Next Article in Journal
Techno-Economic Considerations on Nanocellulose’s Future Progress: A Short Review
Next Article in Special Issue
Flotation Characteristics of Amphibole-Type Oxidized Iron Ore via Reverse Anionic Flotation
Previous Article in Journal
Soy Protein Nanofibers Obtained by Solution Blow Spinning
Previous Article in Special Issue
Changes in Surface Hydrophobicity of Coal Particles and the Formation of Coarse Particle–Bubble Clusters in the Process of High-Intensity Conditioning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Mechanism of MSWI Fly Ash Solidified by Multiple Solid Waste-Based Cementitious Material Using the Rietveld Method

1
School of Civil and Resources Engineering, University of Science and Technology Beijing, Beijing 100083, China
2
BGRIMM Technology Group, State Key Laboratory of Mineral Processing, Beijing 102628, China
3
Cangzhou Municipal Engineering Company Limited, Cangzhou 061000, China
4
Road Materials and Technology Engineering Research Center of Hebei Province, Cangzhou 061000, China
*
Author to whom correspondence should be addressed.
Processes 2023, 11(8), 2311; https://doi.org/10.3390/pr11082311
Submission received: 28 June 2023 / Revised: 22 July 2023 / Accepted: 26 July 2023 / Published: 1 August 2023
(This article belongs to the Special Issue Process Analysis and Carbon Emission of Mineral Separation Processes)

Abstract

:
A novel multiple solid waste-based cementitious material (MSWCM) was developed to immobilize municipal solid waste incineration (MSWI) fly ash. The compressive strength of MSWCM with different ratios of MSWI fly ash reached the standard requirements after curing for 28 days. X-ray powder diffraction (XRD) in combination with the Rietveld method was employed to investigate the content and phase transformation of hydration products. The main hydration products of pure MSWCM paste were C-S-H, hydroxyapatite, ettringite and C-A-S-H. With increases in curing time, the content of ettringite and C-A-S-H increased significantly. The main hydration products of MSWCM paste with MSWI fly ash were C-S-H and Friedel’s salt. The contents increased markedly with increased curing time from 21.8% to 28.0% and from 8.53% to 16.7%, respectively. Additionally, a small amount of PbHPO4 (0.51–0.96%) and lead phosphate Pb3(PO4)2 (0.14–0.51%) were detected, indicating that phosphate had an effective curing effect on lead ions. The results showed that most of the hydration reactions had started at the initial stage of curing and reacted quickly to form a large number of hydration products. The quantitative analyses of hydration products provide essential information for understanding the immobilization mechanism of MSWI fly ash in MSWCM paste.

1. Introduction

Urbanization, population growth and industrialization have led to a rapid increase in municipal solid waste. By 2025, cities worldwide are expected to generate 2.2 billion tons of municipal solid waste [1]. Landfill and incineration are the two main waste disposal methods. Compared with landfill, the incineration method can save land resources. After incineration, the volume of municipal waste can be significantly reduced, and the heat generated by the complete incineration of organic matter can be used for heating and power generation, realizing the resource utilization of waste [2,3,4,5]. Therefore, the incineration method has been commonly used in the treatment of municipal waste. However, it tends to produce a large amount of municipal solid waste incineration (MSWI) fly ash. The level of MSWI has reached 146 million tons in China in 2020 and the production of MSWI fly ash has exceeded 7 million tons [6]. MSWI fly ash is produced during waste incineration. It contains a large amount of hazardous substances, such as leachable heavy metals, dioxins, soluble salts and other harmful components, which may cause harm to the environment and human health [7,8,9,10]. It can used in landfill only after solidification or stabilization [11,12,13].
Cement-based stabilization or solidification technology has been considered as the main treatment option for MSWI fly ash [14,15,16,17]. Ordinary Portland Cement (OPC) is a widely used cementitious binder for the treatment of MSWI fly ash [18,19]. However, OPC will form a strong alkaline environment that is not conducive to the solidification of heavy metals. The compressive strength of OPC–fly ash blocks may decrease significantly in the case of high waste incineration fly ash content. [20,21,22]. In addition, during the production process of Portland cement, some problems will become more prominent, such as pollutant discharge, excessive consumption of energy resources and damage to the balance of the ecological environment [23,24]. Therefore, it is necessary to develop a low carbon environment-friendly binder material.
Recently, due to excellent mechanical properties and good environmental protection, the cementitious material produced from multiple solid wastes, such as metallurgical slags and industrial solid wastes, has been widely used to treat various hazardous wastes including MSWI fly ash [25,26]. Blast furnace slag (BFS) and steel slag (SS) are commonly used metallurgical slags for solid waste-based cementitious material, and the main components are CaO, SiO2 and Al2O3 [27]. BFS and SS can be used as an alternative material for alkaline activator because they are alkaline, which can provide Ca2+ ions and OH ions for the hydration reaction [28]. Desulfurization ash (DA) is a solid waste generated during the dry flue gas desulfurization process, with CaO and SO3 as the main components [29,30]. Phosphoric acid sludge (PAS) is an industrial solid waste obtained from the production of wet phosphoric acid [31,32]. PAS can be used to immobilize MSWI fly ash due to the widespread use of phosphate as a curing agent [33,34,35]. Additionally, Pb can easily react with PO43− and HPO42− to form insoluble phosphate precipitation [33]. Therefore, phosphate shows a good solidification effect on Pb. These industrial solid wastes can be used as the raw materials of multiple solid waste-based cementitious material (MSWCM).
During the hydration process, heavy metals react with MSWCM through physical cementation, chemical adsorption, isomorphic replacement and complex precipitation. Finally, they settle in the cement hydration product in the form of hydroxide or complex [33]. The quantitative and qualitative analyses of mineral phases provide essential information for understanding the immobilization mechanism of MSWI fly ash in MSWCM paste. However, the study of assessing the mineral composition of fly ash and hydration products by conventional X-ray powder diffraction (XRD) has been hampered due to the complex phases and structures, which result in strong overlap of diffraction peaks [36]. Therefore, it is necessary to find a method that can determine the relative content of the mixture phase and hydration products accurately, effectively and conveniently in MSWCM.
XRD in combination with the Rietveld refinement method, is a powerful technique for the characterization of mineral phases. The Rietveld method is basically a profile-fitting method based on the least square approach, which can deal with complex diffraction patterns with strong overlapping peaks [37]. However, few studies have focused on using the Rietveld methods to study changes in the content of hydration products and mineral phases during hydration process.
In this study, blast furnace slag (BFS), steel slag (SS), desulfurization ash (DA) and phosphoric acid sludge (PAS) are used as raw materials to produce a novel multiple solid waste-based cementitious material (MSWCM). The development of the mechanical properties of MSWCM with different ratios of MSWI fly ash, including compressive strength and leachability at different curing times, is investigated. The Rietveld method is employed for the qualitative and quantitative analysis of the MSWI fly ash and of the hydration products in pure MSWCM pastes and pastes containing MSWI fly ash. Meanwhile, the changes in the hydration products and mineral phase content during the hydration process and the immobilization mechanism of MSWI fly ash in MSWCM pastes are also discussed.

