Next Article in Journal
Effect of Different Decontamination Methods on Fracture Resistance, Microstructure, and Surface Roughness of Zirconia Restorations—In Vitro Study
Next Article in Special Issue
The Effect of Process Conditions on Sulfuric Acid Leaching of Manganese Sludge
Previous Article in Journal
Fabrication of Rapid Electrical Pulse-Based Biosensor Consisting of Truncated DNA Aptamer for Zika Virus Envelope Protein Detection in Clinical Samples
Previous Article in Special Issue
Study on the Effect of Temperature on the Crystal Transformation of Microporous Calcium Silicate Synthesized of Extraction Silicon Solution from Fly Ash
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preliminary Study of Preheated Decarburized Activated Coal Gangue-Based Cemented Paste Backfill Material

1
Energy School, Xi’an University of Science and Technology, Xi’an 710054, China
2
Key Laboratory of Western Mines and Hazards Prevention, Ministry of Education of China, Xi’an 710054, China
3
School of Architecture and Civil Engineering, Xi’an University of Science and Technology, Xi’an 710054, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(6), 2354; https://doi.org/10.3390/ma16062354
Submission received: 31 January 2023 / Revised: 8 March 2023 / Accepted: 13 March 2023 / Published: 15 March 2023

Abstract

:
This study proposes a novel idea of the use of coal gangue (CG) activation and preheated decarburized activated coal CG-based cemented paste backfill material (PCCPB) to realize green mining. PCCPB was prepared with preheated decarburized coal CG (PCG), FA, activator, low-dose cement, and water. This idea realized scale disposal and resource utilization of coal CG solid waste. Decarbonization and activation of CG crushed the material to less than 8 mm by preheated combustion technology at a combustion temperature of 900 °C and a decarbonization activation time of 4 min. The mechanism of the effect of different Na2SO4 dosages on the performance of PCCPB was investigated using comprehensive tests (including mechanical property tests, microscopic tests, and leaching toxicity tests). The results show that the uniaxial compressive strength (UCS) of C-S2, C-S3, and C-S4 can meet the requirements of backfill mining, among which the UCS of C-S3 with a curing time of 3 d and 28 d were 0.545 MPa and 4.312 MPa, respectively. Na2SO4 excites PCCPB at different curing time, and the UCS of PCCPB increases and then decreases with the increase in Na2SO4 dosage, and 3% of Na2SO4 had the best excitation effect on the late strength (28 d) of PCCPB. All groups’ (control and CS1-CS4 groups) leachate heavy metal ions met the requirements of groundwater class III standard, and PCCPB had a positive effect on the stabilization/coagulation of heavy metal ions (Mn, Zn, As, Cd, Hg, Pb, Cr, Ba, Se, Mo, and Co). Finally, the microstructure of PCCPB was analyzed using FTIR, TG/DTG, XRD, and SEM. The research is of great significance to promote the resource utilization of coal CG residual carbon and realize the sustainable consumption of coal CG activation on a large scale.

1. Introduction

One of China’s major strategies is sustainable development, and for the coal industry to develop sustainably, reasonable disposal of coal gangue solid waste is necessary. Coal is the ballast stone of China’s energy security, but the coal mining process also produces a series of environmental problems, such as surface damage, coal gangue (CG) emissions, etc. [1,2]. CG accounts for 10–25% of the total coal mining [3], and over the years, China has accumulated more than 7 billion tons of CG solid waste and a growth rate of 150 million t/a [4]. CG stockpiles occupy a large number of land resources and also have potential environmental hazards such as spontaneous combustion and the leaching of heavy metal ions [5,6,7]. Backfill mining as a green mining method can reduce surface damage and effectively dispose of solid waste such as CG [8]. CG has poor cementation properties and is mainly used as aggregate in backfill mining [9,10]. However, the high carbon content of CG leads to poor interfacial bonding of the filler, which affects the performance of the cemented paste backfill (CPB) [11,12]. Borrowing the technical idea of metal mine tailings activation for clinker preparation [13,14,15], decarbonization and activation treatment of CG can fully use CG’s heat and stimulate its activity, which is essential to realize the resourceization and sustainable utilization of CG.
Thermal activation is considered to be the most promising method for destroying the crystal structure of CG and improving its reaction activity [16]. Much research has been carried out to find the best activation process to achieve high chemical reactivity of CG. Organic components and carbon in CG can be burned off after calcination, and kaolinite in CG can also be gradually converted to meta kaolinite in the temperature range of 500–800 °C, thus improving the activity of volcanic ash [17,18,19]. Hao [20] investigated the effect of thermal activation conditions on the mineral phase and structural changes of CG minerals from the Junggar coalfield, Inner Mongolia, China, and showed that the highest volcanic ash activity was reached when calcined at 800 °C for 2 h. Song [21,22] calcined the Xuzhou CG and the results showed that the optimum thermal activation process for CG was 700 °C kept warm for 2 h when it showed the highest reactivity. Frías et al. [23] calcined 5 different samples of CG at different temperatures (500–900 °C) with a holding time of 2 h and used them in cement mortar to replace 12% of cement calcined at 800 °C. The cement mortar with coal CG had 18% higher compressive strength at 28 days than the control group. The results show that the volcanic ash produced by calcination at 600~800 °C has the best activity. However, the required calcination time is 1–2 h, and the long time required for thermal activation seriously restricts the scale of CG thermal activation.
Preheated combustion is an effective way to achieve short-time combustion of low-volatile fuels [24]. A new technology of preheated combustion of pulverized coal was first proposed by the All-Russian Institute of Thermal Engineering [25], where pulverized coal was preheated to about 816 °C and then burned with a minimum fluidization velocity of 0.4 m/s, which ensured pulverized coal burnout. The Institute of Engineering Thermophysics, Chinese Academy of Sciences, proposed a circulating fluidized bed preheating coupled with pulverized coal furnace combustion technology [26,27,28,29], where the circulating fluidized bed ignites and preheats the fuel in the first stage, and stable and efficient combustion of low-volatile-content fuel can be achieved in the pulverized coal furnace in the second stage with a minimum fluidization velocity of 0.67 m/s. Lyu and Wang Shuai et al. [30,31] used a two-stage descending tube furnace to achieve preheated combustion, studied the effects of preheating temperature, residence time, and combustion temperature on combustion exhaustion, and concluded that the pulverized coal could be fully combusted at a preheating time of 0.4 s. Wang Xuebin [32] developed the technology of constant temperature preheat decarburization of low-volatile-matter and low-calorific-value fuels. The 1st generation plant has been in stable operation for 2 years in Ganquanbao, Xinjiang, with an annual treatment of 600,000 tons of gasification slag, a combustion temperature of 900 °C, and a combustion time of 3.7–5.6 min. It has been proved that the preheat combustion technology can burn out the residual carbon of low-volatile-matter fuels in a short time. However, there are few studies on preheat decarbonization of CG, the activity of the CG residue after preheating decarbonization is unknown, and relevant research on the preparation of paste filling material as a gelling material has not been carried out. Therefore, preheat decarbonization activation of coal CG and carrying out preheat decarbonization CG-based CPB material research to promote the resource utilization of CG residue carbon to achieve sustainable consumption of CG activation scale has essential significance.
In order to fill the relevant gap in the literature, this paper prepares a cemented paste backfill with preheated decarburized activated CG (PCG), FA, activator, low-dose cement (most backfill materials use more than 10% cement dosage [33,34,35,36]), and water based on short-time preheated decarburization activation of CG. The innovations of this study are mainly in three aspects: (1) to propose a new idea of CG decarbonization activation and a new PCCPB using the described method; (2) to research the effect law of various Na2SO4 dosages on the performance of novel PCCPB; and (3) to reveal the micromechanism of novel PCCPB by means of FTIR, TG/DTG, XRD, and SEM. The research content provides new methods for decarbonizing and activating CG and large-scale utilization, promoting green and sustainable coal mining.

2. Materials and Methods

2.1. Materials

The raw materials for this experiment are PCG, fly ash (FA), aeolian sand (AS), cement (OPC), activator, and water. PCG and FA served as auxiliary materials, cement as cementitious materials, AS as aggregate, municipal water as mixing water, and Na2SO4 as an activator.

