H, Kyoto, Japan) and differential thermal analysis (DTA) (Instrumentec BP, model RB-3000, Campinas, Brazil), with a heating rate of 12.5 °C·7 min−1, under the air atmosphere and using calcined aluminum oxide (Al2O3) as standard [36,37,43].
The Archimedes method was used to measure water absorption (WA), apparent porosity (AP), and an apparent density (AD). Linear shrinkage (LS) and flexural tensile strength measurements were carried out using three points bending test. Each experiment was repeated 10 times for different samples. The procedures for carrying out these tests are described in other works [38,44,45].

3. Results and Discussion

3.1. Characterization of Raw Materials

Figure 2a,b shows the XRD patterns of the scheelite and quartzite tailings, respectively. The calcite (JCPDS 72-1937) was the main mineralogical phase identified in the scheelite tailing (Figure 2a). Quartz (JCPDS 46-1045), feldspar (JCPDS 09-0465), and mica (JCPDS 83-1808) were identified as minor components. The amount of potassium detected in the chemical analysis comes from mica; (Table 2). For the quartzite tailing (Figure 2b), the mineralogical phases present were mica (JCPDS 83-1808), feldspar (JCPDS 84-0710), and quartz (JCPDS 46-1045). As the quartzite tailing has quartz and feldspar, it can have a physical characteristic similar to the non-plastic raw materials used in the traditional ceramic industries.
Table 2 summarizes the chemical compositions of raw materials (kaolin, plastic clay, and feldspar) and the mine tailings (quartzite and scheelite). SiO2 (45.7% and 54.5%) and Al2O3 (39.5% and 27.5%) were the main oxides identified in the kaolin and plastic clay, respectively. These oxides commonly have origin from the structure of clay minerals and the presence of free silica [46,47]. Also, the high content of K2O (3.9%) was detected in plastic clay. The presence of K2O is important because it is a known flux and acts by lowering the sintering temperature, bringing economic gains to industrial processes. The high content levels of CaO (44.7%) and MgO (2.7%) identified in the scheelite tailing are associated with the presence of calcite, dolomite and SiO2 (20.8%); in the form of quartz as shown in Figure 2a. The presence of calcium oxide (CaO) and magnesium oxide (MgO) is essential because they reduce the refractoriness. The high fire loss measured in the scheelite tailing (15.8%) can be attributed to the thermal decomposition of calcium carbonate and the release of gases. The higher amount of K2O (12.1%) present in feldspar contributes to the formation of the vitreous phase during sintering [46]. The quartzite tailing contained SiO2 (76.5%) and Al2O3 (12.1%) as major constituents, and in smaller proportions, CaO (0.7%), MgO (1.1%) and Fe2O3 (1.5%). For the quartzite tailing, it was also observed that this residue has a silico-aluminous composition, i.e., the sum of the oxides contents SiO2 and Al2O3 is greater than 50%, (SiO2 + Al2O3 > 50%), indicating its potential as a flux agent and to lower the temperature maturation of the ceramic bodies, enabling a reduction in energy consumption [48].
Table 3 shows the particle size distribution of the kaolin, plastic clay, feldspar, and scheelite and quartzite tailings. The scheelite and quartzite tailings presented the most considerable fraction of particles with diameters above 20 μm (70.0% and 50.2% of the accumulated volume, respectively). For the other raw materials, the most considerable fraction of particles has diameters between 2 μm and 20 μm, with 69.0%, 79.9%, and 60.9% of accumulated volume for kaolin, plastic clay, and feldspar, respectively. It is known that the coarser particle size can interfere with sintering reactions kinetics, dramatically influencing the packaging of green samples and the product sintering step. The smaller particle sizes provide greater surface areas and reactivities between the particles, favoring the reactions kinetics and the diffusion process during the phase transformations [37,38,49].
Figure 3a–e shows the TG/DTG data for the kaolin, plastic clay, feldspar, and scheelite and quartzite tailings. Kaolin (Figure 3a) showed a significant loss of mass (13.6%) between 352 °C and 754 °C, probably associated with the dihydroxylation of kaolinite and mica. The TG/DTG curve obtained from the plastic clay (Figure 3b) shows that mass loss occurred in two stages. In the first stage, the mass loss was equal to 3.4% and was observed in the temperature range of 22 °C to 300 °C. The origin of the first thermal event is associated with the loss of free and adsorbed water. A mass loss of 7.3% was observed during the second stage (300–788 °C). This second stage can be attributed to the loss of organic matter and hydroxyl groups present in the clay minerals. The total mass loss for kaolin and plastic clay was 16.1% and 10.9%, respectively. Feldspar (Figure 3c) and quartzite tailing (Figure 3d) showed shallow mass loss (1.6% and 1.5%, respectively), corroborating with the fire loss analysis (see Table 2). In addition, no carbonates, sulfates, organic matter, and clay minerals were found in these materials. The TG/DTG curve of the scheelite tailing (Figure 3e) shows a significant loss of mass (13.9%) between 602 °C and 901 °C. The mass loss of the scheelite tailing is associated with the decomposition of the calcite mineral that releases CO2. The total weight loss observed for the scheelite tailing was 14.8%.