2. Materials and Methods

2.1. Materials

Raw materials for the preparation of multiple solid waste-based cementitious material (MSWCM) included blast furnace slag (BFS), steel slag (SS), desulfurization ash (DA) and phosphoric acid sludge (PAS). The BFS, SS and DA were provided by Wuhan Metal Resources Company in Hubei Province, China. The PAS was collected from China City Environment Protection Engineering Limited Company. The chemical compositions and heavy metal content of the four raw materials were summarized in a previous study [27]. CaO, SiO2 and Al2O3 were the major chemical compositions. Furthermore, the desulfurization ash contained 33.28% SO3, the steel slag contained 26.12% Fe2O3 and the phosphoric acid sludge contained 21.80% P2O5. It had a high content of Pb (2412.81 mg/kg), which was the main heavy metal.

2.2. Orthogonal Experiment and Analysis

2.2.1. Orthogonal Experimental Design

Orthogonal experimental design is a method to study multi-factor and multi-level experiments. An orthogonal experiment was designed to optimize the mass ratio of the raw materials for the production of the MSWCM. The mass proportions of SS, DA and PAS were set as the three factors in the orthogonal experiment. The level settings of each factor are shown in Table 1. The mass proportion of BFS was calculated by [100% − (SS% + DA% + PAS%)]. The standard orthogonal experimental design table L16 (43) was used and a total of 16 themes were tested. The orthogonal experimental design is displayed in Table 2.

2.2.2. Range Analysis Method

In this study, the range analysis method was used to determine the factors’ sensitivity to compressive strength according to the orthogonal experiment [38]. The range method is to subtract the minimum value from the maximum value in the data. In the calculation, the average value ki and the influence degree R are as follows [39]:
k i = K i x  
R = m a x k 1 , k 2 ,   k 3   m i n k 1 , k 2 ,   k 3  
where, ki is the arithmetic mean value of the corresponding test results at level i; Ki is the sum of the corresponding test results; and x is the number of tests on level i. The influence degree R represents the influence of each factor on the index. The higher the R value, the greater the influence of the factor on the experiment results.

2.3. Compressive Strength Test

Compressive strength was measured in an electro-hydraulic universal testing machine (WAW-1000B) in compliance with the GB/T17671 [27] China standard. The measurements were performed on specimens for 3, 7 and 28 days at room temperature and at least three representative samples were tested to obtain the average values.

2.4. Leaching Test

The leaching toxicity of the heavy metals was evaluated by the Solid Waste-Extraction Procedure for Leaching Toxicity—Horizontal Vibration Method (HJ 557-2010). The samples were first crushed to below 3 mm, then 100 g was taken and put into a 2 L volumetric flask, and we added 1000 mL of deionized water. After that, the mixtures were put into a horizontal vibrator for 8 h with a frequency of 110 times per minute. After standing still for 16 h, the concentrations of heavy metals in the leachate were determined by inductively coupled plasma mass spectrometry (ICP-MS).

2.5. X-ray Powder Diffraction (XRD) Analysis

The phase analysis of the samples was determined by the X-ray powder diffraction (XRD) method. The XRD patterns were collected by an Rigaku D/Max-B diffractometer with an X’Celerator detector and using Cu Kα radiation (λ = 0.154056 nm), 40 kV and 40 mA. The measured range of the powder specimens was set from 5 to 80° 2θ, with a 0.04° step size.

2.6. Rietveld Method

The Rietveld method is an advantageous technique for the refinement of crystal structures and phase quantitative analysis. This method was employed to determine the major phases of the MSWI fly ash and the hydration products during solidification in this study. The principle of the Rietveld method is to minimize the difference between the observed and calculated intensities by a least square approach (Equation (3)). The BGMN program was used for the Rietveld refinement [40,41]:
i w i y i obs   y i calc 2 = Minimum  
where yi (obs) and yi (calc) are the observed and calculated intensities at point i, and wi is the weight assigned to each intensity. The intensity at each point is calculated by summing the contributions from all the neighboring Bragg reflections and background scattering:
y i = S k   F hkl 2 ·   G ( Δ 2 θ ) ·   Lp θ ·   M hkl ·   T f ( θ ) ·   A ·   Po hkl + y b i  
where y i intensity of the angular position i in the powder pattern;
S —scale factor of the phase;
F hkl —structure factor;
G( Δ 2 θ)—profile shape function;
Lp ( θ ) —Lorentz and Polarization factor;
M hkl —multiplicity factor;
T f ( θ ) —temperature factor;
A —absorption factor;
Po hkl —correction factor of preferred orientation;
y bi —background intensity.

3. Results and Discussion

3.1. Characterization of MSWI Fly Ash

Municipal solid waste incineration (MSWI) fly ash was collected from MSWI plants in Wuhan, China. The main chemical composition and heavy metal contents are given in Table 3. According to the results, CaO accounted for the largest proportion of MSWI fly ash, accounting for 39.08%, followed by Cl (21.5%) and Na2O (7.03%). This is probably due to the addition of alkali to neutralize the acidic gases generated during incineration, such as HCl and SO2 [42]. The high content of chloride ion may be attributed to the incineration of plastics and halite in food residues, e.g., NaCl [43,44]. The most abundant heavy metals are Pb and Zn, with a content of 1769.9 mg/kg and 5636.2 mg/kg, respectively.
The phase content of the MSWI fly ash was determined by the Rietveld method. Figure 1 presents the Rietveld refinement plot for the MSWI fly ash. The measured patterns showed good agreement with the calculated patterns. The Rwp value was 10.25%. According to the results, the main phases in the MSWI fly ash were NaCl (34.2%), CaClOH (24.6%), Ca(OH)2 (16.5%) and KCl (14.3%) (Table 4). Moreover, CaCO3 and CaSO4 were also detected.

