Next Article in Journal
Interlocking Evaluation of Mesoscopic Skeleton with the Compaction Degree of Hot-Mix Asphalt
Previous Article in Journal
The Influence of Heat Aging Treatments on the Cavitation Erosion Behavior of a Type 6082 Aluminum Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Phosphoric Acid on the Preparation of α-Hemihydrate Gypsum Using Hydrothermal Method

School of Materials and Chemical Engineering, Henan University of Urban Construction, Pingdingshan 467036, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(17), 5878; https://doi.org/10.3390/ma16175878
Submission received: 5 July 2023 / Revised: 20 August 2023 / Accepted: 24 August 2023 / Published: 28 August 2023
(This article belongs to the Section Green Materials)

Abstract

:
The effects of phosphoric acid (H3PO4) in pressurized aqueous solution on the dehydration of CaSO4·2H2O to form α-hemihydrate gypsum (α-HH) phase and the regulation of crystal shape were studied in this paper in order to provide guidance for the low-cost and high-value utilization of phosphogypsum. The results showed that H3PO4 can significantly accelerate the formation rate of the α-HH phase and that it did not participate in the formation of the α-HH phase in the form of eutectic phosphorus during crystalline phase transformation. In terms of crystal shape regulation, H3PO4 can impact the effect of a citric acid crystal regulator on α-HH crystal shape regulation. The more H3PO4 added, the greater the aspect ratio of α-HH. Accordingly, the water consumption and 2 h dry compressive strength of α-HH products were gradually increased and decreased with an increase in H3PO4 content, respectively. Despite this, the compressive strength of α-HH can still meet the requirements of the α20 grade in JC/T 2038-2010 “α High Strength Gypsum” in China when the H3PO4 content was limited to less than 0.4%.

1. Introduction

Phosphogypsum is a by-product of the wet process phosphoric acid industry [1,2,3]. Approximately 5 tons of phosphogypsum are generated per ton of phosphoric acid production [4]. At present, the annual emission of phosphogypsum in the world has exceeded 280 million tons, but the comprehensive utilization rate is only about 10% [5,6]. In recent years, China, as a big country with phosphate mineral resources, produces and dissolves about 70 million tons of phosphogypsum every year, and the utilization rate is only 40% [2]. A large amount of unused phosphogypsum is mainly deposited, which not only occupies a large amount of land resources but also causes serious environmental pollution problems [7,8,9]. Therefore, it is urgent to accelerate the efficient utilization of phosphogypsum.
The main chemical composition of phosphogypsum is calcium sulfate dehydrate (CaSO4·2H2O), making up over 90% of it. Therefore, phosphogypsum can be used to effectively replace natural gypsum to produce α-hemihydrate gypsum (α-HH). This is one of the key ways to realize the high-value utilization of phosphogypsum resources and also one of the most promising and beneficial treatment methods [10,11,12,13]. Moreover, due to the powdery nature of phosphogypsum, it is very suitable to transform it into α-HH by using a hydrothermal method. [12,14,15]. However, unlike natural gypsum, phosphogypsum contains a large number of harmful impurities, such as soluble phosphorus (P2O5), soluble fluorine (F), insoluble matter, organic matter, etc. [16,17,18,19]. Many studies have shown that whether phosphogypsum is used to prepare α-HH, producing gypsum and gypsum products results in various harmful impurities in the phosphogypsum, which have a significant negative impact on its application performance in most cases. However, since the mechanisms of influence on impurities in phosphogypsum are not fully understood, they must be eliminated. In order to avoid the negative impact of the impurity in phosphogypsum, a series of methods can be implemented, such as water washing, lime neutralization, calcination, and other methods to pretreat phosphogypsum before using it. These conventional techniques greatly increase the cost of using phosphorus gypsum and are not conducive to the resource utilization of phosphogypsum [20,21,22]. Therefore, it is of great theoretical value and practical significance to deeply explore the influence of impurities in phosphogypsum on its application performance and put forward control criteria in terms of impurities to realize low-cost and efficient utilization of phosphogypsum resources.
Many studies have shown that soluble P2O5 is the most harmful among the various impurities contained in phosphogypsum. The soluble P2O5 impurities in phosphogypsum mainly come from residual phosphoric acid (H3PO4), which mainly exists in the form of H3PO4, H2PO4 and HPO42− [23,24]. The greatest harm of soluble P2O5 impurities is that it can cause the pH of the reaction liquid system to change when phosphogypsum is used to prepare α-HH, which will inevitably affect the crystallization habit of α-HH and ultimately affect the physical and mechanical properties of the product [25,26]. To avoid the harm of impurities, most current practices include the pre-purification of phosphogypsum before it is used in the production of α- HH, which significantly increases the utilization cost of phosphogypsum and is not conducive to its high-value utilization. Therefore, it is very important to explore the influence law of harmful impurities in phosphogypsum in the α-HH production process and reveal the influence mechanism of impurities, which is the key to realizing green, low-cost, and high-value utilization of phosphogypsum resources. Thus, this study synthesized α-HH by using a hydrothermal method and studied the influence of H3PO4 (P2O5) on the crystal phase transformation of calcium sulfate, the morphology characteristics of α-HH, and the physical–mechanical properties of α-HH during this process. On this basis, the influence mechanisms of H3PO4 on the preparation and performance of hemihydrate gypsum were revealed. The research in this paper can provide guidance for green, low-cost, and high-value utilization of phosphogypsum resources.