Figure 4a–e shows the differential thermal analysis (DTA) data for kaolin, plastic clay, feldspar, and tailings (scheelite and quartzite). In the DTA curve of kaolin (Figure 4a) and plastic clay (Figure 4b), two endothermic and one exothermic event were identified. The endothermic events are related to loss of free water (~92 °C for kaolin and ~108 °C for plastic clay) and loss of hydroxyls (~601 °C for kaolin and ~586 °C for plastic clay) of the octahedral layer present in its structure. The exothermic peak with a maximum of ~975 °C and 987 °C for kaolin and clay, respectively, corresponds to the mullite nucleation with β-quartz release from the amorphous structure created previously [50]. The DTA curve acquired from the quartzite tailing (Figure 4c) showed an endothermic peak at 51 °C and is related to the free water. The exothermic peak at 579 °C is associated with the polymorphic transformation of α-quartz to β-quartz. Still in the quartzite DTA curve, the exothermic peak at 949 °C is related to the structural and characteristic of the crystalline lattice’s destruction. The DTA curve of the scheelite tailing (Figure 4e) shows an exothermic peak at ~465 °C and two endothermic peaks (866 °C and 906 °C) that correspond to the carbonate decomposition.

3.2. Mineralogical Phases and Physical-Mechanical Properties of the Sintered Samples

Figure 5 shows the effect of the sintering temperature on the color of the samples. It was observed that the tones varied between light (1150 °C), yellow (1200 °C), and gray (1250 °C). The color variation was verified throughout all ceramic formulations. It was also found that ceramic formulations with a higher percentage of scheelite tailing presented colors with gray tones at higher temperatures. This color variation is observed because, at temperatures above 1000 °C, oxidation occurs from divalent iron to trivalent iron, responsible for accentuating the samples color for yellow tons. At temperatures above 1100 °C, the percentage of trivalent iron decreases, generating a dark reddish-brown to black color. This explains why there was a decrease in the yellow tone giving rise to a gray color with increased sintering temperature. The samples color is also dependent on the contents coloring oxides (such as Fe2O3, MgO, NaO, etc.). As the ceramic formulation increases, there will be a more significant variation in the sample tone.
Figure 6a,b shows the X-ray diffraction patterns of the ceramic formulations (F0, F1, F2, F3, F4, F5, F6, F7, and F8) sintered at different temperatures (1150 °C and 1250 °C). The following mineralogical phases were identified in all-ceramic formulation investigated: mullite (JCPDS 79-1275), quartz (JCPDS 46-1045), calcite (JCPDS 72-1937), and anorthite (JCPDS 09-0465). The anorthite (CaO.Al2O3.4SiO2) and mullite (Al6Si2O13) are desirable phases in ceramic materials due to their excellent mechanical properties. The increase in temperature from 1150 °C to 1250 °C led to a decrease in the quartz peaks’ intensity, indicating their partial dissolution and the disappearance of the anorthite peak.
Figure 7a–d and Figure 8a–d show the effect of the sintering temperatures (1150 °C, 1200 °C, and 1250 °C) on the physical-mechanical properties (water absorption, linear shrinkage, apparent porosity, and flexural strength) of the F0, F1, F2, F3, F4, F5, F6, and F7 samples uniaxially pressed at 20 MPa and 50 MPa. For all samples and experimental conditions investigated, the linear shrinkage (LS) (Figure 7a and Figure 8a) increased when sintering temperature increases from 1150 °C to 1200 °C. This behavior is probably related to the higher degree of sintering and densification due to the higher production of liquid phase with increasing temperature [51]. However, when the temperature increases from 1200 °C to 1250 °C, the linear shrinkage tends to decrease due to the smaller glass phase expansion. The lowest LS values were observed for ceramic formulations with higher contents of scheelite tailing, i.e., above 2wt% (F4, F5, F6, and F7 samples).