2.1.1. Preheated Decarburized Activated CG

The PCG used in the experiment was prepared using a preheated decarburization 108 device in Ganquanbao, Xinjiang, China. Its preparation process is shown in Figure 1. The device is based on the thermostatic preheated decarburization technology developed by Wang Xuebin [32], and a CG combustion temperature of 900 °C and combustion time of 4.0 min were used in this experiment. The PCG was ground for 10 min in the laboratory using a ball mill (SM-500, Wuxi, China) as this experimental material. Next, PCG’s particle size distribution (PSD) after grinding was detected based on a laser scanner (Malvern 2000; Malvern, UK), as shown in Figure 2. The PSD curves indicated that the experimental coarse particles had relatively low PCG content. Moreover, the chemical properties of the prepared PCG were assessed, and the PCG was tested using an X-ray fluorescence spectrometer (XRF) (ARL spectrometer, Waltham, MA, USA). The chemical composition of PCG was tested (Table 1) and the main chemical constituents were SiO2 (53.55%), Al2O3 (20.41%), and Fe2O3 (8.99%). In addition, PCG’s microscopic morphology and mineral composition were tested by XRD (D8 Advance, Karlsruhe, Germany) and SEM (JSM-7610F, Akishima-shi, Japan). A SEM image of PCG with 2000 times magnification is shown in Figure 3, which indicates that the microscopic morphology of PCG is irregular with an angular shape. The mineralogical composition of PCG is shown in Figure 4. PCG is mainly composed of quartz and mullite. Moreover, the specific surface area of PCG was tested based on a specific surface area meter (BSD-660, Beijing, China) with the result of 3.17 m2/g.

2.1.2. Fly Ash

The FA was bought at the pineapple power plant in Yulin, Shaanxi Province, and the same analytical method as PCG was used for FA. Figure 2 and Table 1 show the d10, d60, and d90 values of FA as 2.24 μm, 33.0 μm, and 135.3 μm, respectively. The PSD curves show that the content of the fine particles is shallow. The results of FA constituent testing are exhibited in Table 1. The main components are Al2O3, SiO2, and Fe2O3 accounting for 86.65%. The microscopic morphology of FA is small and granular. The XRD detecting results are shown in Figure 4. The main minerals of fly ash are aluminosilicate (50–85%), sponge-like vitreous (10–30%), quartz (1–10%), iron oxide (3–25%), carbon particles (1–20%), and sulfate (1–4%).

2.1.3. Aeolian Sand

The AS used in the experiments was taken from Yuyang District, Yulin City. The main light minerals of AS are quartz, feldspar, and calcite, accounting for more than 90%. Heavy minerals are amphibole. Mica and epidote account for more than 5%. Figure 2 and Table 1 show the d10, d60, and d90 values of AS as 8.27 μm, 256.03 μm, and 357.69 μm, respectively. The PSD curves show that the experiment’s acceptable particle content is shallow, and the corresponding gradation is relatively discontinuous. The composition test results of AS can be seen in Table 1. The main components are SiO2 and Al2O3, accounting for 78.1%.

2.1.4. Cement, Activator, and Water

Cement is a cementitious material, and its main mineralogical components are C2S (20%), C3S (50%), C3A (7–15%), C4AF (10–18%), etc. The main chemical composition is shown in Table 1. The main hydration products are calcium hydroxide (CH), calcium silicate hydrate (C-S-H), calcium aluminate hydrate (C-A-H), and calcium alumino-ferrite hydrate (C-A-F-H). The cement type used is ordinary Portland cement (OPC) 42.5, and the basicity of the cement is not more than 0.6%, in line with the Chinese national standard GB175-2007. Na2SO4 was selected as the activator. It was produced based on the standard of GB/T 9853-2008. The mixing water was mixed with municipal tap water.

2.2. Sample Preparation

According to the preliminary experimental study, it was determined that the PCCPB solid mass concentration of this experiment was 78%, the cement dosage was 3% of the total solid mass, the ash–sand ratio was fixed at 0.5, and the PCG, FA, cement, AS, and activator were formulated as shown in Table 2. The numbers in the table indicate different levels.

2.3. Experimental Setup and Method

2.3.1. Mechanical Property Test

Under the condition that PCCPB samples reached curing time (3 d, 7 d, 14 d, and 28 d), the UCS tests were conducted with a condition of 1.0 mm/min based on a DNS100 system (SinoTest, Changchun, China) in terms of the standard of GB/T 50081-2019. All experiments were performed three times, and the mean value of UCS was taken.

2.3.2. Microstructure Test

Fourier-transform infrared spectroscopy (FTIR) tests were performed on PCCPB samples that reached the curing time (3 d, 7 d, 14 d, and 28 d). FTIR tests were performed using a Nicolet iN10 Fourier transform micro-infrared spectrometer (Thermo Fisher, Waltham, MA, USA).
The change in weight loss with temperature was tested for cured 3 d and 28 d PCCPB samples using a TGA5500 device. Nitrogen was applied to block carbonization during the heating process. The heating rate in this experiment was 20 °C/min to 900 °C.
The Bruker X-ray diffractometer was used to test the PCCPB samples with a curing time of 3 d and 28 d. Then, the mineral composition of the samples was determined by comparing JADE6.0 software with powder diffraction file (PDF) cards. After the completion of the UCS test, a sample from the central part of the sample was selected and soaked for 48 h to terminate the hydration process. A scan rate of 5°/min and a spectral range of 5° to 80° were used for the tests.
The microstructure of the curing times of 3 d and 28 d specimens was observed using a JEOL JSM-6460LV SEM device. After the UCS test, the central part of the sample was selected and cut into 2 mm slices, which were put on the sample stage for testing.

2.3.3. Leaching Toxicity Test

The samples were crushed so that all sample particles passed a 3 mm sieve. The “Leaching Toxicity Leaching Method for Solid Wastes Inverted Shaking Method” (HJ557-2010) [37], using pure water as a leaching agent, was used. This standard assesses the likelihood that inorganic contaminants would leach from solid wastes and other solid materials into the surface or groundwater [38]. Figure 5 illustrates the test procedure.

3. Results and Discussion

3.1. Mechanical Performance of PCCPB

3.1.1. Analysis of UCS of PCCPB

The mechanical properties of the PCCPB are critical indicators to evaluate its stability, which can be expressed by its UCS, coefficient of variance (CV) of UCS, and strong growth rate. Table 3 shows the results of the fundamental statistical analysis of the UCS of PCCPB, and the results demonstrate the mean ( X ¯ ), standard deviation (SD), and coefficient of variance (CV). SD and CV measure the degree of dispersion of the data, where CV can be expressed by Equation (1) [39]. The SD of all PCCPB samples was kept below 0.13, and the CV was below 15% [40], indicating that the test data are available.
CV = SD X ¯ × 100 %  
The variation of the UCS of PCCPB samples with the curing time and Na2SO4 dosage is exhibited in Figure 6. The UCS of PCCPB samples is positively related to curing time, which has been shown by other studies [41,42]. As can be seen from Figure 6, the UCS of C-S3 enhanced from 0.545 MPa (3 d) to 2.057 MPa (7 d), 3.284 MPa (14 d), and 4.312 MPa (28 d) after curing 3 d, 7 d, 14 d, and 28 d, respectively. The increase in UCS is caused by a large number of hydration products (C-(A)-S-H, calcium alumina, CH, and silicate, etc.) filling the internal pores, which leads to the higher compressive strength of the material, a conclusion confirmed in Section 3.2.4, the research of Liu et al. also confirmed this conclusion [43,44].
From Figure 6, it can be seen that Na2SO4 has an excitation effect on PCCPB samples at different curing times, and the UCS of PCCPB increases and subsequently decreases during the process of increasing Na2SO4 dosage. Comparing the UCS of PCCPB at 3 d, 7 d, and 14 d, the highest UCS was found in the C-S2 group, indicating that adding 2% Na2SO4 had an apparent enhancement influence on the early strength of PCCPB [45], which indicated that more Na2SO4 in the early reaction process would inhibit the hydration reaction. For the curing time of PCCPB for 28 d, the UCS of C-S1, C-S2, C-S3, and C-S4 enhanced by 17.61%, 33.62%, 49.15%, and 18.89%, respectively, compared with the control group (the rates of increase in UCS of other groups are listed in the following table), and the UCS corresponding to the C-S3 group was the highest. When the addition of Na2SO4 dosage exceeded 3%, its UCS showed a decreasing trend, probably because the Na2SO4 that did not participate in the reaction remained in the pores of the PCCPB samples and produced sulfate erosion of the hydration products later [46].
According to the mechanical requirements of the CPB (3 d ≥ 0.5 MPa and 28 d ≥ 1.0 MPa) [47,48], the mechanical properties of C-S2, C-S3, and C-S4 meet the mechanical requirements of the CPB, and the UCS data of 3 d and 28 d were combined to determine CS3 as the best ratio.

