Next Article in Journal
Exploring the Interplay between Structure and Electronic Behavior across Pressure-Induced Isostructural and Structural Transitions in Weyl-Type Semimetal NbAs
Previous Article in Journal
Design, Manufacturing, Microstructure, and Surface Properties of Brazed Co-Based Composite Coatings Reinforced with Tungsten Carbide Particles
Previous Article in Special Issue
Synthesis and Mechanism Study of Carbon Nanowires, Carbon Nanotubes, and Carbon Pompons on Single-Crystal Diamonds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sustainable Production and Physicochemical Characteristics of Calcium Sulfate Dihydrate Prepared from Waste Eggshells

by
Somkiat Seesanong
1,
Chaowared Seangarun
2,
Banjong Boonchom
2,3,4,*,
Nongnuch Laohavisuti
5,*,
Wimonmat Boonmee
6,
Pesak Rungrojchaipon
4 and
Phairat Phimsirikul
1
1
Office of Administrative Interdisciplinary Program on Agricultural Technology, School of Agricultural Technology, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
2
Material Science for Environmental Sustainability Research Unit, School of Science, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
3
Municipal Waste and Wastewater Management Learning Center, School of Science, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
4
Department of Chemistry, School of Science, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
5
Department of Animal Production Technology and Fishery, School of Agricultural Technology, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
6
Department of Biology, School of Science, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(7), 577; https://doi.org/10.3390/cryst14070577
Submission received: 15 May 2024 / Revised: 6 June 2024 / Accepted: 16 June 2024 / Published: 21 June 2024

Abstract

:
Gypsum products (calcium sulfate dihydrate, CaSO4·2H2O) were synthesized through an eco-friendly and low-cost process by two different renewable calcium carbonate sources (CaCO3), hen and duck eggshell wastes, with product yields obtained of 84.73 and 87.74%, respectively. The X-ray fluorescence results indicated that calcium oxide (CaO) and sulfur trioxide (SO3) are the major elemental components of CaSO4·2H2O prepared from both calcium sources. The Fourier transform infrared results confirmed the vibrational characteristics of SO42− and H2O functional groups in the chemical structure of the prepared samples. The X-ray diffraction patterns of CaSO4·2H2O prepared from both calcium sources confirmed the sample’s crystal structure as well as the chemical formula, after comparing them to the standard powder diffraction file. The crystallite sizes of CaSO4·2H2O products were calculated from the experimental diffraction peak through the Scherrer equation and found to be 19–20 nm. The positive preferential growth (Pg) value highlighted the excellent stability of the synthesized CaSO4·2H2O. The scanning electron microscopic results showed the agglomeration particles of hen- and duck-CaCO3 raw agents, whereas plate-like particles were observed for hen- and duck-CaSO4·2H2O products but the particle sizes were different.

1. Introduction

Calcium sulfate, a ceramic material, is widely used for various applications. It has been used as a binder material for building construction [1], bone graft material [2], periodontal disease treatment [3], endodontic lesions [4], alveolar bone loss [5], maxillary sinus augmentation [6], filler for plastic [7] and rubber [8], coating [9], desiccant [10], coagulant [11], catalyst [12], and bleaching agent [13]. Four calcium sulfate forms have been observed and previously reported: calcium sulfate dihydrate (CaSO4·2H2O, gypsum) [13], α- and β-calcium sulfate hemihydrate (CaSO4·0.5H2O, bassanite) [14], and calcium sulfate anhydrous (CaSO4, anhydrite) [15]. Moreover, two different anhydrous polymorphs, β-CaSO4 (insoluble anhydrite, natural anhydrite form) and γ-CaSO4 (soluble anhydrite), were also reported [16]. The phase transformation, crystal structure, morphology, and growth rate of calcium sulfate depend on different physicochemical conditions, i.e., temperature, pressure, solutions, and other ionic minerals [17,18]. Calcium sulfates can be obtained from different preparation processes, i.e., formed from an aqueous sulfuric acid [19], flue gas desulfurization [20], reverse microemulsion [21], and chemical precipitation [22]. In turn, their characteristics (color, hardness, and density), crystal structure, solubility, optical property (translucence and refractive index), and applications depend on the preparation conditions [18].
Eco-friendly production and consumption for sustainable development are considered important topics. Furthermore, the depletion of the environment is intensifying the need to reduce pollution. In that context, “sustainability” is related to the reliability of resources and ecology, which connects to society, economies, and environments. At present, one-third of the total food we produce is cast off as waste, and this waste affects the environment [23]. Accordingly, the utilization of waste food may form part of the approach to ensuring a sustainable environment. Waste eggshells are one of the most discarded foods, generated from consumption as well as from various industries, i.e., drugs, cosmetics, etc. They offer an important calcium source since the main component of eggshells is calcium carbonate (CaCO3), which varies from 95 to 97% [24], mainly in the form of calcite polymorph > 94 wt%, which is considered a rich natural calcium source [25].
Eggshells have been applied in various fields of materials science, mostly with industrial and biomedical applications [26]. Meng and Deng [27] demonstrated a novel application of waste eggshells as a multifunctional reaction system to regulate the reactants and solution pH. They proposed the idea of transferring waste eggshells into a unique reaction system for the synthesis of Co9S8 nanorod arrays. Furthermore, due to its pore structure, eggshell is an excellent secondary raw material for the preparation of value-added products with a fine powder, such as porous solid catalysts [26]. Since the main chemical component of eggshells is CaCO3, which can then be calcined, CaO may be obtained, one of the most promising metal oxides with high basicity. This basicity property highlights that eggshell-derived CaO can be used as a heterogeneous catalyst in biodiesel production [28]. Previously, eggshell was applied as an adsorbent for the removal of hydrogen sulfide (H2S) from wastewater [29]. Additionally, eggshells are very useful green and economic adsorbents due to their availability and absence of toxic elements. Another novel application of eggshell waste is as a calcium (Ca2+) source to prepare hydroxyapatite (a calcium phosphate ceramic material with the chemical formula Ca10(PO4)6(OH)2) [30]. This calcium phosphate form, hydroxyapatite, has been applied as a biomaterial for bone and dental replacement because of its natural characteristics and its incorporation into bone tissues [31]. Moreover, beyond those already reported, the use of food waste for the production of other compounds will also help reduce the pollution effect of waste on the environment, with the subsequent conversion of the waste into a highly valuable product having the added benefit of reducing the production cost of that product.
In addition to the preparation of hydroxyapatite, the use of eggshells has also been reported as a secondary material for the preparation of other important agricultural- and chemical-based compounds, i.e., monocalcium phosphate (Ca(H2PO4)2·H2O), dicalcium phosphate (CaHPO4⋅2H2O), and tricalcium phosphate (Ca3(PO4)2) [32]. Eggshell was also used as a Ca2+ source in the preparation of calcium sulfate compounds, e.g., CaSO4·2H2O (gypsum) [23], which is an important fertilizer that provides Ca2+ and SO42− nutrients for plants. Gypsum is obtained from two sources: natural and synthetic gypsums. Synthetic gypsum is produced as a waste product or by-product in a range of industrial processes. Recently, a study reported the synthesis of gypsum from a reaction between bassanite, eggshells, and Pila globosa shells with sulfuric acid, which was carried out with different process conditions, acid concentrations, and time lengths [23,24,26,27,28,29].
Egg production in Thailand displays significant growth. According to the statistical data reported by the Information and Communication Technology Center (Department of Livestock Development, Ministry of Agriculture and Cooperatives, Thailand), in 2022, Thailand had more than 66 and 17 million hens and ducks, which produced more than 51 and 12 million hen and duck eggs per day, respectively. These quantities create lots of hen and duck eggshell waste. Against that background, the synthesis of value-added compounds from renewable natural resources is the driving force for environmental sustainability. Bringing the two together, this work aimed to use bio-waste eggshells as raw materials (a Ca2+ source) for the preparation of valuable products (CaSO4·2H2O) for environmental sustainability, specifically products to be used as soil treatments and fertilizers. The demand for fertilizer is greatly influenced by the demand for food, which is ever-growing due to the significantly increasing global population. CaSO4·2H2O is suitable to use as a fertilizer component as it is 150 times more soluble in water than CaCO3 and is used to minimize crusting, lessen erosion, and reduce phytotoxicity, as well as to treat sodium-enriched soil [23]. In this study, the physicochemical characteristics of CaSO4·2H2O samples were investigated by using X-ray fluorescence (XRF, an analytical technique used to determine the elemental composition of the samples), X-ray diffraction (XRD, a technique used for phase identification of the samples), Fourier-transform infrared (FTIR) spectroscopy (a technique used to record an infrared spectrum of absorption or transmission of the samples and confirm the presence of the functional groups), and scanning electron microscopy (SEM, a type of electron microscope that photographs the morphology of a sample by scanning the surface with a focused beam of electrons). This article reports an easy and low-cost process of using two different renewable calcium carbonate sources (CaCO3)—hen and duck eggshell wastes—to produce cheap gypsum products to be used in the agroindustry.