2. Materials and Methods

2.1. Materials

The pure gypsum used in the experiment was taken from Tianjin kemio Chemical Reagent Co., Ltd. (Tianjin, China), where the content of CaSO4·2H2O was ≥99%. The XRD spectrum and particle size distribution of pure gypsum are shown in Figure 1 and Figure 2, respectively. Phosphoric acid was bought from Yantai Shuangshuang Chemical Co., Ltd. (Yantai, China), which had a H3PO4 concentration of 85 wt.%. Citric acid was used as a crystal modifier, purchased from Tianjin Bodi Chemical Co., Ltd. (Tianjin, China). Anhydrous ethanol was used to wash the synthesized α-hemihydrate gypsum (α-HH) samples and immobilize the crystal form, purchased from Yansheng Chemical Co., Ltd. (Tianjin, China). The deionized water was prepared using a lab tower EDI pure water instrument.

2.2. Preparation of α-HH

A hydrothermal reactor (500 mL) was adopted to prepare α-HH products. The specific procedures were as follows:
Firstly, 120 g of pure gypsum was mixed, and 280 g of deionized water was accurately weighed and prepared in the reaction slurry with a slurry concentration of (ca.) 30%. Secondly, H3PO4 and citric acid were measured and mixed with the reaction slurry according to the experimental requirements. Afterward, the prepared slurry was placed in the reactor and sealed, and α-HH was prepared at 130 °C and under 260 r/min stirring conditions. The solid material in the reactor was fully filtered and washed with absolute ethanol immediately after reaching the set reaction time. Finally, the obtained samples were dried at 50 °C for 5 h in an oven to obtain the final α-HH product.

2.3. Experimental Methods

2.3.1. XRD Test

The prepared α-HH samples were analyzed using X, Pert Pro X-ray diffractometer (XRD), produced by Malvern Panalytical in the Netherlands. A Cu Ka emission source was used for testing. The test range was 5–60° at a scanning rate of 10 °/min. This test can clarify the crystal transformation of gypsum in the prepared sample.

2.3.2. Morphology Test

The morphology of the α-HH samples was observed using a CI-L material microscope produced by Nikon. The specific steps were as follows:
Firstly, a small number of samples were taken on the slide, and the samples were dispersed with absolute ethanol. After drying, the samples were placed on the sample table of the material microscope. The magnification of the eyepiece was adjusted to 400× to observe the morphological characteristics of the prepared α-HH, and the diameter and length of the crystal were measured and analyzed.

2.3.3. Infrared Spectrum Test

The chemical groups of prepared α-HH samples were characterized using a Fourier transform infrared spectrophotometer (FTIR; Nicolet iS10; Thermo Fisher Scientific Corporation, Waltham, MA, USA) with a scanning range from 500 cm−1 to 4000 cm−1. The aim of this test was to observe whether H3PO4 was involved in the formation of α-HH in the form of eutectic phosphorus.

2.3.4. Determination of Water of Crystal Content

In order to evaluate the effect of H3PO4 on α-HH conversion, the content of crystalline water in the samples was analyzed. The specific test process was as follows:
A certain number of prepared samples were calcined in an oven at 230 ± 5 °C to a constant weight. According to the change of sample mass before and after calcination, the content of water and crystal in the sample was calculated with Formula (1).
W = m 0 m 1 m 0 × 100 %
where:
  • W —Water of crystallization content, %;
  • m 0 —Mass of sample before calcination, g;
  • m 1 —Mass of sample after calcination, g.