3.1.2. Analysis of Elastic Modulus of PCCPB

The elastic modulus (EM) is a physical and mechanical parameter that directly affects the performance of the PCCPB and is also a necessary parameter for stability analysis, optimization of the structural parameters of the quarry, and numerical simulation. The EM can reflect the degree of bonding between aggregate particles in the PCCPB, and increasing the EM of the PCCPB can reduce the damage through the elastic buffering effect, thus improving the stability of the PCCPB [49].
Figure 7 shows the EM of PCCPB with different Na2SO4 dosages. The effect of Na2SO4 on the EM of PCCPB is similar to that of UCS, where the EM at any curing time increases with Na2SO4 and then decreases, with the highest EM corresponding to the C-S3 group and the most vigorous resistance to deformation of the PCCPB [50]. The EM of the C-S3 group increased by 80.1%, 85.9%, 29.5%, and 52.9% at 3 d, 7 d, 14 d, and 28 d of curing time, respectively, compared with the control group. Under the addition of Na2SO4 of more than 3%, the EM showed a decreasing trend.
Figure 8 reflects the variation of the EM of PCCPB with UCS, which shows that as the UCS increases, the EM increases, and vice versa. The relationship between UCS and EM of PCCPB was analyzed by the regression analysis method, including the square root of compressive strength, cube root of compressive strength, and quadratic function of UCS. The square root of UCS was found to be more suitable to describe the relationship between UCS and EM, and red mud material showed a similar relationship [51]. The final equation derived from the analysis is shown in Equation (2):
E = A × UCS 0 . 5 + B
where E represents the EM (MPa), and A and B are fitting parameters.
The fit coefficient R2 = 0.9424 indicates that a quadratic function using the UCS’s square root is consistent with the inspired relationship between the EM and the UCS of the PCCPB.

3.2. Microstructure of ARFGB

3.2.1. FTIR Results Analysis of PCCPB

The Fourier-transform infrared spectra (FTIR) of the PCCPB samples are shown in Figure 9. The bands corresponding to different Na2SO4 dosages and different curing times of the PCCPB samples shown in Figure 9 are around 3436 cm−1, 1633 cm−1, 1419 cm−1, 1092 cm−1, 871 cm−1, 772 cm−1, 612 cm−1, and 463 cm−1, respectively. The absorption peaks near 3436 cm−1 and 1633 cm−1 are O-H stretching vibration peaks and bending vibration peaks, respectively. These O-H characteristic absorption peaks originate from the structural water produced by the hydration reaction and a small part of the free water in the system [52]. The absorption peak near 468 cm−1 is due to the bending vibration of the Si-O bond [53]. This band represents quartz. The absorption peaks near 612 cm−1 are bending vibrations of Al-O, which correspond to the [Si(OH)5] monomeric structures in the 3D mesh of the gelling material [54]. The absorption peak near 871 cm−1 corresponds to the Si-OH band’s bending vibration [55]. The band at 1419 cm−1 indicates that O-C-O bonds stretch in the presence of carbonates associated with calcite in the sample [56]. The peak near 772 cm−1 corresponds to the symmetric stretching vibration of Si-O-T dissolved in SiO4 tetrahedra; thus, the dissolved Si-Al elements are involved in forming C-(A)-S-H gels; this is also confirmed by Li et al. [17]. The peak near 1092 cm−1 is the stretching vibration peak of asymmetric Si-O-Si or Si-O-Al. These absorption peaks show a structure of the Si-O-Al-O bond interconnection in the system. The [Si(OH)5] and [Al(OH)4] monomer structures in the system are connected by these bonds to form a polymer, which recombines into a three-dimensional network structure of the cementitious material [57].

3.2.2. TG-DTG Results Analysis of PCCPB

Figure 10 exhibits the TG–DTG data for different Na2SO4 dosages and curing times of PCCPB samples. By and large, all TG–DTG curves have similar characteristics with three prominent heat absorption peaks in the temperature ranges of 50–250 °C (zone I) and 400–450 °C (II). The heat absorption peak at 50–250 °C is associated with the evaporation of free water and dehydration of hydration products (C-(A)-S-H and AFt) [58]. The heat absorption peak at 400–450 °C is assigned to CH dehydration, and the decomposition of calcium carbonate and CH occurs at 600–750 °C [59]. In addition, a weaker heat absorption peak in zone IV above 750 °C is associated with the dehydroxylation of silicate minerals in the sample [60,61]; the weight loss reactions in zones I, II, and III also reveal the hydration reaction products of PCCPB, and the results are consistent with Section 3.2.3.

3.2.3. XRD Results Analysis of PCCPB

Figure 11 shows the XRD results of PCCPB. The diffraction peaks of quartz (PDF # 51–1377), ettringite (PDF # 37–1476), calcium hydroxide (PDF # 84–1263), C-(A)-S-H gel (PDF # 34–0002), calcite (PDF # 41–1475), and mullite (PDF # 83–1881) are observed. This is consistent with the results of Mota et al.’s [65] study on cement hydration products in pure cement and sodium sulfate environments. Calcium alumina, calcium hydroxide, and C-(A)-S-H gel are the hydration products generated in the specimens. The primary source of both is from both generation channels. The hydration of cement generates 1, see Equation (3)—the reaction of reactive SiO2 generates 4 [66,67] and the other and Al2O3 in FA and PCG with CH in the hydration reaction of silicate cement, see Equations (5)–(7) [68,69]. The second channel, calcium alumina, is formed by the reaction of Al2O3 with CH to form C-A-H and then with gypsum. The formula for C-A-S-H is given in Equation (8). According to the XRD pattern of the original material, it contains quartz and mullite minerals and does not participate in the hydration reaction [70]. The calcite is derived from both the raw material and the specimen. Calcite is derived from the specimen’s raw material and surface charring.
3 CaO Al 2 O 3 ( C 3 A ) + 3 CaSO 4 2 H 2 O + 30 H 2 O CaO Al 2 O 3 3 CaSO 4 32 H 2 O ( AFt )
3 Ca SiO 2 + xH 2 O xCaO SiO 2 yH 2 O ( C - S - H ) + ( 3 x ) Ca ( OH ) 2
Al 2 O 3 + Ca ( OH ) 2 + xH 2 O CaO Al 2 O 3 xH 2 O ( C - A - H )
3 CaO Al 2 O 3 6 H 2 O + 3 CaSO 4 + 26 H 2 O 3 CaO Al 2 O 3 3 CaSO 4 32 H 2 O ( AFt )
SiO 2 + Ca ( OH ) 2 + xH 2 O CaO SiO 2 xH 2 O ( C - S - H )
Al 2 O 3 + Ca ( OH ) 2 + 2 SiO 2 + xH 2 O CaO Al 2 O 3 2 SiO 2 xH 2 O ( C - A - S - H )
From the result in Figure 11a, quartz, calcite, and mullite diffraction peaks do not change significantly in all plots as the Na2SO4 dosage increases from 0 wt.% to 4 wt.%. The variation is more significant for calcium alumina and C-(A)-S-H gels, where the intensity of the diffraction peaks of these two hydration products first becomes stronger from a weak diffraction peak and then changes to a weak diffraction peak. Specifically, the dosage of Na2SO4 increased from 0 wt.% to 2 wt.%, and the diffraction peak of chalcocite at 2θ was enhanced at 9°. However, the dosage of the Na2SO4 was increased to 4 wt.%, and the diffraction peak of chalcocite in the C-S4 sample was decreased. In addition, a similar phenomenon was observed for the C-(A)-S-H gel diffraction peak at 2θ of 39°. This implies that the diffraction peaks of the hydration products in the C-S2 specimens are the strongest at the curing time of 3 d. This is consistent with the UCS of all specimens at 3 d. This suggests that Na2SO4 as an activator facilitates the generation of calcium alumina and improves early strength development, which agrees with Razali et al. [71] and Zhao et al. [72]. However, with excess Na2SO4, the slurry viscosity increases, leading to the precipitation of C-(A)-S-H gel in the early stage, which inhibits the positive proceeding of hydration [73,74]. In addition, comparing Figure 11a,b, it can be seen that the diffraction peaks of calcite and C-(A)-S-H gels in the curing time of 28 d of PCCPB were significantly enhanced, especially for calcite with 2θ of 16°. The diffraction peaks of calcite with 2θ of 49° were significantly enhanced compared with those in the 3 d samples, especially for the C-S4 sample. Notably, the diffraction peaks of calcium hydroxide with 2θ of 18° and 34° do not change significantly in all samples.
On the one hand, it may be due to the small cement content used in this case of only 3 wt.% of the solid mass, which generates less calcium hydroxide. On the other hand, it may be that the excitation of Na2SO4 promotes the hydrolysis of calcium oxide in FA, replenishing the calcium hydroxide consumed by hydration. In order to analyze the effect of curing time on the hydration products, C-S3 samples were selected for testing, and the curing time of the specimens was 3 d, 7 d, 14 d, and 28 d. The XRD results are shown in Figure 11c. The diffraction peaks of the hydration products (calcium alumina, C-(A)-S-H gel) in the samples changed from weak to firm as the curing time increased from 3 d to 28 d. This indicates that the content of hydration products increases with the curing time, increasing strength. Ouyang et al. [75] also obtained a similar conclusion and found that the diffraction peaks of the hydration products (AFt, calcium alumina, and C-(A)-S-H gel) of the sand-based cemented paste filling material were enhanced with the extension of curing age from 1 d to 7 d and 28 d.