3.2. Orthogonal Experiment Results Analysis

The range analysis method was adopted for the orthogonal experiment results. The R value (Range) is used to divide the impact degree of different factors on the experimental objectives. Figure 2 and Table 5 present the results of the orthogonal experiment and range analysis. According to the K value, the optimal combination of factors and levels is SS2 DA3 PAS1, and their corresponding mass proportions are 32%, 12% and 20%, respectively. The mass proportion of BFS is calculated as 36%. As shown in Table 5, the order of the R value is PAS > SS > DA, indicating that the mass ratio of PAS is the dominant factor affecting the compressive strength, followed by the mass ratios of SS and DA.
Figure 3 shows the effect curves of different raw materials on the compressive strength. Based on the results, the influence of the PAS ratio on the compressive strength is the greatest. The compressive strength decreases with increases in the PAS mass ratio. However, the compressive strength shows a tendency to first increase and then decrease with increases in the DA mass ratio. The effect of the SS mass ratio on the compressive strength is fluctuating.

3.3. Effect of MSWI Fly Ash Proportions on Compressive Strength

The compressive strength of multiple solid waste-based cementitious material (MSWCM) with different amounts of MSWI fly ash is plotted in Figure 3. The compressive strength increased with curing time for all specimens and, after curing for 3, 7 and 28 days, the compressive strength varied from 1.2 to 33.4 MPa, which all met the standard requirements of Chinese technical standards for the solidification/stabilization treatment of MSWI fly ash. The compressive strength of the MSWCM pastes without MSWI fly ash had reached 27.2, 28.6 and 33.4 MPa after curing for 3, 7 and 28 days, respectively. It was noticed that the 7-day compressive strength was 81.4% of the 28-day strength, indicating a rapid early strength development.
According to Figure 4, the compressive strength showed a tendency to decrease with increases in MSWI fly ash content. In comparison with the MSWCM pastes after a 28-day curing period, the compressive strength of the MSWCM pastes decreased by 4.2%, 17.3%, 27.5%, 45.5%, 57.5%, 72.2%, 82.6% and 91.3%, respectively, with an increase in MSWI fly ash content from 10% to 80%. The addition of MSWI fly ash had a negative impact on the compressive strength of the MSWCM pastes, mainly due to the lower activity of MSWI fly ash and the high level of heavy metals, which inhibited the hydration reaction and weakened the compressive strength [42,45,46].

3.4. Leaching Toxicity

The leaching concentrations of As, Cr, Pb and Zn of the MSWCM pastes with MSWI fly ash at 3, 7 and 28 curing days are given in Figure 5. With increases in MSWI fly ash content, the leaching concentration of As, Cr and Pb increased. The pure MSWCM pastes showed a low leaching concentration of these four heavy metals. The maximum leaching concentrations of As, Cr, Pb and Zn were 22.6 µg/L, 12.5 µg/L,113.0 µg/L and 75.6 µg/L, respectively, which were far below the Toxicity Characteristic Leaching Procedure (TCLP) standard requirement and the threshold value specified in Standard GB16889-2008. This indicated that MSWCM can effectively immobilize these heavy metals.
MSWCM after curing with MSWI fly ash can be used as a supplementary cementitious material, replacing or partially replacing traditional cement and other cementitious materials. This type of cementitious material can be widely used in various applications, such as roadbed materials, landfill closure and ground hardening.

3.5. Characterization of MSWCM Hydration Products with and without MSWI Fly Ash

3.5.1. XRD Analysis of Hydration Products

The XRD patterns of the pure MSWCM pastes (MF0) and the MSWCM pastes with 30% MSWI fly ash (MF1) at a curing time of 3, 7 and 28 days are shown in Figure 6 and the main mineral phases have been labelled. According to the diffraction pattern, the diffraction peak with the strongest intensity in the pure MSWCM pastes belonged to calcium carbonate, and the main hydration products were hydroxyapatite, ettringite and calcium silicate hydrate (C-S-H and C-A-S-H).
In contrast, strong peaks for chlorine salts such as halite and sylvite were observed in the samples mixed with 70% MSWCM and 30% MSWI fly ash. Additionally, some new hydrates such as Friedel’s salt (3CaO·Al2O3·CaCl2·10H2O), PbHPO4 and Pb3(PO4)2 were formed. The diffraction peaks intensity of the Friedel’s salt increased significantly with increases in curing time. Meanwhile, the diffraction peaks of the halite (NaCl) and sylvite (KCl) gradually weakened with the increase of Friedel’s salt. It can be deduced that chloride salts might be partially transformed into Friedel’s salt, which indicated that part of the Cl in MSWI fly ash could be encapsulated in the MSWCM solidified body. The generation of PbHPO4 and Pb3(PO4)2 in sample MF1 indicated that the Pb took part in a chemical reaction with phosphate and then formed insoluble lead phosphate minerals. This conclusion was consistent with [42] and confirmed that phosphate shows a good solidification effect on Pb.
Furthermore, calcium-containing hydration products (C-S-H and C-A-S-H) were found in both the MF0 and MF1 samples, and the intensity of these diffraction peaks in sample MF1 is significantly stronger than that in sample MF0. However, no distinct peaks of portlandite were observed in the MF0 and MF1 samples. It is indicated that these hydration products were further hydrated with the components in the MSWI fly ash.