2.3.5. Physical and Mechanical Properties Test

The water consumption of the standard consistency of the prepared α-HH samples was tested according to Gypsum—Determination of physical properties of pure paste (GB/T 17669.4-1999) [27]. The samples were transformed into uniform slurry and molded into test blocks of 40 mm × 40 mm × 40 mm under the water consumption of standard consistency. After 30 min, the castings were demoulded and then cured for 2 h under natural conditions. The prisms were dried to a constant weight at 60 °C and readied for the mechanical properties test. The mechanical strength was tested using an automated breaking and compression resistance tester (WAY-300, Xiyi Building Material Instrument Factory, Wuxi, China).

3. Results and Discussion

3.1. X-ray Diffraction Analysis

The formation of α-HH was a process of dissolution and recrystallization. Under the condition of pressurized water vapor, the dihydrous gypsum in phosphogypsum first removed two-thirds of crystalline water to form a α-HH crystal embryo. This was in the environment of liquid water and quickly dissolved under suitable temperature conditions, making the concentration of α-HH in the liquid phase increase continuously. When it reached saturation, the α-HH rapidly crystallized and formed. However, the acidification effect and impurity effect caused by the raw material characteristics of phosphogypsum were bound to have a certain degree of influence on the thermodynamics and kinetics of the reaction process. This, in turn, would affect the crystal phase transition process, which was obviously directly related to the reaction efficiency and would thus alter the process cost. Therefore, combined with the starting point of this study, the effect of H3PO4 on the α-HH crystal phase transformation process without a crystal regulator was initially studied.
Figure 3 shows the XRD patterns of the samples prepared under the conditions of 0.5 h, 1.0 h and 1.5 h reaction times of the reaction system with different H3PO4 contents. For the reaction system without H3PO4, a considerable amount of α-HH was detected after only 0.5 h of reaction time, but there was still a certain amount of dihydrate gypsum that had not been completely transformed. The characteristic peak of dihydrate gypsum disappeared until the sample was reacted for an extra 1.5 h, indicating that it had completely changed to α-HH at this time. The transition process of the crystal phase can be entirely completed after 0.5 h of reaction time when 0.2% H3PO4 was added, which revealed that the existence of H3PO4 can significantly improve reaction efficiency. With the increase in H3PO4 content, the time of complete conversion of dihydrate gypsum to α-HH remained within 0.5 h without significant change. The main reason why H3PO4 can accelerate the formation rate of α-HH is because H3PO4 can reduce the pH value of the reaction system. This, in turn, increases the supersaturation of α-HH and reduces the surface free energy at the interface between crystal and reaction liquid, which was conducive to the rapid formation of the α-HH phase [25]. At the same time, further analysis of Figure 3, especially Figure 3a, clearly shows that the incorporation of phosphoric acid can also affect the crystallinity of α-HH. The (400) crystal plane of α-HH was parallel to the crystal C-axis, while (204) was a crystal plane perpendicular to its C-axis. With the addition of phosphoric acid, the ratio of (400) crystal plane diffraction peak intensity to (204) crystal plane diffraction peak intensity basically increased. This showed that phosphoric acid can affect the growth rate of different crystal faces, promote the growth along the C-axis direction, and inhibit the growth in the diameter direction so that the crystal morphology was highly fine-needle-like. The results were also consistent with those of crystal morphology analysis.

3.2. Crystal Water Content

The theoretical crystalline water content of dihydrate gypsum and α-HH was 20.9% and 6.21%, respectively. Therefore, the formation of α-HH can be reflected by measuring the content of crystal water in the samples. The crystal water content of each sample was measured, which is shown in Figure 4. For the reaction system without H3PO4, the crystalline water content of the prepared sample decreased significantly from 20.9% to 8.3% after a 0.5 h reaction time. This indicated that a considerable amount of α-HH had been formed at this time. After 1.5 h, the water of crystallization content was 6.22%, which was very close to the theoretical water of crystallization content of α-HH, revealing that the dihydrate gypsum crystal had all transformed into the α-HH phase at this time. For each reaction system with different amounts of H3PO4, the crystal water content of the obtained products was close to the theoretical crystal water content of α-HH after a 0.5 h reaction time, which indicated the crystal transformation process had been fully completed at this time. In general, the results of the crystal water content were consistent with the XRD result. This further supports the experimental conclusion that the presence of H3PO4 can promote the rapid formation of the α-HH phase.