3.2.4. SEM Results Analysis of PCCPB

Figure 12 shows the SEM results of PCCPB samples for the curing times of 3 d and 28 d. It was observed that the microscopic morphology of PCCPB samples showed significant differences depending on curing time and the ratio of Na2SO4. With the increase in Na2SO4, the microscopic morphology of all PCCPB samples showed a pattern from loose to dense and slightly loose. Under the condition of a curing time of 3 d, the control group was the loosest and most porous. When Na2SO4 was increased to 2 wt.%, the microscopic morphology of the C-S2 group changed more, from loose to relatively dense, and the matrix of small granular FA and PCG was reduced. After curing for 28 d, abundant needle-like calcarenite (whose length is greater than 1 μm) and irregularly shaped C-(A)-S-H gels were observed in microscopic images of all samples, which had relatively more hydration products and a more extensive distribution range compared with the 3 d samples. The intercalation of calcarenite between the pores of PCCPB increases the denseness of the specimens. However, the control group has fewer hydration products and a relatively loose structure. The microstructure of PCCPB is mainly composed of C-(A)-S-H gels, fine needle-like AFt crystals, hexagonal lamellar CH, unhydrated AS particles, and PCG. This differs from the hydration products of binders made from blast furnace slag [76,77], and the reported results for the hydration product of pure cement are consistent [78,79]. In addition, the formation of hydration products provides the basis for the UCS of C-(A)-S-H.
Figure 13 shows the SEM results of C-S3 group samples. From Figure 13, it can be seen that the microstructure of the samples becomes dense as the age increases from 3 d to 28 d with the gradual increase in calcium alumina (AFt) and C-(A)-S-H gel. It was also found that the number of unreacted FA and PCG particles decreased. This is because the silica–oxygen tetrahedral monomer and aluminum–oxygen tetrahedral monomer in FA and PCG are involved in the generation of C-(A)-S-H gels and are connected, gradually forming a monolithic structure [80]. The hydrated calcium silica-aluminate structure has a specific skeletal effect [81,82], thus increasing the UCS of the samples with curing time. This is consistent with the findings of the previous UCS and XRD studies. Zhang and Tang et al. [55] reached similar conclusions. For example, Tang et al. [83] studied fly ash–aeolian sand filling materials and found that AFt and C-(A)-S-H gels in the samples increased with the extension of curing age, and the micro-morphology of the samples changed from loose to dense.

3.3. Leaching Toxicity Results Analysis of PCCPB

When using PCCPB as a backfill material, it is necessary to evaluate the impact of PCCPB weight metal ion leaching on groundwater [84]. Table 5 exhibits the leaching data of heavy metal ions from PCCPB in 28 days of curing. The standard limits are taken from HJ/T 300-2007, which show that the heavy metal ions of leachate in the control group and CS1-CS4 group meet the requirements of groundwater class III standard, which indicates that PCCPB meets the environmental safety requirements and PCCPB is a promising technology for heavy metal curing and stabilization. The CS1-CS4 group is lower than the control group, and the CS3 group is the lowest for three main reasons: (1) C-(A)-S-H and CH are formed by PCCPB hydration, as shown in Section 3.2.2 and Section 3.2.3. C-(A)-S-H and CH can immobilize heavy metals on the gelling particles through chemical bonding and physical adsorption/wrapping microstructure [79,85,86]. The lowest CS-4 value is attributed more to C-(A)-S-H and CH products (Section 3.2.3), confirmed by its macroscopic uniaxial compressive strength. (2) Calcareous aluminate neutralizes heavy metal ions within the needles by chemical replacement. There is ample evidence that calcarenite can trap heavy metal elements such as Cd and Cu [87]. (3) The solidification/stabilization of Cr by PCCPB is due to the physical adsorption of hydration products such as C-(A)-S-H gels and sodalite. In addition, hydration refines the pore structure and blocks the transport channels of Cr ions, facilitating solidification/stabilization [88]. Calcite, calcium silicate, calcium hydroxide, and calcium silicate may play a key role in immobilizing As species and heavy metals in PCCPB [89].

4. Conclusions

This study comprehensively evaluated the properties of a novel PCCPB prepared from PCG, FA, activator, low-dose cement, and water. The effects of different activator dosages on the mechanical properties, microstructure, and leaching risk of PCCPB were investigated. Based on the experimental results, the following main conclusions can be drawn:
(1)
The UCS of C-S2, C-S3, and C-S4 can meet the requirements of backfill mining, among which the UCS of C-S3 is 0.545 and 4.312 MPa at 3 d and 28 d, respectively. Na2SO4 has an excitation effect on PCCPB at different curing times, and the UCS of PCCPB increases and then decreases with the increase in Na2SO4. The 3% Na2SO4 has the best excitation effect on the later strength (28 d) of PCCPB.
(2)
The main hydration products of PCCPB are C-(A)-S-H gel and calcium alumina (AFt), and the effect of different Na2SO4 dosages on the content and micromorphology of the gelling products of PCCPB is the main reason for changing its mechanical properties and leaching risk.
(3)
All groups’ (control group and CS1–CS4 group) leachate heavy metal ions meet the groundwater class III standard requirements, and PCCPB meets the environmental safety requirements. Based on the mechanical properties and leaching results, C-S3 was the best ratio.
(4)
Preheated decarburization (combustion temperature 900 °C, time 4 min) is an effective method of CG activation, and with PCG, FA, activator, low-dose cement, and water, preparation of the novel PCCPB is feasible.