2. Experiments

2.1. Materials and their Synthesis

Distilled water prepared by the Department of Chemistry, School of Science, King Mongkut’s Institute of Technology Ladkrabang, Thailand was used throughout this study. Sulfuric acid (H2SO4, 98% w/w, 18 mol/L, AR grade) was purchased from Sigma-Aldrich. Hen and duck eggshells were collected from the bakery shops and canteens at the School of Science. Each eggshell type (duck or hen) was random, soaked, and cleaned with tap and distilled water several times, exposed to sunlight under ambient air for 2 days (drying), and ground with a 450 rpm ball milled process (MTI Corporation, Richmond, CA, USA, TMAX-XQM planetary ball milling) for 120 min using binary mixtures of two zirconium oxide (ZrO2) ball sizes (2 and 10 mm) under dry conditions with an eggshell:ball weight ratio of 1:10, creating hen- or duck-eggshell CaCO3 powders.
For the preparation of calcium sulfate, concentrated H2SO4 (18 mol/L, 98% w/w) was first diluted with distilled water to prepare 10 mol/L (~54% w/w), and diluted H2SO4 was then used as the SO42− source, whereas the eggshell powder was used as the Ca2+ source (CaCO3). We gradually added 20 mL of diluted H2SO4 solution (~0.2 mole) to 20 g of each CaCO3 (~0.2 mole) powdered from duck or hen eggshell. This created a molar ratio between the eggshell powder and H2SO4 of 1:1. The preparation reaction was conducted under ambient conditions without temperature control. The mixed suspension was stirred (C-MAG HS 7, IKA magnetic stirrer, Staufen, Germany) at 100 rpm until no more carbon dioxide (CO2) evolved. Equation (1) demonstrates the chemical reaction between H2SO4 solution and eggshell CaCO3 powder, which is the same reaction as in previous research [23,24,26,27,28,29] but the process conditions, acid concentration, and time consumption are different.
H2SO4(aq) + CaCO3(s) Ca2+(aq) + SO42−(aq) + nH2O(l) + CO2(g)
The calcium sulfate product (calcium sulfate hydrate, CaSO4·nH2O) was precipitated in a powder form, as demonstrated in Equation (2), when the solvent (H2O) was evaporated in an oven (Binder FD 115) at 60 °C for 48 h.
Ca2+(aq) + SO42−(aq) CaSO4·nH2O(s) + mH2O(g)
No additional chemicals were used in the calcium sulfate formation, ensuring a sustainable development by avoiding unnecessary precipitating reagents. The usage of waste eggshells as a raw reacting material also supports the sustainable production of CaSO4·nH2O products. Three repetitions were performed, and an average value was reported. After evaporation and formation had occurred completely, the product (CaSO4·nH2O) was kept in a desiccator for the sample characterization process.

2.2. Characterization

An XRF spectrometer (Bruker, Karlsruhe, Germany, SRS 3400) was employed to investigate the elemental compositions of the raw materials (hen and duck eggshell powders, CaCO3) as well as the synthesized calcium sulfate samples (CaSO4·nH2O) using operation conditions of 100 mA current and 50 kV voltage. An XRD (Bruker, D8 Advance) diffractometer was employed to characterize the crystallographic characteristics of the sample (conditions: 40 mA and 40 kV). The diffraction database (powder diffraction file, PDF) generated by the International Centre for Diffraction Data (ICDD) [33] was used to confirm the experimental diffraction patterns of the sample. The diffraction was recorded from 5 to 60° (2θ angles) with a 0.01° increment and 1 s/step as the scan speed. An FTIR spectrometer (PerkinElmer, Waltham, MA, USA, Spectrum GX) was employed to record the infrared absorption spectra of the sample, and the recorded spectra were used for the identification of the functional groups (i.e., CO32−, SO42−, and H2O) that existed in the sample structure. Each sample was first mixed with spectroscopic-grade potassium bromide powder and then pressed to form a small sample pellet for infrared measurement [34]. Infrared spectra of the samples were recorded at 4000–400 cm−1 (mid-infrared, MIR, region) using a resolution of 2 cm−1 and 64 scans. An SEM (Zeiss, Oberkochen, Germany, LEO VP1450) was used to observe the sample morphology after the gold coating technique.

3. Results and Discussion

3.1. Elemental Composition

The elemental compositions of the raw CaCO3 materials (hen and duck eggshell powders) and their reacted products (CaSO4·nH2O) were measured by the XRF technique, and the results are listed in Table 1. The major elemental component of both hen and duck eggshell powders is CaO, and we recorded contents of 97.0 and 97.4 wt%, respectively. These confirmed the chemical presentation of CaCO3 [23]. However, other compounds (mainly in oxide forms, as shown in Table 1) were also observed with total contents of 3.01 and 2.53 wt% for the hen and duck CaCO3 samples, respectively. These experimental values are in agreement with the results reported in the literature. Hossain et al. [23] used the wavelength-dispersive XRF technique to evaluate the elemental composition of eggshells, and the results showed that CaO was the major component (97.0%). They also observed MgO (~1.08%), P2O5 (~0.457%), and Na2O (~0.17%) as minor compounds. CuO is not present in hen eggshells but is present in duck eggshells, which may be due to hens’ and ducks’ different diets (with different mineral contents) in animal husbandry.
The CaSO4·nH2O products synthesized from hen and duck eggshell powders were composed of two major elemental components: CaO (39.7 wt%) and sulfur trioxide (SO3, 58.8 and 59.40 wt%). Other minor compounds in oxide forms were also observed, with total contents of 1.54 and 0.928 wt% for HES-CaSO4·nH2O and DES-CaSO4·nH2O, respectively. As listed in Table 1, the raw CaCO3 materials (HES and DES) powdered from hen and duck eggshells were found to have an insignificant difference in their CaO (97.00 and 97.40 wt%) contents. These raw CaCO3 materials with the same content of CaO (39.70 wt%), after being reacted with the SO42− source (H2SO4), were observed in both the HES-CaSO4·nH2O and DES-CaSO4·nH2O products. Although both products showed the same content of CaO, a slight difference in the SO3 content (0.6 wt%) was observed, which might have arisen from Ca2+ sources that initially contained different SO3 contents (0.644 and 1.08 wt% for hen and duck eggshells, respectively). The lowest content of minor components (0.928 wt%) indicated the highest purity of the DES-CaSO4·nH2O product. Hence, it could be inferred that the preparation of CaSO4·nH2O by using duck eggshell waste as a renewable source provided the highest purity of the product. In addition, toxic metals, i.e., lead, arsenic, cadmium, and chromium, were not found in the hen and duck eggshell wastes (HES and DES), highlighting their green characteristics for use as eco-friendly and low-cost renewable sources for the production of value-added products like hydrated calcium sulfate (CaSO4·nH2O).