As the sintering temperature increased, the water absorption (WA) and apparent porosity (AP) decreased (Figure 7c and Figure 8c). This may be related to the pores’ filling due to the melting of the fluxing oxides (Fe2O3, K2O, and Na2O) present in the ceramic masses, which leads to greater packaging of the piece due to the formation of a liquid phase [37,38]. In summary, an increase in flexural strength (FS) is observed with an increase in temperature from 1150 °C to 1200 °C (Figure 7d and Figure 8d). This behavior is related to the reduction of porosity in the parts. This effect is desirable since the pores act as stress concentrators. With the temperature increases to 1250 °C, there is a considerable change in the samples’ mechanical behavior. The formulations that present a higher percentage of scheelite residue (above 2wt%) have their flexural strength modules reduced due to the excessive increase in the hardness of the ceramic body, increasing the fragility of the samples. In general, it is observed that the increase in the pressing pressure from 20 MPa to 50 MPa promoted a reduction in the values of linear shrinkage, water absorption, and apparent porosity.
Table 4 summarizes the physical-mechanical properties (water absorption, linear shrinkage, apparent porosity, and flexural strength) measured from F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at different conditions (20 MPa and 50 MPa) and sintered at 1150 °C, 1200 °C, and 1250 °C. The ceramic formulation without scheelite tailing (F0 samples, see Table 1) was defined as the control sample. At 1150 °C, the scheelite-tailing-containing ceramics formulations presented better performance than the F0 sample, and the F5, F6, and F7 samples are potential candidates for the production of ceramic tiles. At 1200 °C, the ceramic formulations with scheelite tailing showed physico-mechanical properties had been performance statistically similar to the F0 sample, except the F3 sample that shown a behavior inferior. At 1250 °C, of the ceramic formulations with a higher content of scheelite tailing (i.e., a higher amount of fluxes), only the F1 and F2 samples showed performance similar to that of the F0 sample, and these samples present potential to be used as porcelain stoneware tile.

4. Conclusions

The incorporation of quartzite and scheelite tailings in ceramic masses proves to be a sustainable and viable alternative, making it possible to obtain ceramic pieces with low water absorption and low porosity, in addition to flexural strength in ranges of values that allow their application for the sustainable production of ceramic tiles and porcelain stoneware.
The incorporation of scheelite tailing by up to 8 wt% did not significantly alter the physical-mechanical properties for samples sintered at 1150 °C. However, the effects were more evident in samples sintered at 1200 °C and 1250 °C. Overall, the results showed that quartzite and scheelite tailings could be considered to be an alternative raw material for use in the ceramic tile and porcelain stoneware industry.

Author Contributions

Experimental data, data curation and formal analysis, D.G.G., F.P.d.C. and J.V.F.; conceptualization, funding acquisition and project administration, G.d.A.N., R.R.M. and L.N.d.L.S.; formal analysis and writing—review and editing, A.M.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), grant number 88882.455332/2019-01.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghiani, G.; Laganà, D.; Manni, E.; Musmanno, R.; Vigo, D. Operations research in solid waste management: A survey of strategic and tactical issues. Comput. Oper. Res. 2014, 44, 22–32. [Google Scholar] [CrossRef]
  2. Machado, T.G.; Umbelino Gomes, U.; Monteiro, F.M.; Valcacer, S.M.; da Silva, G.G. Analysis of the Incorporation of Scheelite Residue in Kaolinitic Clay of Boa Saúde-RN. Mater. Sci. Forum 2012, 727–728, 844–849. [Google Scholar] [CrossRef]