3.3. Infrared Analysis

Due to self-ionization, H3PO4 can exist in the form of H3PO4, H2PO4 and HPO42− in the reaction liquid phase and can combine with Ca2+ ions in the liquid phase to form various forms of calcium phosphate. Calcium phosphate salts, especially those containing HPO42−, had similar lattice constants with calcium sulfate salts. Therefore, it was easy to place SO42− into the α-HH lattice to form a eutectic phosphorus solid solution. Once eutectic phosphorus formed, the hydration and hardening properties of the prepared α-HH can be adversely affected. In order to reveal whether eutectic phosphorus can be produced in the process of α-HH preparation, two kinds of α-HH samples prepared with H3PO4 contents of 0% and 1.0% were analyzed via infrared spectroscopy. The results are shown in Figure 5. The infrared characteristic absorption peak of eutectic phosphorus was usually located near 840 cm−1, and no characteristic peak of eutectic phosphorus was found in the two groups of samples via a comparative test [28]. At the same time, the characteristic peaks of all kinds of calcium phosphate salts were not detected, which may be due to the low amount of phosphate acid incorporation, resulting in the low content of calcium phosphate salts.

3.4. Morphology Analysis

In order to analyze the effect of H3PO4 on the morphology of α-HH, the morphology of each sample prepared under the 1.5 h reaction time was observed, as shown in Figure 6. Without H3PO4, α-HH presented a fine-needle-like crystalline form with a diameter of 2–5 μm, length of 15–80 μm and aspect ratio of 4–16. The addition of H3PO4 did not change the crystal morphology of α-HH but caused a significant increase in the aspect ratio of the α-HH crystal. And with the increase in H3PO4 content, the aspect ratio of the crystal also gradually increased. When H3PO4 content was as high as 1.0%, the aspect ratio of α-HH increased significantly to 44–60. The main reason for this phenomenon was that H+ in the reaction solution can combine with SO42− to form HSO4 with the addition of H3PO4, which may increase the free SO42− concentration in the solution. The growth rate of the (111) crystal plane located at the top of the α-HH crystal can be accelerated. Accordingly, the aspect ratio of the crystal increased.
Many studies have shown that it is necessary to add some crystal modulators to regulate the crystal morphology of α-HH in the process of preparation in order to give good physical and mechanical properties. The performance was best when the aspect ratio of the crystal was close to 1:1. Therefore, an appropriate amount of citric acid was introduced into the reaction system, and the morphology of each α-HH sample prepared is shown in Figure 7. Figure 7a shows the morphology of α-HH prepared by the reaction system without H3PO4 doped and regulated by citric acid. It can be seen by comparison with Figure 7a that the morphology of the α-HH crystal changed from the original fine-needle-like shape to a short column shape after the addition of citric acid. The aspect ratio of the crystal also decreased to 1:1, which indicated citric acid had an excellent crystal-regulating effect. It can be seen from Figure 7b–e that H3PO4 had a negative impact on the crystal-regulating effect of citric acid, and the larger the amount of H3PO4, the more significant the effect. When 0.2% H3PO4 was added, the aspect ratio of α-HH only increased to 2:1. When the content increased to 0.4%, α-HH still showed columnar morphology on the whole, and the aspect ratio only increased to around 6:1. However, the morphology of α-HH showed an obvious fine-needle-like shape when the H3PO4 content was more than 0.8%, and the crystal aspect ratio increased to more than 20:1.
The main reason why phosphoric acid can impact the crystal-regulating effect of citric acid was that H3PO4, as a kind of medium–strong acid, can change the pH of the liquid reaction system, and then cause the ionization process of citric acid molecules. In fact, there would be a three-order dissociation equilibrium in the reaction solution when citric acid (abbreviated as H3Cit), which is categorized as a ternary weak acid, was added to the reaction system. In the first stage, H3Cit was ionized to H2Cit and H+. In the second stage, H2Cit was further ionized to HCit2− and H+. Finally, HCit2− was further ionized to Cit3− and H+. The complexation of H2Cit, HCit2− and Cit3− carboxylate ions was generated due to the ionization of citric acid in the reaction liquid phase with Ca2+ on the crystal surface of the α-HH crystal (111). This was mainly due to the fact that calcium ions located on the (111) crystal plane of α-HH can be complexed with H2Cit, HCit2− and Cit3− complex anions. These were generated due to the ionization of citric acid, which significantly decreased the growth rate of the crystal along the C-axis and made α-HH develop in the short columnar direction [29]. Therefore, when H3PO4 was not added to the reaction system, citric acid could easily ionize and produce more complex anions than in a neutral environment, which significantly regulated the crystal morphology of α-HH. The addition of H3PO4 can significantly increase the acidity of the reaction liquid phase and then affect the degree of dissociation of citric acid molecules. Moreover, the greater the amount of H3PO4, the more significant this effect can become, which leads to a greatly reduced crystal-regulating effect of citric acid.