Author Contributions

Conceptualization, B.Z.; data curation, R.T.; formal analysis, R.T. and C.T.; investigation, L.L., X.S. and W.R.; methodology, R.T.; project administration, B.Z.; resources, L.L. and X.S.; software, B.X., C.T. and R.T.; supervision, R.T.; validation, B.Z. and X.S.; visualization, B.X.; writing—original draft, R.T.; writing—review and editing, C.T. and B.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research has the support of the Natural Science Basic Research Project of Shaanxi Province (Grant No. 2023-JC-QN-0399) and the National Natural Science Foundation of China (Grant No. 52074208).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data used to support the findings of this study are included in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qian, M.; Miao, X.; Xu, J. Green Mining of Coal Resources Harmonizing with Environment. J. China Coal Soc. 2007, 1, 1–7. [Google Scholar]
  2. Xiu, Z.; Wang, S.; Ji, Y.; Wang, F.; Ren, F.; Nguyen, V.T. Loading Rate Effect On the Uniaxial Compressive Strength (UCS) Behavior of Cemented Paste Backfill (CPB). Constr. Build. Mater. 2021, 271, 121526. [Google Scholar] [CrossRef]
  3. Huang, G.; Ji, Y.; Li, J.; Hou, Z.; Dong, Z. Improving Strength of Calcinated Coal Gangue Geopolymer Mortars Via Increasing Calcium Content. Constr. Build. Mater. 2018, 166, 760–768. [Google Scholar] [CrossRef]
  4. Ma, H.; Zhu, H.; Wu, C.; Chen, H.; Sun, J.; Liu, J. Study On Compressive Strength and Durability of Alkali-Activated Coal Gangue-Slag Concrete and its Mechanism. Powder Technol. 2020, 368, 112–124. [Google Scholar] [CrossRef]
  5. Qin, L.; Gao, X. Properties of Coal Gangue-Portland Cement Mixture with Carbonation. Fuel 2019, 245, 1–12. [Google Scholar] [CrossRef]
  6. Wu, Y.; Yu, X.; Hu, S.; Shao, H.; Liao, Q.; Fan, Y. Experimental Study of the Effects of Stacking Modes On the Spontaneous Combustion of Coal Gangue. Process Saf. Environ. Protect. 2018, 123, 39–47. [Google Scholar] [CrossRef]
  7. Guo, S.; Zhang, J.; Li, M.; Zhou, N.; Song, W.; Wang, Z.; Qi, S. A Preliminary Study of Solid-Waste Coal Gangue Based Biomineralization as Eco-Friendly Underground Backfill Material: Material Preparation and Macro-Micro Analyses. Sci. Total Environ. 2021, 770, 145241. [Google Scholar] [CrossRef]
  8. Guo, Z.; Qiu, J.; Jiang, H.; Xing, J.; Sun, X.; Ma, Z. Flowability of Ultrafine-Tailings Cemented Paste Backfill Incorporating Superplasticizer: Insight From Water Film Thickness Theory. Powder Technol. 2021, 381, 509–517. [Google Scholar] [CrossRef]
  9. Zhang, J.; Zhang, Q.; Spearing, A.J.S.; Miao, X.; Guo, S.; Sun, Q. Green Coal Mining Technique Integrating Mining-Dressing-Gas Draining-Backfilling-Mining. Int. J. Min. Sci. Technol. 2017, 27, 17–27. [Google Scholar] [CrossRef]
  10. Zhang, Q.; Wang, Z.; Zhang, J.; Jiang, H.; Wang, Y.; Yang, K.; Tian, X.; Yuan, L. Integrated Green Mining Technology of “Coal Mining-Gangue Washing-Backfilling-Strata Control-System Monitoring”-Taking Tangshan Mine as a Case Study. Environ. Sci. Pollut. Res. Int. 2021, 29, 5798–5811. [Google Scholar] [CrossRef]
  11. Liu, C.; Deng, X.; Liu, J.; Hui, D. Mechanical Properties and Microstructures of Hypergolic and Calcined Coal Gangue Based Geopolymer Recycled Concrete. Constr. Build. Mater. 2019, 221, 691–708. [Google Scholar] [CrossRef]
  12. Bai, G.; Zhu, C.; Liu, C.; Liu, B. An Evaluation of the Recycled Aggregate Characteristics and the Recycled Aggregate Concrete Mechanical Properties. Constr. Build. Mater. 2020, 240, 117978. [Google Scholar] [CrossRef]
  13. Khudyakova, T.M.; Kolesnikova, O.G.; Zhanikulov, N.N.; Botabaev, N.E.; Kenzhibaeva, G.S.; Iztleuov, G.M.; Suigenbaeva, A.Z.; Kutzhanova, A.N.; Ashirbaev, H.A.; Kolesnikova, V.A. Low-Basicity Cement, Problems and Advantages of its Utilization. Refract. Ind. Ceram. 2021, 62, 369–374. [Google Scholar] [CrossRef]
  14. Kolesnikova, O.; Syrlybekkyzy, S.; Fediuk, R.; Yerzhanov, A.; Nadirov, R.; Utelbayeva, A.; Agabekova, A.; Latypova, M.; Chepelyan, L.; Volokitina, I.; et al. Thermodynamic Simulation of Environmental and Population Protection by Utilization of Technogenic Tailings of Enrichment. Materials 2022, 15, 6980. [Google Scholar] [CrossRef]
  15. Kolesnikova, O.; Vasilyeva, N.; Kolesnikov, A.; Zolkin, A. Optimization of Raw Mix Using Technogenic Waste to Produce Cement Clinker. Min. Inf. Anal. Bull. 2022, 60, 103–115. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Ling, T. Reactivity Activation of Waste Coal Gangue and its Impact On the Properties of Cement-Based Materials—A Review. Constr. Build. Mater. 2020, 234, 117424. [Google Scholar] [CrossRef]
  17. Li, C.; Wan, J.; Sun, H.; Li, L. Investigation On the Activation of Coal Gangue by a New Compound Method. J. Hazard. Mater. 2010, 179, 515–520. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Zhang, Z.; Zhu, M.; Cheng, F.; Zhang, D. Decomposition of Key Minerals in Coal Gangues During Combustion in O2/N2 and O2/Co2 Atmospheres. Appl. Therm. Eng. 2018, 148, 977–983. [Google Scholar] [CrossRef]
  19. Xu, B.; Liu, Q.; Ai, B.; Ding, S.; Frost, R.L. Thermal Decomposition of Selected Coal Gangue. J. Therm. Anal. Calorim. 2018, 131, 1413–1422. [Google Scholar] [CrossRef]
  20. Hao, R.; Li, X.; Xu, P.; Liu, Q. Thermal Activation and Structural Transformation Mechanism of Kaolinitic Coal Gangue From Jungar Coalfield, Inner Mongolia, China. Appl. Clay Sci. 2022, 223, 106508. [Google Scholar] [CrossRef]
  21. Li, L.; Zhang, Y.; Zhang, Y.; Sun, J.; Hao, Z. The Thermal Activation Process of Coal Gangue Selected From Zhungeer in China. J. Therm. Anal. Calorim. 2016, 126, 1559–1566. [Google Scholar] [CrossRef]
  22. Song, X.; Gong, C.; Li, D. Study On Structural Characteristic and Mechanical Property of Coal Gangue in Activation Process. J. Chin. Ceram. Soc. 2004, 3, 358–363. [Google Scholar]
  23. Frías, M.; de Rojas, M.I.S.; García, R.; Valdés, A.J.; Medina, C. Effect of Activated Coal Mining Wastes On the Properties of Blended Cement. Cem. Concr. Compos. 2012, 34, 678–683. [Google Scholar] [CrossRef]
  24. Rabovitser, J.; Bryan, B.; Knight, R.; Nester, S.; Ake, T. Development and Testing of a Novel Coal Preheating Technology for Nox Reduction From Pulverized Coal-Fired Boilers. Gas 2003, 1, 4. [Google Scholar]
  25. Man, C.; Zhu, J.; Ouyang, Z.; Liu, J.; Lyu, Q. Experimental Study On Combustion Characteristics of Pulverized Coal Preheated in a Circulating Fluidized Bed. Fuel Process. Technol. 2018, 172, 72–78. [Google Scholar] [CrossRef]
  26. Pan, F.; Zhu, J.; Liu, J. Experimental Study and Numerical Simulation of Preheating Combustion in Circulating Fluidized Bed. Clean Coal Technol. 2021, 27, 180–188. [Google Scholar]
  27. Lyu, Q.; Wang, J.; Zhu, J. Experimental Study On Combustion Characteristics of Anthracite Pulverized Coal Preheated by Circulation Fluidized Bed. Boil. Technol. 2011, 42, 23–27. [Google Scholar]
  28. Zhu, J.; Ouyang, Z.; Lu, Q. An Experimental Study On Nox Emissions in Combustion of Pulverized Coal Preheated in a Circulating Fluidized Bed. Energy Fuels 2013, 27, 7724–7729. [Google Scholar] [CrossRef]
  29. Zhu, S.; Lyu, Q.; Zhu, J.; Liang, C. Experimental Study On No X Emissions of Pulverized Bituminous Coal Combustion Preheated by a Circulating Fluidized Bed. J. Energy Inst. 2018, 92, 247–256. [Google Scholar] [CrossRef]