  3. Lottermoser, B.; Lottermoser, B.G. Introduction to Mine Wastes. In Mine Wastes; Springer: Berlin/Heidelberg, Germany, 2010; pp. 1–41. [Google Scholar]
  4. Jones, H.; Boger, D.V. Sustainability and waste management in the resource industries. Ind. Eng. Chem. Res. 2012, 51, 10057–10065. [Google Scholar] [CrossRef]
  5. Hughes, D.J.; Shimmield, T.M.; Black, K.D.; Howe, J.A. Ecological impacts of large-scale disposal of mining waste in the deep sea. Sci. Rep. 2015, 5, 9985. [Google Scholar] [CrossRef] [Green Version]
  6. Edraki, M.; Baumgartl, T.; Manlapig, E.; Bradshaw, D.; Franks, D.M.; Moran, C.J. Designing mine tailings for better environmental, social and economic outcomes: A review of alternative approaches. J. Clean. Prod. 2014, 84, 411–420. [Google Scholar] [CrossRef]
  7. Kinnunen, P.; Ismailov, A.; Solismaa, S.; Sreenivasan, H.; Räisänen, M.L.; Levänen, E.; Illikainen, M. Recycling mine tailings in chemically bonded ceramics—A review. J. Clean. Prod. 2018, 174, 634–649. [Google Scholar] [CrossRef] [Green Version]
  8. Adiansyah, J.S.; Rosano, M.; Vink, S.; Keir, G. A framework for a sustainable approach to mine tailings management: Disposal strategies. J. Clean. Prod. 2015, 108, 1050–1062. [Google Scholar] [CrossRef] [Green Version]
  9. De Medeiros, P.S.S.; Lira, H.D.L.; Rodriguez, M.A.; Menezes, R.R.; Neves, G.D.A.; Santana, L.N.D.L. Incorporation of quartzite waste in mixtures used to prepare sanitary ware. J. Mater. Res. Technol. 2019, 8, 2148–2156. [Google Scholar] [CrossRef]
  10. Da Silva, V.J.; de Almeida, E.P.; Gonçalves, W.P.; da Nóbrega, R.B.; de Araújo Neves, G.; de Lucena Lira, H.; Menezes, R.R.; de Lima Santana, L.N. Mineralogical and dielectric properties of mullite and cordierite ceramics produced using wastes. Ceram. Int. 2019, 45, 4692–4699. [Google Scholar] [CrossRef]
  11. Silva, R.H.L.; Neves, G.A.; Ferreira, H.C.; Santana, L.N.L.; Nóbrega, A.C.V.; Menezes, R.R. Use of diopside in ceramic masses for sanitary ware. Cerâmica 2019, 65, 1–12. [Google Scholar] [CrossRef] [Green Version]
  12. Medeiros, B.A.; Neves, G.A.; Barbosa, N.P.; Menezes, R.R.; Ferreira, H.C. Mechanical properties of mortar produced with the replacement of natural sand by scheelite residue. Ceramica 2019, 65, 443–451. [Google Scholar] [CrossRef]
  13. Tuskaeva, Z.; Karyaev, S.; Tagirov, T. Molybdenum waste usage as thinning agent in concrete production. Iop Conf. Ser. Mater. Sci. Eng. 2020, 918, 012093. [Google Scholar] [CrossRef]
  14. Da Silva, V.J.; da Silva, M.F.; Gonçalves, W.P.; de Menezes, R.R.; de Araújo Neves, G.; de Lucena Lira, H.; de Lima Santana, L.N. Porous mullite blocks with compositions containing kaolin and alumina waste. Ceram. Int. 2016, 42, 15471–15478. [Google Scholar] [CrossRef]
  15. Alves, H.P.A.; Silva, J.B.; Campos, L.F.A.; Torres, S.M.; Dutra, R.P.S.; Macedo, D.A. Preparation of mullite based ceramics from clay–kaolin waste mixtures. Ceram. Int. 2016, 42, 19086–19090. [Google Scholar] [CrossRef]
  16. Halicka, A.; Ogrodnik, P.; Zegardlo, B. Using ceramic sanitary ware waste as concrete aggregate. Constr. Build. Mater. 2013, 48, 295–305. [Google Scholar] [CrossRef]
  17. Scribot, C.; Maherzi, W.; Benzerzour, M.; Mamindy-Pajany, Y.; Abriak, N.E. A laboratory-scale experimental investigation on the reuse of a modified red mud in ceramic materials production. Constr. Build. Mater. 2018, 163, 21–31. [Google Scholar] [CrossRef]
  18. Parhi, B.R.; Sahoo, S.K.; Sahu, M.; Bishoyi, B.D.; Mishra, S.C.; Bhoi, B.; Mishra, B. Physico-chemical investigations of high iron bauxite for application of refractive and ceramics. Metall. Res. Technol. 2017, 114, 307. [Google Scholar] [CrossRef]
  19. Babisk, M.P.; Amaral, L.F.; Ribeiro, L.D.S.; Vieira, C.M.F.; Do Prado, U.S.; Gadioli, M.C.B.; Oliveira, M.S.; Luz, F.S.D.a.; Monteiro, S.N.; Garcia Filho, F.D.C. Evaluation and application of sintered red mud and its incorporated clay ceramics as materials for building construction. J. Mater. Res. Technol. 2020, 9, 2186–2195. [Google Scholar] [CrossRef]