3.5. Water Consumption for Standard Consistency

Figure 8 shows the influence of H3PO4 on the water consumption for standard consistency of α-HH with and without a citric acid crystal regulator. Without a citric acid crystal regulator, the α-HH crystal was presented in a fine-needle-like shape (Figure 6), and the aspect ratio of the crystal was significantly larger. Accordingly, the water consumption of its standard consistency was considerably greater. The addition of H3PO4 further increased the aspect ratio of α-HH. Therefore, the water consumption of standard consistency of α-HH also gradually increased with the increase in H3PO4 content. After citric acid was added, the water consumption of α-HH at standard consistency decreased significantly under the regulation of its crystal form. For example, for samples without phosphoric acid, the water consumption for standard consistency decreased significantly from 1.1 to 0.35 before and after citric acid incorporation. However, due to how H3PO4 can significantly impact the regulatory effect of citric acid, the water consumption of α-HH standard consistency increased significantly with H3PO4 incorporation, and the higher the H3PO4 content, the higher the water consumption. However, with the same amount of H3PO4, the water requirement of α-HH prepared with a citric acid crystal regulator was significantly lower than that without citric acid.

3.6. Mechanical Properties

The test results of mechanical properties are shown in Figure 9. When the citric acid crystal regulator was not added, the α-HH samples had a fine-needle-like shape, which led to higher water consumption of standard consistency, resulting in a significantly lower strength. With the increase in H3PO4 content, the strength can be further reduced. After the crystal regulation of citric acid, the strength of α-HH was significantly increased due to the significant reduction in water consumption for standard consistency of α-HH. For example, the strength of α-HH samples with 0% H3PO4 content increased significantly from 3.7 MPa to 37.6 MPa before the inclusion of citric acid. According to the Chinese standard (JC/T 2038-2010 “α High Strength Gypsum”) [30], it can meet the requirements of α30 strength grade. Similarly, due to the negative effect of H3PO4, the strength of α-HH decreased obviously with the incorporation of H3PO4. Moreover, the higher the H3PO4 content, the more the strength decreased. However, the strength of α-HH was still high, up to 24.5 MPa when the content of H3PO4 was 0.4%, which meets the requirements of α20 strength grade.

4. Conclusions

This paper mainly studied the effect of H3PO4 on the crystal transformation, crystal regulation and physical mechanical properties of α-HH in the process of hydrothermal synthesis and aimed to provide necessary guidance for the green, low-cost, and high-value utilization of phosphogypsum. The important conclusions are as follows:
(1)
The presence of H3PO4 can accelerate the formation rate of the α-HH phase. The transition time of all gypsum to the α-HH phase can be shortened from 1.5 h to 0.5 h when only 0.2% H3PO4 was added.
(2)
The infrared test results showed that H3PO4 did not enter the α-HH lattice to form eutectic phosphorus solid solution. In addition, the presence of related calcium phosphate was not detected in α-HH, possibly due to the small amount of phosphate content leading to the small amount of calcium phosphate production.
(3)
The addition of H3PO4 can significantly affect the crystal shape of α-HH and significantly weaken the regulatory effect of citric acid on the crystal shape of α-HH, resulting in a significant increase in the aspect ratio of α-HH crystals. Moreover, the higher the H3PO4 content, the more significant the negative impact.
(4)
Due to the significant increase in the aspect ratio of α-HH caused by H3PO4, the standard consistency water consumption of α-HH gradually increased with the increase in H3PO4 content. Accordingly, the strength of the hardened body of α-HH gradually decreased. However, the prepared α-HH can still meet the requirements of α20 high-strength gypsum when the H3PO4 content was less than 0.4%.
(5)
It can be inferred from this study that the soluble P2O5 impurities present in phosphogypsum can effectively promote the rapid formation of the α-HH phase when phosphogypsum is used for hydrothermal preparation of high-strength gypsum, but it can have a significant negative effect on the crystal regulation of α-HH. Therefore, the screening of suitable crystallization agents to avoid the adverse effects of soluble P2O5 will be one of the focuses of future research in this field.