3.2. Reaction Times and Product Yields

The “reaction time” was considered the time taken to dissolve eggshell CaCO3 powders in H2SO4 solution completely, then generate a calcium sulfate (Ca2+-SO42−) salt solution, as demonstrated in Equation (1). Table 2 illustrates the reaction time recorded during the generation of salt solution.
As described in Section 2.1, to achieve the complete chemical reaction, the molar ratio required between eggshell CaCO3 powder and H2SO4 solution was 1:1, forming Ca2+-SO42−-H2O solution. We observed that the temperature of the reaction increased, demonstrating that the dissolution process is an exothermic reaction. As can be seen in Table 2, there was an insignificant difference between the reaction times observed for the hen and duck eggshell systems; however, the results indicated that duck eggshell waste should be selected as the Ca2+ source used to rapidly prepare Ca2+-SO42− solution in a large-scale system.
As shown in Equations (1) and (2), the yields (Yp/%) of CaSO4·nH2O products prepared from eggshell CaCO3 raw materials were computed by using Equation (3) [29]:
Y P = w t o c s w t t c s   × 100 %
where wtocs and wttcs (g) are the weights of CaSO4·nH2O observed and in theory, respectively.
Table 2 lists the yields of CaSO4·nH2O products synthesized from the two different eggshell wastes and 50% w/w H2SO4. The yields of the prepared CaSO4·nH2O products derived from hen and duck eggshells are 84.73 and 87.74%, respectively. The product yields reported in this work are higher than that reported in a previous work (<80%) [29]. These results highlight that to achieve the highest CaSO4·nH2O yield on a plant scale with a short reaction time, duck eggshell waste should be selected as the starting material for the CaSO4·nH2O preparation. The reaction times of the gypsum products synthesized from the two kinds of eggshell wastes in this work are shorter than those reported in previous research (24 h) [23,24].

3.3. Vibrational Spectroscopy

The vibrational characteristics of the functional groups of the samples were observed and identified using an FTIR spectrometer, and the spectra of CaCO3 and CaSO4·nH2O in the MIR range (4000−400 cm−1) are demonstrated in Figure 1 and Figure 2, respectively. The functional group associated with the eggshell CaCO3 samples is the carbonate (CO32−) anion, whereas sulfate (SO42−) and water (H2O) are the functional groups of the CaSO4·nH2O products. Therefore, the vibrational characteristics of the samples were characterized and mainly assigned based on the asymmetric stretching, symmetric stretching, and bending vibrational modes of CO32−, SO42−, and H2O.
CaCO3 can crystallize in three anhydrous polymorphs, e.g., calcite, aragonite, and vaterite, [35]; however, calcite is the most stable form under ambient conditions (room temperature and 1 atm) [36], while aragonite is slightly observed [37] and vaterite is formed under high specific conditions [38]. The FTIR spectroscopic technique was previously applied in the literature to quantify the vibrational modes of calcite and aragonite CaCO3 polymorphs [37]. Therefore, the spectroscopic technique was selected in this work to identify the crystal structure (polymorph) of CaCO3 milled from hen and duck eggshells, and the infrared spectra of CaCO3 samples are demonstrated in Figure 1. The vibrational bands were considered based on the vibrational positions of trigonal planar CO32−, which consists of four absorption modes: asymmetric (υ3) and symmetric (υ1) stretching modes between C and O atoms, and out-of-plane (υ2) and in-plane (υ4) bending modes between C and two O atoms. Another two vibrational characteristic combinations—between υ1 and υ3, and between υ1 and υ4—were also previously reported [37].
Calcite and aragonite CaCO3 polymorphs show slightly different absorption positions overall, but various absorption positions between calcite and aragonite overlap [36,37]. Previously, three absorption positions of 670, 1021, and 1094 cm−1 of hen eggshell CaCO3 powder or 668, 1019, and 1093 cm−1 of duck eggshell CaCO3 powder were assigned as the vibrational characteristics of an aragonite CaCO3 polymorph [36,37], whereas, in this study, other absorption positions were assigned as the vibrational characteristics of calcite CaCO3. This revealed that both eggshell waste powders mainly contained the calcite CaCO3 polymorph, with a very small aragonite CaCO3 amount. The vibrational characteristics of the milled CaCO3 samples are listed in Table 3.
Figure 2 shows the infrared spectra of the CaSO4·nH2O products derived from hen and duck eggshell CaCO3 raw materials. The vibrational modes (asymmetric stretching, symmetric stretching, and bending) of SO42− and H2O were mainly considered and assigned.
Takahashi et al. [39] carried out factor group analysis of gypsum (CaSO4·2H2O) and reported that the symmetric stretching vibration (υ1 (A1, single mode, one component)) of the SO42− ions belongs to the Au symmetric species. The bending vibration (υ2 (E, doubly degenerate mode, two components)) also belongs to Au symmetry. Both the asymmetric stretching (υ3 (F2, triply degenerate mode, three components)) and bending (υ4 (F2, triply degenerate mode, three components)) vibrations belong to two symmetric species: Au (one component) and Bu (two components). This previously obtained information was used for the assignment of the vibrational characteristics of the synthesized CaSO4·2H2O. All observed vibrations of CaSO4·nH2O products derived from hen and duck eggshell CaCO3 powders, as listed in Table 4, were interpreted. Two vibrational positions at 510 and 456 cm−1 were assigned as the υ2 mode of SO42− [39]. In addition, these two bands were also assigned as the vibrational modes (rocking, wagging, and twisting) of H2O [40]. Three vibrational positions at 674, 612, and 594 cm−1 were assigned as the υ4 mode of SO42−. This observation is in agreement with the vibrational characteristics of CaSO4·2H2O reported by Kayabaş and Yildirim [41]. The peak at 674 cm−1 is prominent, whereas the two peaks at 612 and 594 cm−1 are weaker. This phenomenon is in agreement with the results reported by Takahashi et al. [39]. The intensities of these three peaks exhibit remarkable polarization dependence, indicating that the direction of the transition moment of the 674 cm−1 peak is parallel, while the transition moments of the 612 and 594 cm−1 peaks are perpendicular to the c-axis. Three modes—υ1 (symmetric O−H stretching), υ2 (H−O−H bending), and υ3 (asymmetric O−H stretching)—of H2O in CaSO4·2H2O crystals are in the Au and Bu symmetry species. A peak at 3412 cm−1 is assigned as both the asymmetric (υ3) and symmetric (υ1) O−H stretching modes (belonging to Bu symmetry) of H2O, whereas a peak at 1642 cm−1 is assigned as the H−O−H bending (υ2) mode of H2O. All these observed vibrational bands are in agreement with results reported by Salvadori et al. [42] and Hossain et al. [23], which confirm the presence of SO42− and H2O. Other vibrational modes and their positions of SO42− and H2O, as well as CO2 (adsorbed gas on the surface of the sample), are listed in Table 4.