  20. Alekseev, K.; Mymrin, V.; Avanci, M.A.; Klitzke, W.; Magalhães, W.L.E.; Silva, P.R.; Catai, R.E.; Silva, D.A.; Ferraz, F.A. Environmentally clean construction materials from hazardous bauxite waste red mud and spent foundry sand. Constr. Build. Mater. 2019, 229, 116860. [Google Scholar] [CrossRef]
  21. Xu, X.; Song, J.; Li, Y.; Wu, J.; Liu, X.; Zhang, C. The microstructure and properties of ceramic tiles from solid wastes of Bayer red muds. Constr. Build. Mater. 2019, 212, 266–274. [Google Scholar] [CrossRef]
  22. Wang, W.; Sun, K.; Liu, H. Effects of different aluminum sources on morphologies and properties of ceramic floor tiles from red mud. Constr. Build. Mater. 2020, 241, 118119. [Google Scholar] [CrossRef]
  23. Gonzalez-Triviño, I.; Pascual-Cosp, J.; Moreno, B.; Benítez-Guerrero, M. Manufacture of ceramics with high mechanical properties from red mud and granite waste. Mater. De Constr. 2019, 69, 180. [Google Scholar] [CrossRef]
  24. Menezes, R.R.; Ferreira, H.S.; Neves, G.A.; de L Lira, H.; Ferreira, H.C. Use of granite sawing wastes in the production of ceramic bricks and tiles. J. Eur. Ceram. Soc. 2005, 25, 1149–1158. [Google Scholar] [CrossRef]
  25. Hojamberdiev, M.; Eminov, A.; Xu, Y. Utilization of muscovite granite waste in the manufacture of ceramic tiles. Ceram. Int. 2011, 37, 871–876. [Google Scholar] [CrossRef]
  26. De Azevedo, A.R.G.; Marvila, M.T.; da Silva Barroso, L.; Zanelato, E.B.; Alexandre, J.; de Castro Xavier, G.; Monteiro, S.N. Effect of Granite Residue Incorporation on the Behavior of Mortars. Materials 2019, 12, 1449. [Google Scholar] [CrossRef] [Green Version]
  27. Lima, R.C.O.; Lira, H.L.; Neves, G.A.; Silva, M.C.; França, K.B. Study of the influence of granite residue in different compositions to prepare ceramic membranes. Mater. Sci. Forum 2014, 798–799, 542–547. [Google Scholar] [CrossRef]
  28. Barros, M.M.; de Oliveira, M.F.L.; da Conceição Ribeiro, R.C.; Bastos, D.C.; de Oliveira, M.G. Ecological bricks from dimension stone waste and polyester resin. Constr. Build. Mater. 2020, 232, 117252. [Google Scholar] [CrossRef]
  29. Souza, A.J.; Pinheiro, B.C.A.; Holanda, J.N.F. Recycling of gneiss rock waste in the manufacture of vitrified floor tiles. J. Environ. Manag. 2010, 91, 685–689. [Google Scholar] [CrossRef] [PubMed]
  30. Campos, M.; Velasco, F.; Martínez, M.A.; Torralba, J.M. Recovered slate waste as raw material for manufacturing sintered structural tiles. J. Eur. Ceram. Soc. 2004, 24, 811–819. [Google Scholar] [CrossRef]
  31. De Medeiros, P.S.S.; Santana, L.N.L.; Silva, V.J.; Neves, G.A.; Lira, H.L. Evaluation of the potential of using quartzite residue in mass for the production of sanitary ware. Mater. Sci. Forum 2016, 869, 181–185. [Google Scholar] [CrossRef]
  32. Barros, S.V.A.; Marciano, J.E.A.; Ferreira, H.C.; Menezes, R.R.; Neves, G.D.A. Addition of quartzite residues on mortars: Analysis of the alkali aggregate reaction and the mechanical behavior. Constr. Build. Mater. 2016, 118, 344–351. [Google Scholar] [CrossRef]
  33. Oliveira, S.S.L.; Oliveira, S.S.L.; da Silva Barbosa Ferreira, R.; de Lucena Lira, H.; de Lima Santana, L.N.; Araújo, E.M. Development of hollow fiber membranes with alumina and waste of quartzite. Mater. Res. 2019, 22, 20190171. [Google Scholar] [CrossRef] [Green Version]
  34. Torres, P.; Manjate, R.S.; Quaresma, S.; Fernandes, H.R.; Ferreira, J.M.F. Development of ceramic floor tile compositions based on quartzite and granite sludges. J. Eur. Ceram. Soc. 2007, 27, 4649–4655. [Google Scholar] [CrossRef]
  35. Junkes, J.A.; Prates, P.B.; Hotza, D.; Segadães, A.M. Combining mineral and clay-based wastes to produce porcelain-like ceramics: An exploratory study. Appl. Clay Sci. 2012, 69, 50–57. [Google Scholar] [CrossRef]
  36. Fernandes, J.V.; Rodrigues, A.M.; Menezes, R.R.; de A Neves, G. Adsorption of Anionic Dye on the Acid-Functionalized Bentonite. Materials 2020, 13, 3600. [Google Scholar] [CrossRef] [PubMed]