3.5.2. Immobilization Mechanism Analysis using Rietveld Method

Although the main mineral composition can be obtained by traditional XRD analysis, it is difficult to identify minor minerals and their content, and the accuracy of the results is not high. The qualitative and quantitative phases of the MSWCM pastes with MSWI fly ash are crucial for understanding the hydration process and the hydration mechanism. Therefore, the Rietveld method was employed in this work to investigate the mineral composition and content and the mineral phase transformation of the hydration products.
The mineralogical compositions and quantitative analysis of the hydration products in the MSWCM pastes (MF0) without or with MSWI fly ash (samples MF0 and MF1) at a curing time of 3, 7 and 28 days are presented in Table 6 and Table 7. The quality of the Rietveld fitting was evaluated by the weighted profile factor Rwp and the expected factor Rexp. Figure 7 shows the refinement results of samples MF0 and MF1 after curing for 28 days. The calculated patterns of these two samples showed comparatively good agreement with the measured patterns. The refinements of samples MF0 and MF1 after curing different days achieved Rwp values in the range of 7.35% to 7.55% and 7.70% to 7.81%, and Rexp values in the range of 5.85% to 5.90% and 6.01% to 6.12%, respectively. It is indicated that the refinement results for these samples were acceptable and reasonable.
  • Pure MSWCM Paste
As shown in Table 6, the pure MSWCM paste without MSWI fly ash (sample MF0) contained C2S (17.85–20.55%) and calcite (12.59–14.71%). The main hydration products of sample MF0 were C-S-H (Ca5Si6O16(OH)·4H2O), hydroxyapatite (Ca5(PO4)3(OH)) and ettringite (3CaO·Al2O3·3CaSO4·32H2O).
The hydration reaction of C2S and C3S in the raw material produced calcium silicate hydrate (C-S-H) and calcium hydroxide (Equations (5) and (6)). The hydration reaction of C2S is much slower than that of C3S. Therefore, after 28 days of curing, there was still unreacted C2S. The content of C-S-H and hydroxyapatite at different curing times was in the range of 19.8–24.7% and 11.29–19.87%, respectively. It can be seen from the results that a large amount of C-S-H and hydroxyapatite have been formed when the curing time is 3 days. This is very beneficial to the improvement of compressive strength and the solidification of MSWI fly ash. The generated alkaline activator Ca(OH)2 can destroy and dissolve the silica vitreous structure and increase the amount of Ca2+, Al3+ and SiO42−. This may accelerate the hydration reaction, resulting in higher early compressive strength. With the continuous accumulation of various ions in the liquid phase, the solution reaches the supersaturated state of new hydration products, thus generating new hydration products. For example, Ca(OH)2 further reacts with Al2O3 and CaSO4 to form ettringite and a small amount of C-A-S-H (3CaO·Al2O3·2SiO2·2H2O) (Equations (7) and (8)). The generated C-A-S-H content was 2.35–4.71%. Compared with the C-A-S-H, the content of generated ettringite was relatively large, and with the prolonging of the curing time the content increased significantly. Its content were 4.15%, 17.9% and 16.9% at a curing time of 3, 7 and 28 days, respectively. It indicated that the hydration process was constantly evolving.
In addition, PO43− produced from phosphoric acid sludge (PAS) reacted with Ca2+ in the aqueous solution to form a new hydration product, hydroxyapatite (Ca5(PO4)3(OH)) (Equation (9)). The content of hydroxyapatite was 19.87%, 13.40% and 11.29% at a curing time of 3, 7 and 28 days, respectively. The possible hydration reactions that occurred in the pure MSWCM paste samples were listed as follows:
2 C 3 S + 6 H 2 O     3 C - S - H + 3 Ca ( OH ) 2
2 C 2 S + 4 H 2 O     3 C - S - H + Ca ( OH ) 2
3 Ca ( OH ) 2 + Al 2 O 3 + 3 CaSO 4 · H 2 O + 23 H 2 O     3 CaO · Al 2 O 3 · 3 CaSO 4 · 32 H 2 O
Ca ( OH ) 2 + Al 2 O 3   + SiO 2 + H 2 O     CaO · SiO 2 · Al 2 O 3 · ( n + 1 ) H 2 O   ( C - A - S - H )
5 Ca 2 + + 4 OH - + 3 HPO 4 2 -     Ca 5 PO 4 3 OH + H 2 O
Some heavy metal oxides were detected at very low levels, such as PbO (0.17–0.3%) and PbO2 (0.43–0.52%). And, after curing for 28 days, a small amount of mayenite (12CaO·7Al2O3) (0.55%) appeared.
Table 6. Quantitative results of sample MF0 using the Rietveld method.
Table 6. Quantitative results of sample MF0 using the Rietveld method.
MineralsChemical FormulaRietveld Quantitative Results
Contents/Mass %
3 d7 d28 d
Rwp = 7.55%
Rexp = 5.88%
Rwp = 7.35%
Rexp = 5.85%
Rwp = 7.36%
Rexp = 5.90%
CalciteCaCO314.70 (1) a13.78 (1)12.59 (1)
Ettringite3CaO·Al2O3·3CaSO4·32H2O4.15 (1)17.90 (1)16.90 (1)
C2S2CaO·SiO220.10 (1)20.55 (1)17.85 (1)
C-S-HCa5Si6O16(OH)·4H2O24.70 (2)17.80 (2)19.80 (2)
GypsumCaSO4·2H2O4.08 (1)3.09 (1)4.23 (1)
HydroxyapatiteCa5(PO4)3(OH)19.87 (1)13.40 (1)11.29 (1)
PortlanditeCa(OH)2000
Merwinite12CaO·7Al2O39.79 (1)9.37 (1)11.44 (1)
MayeniteCa12Al14O3300.39 (2)0.55 (2)
PbO-0.30 (1)0.16 (1)0.21 (1)
PbO2-00.48 (1)0.47 (1)
C-A-S-H3CaO·Al2O3·2SiO2·2H2O2.35 (1)3.07 (1)4.71 (1)
sum 100.04100.00100.04
a Standard deviation.
  • MSWCM Paste with MSWI Fly Ash
The mineral phase quantitative analysis of MSWCM paste with MSWI fly ash (sample MF1) under different curing times using the Rietveld method is listed in Table 7. It can be seen from the results that the phase composition is relatively complex. The dominant mineral phases were C-S-H, halite, sylvite, Friedel’s salt, K3PO4, hydroxyapatite, AlPO4, calcite and gypsum, while the minor mineral phases were portlandite, periclase, mayenite, ettringite, grossite, katoite, PbO2, CaCl2, PbHPO4 and lead phosphate (Table 7).
The main hydration products of sample MF1 were C-S-H (Ca5Si6O16(OH)·4H2O) and Friedel’s salt (3CaO·Al2O3·CaCl2·10H2O). The content of C-S-H and Friedel’s salt increased significantly with increases in curing time and their contents were 21.8%, 26.2% and 28.0% and 8.53%, 12.68% and 16.17% after the curing time of 3, 7 and 28 days, respectively. This result is consistent with the intensity variation of the diffraction peaks (Figure 6). At the same time, the amount of halite (NaCl) and sylvite (KCl) is decreasing, further illustrating the effect of Friedel’s salt on the solidification of the chloride ions in the MSWI fly ash. C-S-H has been proved to immobilize heavy metals (such as Zn, Cd and Pb) effectively [33]. With increases in curing time, the content of C-S-H increased, the stabilization efficiency should be better. This is consistent with the leaching results shown above (Figure 5). Additionally, the more C-S-H generated, the more beneficial it is to improve the compressive strength. This is consistent with the growth trend in Figure 3. In addition, there was a small amount of C-A-S-H (3CaO·Al2O3·2SiO2·2H2O) generated, the content of which is 2.54–4.22% at different curing times.
Halite (NaCl) and sylvite (KCl) were the main phases in the MSWI fly ash (Figure 1). NaCl can react with calcium aluminate hydrate to form Friedel’s salt (3CaO·Al2O3·CaCl2·10H2O) (Equation (10)) and potassium ions may form potassium phosphate (K3PO4) (Equation (11)). The content of halite and sylvite decreased with increases in curing days (Table 7). This suggested that some chloride ions in MSWI fly ash could be encapsulated in MSWCM paste by a chemical bonding effect.
A small amount of PbHPO4 (0.51–0.96%) and lead phosphate Pb3(PO4)2 (0.14–0.51%) were detected. Irene et al. and Ren et al. [35,47] found that the presence of phosphate was beneficial for the formation of insoluble lead phosphate salts (Equations (12) and (13)). Compared with other heavy metal ions, such as Zn, Cr and As, the phosphate ions reacted preferentially with Pb [48]. Due to the very low solubility of lead phosphate salts, Pb can be stably solidified in MSWCM paste and have a good curing effect. The quantitative change of the main hydration products is shown in Figure 8.
Al 2 O 3 + 4 CaO + 2 NaCl + 11 H 2 O     3 CaO · Al 2 O 3 · CaCl 2 · 10 H 2 O + NaOH
PO 4 3 - + 3 OH - + 3 K +   K 3 PO 4 + 3 H 2 O
2 PO 4 3 - + 3 Pb 2 +   Pb 3 PO 4 2
HPO 4 2 - + Pb 2 +   PbHPO 4
Table 7. Quantitative results of sample MF1 using the Rietveld method.
Table 7. Quantitative results of sample MF1 using the Rietveld method.
MineralsChemical FormulaRietveld Quantitative Results
Contents/Mass %
3 d7 d28 d
Rwp = 7.81%
Rexp = 6.07%
Rwp = 7.77%
Rexp = 6.12%
Rwp = 7.70%
Rexp = 6.12%
PortlanditeCa(OH)20.96 (2) a0.17 (1)0
C-S-HCa5Si6O16(OH)·4H2O21.8 (1)26.2 (1)28.0 (1)
PericlaseMgO0.49 (2)2.32 (2)1.43 (3)
HaliteNaCl11.22 (3)12.27 (3)6.76 (3)
SylviteKCl4.57 (2)5.05 (2)1.62 (2)
MayeniteCa12Al14O330.86 (4)00
GypsumCaSO4·2H2O5.52 (4)5.17 (4)5.79 (4)
Ettringite3CaO·Al2O3·3CaSO4·32H2O02.16 (1)0.49 (1)
Friedel‘s salt3CaO·Al2O3·CaCl2·10H2O8.53 (5)12.68 (5)16.17 (6)
GrossiteCaAl4O71.36 (3)2.09 (5)1.08 (3)
CalciteCaCO36.57 (5)5.22 (5)9.84 (5)
K3PO4-8.56 (3)6.54 (3)8.62 (3)
PbO2-0.68 (1)0.27 (1)0.41 (1)
AlPO4-7.07 (1)7.57 (1)6.71 (1)
HydroxyapatiteCa5(PO4)3(OH)7.95 (5)6.68 (45)8.2 (6)
CaCl2-0.28 (1)0.52 (1)0.39 (1)
PbHPO4-0.96 (1)0.73 (1)0.51 (1)
Pb3(PO4)2-0.19 (1)0.14 (1)0.51 (1)
C-A-S-H3CaO·Al2O3·2SiO2·2H2O2.54 (5)4.22 (4)3.44 (4)
CarnalliteKCl·MgCl2·6H2O9.87 (6)--
Sum 100.0100.0100.0
a Standard deviation.