Author Contributions

Writing and preparing of the original draft, reviewing and editing, J.Z.; writing—review and editing, X.W., P.H., B.J., X.Z. and Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been vigorously supported by the Key Public Welfare Special Project of Henan Province (No. 201300311000).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saadaoui, E.; Ghazel, N.; Ben Romdhane, C.; Massoudi, N. Phosphogypsum: Potential uses and problems—A review. Int. J. Environ. Stud. 2017, 74, 558–567. [Google Scholar] [CrossRef]
  2. Rashad, A.M. Phosphogypsum as a construction material. J. Clean. Prod. 2017, 166, 732–743. [Google Scholar] [CrossRef]
  3. Tayibi, H.; Choura, M.; López, F.A.; Alguacil, F.J.; López-Delgado, A. Environmental impact and management of phosphogypsum. J. Environ. Manag. 2009, 90, 2377–2386. [Google Scholar] [CrossRef] [PubMed]
  4. Guan, Q.; Sui, Y.; Zhang, F.; Yu, W.; Bo, Y.; Wang, P.; Peng, W.; Jin, J. Preparation of α-calcium sulfate hemihydrate from industrial by-product gypsum: A review. Physicochem. Probl. Miner. Process. 2021, 57, 168–181. [Google Scholar] [CrossRef]
  5. Islam, G.S.; Chowdhury, F.H.; Raihan, M.T.; Amit, S.K.S.; Islam, M.R. Effect of Phosphogypsum on the Properties of Portland Cement. Procedia Eng. 2017, 171, 744–751. [Google Scholar] [CrossRef]
  6. Jiang, G.; Wu, A.; Wang, Y.; Lan, W. Low cost and high efficiency utilization of hemihydrate phosphogypsum: Used as binder to prepare filling material. Constr. Build. Mater. 2018, 167, 263–270. [Google Scholar] [CrossRef]
  7. Wei, Z.; Deng, Z. Research hotspots and trends of comprehensive utilization of phosphogypsum: Bibliometric analysis. J. Environ. Radioact. 2022, 242, 106778. [Google Scholar] [CrossRef] [PubMed]
  8. Chernysh, Y.; Yakhnenko, O.; Chubur, V.; Roubík, H. Phosphogypsum Recycling: A Review of Environmental Issues, Current Trends, and Prospects. Appl. Sci. 2021, 11, 1575. [Google Scholar] [CrossRef]
  9. Ding, C.; Sun, T.; Shui, Z.; Xie, Y.; Ye, Z. Physical properties, strength, and impurities stability of phosphogypsum-based cold-bonded aggregates. Constr. Build. Mater. 2022, 331, 127307. [Google Scholar] [CrossRef]
  10. Du, M.; Wang, J.; Dong, F.; Wang, Z.; Yang, F.; Tan, H.; Fu, K.; Wang, W. The study on the effect of flotation purification on the performance of α-hemihydrate gypsum prepared from phosphogypsum. Sci. Rep. 2022, 12, 95. [Google Scholar] [CrossRef] [PubMed]
  11. Guan, Q.; Sui, Y.; Yu, W.; Bu, Y.; Zeng, C.; Liu, C.; Zhang, Z.; Gao, Z.; Chi, Z. Deep removal of phosphorus and synchronous preparation of high-strength gypsum from phos-phogypsum by crystal modification in NaCl-HCl solutions. Sep. Purif. Technol. 2022, 298, 121592. [Google Scholar] [CrossRef]
  12. Li, X.; Gao, W. Conversion of phosphogypsum into α-hemihydrate in the presence of potassium acid phthalate and Ca2+: Ex-perimental and DFT studies. Colloids Surf. A Physicochem. Eng. Asp. 2022, 652, 129906. [Google Scholar] [CrossRef]
  13. Ma, B.; Lu, W.; Su, Y.; Li, Y.; Gao, C.; He, X. Synthesis of α-hemihydrate gypsum from cleaner phosphogypsum. J. Clean. Prod. 2018, 195, 396–405. [Google Scholar] [CrossRef]