  37. Da Costa, F.P.; Morais CR da, S.; Pinto, H.C.; Rodrigues, A.M. Microstructure and physico-mechanical properties of Al2O3-doped sustainable glass-ceramic foams. Mater. Chem. Phys. 2020, 256, 123612. [Google Scholar] [CrossRef]
  38. Pereira da Costa, F.; Rodrigues da Silva Morais, C.; Rodrigues, A.M. Sustainable glass-ceramic foams manufactured from waste glass bottles and bentonite. Ceram. Int. 2020, 46, 17957–17961. [Google Scholar] [CrossRef]
  39. Rodrigues, A.M.; Rino, J.P.; Pizani, P.S.; Zanotto, E.D. Structural and dynamic properties of vitreous and crystalline barium disilicate: Molecular dynamics simulation and Raman scattering experiments. J. Phys. D Appl. Phys. 2016, 49, 435301. [Google Scholar] [CrossRef]
  40. Moulton, B.J.A.; Rodrigues, A.M.; Sampaio, D.V.; Silva, L.D.; Cunha, T.R.; Zanotto, E.D.; Pizani, P.S. The origin of the unusual DSC peaks of supercooled barium disilicate liquid. CrystEngComm 2019, 21, 2768–2778. [Google Scholar] [CrossRef]
  41. Avila, P.R.T.; Rodrigues, A.M.; Guimarães, M.C.R.; Walczak, M.; Menezes, R.R.; de A Neves, G.; Pinto, H.C. Nitrogen-Enriched Cr1−xAlxN Multilayer-Like Coatings Manufactured by Dynamic Glancing Angle Direct Current Magnetron Sputtering. Materials 2020, 13, 3650. [Google Scholar] [CrossRef]
  42. Soares, V.O.; Rodrigues, A.M. Improvements on sintering and thermal expansion of lithium aluminum silicate glass-ceramics. Ceram. Int. 2020, 46, 17430–17436. [Google Scholar] [CrossRef]
  43. Rodrigues, A.M.; Silva, L.D.; Zhang, R.; Soares, V.O. Structural effects on glass stability and crystallization. CrystEngComm 2018, 20, 2278–2283. [Google Scholar] [CrossRef]
  44. Fischer, J.; Stawarczyk, B.; Hämmerle, C.H.F. Flexural strength of veneering ceramics for zirconia. J. Dent. 2008, 36, 316–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Ma, B.; Li, Y.; Liu, G.; Liang, D. Preparation and properties of Al2O3-MgAl2O4 ceramic foams. Ceram. Int. 2015, 41, 3237–3244. [Google Scholar] [CrossRef]
  46. Mahmoudi, S.; Bennour, A.; Meguebli, A.; Srasra, E.; Zargouni, F. Characterization and traditional ceramic application of clays from the Douiret region in South Tunisia. Appl. Clay Sci. 2016, 127–128, 78–87. [Google Scholar] [CrossRef]
  47. De Aza, A.H.; Turrillas, X.; Rodriguez, M.A.; Duran, T.; Pena, P. Time-resolved powder neutron diffraction study of the phase transformation sequence of kaolinite to mullite. J. Eur. Ceram. Soc. 2014, 34, 1409–1421. [Google Scholar] [CrossRef]
  48. Menezes, R.R.; de A Neves, G.; Ferreira, H.C. State of the art about the use of wastes as alternative to ceramic raw materials. Rev. Bras. De Eng. Agrícola E Ambient. 2002, 6, 303–313. [Google Scholar] [CrossRef]
  49. Chen, C.Y.; Tuan, W.H. The processing of kaolin powder compact. Ceram. Int. 2001, 27, 795–800. [Google Scholar] [CrossRef]
  50. Issaoui, M.; Limousy, L.; Lebeau, B.; Bouaziz, J.; Fourati, M. Design and characterization of flat membrane supports elaborated from kaolin and aluminum powders. Comptes Rendus Chim. 2016, 19, 496–504. [Google Scholar] [CrossRef]
  51. Bennour, A.; Mahmoudi, S.; Srasra, E.; Boussen, S.; Htira, N. Composition, firing behavior and ceramic properties of the Sejnène clays (Northwest Tunisia). Appl. Clay Sci. 2015, 115, 30–38. [Google Scholar] [CrossRef]
Figure 1. Currais Novos City in Rio Grande do Norte State (Brazil), where the largest scheelite mine in South America (Brejuí mine) is located.
Figure 1. Currais Novos City in Rio Grande do Norte State (Brazil), where the largest scheelite mine in South America (Brejuí mine) is located.
Sustainability 12 09417 g001
Figure 2. X-ray diffraction (XRD) patterns measured from scheelite (a) and quartzite tailings (b) (⛛ Mica, ● Quartz, □ Feldspar, ♦ Calcite).
Figure 2. X-ray diffraction (XRD) patterns measured from scheelite (a) and quartzite tailings (b) (⛛ Mica, ● Quartz, □ Feldspar, ♦ Calcite).