4. Conclusions

This study proposed the preparation of multiple solid waste-based cementitious material (MSWCM) from blast furnace slag (BFS), steel slag (SS), desulfurization ash (DA) and phosphoric acid sludge (PAS). The Rietveld method was applied for the qualitative and quantitative analysis of the hydration products of pure MSWCM pastes and MSWCM pastes with MSWI fly ash. The immobilization mechanism of MSWI fly ash by MSWCM binders was explored in depth. The main conclusions can be drawn as follows:
(a)
The optimal mass ratios of BFS:SS:DA:PAS for MSWCM preparation were 36:32:12:20 (w.t.%). The compressive strength of MSWCM with 80% MSWI fly ash content was 2.9 MPa during the 28-day curing period, which met the standard requirements. The leachability of As, Pb, Cr and Zn were far below Toxicity Characteristic Leaching Procedure (TCLP) standard requirement and the threshold value specified in Standard GB16889-2008.
(b)
The XRD analysis showed that the phase composition of the hydrated product was relatively complex. The most obvious diffraction peak in the pure MSWCM pastes was calcium carbonate and the main hydration products were hydroxyapatite, ettringite and calcium silicate hydrate (C-S-H and C-A-S-H). In contrast, strong peaks for chlorine salts such as halite and sylvite were observed in the mixed sample of 70% MSWCM and 30% MSWI fly ash. Additionally, small amounts of lead phosphate salts (PbHPO4 and Pb3(PO4)2) were formed.
(c)
The refinements of the samples MF0 and MF1 got low Rwp and Rexp values, which indicated that the refinement results were acceptable and reasonable. Most of the hydration reactions had started at the initial stage of curing and reacted quickly to form a large amount of hydration products. For the pure MSWCM paste, the main hydration products were C-S-H, hydroxyapatite and ettringite. With increased curing time, the content of ettringite increased obviously. For the MSWCM paste with MSWI fly ash, the content of C-S-H and Friedel’s salt increased significantly with increased curing time. During a curing period of 3 days to 28 days, the content of C-S-H and Friedel’s salt increased by 22% and 90%, respectively. Phosphate ions reacted with Pb to form a small amount of PbHPO4 and Pb3(PO4)2. It was further confirmed that phosphate had a good curing effect on lead ions. The hydration process could be deduced and evidenced by the changes in the content of the hydration products.