3.4. Crystallographic Characteristics

As described in the vibrational spectroscopic results, calcite and aragonite polymorphic crystals of CaCO3 were observed. Therefore, another powerful technique, XRD, was used for the characterization of different CaCO3 polymorphs [36,37]. The diffraction patterns (relationships between the XRD intensity and 2θ angle) of hen and duck eggshell waste powders are demonstrated in Figure 3a.
Similar diffraction patterns between hen and duck eggshell waste powders were observed (Figure 3a), indicating that both materials (hen and duck eggshell powders) had the same crystallographic characteristics. All diffraction peaks were consistent with the standard diffraction data, PDF-ICDD #47-1743 [43], revealing that only the calcite CaCO3 polymorph was observed, which is the natural polymorph of CaCO3 present in the eggshell [44]. No diffraction peak of the aragonite CaCO3 polymorph was observed. This result might have been since there was a very small amount of aragonite CaCO3, which was not observed in the XRD technique [44]. Smooth diffraction baselines with high diffraction intensity peaks, as exhibited in Figure 3a, indicated the high crystallinity of CaCO3 powders obtained from both hen and duck eggshells. Figure 3b shows the crystal structure of CaCO3 in the calcite polymorph [15]. The lattice parameters a and c of calcite CaCO3 are 4.9896 and 17.0610 Å, with a cell volume of 367.8537 Å3 [37]. The crystallite sizes (Sc) of the samples were calculated by using the main diffraction (hkl lattice plane: 104; 2θ angles: 29.40° and 29.41° for hen and duck eggshell CaCO3 samples) through the Scherrer equation (Equation (4)) [45].
S c = 0.89 λ β   c o s θ
where λ is the X-ray wavelength (1.5406 Å), β is the full width at half-maximum (FWHM) value of the selected XRD diffraction peak, and θ is the angle of the selected diffraction peak.
The calculated crystallite sizes for the (104) calcite-lattice plane are listed in Table 5 for hen and duck eggshell waste CaCO3 powders, respectively.
Other crystallographic data (observed 2θ positions, diffraction intensities, the standard d-spacing values, and hkl lattice planes [46]) of hen and duck eggshell CaCO3 powders are listed in Table 6. These obtained data—specifically the higher intensities of all diffraction peaks, except for the (116) lattice plane—indicated that good crystallinity was obtained for the duck eggshell CaCO3 sample.
The diffraction patterns of the products obtained from the calcination process of raw hen and duck eggshell waste powders, as demonstrated in Figure 4a, showed three crystalline phases, consisting of CaO (lime), calcium hydroxide (Ca(OH)2, portlandite), and calcite CaCO3. Diffraction peaks at 32.47°, 37.62°, and 54.10° (Figure 4a) were assigned to the (111), (200), and (220) lattice planes of CaO based on PDF-ICDD #74-1226 [47], and Figure 4b shows the crystal structure of CaO [48]. Conventionally, a small fraction of CaO material can adsorb atmospheric H2O, resulting in the formation of Ca(OH)2 crystals due to the exposure of CaO to the air. The presence of Ca(OH)2 was confirmed by PDF-ICDD #01-1079 with 2θ results of 18.26, 34.23, and 47.37° for the (001), (101), and (102) lattice planes of Ca(OH)2 [49]. A diffraction peak at 29.66° (for both hen-and duck CaCO3) was observed, which is the highest diffraction intensity of calcite CaCO3, corresponding to the (104) lattice plane [37]. The crystallite sizes of CaO samples calculated from the Scherrer equation are listed in Table 5.
The experimental diffraction pattern and the crystal structure [50] of the synthesized CaSO4·nH2O product are demonstrated in Figure 5a,b, respectively. All diffraction peaks (Figure 5a) observed for the synthesized products were consistent with the standard diffraction of PDF-ICDD #33-0311, demonstrating the diffraction characteristics of gypsum (CaSO4·2H2O) [23,51].
Given the similarity of the diffraction patterns of the hen and duck CaSO4·2H2O products, we surmised that both prepared products had the same structural and chemical characteristics, which were those of gypsum (CaSO4·2H2O). The characteristic peaks of other phases were not observed in the diffraction patterns, indicating that CaSO4·2H2O might be formed as a single phase or that other phases with small amounts were not observed by the XRD technique. The observed 2θ positions, diffraction intensities, standard d-spacing values, and hkl lattice planes of hen- and duck-eggshell-derived CaSO4·2H2O samples are listed in Table 7.
The crystallite sizes calculated by using Scherrer’s equation for the hen and duck CaSO4·2H2O samples are listed in Table 5. Pinto et al. [52] synthesized pure CaSO4·2H2O by using CaCl2 and Na2SO4 as the Ca2+ and SO42− sources, respectively. The lattice parameters of CaSO4·2H2O were then calculated and found to comprise a, b, and c values of 5.675, 15.214, and 6.284 Å, respectively. CaSO4·2H2O crystallizes in a monoclinic crystal system with the space group C2/c (#15) [23]. The relative intensity (Ri), an important parameter, was also calculated in this work from the experimental XRD data. The Ri for the highest diffraction intensity ((020) plane) was calculated from the other three strong lattice planes—(021), (040), and (041)—using Equation (5) [23]:
R i = I n t 020 I n t 021 + I n t 040 + I n t 041
where Int(020), Int(021), Int(040), and Int(041) are the intensities of the diffraction peaks of the (020), (021), (040), and (041) lattice planes, respectively.
The Ri values of the hen and duck CaSO4·2H2O samples were 1.23 and 1.25, respectively. In order to compare the Ri values, the Ri of the diffraction database (PDF-ICDD #33-0311) was also calculated for the same lattice planes and found to be 0.52. The preferential growth (Pg) parameter of the (020) plane was then calculated using Equation (6) [23]:
P g = R i s s R i s d R i s d
where Riss and Risd are the relative intensities of the synthesized sample (CaSO4·2H2O) and the standard data (PDF-ICDD #33-0311, Ri = 0.52), respectively.
The Pg values for the synthesized hen and duck CaSO4·2H2O samples were calculated and found to be 4.54 and 4.62, respectively. The positive values we obtained suggested that the growth along the (020) lattice plane was stable and favorable, indicating good crystal stability [23].
Figure 6a demonstrates the diffraction patterns of products obtained from the calcination step (900 °C, 2 h) of CaSO4·2H2O. All diffraction peaks show the structural characteristics of the anhydrite (CaSO4) phase. As described in the Section 1, two polymorphs of CaSO4 were reported, β-CaSO4 and γ-CaSO4, which are the stable and metastable forms, respectively [16]. The hydration rate of the γ polymorph (CaSO4(s) + nH2O(g) CaSO4·nH2O(s)) is greater than that of the β polymorph; consequently, if the γ form comes in contact with atmospheric water, its rehydration process will occur, preventing the formation of the γ form [53].
All diffraction peaks of both CaSO4 samples, obtained from hen- and duck-eggshell-derived CaSO4·2H2O, were consistent with PDF-ICDD #72-0916. This standard diffraction confirmed that the product of the calcination process of CaSO4·2H2O is the orthorhombic CaSO4 phase with β polymorphic crystals. The crystal structure of β-CaSO4 [54] is shown in Figure 6b. The crystallite sizes of β-CaSO4 samples calculated from the Scherrer equation are listed in Table 5.

3.5. Morphological Characteristics

The morphological characteristics of hen and duck eggshell CaCO3 raw materials and their reaction products (CaSO4·2H2O) were imaged with a magnification of 15 kx, and the resulting micrographs of the samples are presented in Figure 7. The micrographs of the CaCO3 particles obtained from the milling process of raw hen and duck eggshell powders present an agglomeration of CaCO3 particles, causing different particle sizes. These different particle sizes were obtained since milling is a mechanical process, commonly employed for material size reduction through centrifugal forces [55]. During the milling process, some fractions of eggshells were milled completely, resulting in small powdered particles, whereas other fractions were not milled completely, resulting in large powdered particles. In addition, during the rotation process, small particles were randomly moved and located on the surface of large particles.
The micrographs of the CaSO4·2H2O samples synthesized from the reaction between hen or duck eggshell CaCO3 raw material and H2SO4, as demonstrated in Figure 7, show plate-like particles. This finding is not consistent with the results reported in the literature. Mbogoro et al. [56] used commercial CaCl2·2H2O and Na2SO4 as Ca2+ and SO42− sources, respectively, to form CaSO4·2H2O crystals. The same concentration of Ca2+ and SO42− salts of 0.56 mol/L was prepared as stock solutions and then those were employed to prepare growth solutions with five different Ca2+:SO42− free-ion ratios of 2.5:18.8, 3.4:13.7, 6.9:6.8, 13.6:3.4, and 17.8:2.5, corresponding to the ionic ratios (r) of 0.13, 0.25, 1.01, 4.00, and 7.12, respectively. At low ionic ratios (r < 1, SO42−-rich solutions), plate-like CaSO4·2H2O crystals were observed, whereas needle-like crystals were observed at an equal stoichiometry (r = 1) and at high ionic ratios (r < 1, Ca2+-rich solutions) [56]. Moreover, Madeja et al. [57] also used commercial CaCl2·2H2O and Na2SO4 solutions to prepare CaSO4·2H2O using titration precipitation and then investigated the influences of various additives (small molecules, homo- and co-polymers) on the morphological characteristics of CaSO4·2H2O. Among the additives, a co-polymer of vinylpyrrolidone (PVP) and acrylic acid (PAA) acted as an important modifier for gypsum development. CaSO4·2H2O particles with tilted stacking edges and pseudo-hexagonal plates were obtained [57]. In this work, irregular plate-like CaSO4·2H2O particles were observed by using an ionic ratio between Ca2+ (hen or duck eggshell CaCO3) and SO42− (H2SO4) of 1 (as described in “Materials and their Synthesis”, Equations (1) and (2)) without an additive addition step. These results demonstrate that different formation conditions cause different morphological characteristics [56,57].