  30. Lv, Z.; Xiong, X.; Yu, S.; Tan, H.; Xiang, B.; Huang, J.; Peng, J.; Li, P. Experimental Investigation On No Emission of Semi-Coke Under High Temperature Preheating Combustion Technology. Fuel 2021, 283, 119293. [Google Scholar] [CrossRef]
  31. Wang, S.; Gong, Y.; Niu, Y.; Hui, S. Study On No Formation During Preheating-Combustion Coupling of Pulverized Coal. Proc. CSEE 2020, 40, 2951–2959. [Google Scholar]
  32. Shi, Z.; Wang, G.; Wang, X.; Chen, Y.; Yu, W.; Miao, Y.; Peng, Y.; Tan, H. Study On Combustion Characteristics of Preheating Decarburization Process for Fine Slag in Coal Gasification. Clean Coal Technol. 2021, 27, 105–110. [Google Scholar]
  33. Shao, X.; Tian, C.; Li, C.; Fang, Z.; Zhao, B.; Xu, B.; Ning, J.; Li, L.; Tang, R. The Experimental Investigation On Mechanics and Damage Characteristics of the Aeolian Sand Paste-Like Backfill Materials Based On Acoustic Emission. Materials 2022, 15, 7235. [Google Scholar] [CrossRef] [PubMed]
  34. Zhou, N.; Dong, C.; Ouyang, S.; Deng, X.; Du, E. Feasibility Study and Performance Optimization of Sand-Based Cemented Paste Backfill Materials. J. Clean Prod. 2020, 259, 120798. [Google Scholar] [CrossRef]
  35. Deng, X.; Zhang, J.; Klein, B.; Zhou, N.; De Wit, B. Experimental Characterization of the Influence of Solid Components On the Rheological and Mechanical Properties of Cemented Paste Backfill. Int. J. Miner. Proc. 2017, 168, 116–125. [Google Scholar] [CrossRef]
  36. Shao, X.; Sun, J.; Xin, J.; Zhao, B.; Sun, W.; Li, L.; Tang, R.; Tian, C.; Xu, B. Experimental Study On Mechanical Properties, Hydration Kinetics, and Hydration Product Characteristics of Aeolian Sand Paste-Like Materials. Constr. Build. Mater. 2021, 303, 124601. [Google Scholar] [CrossRef]
  37. HJ 557-2010; Institute of Solid Waste Pollution Control Technology, Solid Waste-Extraction Procedure for Leaching Toxicity-Horizontal Vibration Method. China Environmental Science Press: Beijing, China, 2010.
  38. Huang, Z.; Su, X.; Zhang, J.; Luo, D.; Chen, Y.; Li, H. Study On Leaching of Heavy Metals From Red Mud-Phosphogypsum Composite Materials. Inorg. Chem. Ind. 2022, 54, 133–140. [Google Scholar]
  39. Bhattacharyya, G.K.; Johnson, R.A. Statistical Concepts and Methods; Wiley & Sons: Hoboken, NJ, USA, 1977. [Google Scholar]
  40. Behera, S.K.; Ghosh, C.N.; Mishra, D.P.; Singh, P.; Mishra, K.; Buragohain, J.; Mandal, P.K. Strength Development and Microstructural Investigation of Lead-Zinc Mill Tailings Based Paste Backfill with Fly Ash as Alternative Binder. Cem. Concr. Compos. 2020, 109, 103553. [Google Scholar] [CrossRef]
  41. Qi, T.; Feng, G.; Guo, Y.; Zhang, Y.; Ren, A.; Kang, L.; Guo, J. Experimental Study On the Changes of Coal Paste Backfilling Material Performance During Hydration Process. J. Min. Saf. Eng. 2015, 32, 42–48. [Google Scholar]
  42. Shao, X.; Wang, L.; Li, X.; Fang, Z.; Zhao, B.; Tao, Y.; Liu, L.; Sun, W.; Sun, J. Study On Rheological and Mechanical Properties of Aeolian Sand-Fly Ash-Based Filling Slurry. Energies 2020, 13, 1266. [Google Scholar] [CrossRef] [Green Version]
  43. Liu, L.; Fang, Z.; Qi, C.; Zhang, B.; Guo, L.; Song, K. Experimental Investigation On the Relationship Between Pore Characteristics and Unconfined Compressive Strength of Cemented Paste Backfill. Constr. Build. Mater. 2018, 179, 254–264. [Google Scholar] [CrossRef]
  44. Liu, L.; Xin, J.; Qi, C.; Jia, H.; Song, K. Experimental Investigation of Mechanical, Hydration, Microstructure and Electrical Properties of Cemented Paste Backfill. Constr. Build. Mater. 2020, 263, 120137. [Google Scholar] [CrossRef]
  45. Phuong, T.B.; Yuko, O.; Kenji, K. Effect of Sodium Sulfate Activator On Compressive Strength and Hydration of Fly-Ash Cement Pastes. J. Mater. Civ. Eng. 2020, 32, 04020117. [Google Scholar]
  46. Li, W.; Fall, M. Sulphate Effect On the Early Age Strength and Self-Desiccation of Cemented Paste Backfill. Constr. Build. Mater. 2016, 106, 296–304. [Google Scholar] [CrossRef]
  47. Ercikdi, B.; Baki, H.; İzki, M. Effect of Desliming of Sulphide-Rich Mill Tailings On the Long-Term Strength of Cemented Paste Backfill. J. Environ. Manag. 2013, 115, 5–13. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, L.; Xin, J.; Feng, Y.; Zhang, B.; Song, K.I. Effect of the Cement–Tailing Ratio On the Hydration Products and Microstructure Characteristics of Cemented Paste Backfill. Arab. J. Sci. Eng. 2019, 44, 6547–6556. [Google Scholar] [CrossRef]
  49. Rao, Y.; Shao, Y.; Huang, Y.; Sun, X.; Li, Y.; Li, J. Effect of Superplasticizer On the Elastic Modulus of Super Fine Tailings Filling Body. Min. Res. Dev. 2016, 36, 31–35. [Google Scholar]
  50. Li, X.; Liu, C. Mechanical Properties and Damage Constitutive Model of High Water Material at Different Loading Rates. Adv. Eng. Mater. 2018, 20, 1701098. [Google Scholar] [CrossRef]
  51. Choe, G.; Kang, S.; Kang, H. Mechanical Properties of Concrete Containing Liquefied Red Mud Subjected to Uniaxial Compression Loads. Materials. 2020, 13, 854. [Google Scholar] [CrossRef] [Green Version]
  52. Andoni, A.; Delilaj, E.; Ylli, F.; Taraj, K.; Korpa, A.; Xhaxhiu, K.; Çomo, A. FTIR Spectroscopic Investigation of Alkali-Activated Fly Ash: Atest Study. Zaštita Mater. 2018, 59, 539–542. [Google Scholar] [CrossRef]
  53. Criado, M.; Fernández-Jiménez, A.; Palomo, A. Alkali Activation of Fly Ash: Effect of the SiO2/Na2O Ratio Part I: FTIR Study. Microporous Mesoporous Mat. 2007, 106, 180–191. [Google Scholar] [CrossRef]
  54. Nguyen, H.T. Evaluation On Formation of Aluminosilicate Network in Ternary-Blended Geopolymer Using Infrared Spectroscopy. Solid State Phenom. 2019, 296, 99–104. [Google Scholar] [CrossRef]
  55. Zhang, N.; Li, H.; Zhao, Y.; Liu, X. Hydration Characteristics and Environmental Friendly Performance of a Cementitious Material Composed of Calcium Silicate Slag. J. Hazard. Mater. 2016, 306, 67–76. [Google Scholar] [CrossRef]
  56. Xin, J.; Liu, L.; Xu, L.; Wang, J.; Yang, P.; Qu, H. A Preliminary Study of Aeolian Sand-Cement-Modified Gasification Slag-Paste Backfill: Fluidity, Microstructure, and Leaching Risks. Sci. Total Environ. 2022, 830, 154766. [Google Scholar] [CrossRef]
  57. Lecomte, I.; Henrist, C.; Liégeois, M.; Maseri, F.; Rulmont, A.; Cloots, R. (Micro)-Structural Comparison Between Geopolymers, Alkali-Activated Slag Cement and Portland Cement. J. Eur. Ceram. Soc. 2006, 26, 3789–3797. [Google Scholar] [CrossRef]
  58. Ouellet, S.; Bussière, B.; Aubertin, M.; Benzaazoua, M. Microstructural Evolution of Cemented Paste Backfill: Mercury Intrusion Porosimetry Test Results. Cem. Concr. Res. 2007, 37, 1654–1665. [Google Scholar] [CrossRef]
  59. Abdul-Hussain, N.; Fall, M. Unsaturated Hydraulic Properties of Cemented Tailings Backfill that Contains Sodium Silicate. Eng. Geol. 2011, 123, 288–301. [Google Scholar] [CrossRef]
  60. Liu, S.; Wang, L.; Li, Q.; Song, J. Hydration Properties of Portland Cement-Copper Tailing Powder Composite Binder. Constr. Build. Mater. 2020, 251, 118882. [Google Scholar] [CrossRef]
  61. Scrivener, K.L.; John, V.M.; Gartner, E.M. Eco-Efficient Cements: Potential Economically Viable Solutions for a Low-CO2 Cement-Based Materials Industry. Cem. Concr. Res. 2018, 114, 2–26. [Google Scholar] [CrossRef]
  62. Quanji, Z. Thixotropic Behavior of Cement-Based Materials: Effect of Clay and Cement Types. Master’s Thesis, Iowa State University, Ames, IA, USA, 2010. [Google Scholar]
  63. Zhang, S.; Niu, D.; Luo, D. Enhanced Hydration and Mechanical Properties of Cement-Based Materials with Steel Slag Modified by Water Glass. J. Mater. Res. Technol. 2022, 21, 1830–1842. [Google Scholar] [CrossRef]