Author Contributions

Conceptualization, X.W. and P.F.; methodology, investigation, data curation, X.W., W.D., J.S. and M.X.; writing—original draft preparation, X.W.; writing—review and editing, X.W. and P.F.; supervision, project administration, P.F.; funding acquisition, X.W. and P.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Open Foundation of State Key Laboratory of Mineral Processing (grant numbers BGRIMM-KJSKL-2022-22), the Key Research and Development Program of Hebei Province (grant numbers 22373809D) and the National Key Research and Development Program of China (grant numbers 2018YFC1900604).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Song, J.; Sun, Y.; Jin, L. PESTEL analysis of the development of the waste-to-energy incineration industry in China. Renew. Sustain. Energ. Rev. 2017, 80, 276–289. [Google Scholar] [CrossRef]
  2. Liu, A.; Ren, F.; Lin, W.L.Y.; Wang, J.Y. A review of municipal solid waste environmental standards with a focus on incinerator residues. Int. J. Sustain. Built Environ. 2015, 4, 165–188. [Google Scholar] [CrossRef] [Green Version]
  3. Moon, M.H.; Kang, D.; Lim, H.; Oh, J.E.; Chang, Y.S. Continuous fractionation of fly ash particles by SPLITT for the investigation of PCDD/Fs levels in different sizes of insoluble particles. Environ. Sci. Technol. 2002, 36, 4416–4423. [Google Scholar] [CrossRef] [PubMed]
  4. Guo, Y.; Glad, T.; Zhong, Z.Z.; He, R.N.; Tian, J.P.; Chen, L.J. Environmental life-cycle assessment of municipal solid waste incineration stocks in Chinese industrial parks. Resour. Conserv. Recycl. 2018, 139, 387–395. [Google Scholar] [CrossRef]
  5. Cucchiella, F.; D’Adamo, I.; Gastaldi, M. Sustainable waste management: Waste to energy plant as an alternative to landfill. Energy Convers. Manag. 2017, 131, 18–31. [Google Scholar] [CrossRef]
  6. Zhu, Y.; Ye, Z.; Ren, L.; Zhong, Y.; Zhou, W. Experimental Study on Municipal Solid Waste Incineration Fly Ash in Conjunction with Construction Waste to Burn Ceramics. Environ. Eng. 2023, 1–11. Available online: http://kns.cnki.net/kcms/detail/11.2097.X.20230411.1337.010.html (accessed on 20 June 2023).
  7. Zhang, Y.; Wang, L.; Chen, L.; Ma, B.; Zhang, Y.; Ni, W.; Tsang, D.C.W. Treatment of municipal solid waste incineration fly ash: State-of-the-art technologies and future perspectives. J. Hazard. Mater. 2021, 411, 125–132. [Google Scholar] [CrossRef]
  8. Nikravan, M.; Ramezanianpour, A.; Maknoon, R. Study on physiochemical properties and leaching behavior of residual ash fractions from a municipal solid waste incinerator (MSWI) plant. J. Environ. Manag. 2020, 260, 110042. [Google Scholar] [CrossRef]
  9. Loginova, E.; Proskurnin, M.; Brouwers, H.J.H. Municipal solid waste incineration (MSWI) fly ash composition analysis: A case study of combined chelatant-based washing treatment efficiency. J. Environ. Manag. 2019, 235, 480–488. [Google Scholar] [CrossRef]
  10. Lombardi, F.; Mangialardi, T.; Piga, L. Mechanical and leaching properties of cement solidified hospital solid waste incinerator fly ash. Waste Manag. 1998, 18, 99–106. [Google Scholar] [CrossRef]
  11. Luo, H.W.; Cheng, Y.; He, D.; Yang, E. Review of leaching behavior of municipal solid waste incineration (MSWI) ash. Sci. Total Environ. 2019, 668, 90–103. [Google Scholar] [CrossRef]
  12. Bie, R.S.; Chen, P.; Song, X.F.; Ji, X.Y. Characteristics of municipal solid waste incineration fly ash with cement solidificationtreatment. J. Energy Inst. 2016, 89, 704–712. [Google Scholar] [CrossRef]
  13. Quina, M.J.; Bordado, J.C.; Quinta-Ferreira, R.M. Treatment and use of air pollution control residues from MSW incineration: An overview. Waste Manag. 2008, 28, 2097–2121. [Google Scholar] [CrossRef] [Green Version]
  14. Du, B.; Li, J.; Wen, F.; Liu, J. Comparison of long-term stability under natural ageing between cement solidified and chelator-stabilised MSWI fly ash. Environ. Pollut. 2019, 250, 68–78. [Google Scholar] [CrossRef]
  15. Ma, W.; Chen, D.; Pan, M.; Gu, T.; Zhong, L.; Chen, G.; Yan, B.; Cheng, Z. Performance of chemical chelating agent stabilization and cement solidification on heavy metals in mswi fly ash: A comparative study. J. Environ. Manag. 2019, 247, 169–177. [Google Scholar] [CrossRef]
  16. Leist, M.; Casey, R.J.; Caridi, D. The management of arsenic wastes: Problems and prospects. J. Hazard. Mater. 2000, 76, 125–138. [Google Scholar] [CrossRef]
  17. Chen, L.; Wang, L.; Zhang, Y.; Ruan, S.; Mechtcherine, V.; Tsang, D.C.W. Roles of biochar in cement-based stabilization/solidification of municipal solid waste incineration fly ash. Chem. Eng. J. 2022, 430, 132972. [Google Scholar] [CrossRef]
  18. Gineys, N.; Aouad, G.; Damidot, D. Managing trace elements in Portland cement—Part I: Interactions between cement paste and heavy metals added during mixing as soluble salts. Cem. Concr. Compos. 2010, 32, 563–570. [Google Scholar] [CrossRef]
  19. Tang, P. Stabilization/solidification of municipal solid waste incineration bottom ash. In Low Carbon Stabilization and Solidification of Hazardous Wastes; Tsang, D.C.W., Wang, L., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 157–174. [Google Scholar]