4. Conclusions

Gypsum (CaSO4·2H2O) products were synthesized successfully using two kinds of eggshells (hen and duck) with a short time period and easy process, generating low-cost products. The experimental and standard diffraction patterns confirmed that the value-added product from the eggshell waste powders is CaSO4·2H2O (gypsum). When CaSO4·2H2O is thermally treated, it can completely dehydrate to form CaSO4 (anhydrite, β polymorph). The diffraction patterns also indicated that the sample phase will change from a monoclinic (CaSO4·2H2O) to an orthorhombic (CaSO4) crystal structure. The morphological characteristics of CaCO3 show an irregular shape with different agglomeration particle sizes, whereas plate-like particles were observed for CaSO4·2H2O. The crystallite sizes calculated from the diffraction peak with the highest intensity for CaCO3, CaO, CaSO4·2H2O, and CaSO4 were 18−21, 7−8, 19−20, and 11−12 nm, respectively.
This article has reported an easy and low-cost method with no environmental effect, which may be applied to produce cheap CaSO4·2H2O products to be used in the agroindustry, forming part of zero-waste sustainability efforts.

Author Contributions

C.S. and W.B.: investigation, experiment, data curation, writing—original draft, review and editing. P.P. and S.S.: data curation, writing—review and editing. B.B. and N.L.: conceptualization, investigation, writing—review and editing, supervision. P.R.: conceptualization, data curation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work is a result of the project entitled “Conversion of shell/eggshell biowastes for sustainable environmental remediation” (Grant No. RE-KRIS/FF67/030) of King Mongkut’s Institute of Technology Ladkrabang (KMITL), which has received funding support from the NSRE.