  14. Fan, H.; Song, X.; Liu, T.; Xu, Y.; Yu, J. Effect of Al3+ on crystal morphology and size of calcium sulfate hemihydrate: Experimental and molecular dynamics simulation study. J. Cryst. Growth 2018, 495, 29–36. [Google Scholar] [CrossRef]
  15. Garg, M.; Jain, N.; Singh, M. Development of alpha plaster from phosphogypsum for cementitious binders. Constr. Build. Mater. 2009, 23, 3138–3143. [Google Scholar] [CrossRef]
  16. Feldmann, T.; Demopoulos, G.P. Influence of Impurities on Crystallization Kinetics of Calcium Sulfate Dihydrate and Hemihy-drate in Strong HCl-CaCl2 Solutions. Ind. Eng. Chem. Res. 2013, 52, 6540–6549. [Google Scholar] [CrossRef]
  17. Lv, X.; Xiang, L. The Generation Process, Impurity Removal and High-Value Utilization of Phosphogypsum Material. Nanomaterials 2022, 12, 3021. [Google Scholar] [CrossRef]
  18. Cai, Q.; Jiang, J.; Ma, B.; Shao, Z.; Hu, Y.; Qian, B.; Wang, L. Efficient removal of phosphate impurities in waste phosphogypsum for the production of cement. Sci. Total. Environ. 2021, 780, 146600. [Google Scholar] [CrossRef]
  19. Huang, Y.; Qian, J.; Kang, X.; Yu, J.; Fan, Y.; Dang, Y.; Zhang, W.; Wang, S. Belite-calcium sulfoaluminate cement prepared with phosphogypsum: Influence of P2O5 and F on the clinker formation and cement performances. Constr. Build. Mater. 2019, 203, 432–442. [Google Scholar] [CrossRef]
  20. Liu, Y.; Zhang, Q.; Chen, Q.; Qi, C.; Su, Z.; Huang, Z. Utilisation of Water-Washing Pre-Treated Phosphogypsum for Cemented Paste Backfill. Minerals 2019, 9, 175. [Google Scholar] [CrossRef]
  21. Liu, S.; Fang, P.; Ren, J.; Li, S. Application of lime neutralised phosphogypsum in supersulfated cement. J. Clean. Prod. 2020, 272, 122660. [Google Scholar] [CrossRef]
  22. Li, J.; Peng, X.; Zheng, J.; Mao, M.; Sun, X.; Wang, J.; Li, X.; Chai, X.; Lin, Z.; Liu, W. Simultaneous removal of phosphorus and organic contaminants from phosphogypsum using hydrothermal method for gypsum resource regeneration. J. Environ. Chem. Eng. 2022, 10, 108441. [Google Scholar] [CrossRef]
  23. Aliedeh, M.A. Factorial Design Study of P2O5 Reduction for Jordanian Phosphogypsum Using Sulfuric and Nitric Acids So-lutions. J. Chem. Technol. Metall. 2018, 53, 437–450. [Google Scholar]
  24. Al-Thyabat, S.; Zhang, P. REE extraction from phosphoric acid, phosphoric acid sludge, and phosphogypsum. Miner. Process. Extr. Metall. 2015, 124, 143–150. [Google Scholar] [CrossRef]
  25. Li, Z.; Liu, Y.; Xing, D.; Wang, B.; Liu, L.; Tang, J. Effect of maleic acid and pH on the preparation of α-calcium sulfate hemihydrate from phosphogypsum in Mg (NO3)2 solution. J. Mater. Cycles Waste Manag. 2022, 24, 143–154. [Google Scholar] [CrossRef]
  26. Chen, X.; Wang, Q.; Wu, Q.; Xie, X.; Tang, S.; Yang, G.; Luo, L.; Yuan, H. Hydration reaction and microstructural characteristics of hemihydrate phosphogypsum with variable pH. Constr. Build. Mater. 2022, 316, 125891. [Google Scholar] [CrossRef]
  27. GB/T 17669.4-1999; Gypsum Plasters-Determination of Physical Properties of Pure Paste. State Quality and Technical Supervision: Beijing, China, 1999.
  28. Ölmez, H.; Yilmaz, V. Infrared study on the refinement of phosphogypsum for cements. Cem. Concr. Res. 1988, 18, 449–454. [Google Scholar] [CrossRef]