  20. Wang, L.; Chen, L.; Cho, D.W.; Tsang, D.C.W.; Yang, J.; Hou, D.; Baek, K.; Kua, H.W.; Poon, C.S. Novel synergy of Si-rich minerals and reactive MgO for stabilization/solidification of contaminated sediment. J. Hazard. Mater. 2019, 365, 695–706. [Google Scholar] [CrossRef]
  21. Wang, L.; Yu, K.; Li, J.S.; Tsang, D.C.W.; Poon, C.S.; Yoo, J.C.; Baek, K.; Ding, S.; Hou, D.; Dai, J.G. Low-carbon and low-alkalinity stabilization/solidification of high-Pb contaminated soil. Chem. Eng. J. 2018, 351, 418–427. [Google Scholar] [CrossRef]
  22. Dutre, V.; Vandecasteele, C. Immobilization mechanism of arsenic in waste solidified by using cement and lime. Environ. Sci. Technol. 1998, 32, 2782–2787. [Google Scholar] [CrossRef]
  23. Arthur, S.; Saitoh, M.; Pal, S.K. Advances in Civil Engineering: Select Proceedings of ICACE 2020; Springer Nature: Berlin/Heidelberg, Germany, 2021; Volume 184, pp. 55–63. [Google Scholar]
  24. Abdollahi, S.; Zarei, Z. Reduction of CO2 emission and production costs by using pozzolans in Lamerd cement factory. Iran. J. Chem. Chem. Eng. 2018, 37, 223–230. [Google Scholar]
  25. Ferreira, C.; Ribeiro, A.; Ottosen, L. Possible applications for municipal solid waste fly ash. J. Hazard. Mater. 2003, 96, 201–216. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, Y.Y.; Zhang, S.Q.; Ni, W.; Yan, Q.H.; Gao, W.; Li, Y.Y. Immobilisation of high-arsenic-containing tailings by using metallurgical slag-cementing materials. Chemosphere 2019, 223, 117–123. [Google Scholar] [CrossRef] [PubMed]
  27. Deng, W.; Fu, P.; Fang, G.; Zhu, W.; Li, S.; Wang, X.; Xue, T.; Chen, Y. Solidification/Stabilization of MSWI Fly Ash Using a Novel Metallurgical Slag-Based Cementitious Material. Minerals 2022, 12, 599. [Google Scholar] [CrossRef]
  28. Lampris, C.; Stegemann, J.A.; Cheeseman, C.R. Chloride leaching from air pollution control residues solidified using ground granulated blast furnace slag. Chemosphere 2008, 73, 1544–1549. [Google Scholar] [CrossRef]
  29. Xia, Z.R.; Lin, J.; Chi, Y.W.; Wang, J.D. Hydrothermal solidification mechanism of calcareous materials and resource utilization of desulfurization ash. Glass Phys. Chem. 2020, 46, 53–63. [Google Scholar]
  30. Wang, C.Q.; Tan, K.F.; Xu, X.X.; Wang, P.X. Effect of activators, admixtures and temperature on the early hydration performance of desulfurization ash. Constr. Build. Mater. 2014, 70, 322–331. [Google Scholar] [CrossRef]
  31. Jang, G.G.; Ladshaw, A.; Keum, J.K.; Zhang, P.; Tsouris, C. Continuous-flow centrifugal solid/liquid separation for the recovery of rare-earth elements containing particles from phosphoric acid sludge. Ind. Eng. Chem. Res. 2020, 59, 21901–21913. [Google Scholar] [CrossRef]
  32. Sugiyama, S.; Imanishi, K.; Shimoda, N.; Liu, J.C.; Satou, H.; Yamamoto, T. Recovery of phosphoric acid and calcium phosphate from dephosphorization slag. J. Chem. Eng. Jpn. 2021, 54, 467–471. [Google Scholar] [CrossRef]
  33. Fan, C.; Wang, B.; Zhang, T. Review on Cement Stabilization/Solidification of Municipal Solid Waste Incineration Fly Ash. Adv. Mater. Sci. Eng. 2018, 2018, 5120649. [Google Scholar] [CrossRef] [Green Version]
  34. Su, Y.; Yang, J.; Liu, D.; Zhen, S.; Lin, N.; Zhou, Y. Effects of municipal solid waste incineration fly ash on solidification/stabilization of cd and pb by magnesium potassium phosphate cement. J. Environ. Chem. Eng. 2016, 4, 259–265. [Google Scholar] [CrossRef]
  35. Buj, I.; Torras, J.; Rovira, M.; Pablo, J.D. Leaching behaviour of magnesium phosphate cements containing high quantities of heavy metals. J. Hazard. Mater. 2010, 175, 789–794. [Google Scholar] [CrossRef]
  36. Mahieux, P.Y.; Aubert, J.E.; Cyr, M.; Coutand, M.; Husson, B. Quantitative mineralogical composition of complex mineral wastes-contribution of the Rietveld method. Waste Manag. 2010, 30, 378–388. [Google Scholar] [CrossRef]
  37. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  38. Xia, S.; Lin, R.; Cui, X.; Shan, J. The application of orthogonal test method in the parameters optimization of PEMFC under steady working condition. Int. J. Hydrogen Energy 2016, 41, 11380–11390. [Google Scholar] [CrossRef]
  39. Taguchi, G.; Phadke, M.S. Quality Engineering through Design Optimization. In Quality Control, Robust Design, and the Taguchi Method; Springer: Berlin/Heidelberg, Germany, 1989; pp. 77–96. [Google Scholar]
  40. Taut, T.; Kleeberg, R.; Bergmann, J. The new seifert Rieveld program BGMN and its application to quantitative phase analysis. Mater. Struct. 1998, 5, 57–66. [Google Scholar]
  41. Bergmann, J.; Friedel, P.; Kleeberg, R. BGMN-a new fundamental parameters based Rietveld program for laboratory X-ray sources; it’s use in quantitative analysis and structure investigations. CPD Newsl. 1998, 20, 5–8. [Google Scholar]
  42. Fan, C.; Wang, B.; Qi, Y.; Liu, Z. Characteristics and leaching behavior of MSWI fly ash in novel solidification/stabilization binders. Waste Manag. 2021, 131, 277–285. [Google Scholar] [CrossRef]
  43. Ren, P.; Ling, T.C. Roles of chlorine and sulphate in MSWIFA in GGBFS binder: Hydration, mechanical properties and stabilization considerations. Environ. Pollut. 2021, 284, 117175. [Google Scholar] [CrossRef]
  44. Forestier, L.; Libourel, G. Characterization of flue gas residues from municipal solid waste combustors. Environ. Sci. Technol. 1998, 32, 2250–2256. [Google Scholar] [CrossRef]