  64. Feng, J.; Sun, J. A Comparison of the 10-Year Properties of Converter Steel Slag Activated by High Temperature and an Alkaline Activator. Constr. Build. Mater. 2020, 234, 116948. [Google Scholar] [CrossRef]
  65. Mota, B.; Matschei, T.; Scrivener, K. Impact of NaOH and Na2SO4 On the Kinetics and Microstructural Development of White Cement Hydration. Cem. Concr. Res. 2018, 108, 172–185. [Google Scholar] [CrossRef]
  66. Sun, Q.; Tian, S.; Sun, Q.; Li, B.; Cai, C.; Xia, Y.; Wei, X.; Mu, Q. Preparation and Microstructure of Fly Ash Geopolymer Paste Backfill Material. J. Clean Prod. 2019, 225, 376–390. [Google Scholar] [CrossRef]
  67. Kong, X.; Lu, Z.; Zhang, C. Recent Development On Understanding Cement Hydration Mechanism and Effects of Chemical Admixtures On Cement Hydration. J. Chin. Ceram. Soc. 2017, 45, 274–281. [Google Scholar]
  68. Justnes, H.; Sellevold, E.J.; Lundevall, G. High Strength Concrete Binders. Part a: Reactivity and Composition of Cement Pastes with and without Condensed Silica Fume. In Proceedings of the 4th International Conference on Fly Ash, Silica Fume, Slag and Natural Pozzolana in Concrete, Istanbul, Turkey, 3–8 May 1992. [Google Scholar]
  69. Kipkemboi, B.; Zhao, T.; Miyazawa, S.; Sakai, E.; Hirao, H. Effect of C3S Content of Clinker On Properties of Fly Ash Cement Concrete. Constr. Build. Mater. 2020, 240, 117840. [Google Scholar] [CrossRef]
  70. Liu, L.; Ruan, S.; Qi, C.; Zhang, B.; Tu, B.; Yang, Q.; Song, K.I.I.L. Co-Disposal of Magnesium Slag and High-Calcium Fly Ash as Cementitious Materials in Backfill. J. Clean Prod. 2021, 279, 123684. [Google Scholar] [CrossRef]
  71. Razali, N.N.; Sukardi, M.A.; Sopyan, I.; Mel, M.; Salleh, H.M.; Rahman, M.M. The Effects of Excess Calcium On the Handling and Mechanical Properties of Hydrothermal Derived Calcium Phosphate Bone Cement. IOP Conf. Ser. Mater. Sci. Eng. 2018, 290, 012053. [Google Scholar] [CrossRef]
  72. Zhao, Y.; Qiu, J.; Zhang, S.; Guo, Z.; Ma, Z.; Sun, X.; Xing, J. Effect of Sodium Sulfate On the Hydration and Mechanical Properties of Lime-Slag Based Eco-Friendly Binders. Constr. Build. Mater. 2020, 250, 118603. [Google Scholar] [CrossRef]
  73. Somna, K.; Jaturapitakkul, C.; Kajitvichyanukul, P.; Chindaprasirt, P. Naoh-Activated Ground Fly Ash Geopolymer Cured at Ambient Temperature. Fuel 2011, 90, 2118–2124. [Google Scholar] [CrossRef]
  74. Salih, M.A.; Ali, A.A.A.; Farzadnia, N. Characterization of Mechanical and Microstructural Properties of Palm Oil Fuel Ash Geopolymer Cement Paste. Constr. Build. Mater. 2014, 6, 592–603. [Google Scholar] [CrossRef]
  75. Ouyang, S.; Huang, Y.; Zhou, N.; Li, J.; Gao, H.; Guo, Y. Experiment On Hydration Exothermic Characteristics and Hydration Mechanism of Sand-Based Cemented Paste Backfill Materials. Constr. Build. Mater. 2022, 318, 125870. [Google Scholar] [CrossRef]
  76. Roy, D.M.; Ldorn, G.M. Hydration, Structure, and Properties of Blast Furnace Slag Cements, Mortars, and Concrete. J. Proc. 1982, 79, 444–457. [Google Scholar]
  77. Rachid, H.S.; Walid, M.; Mahfoud, B.; Richard, L.; Keith, T.; Agnes, Z.; NorEdine, A. Mechanical Properties and Microstructure of Low Carbon Binders Manufactured From Calcined Canal Sediments and Ground Granulated Blast Furnace Slag (GGBS). Sustainability 2021, 13, 9057. [Google Scholar]
  78. Poon, C.; Azhar, S.; Anson, M.; Wong, Y. Strength and Durability Recovery of Fire-Damaged Concrete After Post-Fire-Curing. Cem. Concr. Res. 2001, 31, 1307–1318. [Google Scholar] [CrossRef]
  79. Tang, P.; Chen, W.; Xuan, D.; Cheng, H.; Poon, C.S.; Tsang, D.C.W. Immobilization of Hazardous Municipal Solid Waste Incineration Fly Ash by Novel Alternative Binders Derived From Cementitious Waste. J. Hazard. Mater. 2020, 393, 122386. [Google Scholar] [CrossRef]
  80. Ding, Z.; Zhou, J.; Su, Q.; Wang, Q.; Sun, H. Mechanical Properties of Geopolymer Recycled Aggregate Concrete. J. Shenyang Jianzhu Univ. Nat. Sci. 2021, 37, 138–146. [Google Scholar]
  81. Ding, Z.; Hong, X.; Zhu, J.; Tian, B.; Fang, Y. Alkali-Activated Red Mud-Slag Cementitious Materials. J. Chin. Electron Microsc. Soc. 2018, 37, 145–153. [Google Scholar]
  82. Liu, J.; Li, Z.; Zhang, M.; Wang, S.; Hai, R. Mechanical Property and Polymerization Mechanism of Red Mud Geopolymer Cement. J. Build. Mater. 2022, 25, 178–183. [Google Scholar]
  83. Tang, R.; Zhao, B.; Li, C.; Xin, J.; Xu, B.; Tian, C.; Ning, J.; Li, L.; Shao, X. Experimental Study On the Effect of Fly Ash with Ammonium Salt Content On the Properties of Cemented Paste Backfill. Constr. Build. Mater. 2023, 369, 130513. [Google Scholar] [CrossRef]
  84. Türkel, S. Long-Term Compressive Strength and some Other Properties of Controlled Low Strength Materials Made with Pozzolanic Cement and Class C Fly Ash. J. Hazard. Mater. 2006, 137, 261–266. [Google Scholar] [CrossRef]
  85. Chen, X.; Guo, Y.; Ding, S.; Zhang, H.; Xia, F.; Wang, J.; Zhou, M. Utilization of Red Mud in Geopolymer-Based Pervious Concrete with Function of Adsorption of Heavy Metal Ions. J. Clean Prod. 2019, 207, 789–800. [Google Scholar] [CrossRef]
  86. Li, B.; Zhang, S.; Li, Q.; Li, N.; Yuan, B.; Chen, W.; Brouwers, H.J.H.; Yu, Q. Uptake of Heavy Metal Ions in Layered Double Hydroxides and Applications in Cementitious Materials: Experimental Evidence and First-Principle Study. Constr. Build. Mater. 2019, 222, 96–107. [Google Scholar] [CrossRef]
  87. Peng, D.; Wang, Y.; Liu, X.; Tang, B.; Zhang, N. Preparation, Characterization, and Application of an Eco-Friendly Sand-Fixing Material Largely Utilizing Coal-Based Solid Waste. J. Hazard. Mater. 2019, 373, 294–302. [Google Scholar] [CrossRef] [PubMed]
  88. Luo, Z.; Zhi, T.; Liu, L.; Mi, J.; Zhang, M.; Tian, C.; Si, Z.; Liu, X.; Mu, Y. Solidification/Stabilization of Chromium Slag in Red Mud-Based Geopolymer. Constr. Build. Mater. 2022, 316, 125813. [Google Scholar] [CrossRef]
  89. Bah, A.; Jin, J.; Ramos, A.O.; Bao, Y.; Ma, M.; Li, F. Arsenic(V) Immobilization in Fly Ash and Mine Tailing-Based Geopolymers: Performance and Mechanism Insight. Chemosphere 2022, 306, 135636. [Google Scholar] [CrossRef]
Figure 1. Equipment preparation process. (a) Site operation drawing of equipment; (b) flow chart of preheating decarburization.
Figure 1. Equipment preparation process. (a) Site operation drawing of equipment; (b) flow chart of preheating decarburization.
Materials 16 02354 g001
Figure 2. Particle size distribution of PCG, FA, and AS.
Figure 2. Particle size distribution of PCG, FA, and AS.
Materials 16 02354 g002
Figure 3. SEM result. (a) PCG; (b) FA.
Figure 3. SEM result. (a) PCG; (b) FA.
Materials 16 02354 g003
Figure 4. XRD result of raw materials.
Figure 4. XRD result of raw materials.
Materials 16 02354 g004
Figure 5. Experimental flowchart.
Figure 5. Experimental flowchart.
Materials 16 02354 g005
Figure 6. Effect of Na2SO4 dosage on the UCS of PCCPB.
Figure 6. Effect of Na2SO4 dosage on the UCS of PCCPB.
Materials 16 02354 g006
Figure 7. The EM of PCCPB with different Na2SO4 dosages.
Figure 7. The EM of PCCPB with different Na2SO4 dosages.
Materials 16 02354 g007
Figure 8. Fitting relationship between elastic modulus and UCS of PCCPB.
Figure 8. Fitting relationship between elastic modulus and UCS of PCCPB.
Materials 16 02354 g008
Figure 9. FTIR results of PCCPB samples. (a) 3 d; (b) 28 d; (c) C-S2.