Data Availability Statement

All data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors would like to thank the Scientific Instruments Center KMITL for supporting the TGA, FTIR, XRD, and SEM techniques.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Aranda, M.A.; Artioli, I.; Bier, T.; Àngeles, G.; Freyer, D.; Kaden, R.; Lothenbach, B.; Pritzel, C.; Raab, B.; Stöber, S. Cementitious Materials: Composition, Properties, Application; Walter de Gruyter GmbH & Co KG: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  2. Kumar, Y.; Nalini, K.; Menon, J.; Patro, D.K.; Banerji, B. Calcium sulfate as bone graft substitute in the treatment of osseous bone defects, a prospective study. J. Clin. Diagn. Res. 2013, 7, 2926–2928. [Google Scholar] [CrossRef] [PubMed]
  3. Mohammed, A.A.; Elsherbini, A.M.; Ibrahim, F.M.; El-Meadawy, S.M.; Youssef, J.M. Biological effect of the nanocrystalline calcium sulfate bone graft in the periodontal regeneration. J. Oral Biol. Craniofacial Res. 2021, 11, 47–52. [Google Scholar] [CrossRef] [PubMed]
  4. Mukherji, A. Management of periradicular lesion using calcium sulfate graft: An unique case report. CHRISMED J. Health Res. 2017, 4, 268. [Google Scholar] [CrossRef]
  5. Subramaniam, S.; Fang, Y.-H.; Sivasubramanian, S.; Lin, F.-H.; Lin, C. Hydroxyapatite-calcium sulfate-hyaluronic acid composite encapsulated with collagenase as bone substitute for alveolar bone regeneration. Biomaterials 2016, 74, 99–108. [Google Scholar] [CrossRef]
  6. Guarnieri, R.; Bovi, M. Maxillary sinus augmentation using prehardened calcium sulfate: A case report. Int. J. Periodont. Rest. 2002, 22, 503–508. [Google Scholar]
  7. Murariu, M.; Da Silva Ferreira, A.; Bonnaud, L.; Dubois, P. Calcium sulfate as high-performance filler for polylactide (PLA) or how to recycle gypsum as by-product of lactic acid fermentation process. Compos. Interfaces 2009, 16, 65–84. [Google Scholar] [CrossRef]
  8. Wang, J.; Pan, X.; Xue, Y.; Cang, S. Studies on the application properties of calcium sulfate whisker in silicone rubber composites. J. Elastomers Plast. 2012, 44, 55–66. [Google Scholar] [CrossRef]
  9. Al-Hadhrami, L.M.; Quddus, A.; Al-Otaibi, D.A. Calcium sulfate scale deposition on coated carbon steel and titanium. Desalination Water Treat. 2013, 51, 2521–2528. [Google Scholar] [CrossRef]
  10. Burfield, D.R. Desiccant efficiency in solvent and reagent drying. 9. A reassessment of calcium sulfate as a drying agent. J. Org. Chem. 1984, 49, 3852–3854. [Google Scholar] [CrossRef]
  11. Wenjing, L.; Yue, Z.; Cen, Z.; Di, C.; Chaogeng, X. Influence of calcium sulfate incorporated with gluconolactone coagulant on the quality of whole soybean flour tofu. Food Chem. X 2022, 17, 100527. [Google Scholar]
  12. Zuhaimi, N.A.S.; Indran, V.P.; Deraman, M.A.; Mudrikah, N.F.; Maniam, G.P.; Taufiq-Yap, Y.H.; Rahim, M.H.A. Reusable gypsum based catalyst for synthesis of glycerol carbonate from glycerol and urea. Appl. Catal. A Gen. 2015, 502, 312–319. [Google Scholar] [CrossRef]
  13. Chakawa, D.; Nkala, M.; Hlabangana, N.; Muzenda, E. The use of calcium sulphate dihydrate (CaSO4·2H2O) as a bleaching agent for crude soya bean vegetable oil. Procedia Manuf. 2019, 35, 802–807. [Google Scholar] [CrossRef]
  14. Yin, S.; Yang, L. α or β?-hemihydrates transformed from dihydrate calcium sulfate in a salt-mediated glycerol–water solution. J. Cryst. Growth. 2020, 550, 125885. [Google Scholar] [CrossRef]
  15. Ishizawa, N.; Setoguchi, H.; Yanagisawa, K. Structural evolution of calcite at high temperatures: Phase V unveiled. Sci. Rep. 2013, 3, 2832. [Google Scholar] [CrossRef] [PubMed]
  16. Schmid, T.; Jungnickel, R.; Dariz, P. Insights into the CaSO4–H2O system: A Raman-spectroscopic study. Minerals 2020, 10, 115. [Google Scholar] [CrossRef]
  17. Freyer, D.; Voigt, W. Crystallization and phase stability of CaSO4 and CaSO4–based salts. Monatshefte Chemie 2003, 134, 693–719. [Google Scholar] [CrossRef]
  18. Unjan, W.; Tangboriboon, N. Bio-calcium sulfate preparation from duck eggshell via chemical reaction process. Int. J. Curr. Adv. Res. 2017, 6, 2454–2459. [Google Scholar] [CrossRef]
  19. Wang, Y.; Mao, X.; Chen, C.; Wang, W.; Dang, W. Effect of sulfuric acid concentration on morphology of calcium sulfate hemihydrate crystals. Mater. Res. Express. 2020, 7, 105501. [Google Scholar] [CrossRef]
  20. Bakshi, P.; Pappu, A.; Bharti, D.K. Transformation of flue gas desulfurization (FGD) gypsum to β-CaSO4·0.5H2O whiskers using facile water treatment. Mater. Lett. 2022, 308, 131177. [Google Scholar] [CrossRef]
  21. Kong, B.; Yu, J.; Savino, K.; Zhu, Y.; Guan, B. Synthesis of α-calcium sulfate hemihydrate submicron-rods in water/n-hexanol/CTAB reverse microemulsion. Colloids Surf. A Physicochem. Eng. Asp. 2012, 409, 88–93. [Google Scholar] [CrossRef]
  22. Van Driessche, A.; Stawski, T.; Kellermeier, M. Calcium sulfate precipitation pathways in natural and engineered environments. Chem. Geol. 2019, 530, 119274. [Google Scholar] [CrossRef]
  23. Hossain, M.S.; Shaikh, M.A.A.; Ahmed, S. Synthesis of Gypsum fertilizer from waste eggshells for sustainable environment. Mater. Adv. 2023, 4, 240–247. [Google Scholar] [CrossRef]
  24. Hossain, M.S.; Ahmed, S. Synthesis of nano-crystallite gypsum and bassanite from waste Pila globose shells: Crystallographic Characterization. RSC Adv. 2022, 12, 25096–25105. [Google Scholar] [CrossRef]
  25. Lee, W.-D.; Kothari, D.; Niu, K.-M.; Lim, J.-M.; Park, D.-H.; Ko, J.; Eom, K.; Kim, S.-K. Superiority of coarse eggshell as a calcium source over limestone, cockle shell, oyster shell, and fine eggshell in old laying hens. Sci. Rep. 2021, 11, 13225. [Google Scholar] [CrossRef]
  26. Maslyk, M.; Dallos, Z.; Koziol, M.; Seiffert, S.; Hieke, T.; Petrović, K.; Kolb, U.; Mondeshki, M.; Tremel, W. A Fast and Sustainable Route to Bassanite Nanocrystals from Gypsum. Adv. Funct. Mater. 2022, 32, 2111852. [Google Scholar] [CrossRef]
  27. Kamarou, M.; Natalia Korob, N.; Romanovski, V. Structurally controlled synthesis of synthetic gypsum derived from industrial wastes: Sustainable approach. J. Chem. Technol. Biotechnol. 2021, 96, 3134–3141. [Google Scholar] [CrossRef]
  28. Kamarou, M.; Kuzmenkov, M.; Korob, N.; Kwapinski, W.; Romanovski, V. Structurally controlled synthesis of calcium sulphate dihydrate from industrial wastes of spent sulphuric acid and limestone. Environ. Technol. Innov. 2020, 17, 100582. [Google Scholar] [CrossRef]
  29. Kumar, A.; Sharma, S.; Sharma, A.K. Synthesis of Gypsum Utilizing Waste Eggshell. Int. J. Sci. Res. Dev. 2021, 9, 305–309. [Google Scholar]
  30. Umesh, M.; Choudhury, D.D.; Shanmugam, S.; Ganesan, S.; Alsehli, M.; Elfasakhany, A.; Pugazhendhi, A. Eggshells biowaste for hydroxyapatite green synthesis using extract piper betel leaf-evaluation of antibacterial and antibiofilm activity. Environ. Res. 2021, 200, 111493. [Google Scholar] [CrossRef]
  31. Laohavisuti, N.; Boonchom, B.; Boonmee, W.; Chaiseeda, K.; Seesanong, S. Simple recycling of biowaste eggshells to various calcium phosphates for specific industries. Sci. Rep. 2021, 11, 15143. [Google Scholar] [CrossRef]
  32. Wong-Ng, W.; McMurdie, H.; Hubbard, C.; Mighell, A.D. JCPDS-ICDD research associateship (cooperative program with NBS/NIST). J. Res. Natl. Inst. Stand. Technol. 2001, 106, 1013–1028. [Google Scholar] [CrossRef]
  33. Sronsri, C.; Boonchom, B. Synthesis, characterization, vibrational spectroscopy, and factor group analysis of partially metal-doped phosphate materials. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 194, 230–240. [Google Scholar] [CrossRef]