  45. Chi, M.; Huang, R. Binding mechanism and properties of alkali-activated fly ash/slag mortars. Constr. Build. Mater. 2013, 40, 291–298. [Google Scholar] [CrossRef]
  46. Patil, A.A.; Chore, H.S.; Dodeb, P.A. Effect of curing condition on strength of geopolymer concrete. Adv. Concrete Constr. 2014, 2, 29–37. [Google Scholar] [CrossRef] [Green Version]
  47. Ren, J.; Zhang, Z.; Wang, M.; Guo, G.; Du, P.; Li, F. Phosphate-induced differences in stabilization efficiency for soils contaminated with lead, zinc, and cadmium. Front. Environ. Sci. Eng. 2018, 12, 10. [Google Scholar] [CrossRef]
  48. Cao, X.; Wang, W.; Ma, R.; Sun, S.; Lin, J. Solidification/stabilization of Pb(2+) and Zn(2+) in the sludge incineration residue-based magnesium potassium phosphate cement: Physical and chemical mechanisms and competition between coexisting ions. Environ. Pollut. 2019, 253, 171–180. [Google Scholar] [CrossRef]
Figure 1. Rietveld refinement plot for MSWI fly ash.
Figure 1. Rietveld refinement plot for MSWI fly ash.
Processes 11 02311 g001
Figure 2. The result of orthogonal experiment.
Figure 2. The result of orthogonal experiment.
Processes 11 02311 g002
Figure 3. Effect of different raw materials on compressive strength.
Figure 3. Effect of different raw materials on compressive strength.
Processes 11 02311 g003
Figure 4. Compressive strength of MSWCM pastes with MSWI fly ash.
Figure 4. Compressive strength of MSWCM pastes with MSWI fly ash.
Processes 11 02311 g004
Figure 5. Leaching concentrations of heavy metals in MSWCM solidified body with different MSWI fly ash content.
Figure 5. Leaching concentrations of heavy metals in MSWCM solidified body with different MSWI fly ash content.
Processes 11 02311 g005
Figure 6. XRD patterns of MF0 and MF1 samples with different curing times. The main peaks are labelled as follows: 1-Friedel’s salt (3CaO·Al2O3·CaCl2·10H2O); 2-Halite (NaCl); 3-Sylvite (KCl); 4-Gypsum (CaSO4·2H2O); 5-Calcite (CaCO3); 6-C-S-H; 7-Hydroxyapatite (Ca5(PO4)3(OH)); 8-C-A-S-H; 9-PbO2; 10-PbHPO4; 11-C2S (2CaO·SiO2); 12-Ettringite (3CaO·Al2O3·3CaSO4·32H2O); 13-Pb3(PO4)2; 14-Carnallite (KCl·MgCl2·6H2O).
Figure 6. XRD patterns of MF0 and MF1 samples with different curing times. The main peaks are labelled as follows: 1-Friedel’s salt (3CaO·Al2O3·CaCl2·10H2O); 2-Halite (NaCl); 3-Sylvite (KCl); 4-Gypsum (CaSO4·2H2O); 5-Calcite (CaCO3); 6-C-S-H; 7-Hydroxyapatite (Ca5(PO4)3(OH)); 8-C-A-S-H; 9-PbO2; 10-PbHPO4; 11-C2S (2CaO·SiO2); 12-Ettringite (3CaO·Al2O3·3CaSO4·32H2O); 13-Pb3(PO4)2; 14-Carnallite (KCl·MgCl2·6H2O).
Processes 11 02311 g006
Figure 7. Rietveld refinement of samples MF0 and MF1 after curing for 28 days.
Figure 7. Rietveld refinement of samples MF0 and MF1 after curing for 28 days.
Processes 11 02311 g007
Figure 8. The quantitative change of the main hydration products of MSWCM paste with MSWI fly ash.
Figure 8. The quantitative change of the main hydration products of MSWCM paste with MSWI fly ash.
Processes 11 02311 g008
Table 1. Factors and levels for orthogonal test.
Table 1. Factors and levels for orthogonal test.
LevelsFactors
Mass Proportions
SS/%DA/%PAS/%
128820
2321026
3361232
4401438
Table 2. Experimental design schemes of orthogonal test.
Table 2. Experimental design schemes of orthogonal test.
Sample IDLevelsMass Proportion of BFS/%
SSDAPAS
O111144
O212236
O313328
O414420
O521234
O622138
O723418
O824322
O931324
O1032416
O1133132
O1234224
O1341414
O1442318
O1543222
O1644126
Table 3. Chemical composition and heavy metal content of MSWI fly ash.
Table 3. Chemical composition and heavy metal content of MSWI fly ash.
Oxide (%)CaOSiO2Al2O3ClSO3Fe2O3Na2OK2OMgO
39.081.940.3129.177.371.467.0310.580.56
Heavy metal (mg/kg)PbZnAsCrHg
1769.95636.245.734.32.4
Table 4. Results of the Rietveld refinement of MSWI fly ash.
Table 4. Results of the Rietveld refinement of MSWI fly ash.
HaliteSylvingCalcium Chloride HydroxidePortlanditeCalciteGypsumRwp/%Rexp/%
NaClKClCaClOHCa(OH)2CaCO3CaSO4·2H2O
Phase content/w.t.%34.2 (2) a14.3 (1)24.6 (2)16.5 (1)7.6 (5)2.8 (7)10.255.26
a Standard deviation.
Table 5. The result of range analysis.
Table 5. The result of range analysis.
LevelsCompressive Strength/Mpa
SSDAPAS
K181.874.995
K283.780.781.4
K373.181.371.7
K475.877.566.3
k120.4518.7323.75
k220.9320.1820.35
k318.2820.3317.93
k418.9519.3816.58
Range2.651.67.17
RankingPAS > SS > DA
Optimum themeSS2 DA3 PAS1
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

Wang, X.; Fu, P.; Deng, W.; Shi, J.; Xu, M. Study on Mechanism of MSWI Fly Ash Solidified by Multiple Solid Waste-Based Cementitious Material Using the Rietveld Method. Processes 2023, 11, 2311. https://doi.org/10.3390/pr11082311

AMA Style

Wang X, Fu P, Deng W, Shi J, Xu M. Study on Mechanism of MSWI Fly Ash Solidified by Multiple Solid Waste-Based Cementitious Material Using the Rietveld Method. Processes. 2023; 11(8):2311. https://doi.org/10.3390/pr11082311

Chicago/Turabian Style

Wang, Xiaoli, Pingfeng Fu, Wei Deng, JinJin Shi, and Miao Xu. 2023. "Study on Mechanism of MSWI Fly Ash Solidified by Multiple Solid Waste-Based Cementitious Material Using the Rietveld Method" Processes 11, no. 8: 2311. https://doi.org/10.3390/pr11082311

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