Sustainability 12 09417 g002
Figure 3. Thermogravimetry (TG) and derivative thermogravimetry (DTG) curves of the (a) kaolin, (b) plastic clay, (c) feldspar, (d) quartzite tailing, and (e) scheelite tailing. All samples were heated at 12.5 °C·min−1.
Figure 3. Thermogravimetry (TG) and derivative thermogravimetry (DTG) curves of the (a) kaolin, (b) plastic clay, (c) feldspar, (d) quartzite tailing, and (e) scheelite tailing. All samples were heated at 12.5 °C·min−1.
Sustainability 12 09417 g003
Figure 4. Differential thermal analysis (DTA) curves obtained from (a) kaolin, (b) plastic clay, (c) feldspar, (d) quartzite tailing, and (e) scheelite tailing. All samples were heated with a heating rate of 12.5 °C·min−1.
Figure 4. Differential thermal analysis (DTA) curves obtained from (a) kaolin, (b) plastic clay, (c) feldspar, (d) quartzite tailing, and (e) scheelite tailing. All samples were heated with a heating rate of 12.5 °C·min−1.
Sustainability 12 09417 g004
Figure 5. The effect of the sintering temperature (1150 °C, 1200 °C, and 1250 °C) at the sample´s color on the ceramic formulations investigated in this work.
Figure 5. The effect of the sintering temperature (1150 °C, 1200 °C, and 1250 °C) at the sample´s color on the ceramic formulations investigated in this work.
Sustainability 12 09417 g005
Figure 6. XRD patterns of the F0, F1, F2, F3, F4, F5, F6, and F7 samples sintered at 1150 °C (a), and 1250 °C (b) (٭ mullite, ● quartz, ♦ calcite, Δ anorthite).
Figure 6. XRD patterns of the F0, F1, F2, F3, F4, F5, F6, and F7 samples sintered at 1150 °C (a), and 1250 °C (b) (٭ mullite, ● quartz, ♦ calcite, Δ anorthite).
Sustainability 12 09417 g006
Figure 7. Sintering temperature dependence of the physical-mechanical properties of the F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at 20 MPa. Linear shrinkage (a), water absorption (b), apparent porosity (c), and flexural strength (d).
Figure 7. Sintering temperature dependence of the physical-mechanical properties of the F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at 20 MPa. Linear shrinkage (a), water absorption (b), apparent porosity (c), and flexural strength (d).
Sustainability 12 09417 g007
Figure 8. Sintering temperature dependence of the physical-mechanical properties of the F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at 50 MPa. Linear shrinkage (a), water absorption (b), apparent porosity (c), and flexural strength (d).
Figure 8. Sintering temperature dependence of the physical-mechanical properties of the F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at 50 MPa. Linear shrinkage (a), water absorption (b), apparent porosity (c), and flexural strength (d).
Sustainability 12 09417 g008
Table 1. Nomenclature and nominal compositions (wt%) of the ceramic formulations studied in this research.
Table 1. Nomenclature and nominal compositions (wt%) of the ceramic formulations studied in this research.
Raw MaterialsCeramic Formulations (wt%)
F0F1F2F3F4F5F6F7
Kaolin2727272727272727
Plastic clay2929292929292929
Feldspar3332.53231.531292725
Scheelite tailing00.51.01.52.04.06.08.0
Quartzite tailing1111111111111111
Table 2. Chemical compositions of the raw materials (kaolin, plastic clay, and feldspar) and the mine tailings (quartzite and scheelite) used in the new ceramic formulation studied.
Table 2. Chemical compositions of the raw materials (kaolin, plastic clay, and feldspar) and the mine tailings (quartzite and scheelite) used in the new ceramic formulation studied.
Raw Materials Oxides
SiO2Al2O3F2O3K2OMgOCaONa2OOthersLF 1
Kaolin45.739.50.50.90.213.3
Plastic clay54.527.52.63.91.50.81.57.8
Quartzite tailing76.512.11.51.10.73.02.4
Feldspar62.019.912.12.72.71.6
Scheelite tailing20.87.16.80.52.744.72.115.8
1 LF: Loss to fire determined by burning at 1000 °C, after drying at 110 °C.
Table 3. Summary of the particle size distribution of the kaolin, plastic clay, feldspar, and mine tailings (quartzite and scheelite).
Table 3. Summary of the particle size distribution of the kaolin, plastic clay, feldspar, and mine tailings (quartzite and scheelite).
Raw MaterialsFine
(x < 2 µm)
Medium
(2 µm < x < 20 µm)
Thick
(x > 20µm)
D10 (µm)D50 (µm)D90 (µm)Dm (µm)
Kaolin26.6%69.0%4.4%0.63.215.05.6
Plastic clay20.1%79.9%0.93.710.04.6
Quartzite tailing8.9%40.9%50.2%4.528.257.130.3
Feldspar21.7%60.9%17.4%0.96.630.811.4
Scheelite tailing3.6%26.4%70.0%4.232.064.133.6
x = accumulated fraction; D10 = 10% diameter; D50 = 50% diameter; D90 = Diameter at 90%; Dm = Average diameter.