  34. Morse, J.W.; Arvidson, R.S.; Lüttge, A. Calcium carbonate formation and dissolution. Chem. Rev. 2007, 107, 342–381. [Google Scholar] [CrossRef]
  35. Sronsri, C.; Sittipol, W.; U-yen, K. Performance of CaO catalyst prepared from magnetic-derived CaCO3 for biodiesel production. Fuel 2021, 304, 121419. [Google Scholar] [CrossRef]
  36. Sronsri, C.; U-yen, K.; Sittipol, W. Quantitative analysis of calcium carbonate formation in magnetized water. Mater. Chem. Phys. 2020, 245, 122735. [Google Scholar] [CrossRef]
  37. Nahi, O.; Kulak, A.N.; Zhang, S.; He, X.; Aslam, Z.; Ilett, M.A.; Ford, I.J.; Darkins, R.; Meldrum, F.C. Polyamines promote aragonite nucleation and generate biomimetic structures. Adv. Sci. 2022, 10, 2203759. [Google Scholar] [CrossRef]
  38. Takahashi, H.; Maehara, I.; Kaneko, N. Infrared reflection spectra of gypsum. Spectrochim. Acta A Mol. Biomol. Spectrosc. 1983, 39, 449–455. [Google Scholar] [CrossRef]
  39. Cygan, R.T.; Daemen, L.L.; Ilgen, A.G.; Krumhansl, J.L.; Nenoff, T.M. Inelastic neutron scattering and molecular simulation of the dynamics of interlayer water in smectite clay minerals. J. Phys. Chem. C 2015, 119, 28005–28019. [Google Scholar] [CrossRef]
  40. Kayabaş, A.; Yildirim, E. New approaches with ATR-FTIR, SEM, and contact angle measurements in the adaptation to extreme conditions of some endemic Gypsophila L. taxa growing in gypsum habitats. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 270, 120843. [Google Scholar] [CrossRef]
  41. Salvadori, B.; Errico, V.; Mauro, M.; Melnik, E.; Dei, L. Evaluation of gypsum and calcium oxalates in deteriorated mural paintings by quantitative FTIR spectroscopy. Spectrosc. Lett. 2003, 36, 501–513. [Google Scholar] [CrossRef]
  42. Render, D.; Samuel, T.; King, H.; Vig, M.; Jeelani, S.; Babu, R.J.; Rangari, V. Biomaterial-derived calcium carbonate nanoparticles for enteric drug delivery. J. Nanomater. 2016, 2016, 3170248. [Google Scholar] [CrossRef]
  43. Gautron, J.; Stapane, L.; Le Roy, N.; Nys, Y.; Rodriguez-Navarro, A.; Hincke, M. Avian eggshell biomineralization: An update on its structure, mineralogy and protein tool kit. BMC Mol. Cell Biol. 2021, 22, 11. [Google Scholar] [CrossRef]
  44. Monshi, A.; Foroughi, M.R.; Monshi, M.R. Modified Scherrer equation to estimate more accurately nano-crystallite size using XRD. World J. Nano Sci. Eng. 2012, 2, 154–160. [Google Scholar] [CrossRef]
  45. Sitepu, H. Texture and structural refinement using neutron diffraction data from molybdite (MoO3) and calcite (CaCO3) powders and a Ni-rich Ni50.7Ti49.30 alloy. Powder Diffr. 2009, 24, 315–326. [Google Scholar] [CrossRef]
  46. Liu, J.; Liu, M.; Chen, S.; Wang, B.; Chen, J.; Yang, D.-P.; Zhang, S.; Du, W. Conversion of Au(III)-polluted waste eggshell into functional CaO/Au nanocatalyst for biodiesel production. Green Energy Environ. 2022, 7, 352–359. [Google Scholar] [CrossRef]
  47. Fiquet, G.; Richet, P.; Montagnac, G. High-temperature thermal expansion of lime, periclase, corundum and spinel. Phys. Chem. Miner. 1999, 27, 103–111. [Google Scholar] [CrossRef]
  48. Yan, F.; Jiang, J.; Zhao, M.; Tian, S.; Li, K.; Li, T. A green and scalable synthesis of highly stable Ca-based sorbents for CO2 capture. J. Mater. Chem. A 2015, 3, 7966–7973. [Google Scholar] [CrossRef]
  49. Comodi, P.; Nazzareni, S.; Zanazzi, P.F.; Speziale, S. High-pressure behavior of gypsum: A single-crystal X-ray study. Am. Miner. 2008, 93, 1530–1537. [Google Scholar] [CrossRef]
  50. Chen, Q.; Jiang, G.; Jia, C.; Wang, H.; Guan, B. A facile method to control the structure and morphology of α-calcium sulfate hemihydrate. CrystEngComm 2015, 17, 8549–8554. [Google Scholar] [CrossRef]
  51. Pinto, A.J.; Carneiro, J.; Katsikopoulos, D.; Jiménez, A.; Prieto, M. The link between brushite and gypsum: Miscibility, dehydration, and crystallochemical behavior in the CaHPO4·2H2O–CaSO4·2H2O system. Cryst. Growth Des. 2012, 12, 445–455. [Google Scholar] [CrossRef]
  52. Christensen, A.N.; Olesen, M.; Cerenius, Y.; Jensen, T.R. Formation and transformation of five different phases in the CaSO4−H2O system: Crystal structure of the subhydrate β-CaSO4·0.5H2O and soluble anhydrite CaSO4. Chem. Mater. 2008, 20, 2124–2132. [Google Scholar] [CrossRef]
  53. Antao, S.M. Crystal-structure analysis of four mineral samples of anhydrite, CaSO4, using synchrotron high-resolution powder X-ray diffraction data. Powder Diffr. 2011, 26, 326–330. [Google Scholar] [CrossRef]
  54. Borhan, M.Z.; Ahmad, R.; Rusop, M.; Abdullah, S. Optimization of ball milling parameters to produce Centella asiatica herbal nanopowders. J. Nanostructure Chem. 2013, 3, 79. [Google Scholar] [CrossRef]
  55. Mbogoro, M.M.; Peruffo, M.; Adobes-Vidal, M.; Field, E.L.; O’Connell, M.A.; Unwin, P.R. Quantitative 3D visualization of the growth of individual gypsum microcrystals: Effect of Ca2+: SO42– ratio on kinetics and crystal morphology. J. Phys. Chem. C 2017, 121, 12726–12734. [Google Scholar] [CrossRef]
  56. Madeja, B.; Avaro, J.; Van Driessche, A.E.; Rückel, M.; Wagner, E.; Cölfen, H.; Kellermeier, M. Tuning the growth morphology of gypsum crystals by polymers. Cem. Concr. Res. 2023, 164, 107049. [Google Scholar] [CrossRef]
  57. Wagner, M.; Decker, M.; Kunther, W.; Machner, A.; Beddoe, R.E.; Heisig, A.; Heinz, D. Gypsum formation mechanisms and their contribution to crystallisation pressure in sulfate resistant hardened cement pastes during early external sulfate attack at low sulfate concentrations. Cem. Concr. Res. 2023, 168, 107138. [Google Scholar] [CrossRef]
Figure 1. Infrared (IR) spectra of the hen- and duck-eggshell CaCO3 powders in the wavenumber range of 4000−400 cm−1.
Figure 1. Infrared (IR) spectra of the hen- and duck-eggshell CaCO3 powders in the wavenumber range of 4000−400 cm−1.
Crystals 14 00577 g001
Figure 2. Infrared (IR) spectra of CaSO4·nH2O products derived from hen and duck eggshell CaCO3 raw materials: Hen-CaSO4·nH2O and Duck-CaSO4·nH2O.
Figure 2. Infrared (IR) spectra of CaSO4·nH2O products derived from hen and duck eggshell CaCO3 raw materials: Hen-CaSO4·nH2O and Duck-CaSO4·nH2O.
Crystals 14 00577 g002
Figure 3. (a) X-ray diffraction (XRD) patterns of hen and duck eggshell waste CaCO3 powders; (b) crystal structure of CaCO3 in calcite polymorph, which consists of Ca2+ (CaO6 octahedron) and CO32− (trigonal planar) units.
Figure 3. (a) X-ray diffraction (XRD) patterns of hen and duck eggshell waste CaCO3 powders; (b) crystal structure of CaCO3 in calcite polymorph, which consists of Ca2+ (CaO6 octahedron) and CO32− (trigonal planar) units.
Crystals 14 00577 g003
Figure 4. (a) X-ray diffraction (XRD) patterns of calcined product (CaO, lime) obtained from the calcination process (900 °C, 2 h) of hen and duck eggshell waste (CaCO3) powders; (b) crystal structure of CaO, which consists of Ca2+ and O2− units (CaO6 octahedron).
Figure 4. (a) X-ray diffraction (XRD) patterns of calcined product (CaO, lime) obtained from the calcination process (900 °C, 2 h) of hen and duck eggshell waste (CaCO3) powders; (b) crystal structure of CaO, which consists of Ca2+ and O2− units (CaO6 octahedron).
Crystals 14 00577 g004
Figure 5. (a) X-ray diffraction (XRD) patterns of CaSO4·2H2O (gypsum) prepared from the reaction between hen or duck eggshell waste (CaCO3) powder and H2SO4; (b) crystal structure of CaSO4·2H2O, which consists of Ca2+ (CaO8 polyhedron), SO42− (tetrahedron), and H2O units.