Figure 9. FTIR results of PCCPB samples. (a) 3 d; (b) 28 d; (c) C-S2.
Materials 16 02354 g009
Figure 10. TG-DTG results of PCCPB samples. (a) 3d sample mass and temperature relationship (b) 3d sample mass loss rate of change and temperature (c) 28d sample mass and temperature relationship (d) 28d sample mass loss rate of change and temperature Table 4 reflects the weight loss rates of PCCPB with various Na2SO4 dosages, and the magnitude of the weight loss rate indirectly reflects the number of hydration products in a given temperature interval [62]. In the I zone, the weight loss of free water, C-(A)-S-H, and AFt in PCCPB was the first to increase and then to decrease with the increase in Na2SO4 dosage. The most significant weight loss in the PCCPB samples for curing time of 3 d was C-S2, which was 3.45% (zone I), of which the free water weight loss represented a relatively large ratio. The most significant weight loss in the PCCPB samples for a curing time of 28 d was C-S3, which was 3.84% (zone I), with C-(A)-S-H and AFt accounting for a relatively large proportion of the weight loss. This is consistent with the findings in Section 3.1 and Section 3.2.4. In addition, mass loss in zone I was found to be higher for all samples at 28 d than at 3 d, indicating that PCCPB continued hydration with increasing age. This is consistent with the studies of Zhang et al. [63] and Feng et al. [64].
Figure 10. TG-DTG results of PCCPB samples. (a) 3d sample mass and temperature relationship (b) 3d sample mass loss rate of change and temperature (c) 28d sample mass and temperature relationship (d) 28d sample mass loss rate of change and temperature Table 4 reflects the weight loss rates of PCCPB with various Na2SO4 dosages, and the magnitude of the weight loss rate indirectly reflects the number of hydration products in a given temperature interval [62]. In the I zone, the weight loss of free water, C-(A)-S-H, and AFt in PCCPB was the first to increase and then to decrease with the increase in Na2SO4 dosage. The most significant weight loss in the PCCPB samples for curing time of 3 d was C-S2, which was 3.45% (zone I), of which the free water weight loss represented a relatively large ratio. The most significant weight loss in the PCCPB samples for a curing time of 28 d was C-S3, which was 3.84% (zone I), with C-(A)-S-H and AFt accounting for a relatively large proportion of the weight loss. This is consistent with the findings in Section 3.1 and Section 3.2.4. In addition, mass loss in zone I was found to be higher for all samples at 28 d than at 3 d, indicating that PCCPB continued hydration with increasing age. This is consistent with the studies of Zhang et al. [63] and Feng et al. [64].
Materials 16 02354 g010
Figure 11. XRD results of PCCPB. (a) 3 d; (b) 28 d; (c) C-S3.
Figure 11. XRD results of PCCPB. (a) 3 d; (b) 28 d; (c) C-S3.
Materials 16 02354 g011aMaterials 16 02354 g011b
Figure 12. SEM results of PCCPB.
Figure 12. SEM results of PCCPB.
Materials 16 02354 g012
Figure 13. SEM results of C-S3 group samples.
Figure 13. SEM results of C-S3 group samples.
Materials 16 02354 g013
Table 1. Main chemical compositions of PCG, FA, AS, and OPC.
Table 1. Main chemical compositions of PCG, FA, AS, and OPC.
Chemical CompositionOPCASPCGFA
Al2O35.53%10.30%20.41%17.14%
SiO222.36%67.80%53.55%41.01%
CaO65.08%5.30%3.76%14.03%
Fe2O33.46%5.80%8.99%14.47%
K2O0.62%7.50%1.62%5.36%
Mg2O1.27%1.78%3.65%4.31%
TiO20.59%0.35%0.73%1.60%
SO3//1.27%0.711%
Others1.09%1.17%5.62%0.83%
Table 2. Mixing proportions of PCCPB.
Table 2. Mixing proportions of PCCPB.
Group NumberConcentration (CO) a,
%
Cement Dosage (CD) b, wt.%AS Dosage (AD) d, wt.%FA/PCGFA+PCG
Content (FC) c, wt.%
Activator Content (AC) e, wt.%Activator Type
Control78%3%507/347//
C-S1 f461Na2SO4
C-S2452
C-S3443
C-S4434
a CO: M A D + M C D + M F C + M A C M A D + M C D + M F C + M A C + M water × 100 % b CD: M C D M A D + M C D + M F C + M A C × 100 % . c FC: M F A + M P C G M A D + M C D + M F A + M P C G + M A C × 100 % . d AD: M A D M A D + M C D + M F C + M A C × 100 % . e AC: M A C M A D + M C D + M F C + M A C × 100 % . f C-S1: represents a concentration of 78%, an AS dosage of 50%, a cement dosage of 3%, a Na2SO4 content of 1%, an FA and PCG content of 46%, and an FA/PCG ratio of 7/3.
Table 3. The statistical analysis of UCS results.
Table 3. The statistical analysis of UCS results.
SampleUCS, MPaR, %UCS, MPaR,%UCS, MPaR,%UCS, MPaR, %
3 d7 d14 d28d
No. of Sample (n): 3No. of Sample (n): 3No. of Sample (n): 3No. of Sample (n): 3
MeanSDCV MeanSDCV MeanSDCV MeanSDCV
Control0.1190.0054.12/0.7990.0374.68/1.80.0864.79/2.8910.0852.96/
C-S10.3760.0153.99215.972.4380.0522.13205.132.9240.1093.7262.443.40.0752.2217.61
C-S20.5920.0315.28397.482.4710.0642.61209.263.4580.0361.0592.113.8630.1273.3033.62
C-S30.5450.0081.44357.982.0570.038 1.83157.453.2840.0662.0282.444.3120.0952.2049.15
C-S40.5270.0132.56342.861.680.0955.67110.263.1710.0892.8076.173.4370.0732.1318.89
Table 4. Weight loss ratio of PCCPB with different Na2SO4 dosages.
Table 4. Weight loss ratio of PCCPB with different Na2SO4 dosages.
Temperature Range, °CWeight Change (3 d, 28 d), %
ControlC-S1C-S2C-S3C-S4
I (30–250)−2.25, −3.19−3.11, −3.76−3.45, −3.64−3.3, −3.84−3.31, −3.43
II (250–650)−2.87, −1.06−1.66, −1.47−1.68, −1.34−1.75, −1.24−1.57, −2.72
III (650–750)−0.80, −1.11−0.31, −0.20−0.19, −0.12−0.22, −0.35−0.16, −0.45
IV (750–900)−0.42, −0.20−0.04, −0.51−0.03, −0.27−0.03, −0.35−0.06, −0.12
Table 5. Leaching results of ARFGB.
Table 5. Leaching results of ARFGB.
Control, mg/LC-S1, mg/LC-S2, mg/LC-S3, mg/LC-S4, mg/LLimit, mg/L
MnND 1.2 × 10−4ND 1.2 × 10−4ND 1.2 × 10−4ND 1.2 × 10−4ND 1.2 × 10−40.1
ZnND 6.7 × 10−4ND 6.7 × 10−4ND 6.7 × 10−4ND 6.7 × 10−4ND 6.7 × 10−41
As7.5 × 10−36.8 × 10−36.2 × 10−34.7 × 10−37.0 × 10−30.01
CdND 5.0 × 10−5ND 5.0 × 10−5ND 5.0 × 10−5ND 5.0 × 10−5ND 5.0 × 10−50.005
Hg1.8 × 10−41.4 × 10−41.1 × 10−48.1 × 10−51.3 × 10−40.001
PbND 9.0 × 10−5ND 9.0 × 10−5ND 9.0 × 10−5ND 9.0 × 10−5ND 9.0 × 10−50.01
Cr0.018.4 × 10−36.2 × 10−34.3 × 10−35.2 × 10−30.05
CuND 8.0 × 10−5ND 8.0 × 10−5ND 8.0 × 10−5ND 8.0 × 10−5ND 8.0 × 10−51
BaND 2.0 × 10−4ND 2.0 × 10−4ND 2.0 × 10−4ND 2.0 × 10−4ND 2.0 × 10−40.7
NiND 6.0 × 10−5ND 6.0 × 10−5ND 6.0 × 10−5ND 6.0 × 10−5ND 6.0 × 10−50.02
AgND 4.0 × 10−5ND 4.0 × 10−5ND 4.0 × 10−5ND 4.0 × 10−5ND 4.0 × 10−50.05
Se2.3 × 10−31.8 × 10−31.5 × 10−31.1 × 10−31.6 × 10−30.01
Mo0.020.0170.0140.0120.0130.07
Sb1.9 × 10−31.6 × 10−31.4 × 10−31.1 × 10−31.3 × 10−30.005
CoND 3.0 × 10−51ND 3.0 × 10−5ND 3.0 × 10−5ND 3.0 × 10−5ND 3.0 × 10−50.05
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

Tang, R.; Zhao, B.; Tian, C.; Xu, B.; Li, L.; Shao, X.; Ren, W. Preliminary Study of Preheated Decarburized Activated Coal Gangue-Based Cemented Paste Backfill Material. Materials 2023, 16, 2354. https://doi.org/10.3390/ma16062354

AMA Style

Tang R, Zhao B, Tian C, Xu B, Li L, Shao X, Ren W. Preliminary Study of Preheated Decarburized Activated Coal Gangue-Based Cemented Paste Backfill Material. Materials. 2023; 16(6):2354. https://doi.org/10.3390/ma16062354

Chicago/Turabian Style

Tang, Renlong, Bingchao Zhao, Chuang Tian, Baowa Xu, Longqing Li, Xiaoping Shao, and Wuang Ren. 2023. "Preliminary Study of Preheated Decarburized Activated Coal Gangue-Based Cemented Paste Backfill Material" Materials 16, no. 6: 2354. https://doi.org/10.3390/ma16062354

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