  29. Li, X.; Zhang, Q.; Ke, B.; Wang, X.; Li, L.; Li, X.; Mao, S. Insight into the effect of maleic acid on the preparation of α-hemihydrate gypsum from phosphogypsum in Na2SO4 solution. J. Cryst. Growth 2018, 493, 34–40. [Google Scholar] [CrossRef]
  30. JC/T 2038-2010; α-High Strength Gypsum Plaster. Ministry of Industry and Information Technology: Beijing, China, 2010.
Figure 1. XRD spectrum of pure gypsum for the experiment.
Figure 1. XRD spectrum of pure gypsum for the experiment.
Materials 16 05878 g001
Figure 2. Particle size distribution of pure gypsum for the experiment.
Figure 2. Particle size distribution of pure gypsum for the experiment.
Materials 16 05878 g002
Figure 3. XRD patterns of each sample under different H3PO4 incorporations and reaction durations ((a): 0.5 h; (b): 1 h; (c): 1.5 h).
Figure 3. XRD patterns of each sample under different H3PO4 incorporations and reaction durations ((a): 0.5 h; (b): 1 h; (c): 1.5 h).
Materials 16 05878 g003aMaterials 16 05878 g003b
Figure 4. The test results of crystal water content.
Figure 4. The test results of crystal water content.
Materials 16 05878 g004
Figure 5. Infrared spectrum test results.
Figure 5. Infrared spectrum test results.
Materials 16 05878 g005
Figure 6. Morphology of α-HH with different H3PO4 content ((a): 0% H3PO4; (b): 0.2% H3PO4; (c): 0.4% H3PO4; (d): 0.8%H3PO4; (e): 1.0% H3PO4).
Figure 6. Morphology of α-HH with different H3PO4 content ((a): 0% H3PO4; (b): 0.2% H3PO4; (c): 0.4% H3PO4; (d): 0.8%H3PO4; (e): 1.0% H3PO4).
Materials 16 05878 g006
Figure 7. Effect of phosphoric acid on the morphology of samples after citric acid crystallization ((a): Citric acid + 0% H3PO4; (b): Citric acid + 0.2% H3PO4; (c): Citric acid + 0.4% H3PO4; (d): Citric acid + 0.8% H3PO4; (e): Citric acid + 1.0% H3PO4).
Figure 7. Effect of phosphoric acid on the morphology of samples after citric acid crystallization ((a): Citric acid + 0% H3PO4; (b): Citric acid + 0.2% H3PO4; (c): Citric acid + 0.4% H3PO4; (d): Citric acid + 0.8% H3PO4; (e): Citric acid + 1.0% H3PO4).
Materials 16 05878 g007
Figure 8. Effect of H3PO4 on the water consumption for standard consistency of α-HH.
Figure 8. Effect of H3PO4 on the water consumption for standard consistency of α-HH.
Materials 16 05878 g008
Figure 9. Effect of H3PO4 on 2 h dry compressive strength of α-HH.
Figure 9. Effect of H3PO4 on 2 h dry compressive strength of α-HH.
Materials 16 05878 g009
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

Zhang, J.; Wang, X.; Hou, P.; Jin, B.; Zhang, X.; Li, Z. Effect of Phosphoric Acid on the Preparation of α-Hemihydrate Gypsum Using Hydrothermal Method. Materials 2023, 16, 5878. https://doi.org/10.3390/ma16175878

AMA Style

Zhang J, Wang X, Hou P, Jin B, Zhang X, Li Z. Effect of Phosphoric Acid on the Preparation of α-Hemihydrate Gypsum Using Hydrothermal Method. Materials. 2023; 16(17):5878. https://doi.org/10.3390/ma16175878

Chicago/Turabian Style

Zhang, Jianwu, Xiao Wang, Pengtao Hou, Biao Jin, Xiaoting Zhang, and Zhixin Li. 2023. "Effect of Phosphoric Acid on the Preparation of α-Hemihydrate Gypsum Using Hydrothermal Method" Materials 16, no. 17: 5878. https://doi.org/10.3390/ma16175878

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