Figure 5. (a) X-ray diffraction (XRD) patterns of CaSO4·2H2O (gypsum) prepared from the reaction between hen or duck eggshell waste (CaCO3) powder and H2SO4; (b) crystal structure of CaSO4·2H2O, which consists of Ca2+ (CaO8 polyhedron), SO42− (tetrahedron), and H2O units.
Crystals 14 00577 g005
Figure 6. (a) X-ray diffraction (XRD) patterns of CaSO4 (anhydrite) obtained from the calcination process (900 °C, 2 h) of CaSO4·2H2O; (b) crystal structure of β-CaSO4, which consists of Ca2+ (CaO8 polyhedron) and SO42− (tetrahedron) units.
Figure 6. (a) X-ray diffraction (XRD) patterns of CaSO4 (anhydrite) obtained from the calcination process (900 °C, 2 h) of CaSO4·2H2O; (b) crystal structure of β-CaSO4, which consists of Ca2+ (CaO8 polyhedron) and SO42− (tetrahedron) units.
Crystals 14 00577 g006
Figure 7. Scanning electron microscopic (SEM) images of CaCO3 powders obtained from the milling process of hen and duck eggshell wastes, and the products (CaSO4·2H2O, gypsum) prepared from the reaction between hen or duck eggshell waste powder and H2SO4.
Figure 7. Scanning electron microscopic (SEM) images of CaCO3 powders obtained from the milling process of hen and duck eggshell wastes, and the products (CaSO4·2H2O, gypsum) prepared from the reaction between hen or duck eggshell waste powder and H2SO4.
Crystals 14 00577 g007
Table 1. Elemental composition of raw CaCO3 materials (hen eggshell: HES; duck eggshell: DES) and CaSO4·nH2O products derived from HES and DES (three repetitions were conducted, and average values are reported).
Table 1. Elemental composition of raw CaCO3 materials (hen eggshell: HES; duck eggshell: DES) and CaSO4·nH2O products derived from HES and DES (three repetitions were conducted, and average values are reported).
Elemental CompositionElemental Contents/wt%
HESDESHES-CaSO4·nH2ODES-CaSO4·nH2O
Calcium oxideCaO97.0097.4039.7039.70
Sulfur trioxideSO30.641.0858.8059.40
Sodium oxideNa2O0.210.170.100.07
Magnesium oxideMgO0.930.370.760.29
Aluminum oxideAl2O30.130.040.060.02
Silicon dioxideSiO20.360.130.170.05
Phosphorus pentoxideP2O50.510.620.340.40
Potassium oxideK2O0.100.050.050.03
Ferric oxideFe2O30.070.0340.030.03
Strontium oxideSrO0.060.010.020.03
ChlorineCl0.020.02
Cupric oxideCuO0.01
Table 2. Reaction times and product yields of CaSO4·nH2O derived from hen or duck eggshell powders and 50% w/w H2SO4 (three repetitions were conducted, and average values are reported).
Table 2. Reaction times and product yields of CaSO4·nH2O derived from hen or duck eggshell powders and 50% w/w H2SO4 (three repetitions were conducted, and average values are reported).
SamplesReaction Time/sProduct Yields/%
Hen-eggshell-derived CaSO4·nH2O6284.73
Duck-eggshell-derived CaSO4·nH2O5587.74
Table 3. Vibrational characteristics and their absorption positions of the hen eggshell CaCO3 (hen-CaCO3) and duck eggshell CaCO3 (duck-CaCO3) powders.
Table 3. Vibrational characteristics and their absorption positions of the hen eggshell CaCO3 (hen-CaCO3) and duck eggshell CaCO3 (duck-CaCO3) powders.
Vibrational CharacteristicsPositions/cm−1
Hen-CaCO3Duck-CaCO3
CO32−
Asymmetric C−O stretching (υ3)1421, 12621420, 1259
Symmetric C−O stretching (υ1)1094, 10211093, 1019
O−C−O out-of-plane bending (υ2)876, 803876, 800
O−C−O in-plane bending (υ4)713, 670710, 668
υ2 + υ3 + υ4 combination2964, 28762962, 2873
υ1 + υ3 combination25162513
υ1 + υ4 combination17971796
H2O
Asymmetric (υ3) and symmetric (υ1) O−H stretching34273291
H−O−H bending (υ2)16541656
CO2
Asymmetric C-O stretching2367, 23432361, 2342
Table 4. Vibrational characteristics and their absorption positions for CaSO4·nH2O products derived from hen and duck eggshell CaCO3 raw materials: hen-CaSO4·nH2O and duck-CaSO4·nH2O.
Table 4. Vibrational characteristics and their absorption positions for CaSO4·nH2O products derived from hen and duck eggshell CaCO3 raw materials: hen-CaSO4·nH2O and duck-CaSO4·nH2O.
Vibrational CharacteristicsPositions/cm−1
Hen-CaSO4·nH2ODuck-CaSO4·nH2O
SO42−
Asymmetric S−O stretching (υ3)1153, 11311156, 1135
Symmetric S−O stretching (υ1)10141013
υ1 + υ3 combination21372139
O−S−O bending (υ4)674, 612, 594675, 610, 596
O−S−O bending (υ2)510, 456509, 451
H2O
Asymmetric (υ3) and symmetric (υ1) O−H stretching34123407
H−O−H bending (υ2)16421640
Vibrations (rocking, wagging, twisting)885, 850, 510, 456886, 851, 509, 451
CO2
Asymmetric C–O stretching23602361, 2342
Table 5. Crystallite sizes (Sc) of the prepared samples calculated from lattice planes (hkl), full width at half-maximum (FWHM, β), and diffraction peaks (2θ, θ).
Table 5. Crystallite sizes (Sc) of the prepared samples calculated from lattice planes (hkl), full width at half-maximum (FWHM, β), and diffraction peaks (2θ, θ).
Sampleshkl Planesβ2θθSc/nm
Hen-CaCO3(104)0.1376329.4014.70018.67
Duck-CaCO30.1245129.4114.70520.48
Hen-CaO(220)0.1818837.6618.8307.54
Duck-CaO0.1706237.6218.8108.04
Hen-CaSO4·2H2O(020)0.0777411.675.83519.57
Duck-CaSO4·2H2O0.0786311.685.84019.30
Hen-CaSO4(200)0.1170125.4212.71011.84
Duck-CaSO40.1242525.4712.73511.19
Table 6. Crystallographic data of hen and duck eggshell CaCO3 powders.
Table 6. Crystallographic data of hen and duck eggshell CaCO3 powders.
2θDiffraction Intensity/a.u.d-Spacing/Åhkl Planes/-
Hen-CaCO3Duck-CaCO3Hen-CaCO3Duck-CaCO3
23.4023.08858796543.8556(012)
29.4029.41143,135147,7103.0364(104)
31.4931.43271730902.8448(006)
35.9735.9910,13614,0822.4951(110)
39.4239.4216,78919,1402.2851(113)
43.1743.1713,45320,0072.0948(202)
47.1347.12379554201.9278(024)
47.5247.4912,09816,6311.9131(018)
48.5348.5016,04715,1251.8758(116)
56.5856.59180142861.6260(211)
57.4457.40522369351.6043(122)
Table 7. Crystallographic data of CaSO4·2H2O samples derived from hen and eggshell CaCO3 raw materials.
Table 7. Crystallographic data of CaSO4·2H2O samples derived from hen and eggshell CaCO3 raw materials.
2θDiffraction Intensity/a.u.d-Spacing/Åhkl Planes/-
Hen-CaSO4·2H2ODuck-CaSO4·2H2OHen-CaSO4·2H2ODuck-CaSO4·2H2O
11.6711.68269910517.6000(020)
20.7920.7810303944.2811(021)
23.4123.433601293.8000(040)
29.1429.178043163.0642(041)
31.1531.173901392.8730(−221)
32.1332.1984322.7875(−112)
33.4133.452791142.6828(220)
34.5734.6466302.5964(−151)
36.0236.1197332.4957(−202)
36.6736.6055342.4521(022)
40.6940.69121592.2174(151)
43.3943.64136622.0860(−152)
45.5445.5651291.9912(−171)
47.9447.96132531.8985(080)
50.4050.3594471.8113(062)
55.2455.2047241.6643(−262)
56.8056.8085481.6207(190)
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

Seesanong, S.; Seangarun, C.; Boonchom, B.; Laohavisuti, N.; Boonmee, W.; Rungrojchaipon, P.; Phimsirikul, P. Sustainable Production and Physicochemical Characteristics of Calcium Sulfate Dihydrate Prepared from Waste Eggshells. Crystals 2024, 14, 577. https://doi.org/10.3390/cryst14070577

AMA Style

Seesanong S, Seangarun C, Boonchom B, Laohavisuti N, Boonmee W, Rungrojchaipon P, Phimsirikul P. Sustainable Production and Physicochemical Characteristics of Calcium Sulfate Dihydrate Prepared from Waste Eggshells. Crystals. 2024; 14(7):577. https://doi.org/10.3390/cryst14070577

Chicago/Turabian Style

Seesanong, Somkiat, Chaowared Seangarun, Banjong Boonchom, Nongnuch Laohavisuti, Wimonmat Boonmee, Pesak Rungrojchaipon, and Phairat Phimsirikul. 2024. "Sustainable Production and Physicochemical Characteristics of Calcium Sulfate Dihydrate Prepared from Waste Eggshells" Crystals 14, no. 7: 577. https://doi.org/10.3390/cryst14070577

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