Table 4. Physical-mechanical properties measured from F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at different conditions (20 MPa and 50 MPa) and sintered at 1150 °C, 1200 °C, and 1250 °C.
Table 4. Physical-mechanical properties measured from F0, F1, F2, F3, F4, F5, F6, and F7 samples pressed at different conditions (20 MPa and 50 MPa) and sintered at 1150 °C, 1200 °C, and 1250 °C.
Samples Pressed at 20 MPa
SamplesLinear Shrinkage (%)Water Absorption (%)Apparent Porosity (%)Flexural Strength (MPa)
1150 (°C)1200 (°C)1250 (°C)1150 (°C)1200 (°C)1250 (°C)1150 (°C)1200 (°C)1250 (°)1150 (°C)1200 (°C)1250 (°C)
F05.257.808.377.262.370.0515.655.570.2221.7133.7542.84
Error0.290.110.130.600.140.271.190.800.657.723.475.81
F15.467.927.886.511.430.1414.243.480.4625.6336.2542.84
Error0.370.200.250.780.140.391.530.760.772.752.625.81
F25.477.937.876.691.670.2214.463.940.5325.6331.9438.81
Error0.450.170.360.810.500.251.570.790.601.473.805.26
F35.868.487.797.124.310.3915.233.790.8825.3529.7836.57
Error0.210.660.290.870.110.231.510.760.522.022.053.90
F45.678.926.686.231.720.1813.674.040.4026.6834.1331.39
Error0.270.150.320.550.130.141.120.700.312.052.573.08
F56.158.477.245.882.160.2312.764.930.4928.2936.4430.35
Error0.340.110.480.480.120.220.940.990.482.122.2611.09
F65.386.534.104.341.660.099.663.790.1932.0632.2828.38
Error0.240.140.180.450.120.160.900.740.321.726.042.35
F74.576.195.135.812.150.3212.244.680.6529.2130.8627.63
Error0.200.110.150.590.140.301.280.770.611.911.831.70
Samples Pressed at 50 MPa
F05.347.217.044.461.010.1410.162.450.3527.6132.9145.40
Error0.220.130.110.300.100.160.640.610.391.362.193.29
F15.227.186.524.920.790.0711.071.880.3830.5030.7738.37
Error0.400.180.310.650.130.291.350.690.692.432.342.55
F25.577.016.494.071.340.069.293.190.3430.7134.3138.36
Error0.260.130.290.370.130.350.790.660.421.582.454.04
F35.517.746.135.051.440.3711.253.410.8328.7333.3234.09
Error0.380.170.360.530.110.281.070.660.611.942.352.19
F45.687.785.553.861.390.278.813.270.6031.8936.8936.13
Error0.260.130.630.500.100.191.060.590.413.622.932.48
F55.896.704.793.741.300.578.493.031.2134.8035.9132.72
Error0.210.380.270.230.130.320.500.610.671.722.112.13
F65.165.102.882.580.420.135.940.970.3734.8132.2828.02
Error0.130.220.320.400.180.180.890.690.282.672.342.57
F74.594.944.074.432.340.299.724.840.6931.2231.4724.95
Error0.260.110.320.590.140.391.300.560.372.442.085.97
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fernandes, J.V.; Guedes, D.G.; da Costa, F.P.; Rodrigues, A.M.; Neves, G.d.A.; Menezes, R.R.; Santana, L.N.d.L. Sustainable Ceramic Materials Manufactured from Ceramic Formulations Containing Quartzite and Scheelite Tailings. Sustainability 2020, 12, 9417. https://doi.org/10.3390/su12229417

AMA Style

Fernandes JV, Guedes DG, da Costa FP, Rodrigues AM, Neves GdA, Menezes RR, Santana LNdL. Sustainable Ceramic Materials Manufactured from Ceramic Formulations Containing Quartzite and Scheelite Tailings. Sustainability. 2020; 12(22):9417. https://doi.org/10.3390/su12229417

Chicago/Turabian Style

Fernandes, Jucielle Veras, Danyelle Garcia Guedes, Fabiana Pereira da Costa, Alisson Mendes Rodrigues, Gelmires de Araújo Neves, Romualdo Rodrigues Menezes, and Lisiane Navarro de Lima Santana. 2020. "Sustainable Ceramic Materials Manufactured from Ceramic Formulations Containing Quartzite and Scheelite Tailings" Sustainability 12, no. 22: 9417. https://doi.org/10.3390/su12229417

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