Next Article in Journal
Study of Coarse Particle Types, Structure and Crystallographic Orientation Relationships with Matrix in Cu-Cr-Zr-Ni-Si Alloy
Next Article in Special Issue
Optimizing Struvite Crystallization at High Stirring Rates
Previous Article in Journal
Effect of Lanthanum Oxide on the Microstructure and Properties of Ti-6Al-4V Alloy during CMT-Additive Manufacturing
Previous Article in Special Issue
Long-Term Stability of Novel Crucible Systems for the Growth of Oxygen-Free Czochralski Silicon Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Foreign Salts and Antiscalants on Calcium Carbonate Crystallization

by
Raghda Hamdi
1,* and
Mohamed Mouldi Tlili
2
1
Department of Chemistry, College of Science and Humanities in Al-Kharj, Prince Sattam bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
2
Laboratory of Desalination and Natural Water Valorisation (LADVEN), Water Researches and Technologies Center (CERTE), Techno-Park Borj Cedria, BP 273, Soliman 8020, Tunisia
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(3), 516; https://doi.org/10.3390/cryst13030516
Submission received: 28 February 2023 / Revised: 13 March 2023 / Accepted: 15 March 2023 / Published: 17 March 2023
(This article belongs to the Special Issue Crystallization Process and Simulation Calculation)

Abstract

:
For more than a century, crystallization has remained a chief research topic. One of the most undesirable crystallization phenomena is the formation of calcium carbonate scale in drinking and industrial water systems. In this work, the influence of chemical additives on CaCO3 formation—in either nucleation, crystal growth, or inhibition processes—is investigated by using the CO2-degasification method. Chemical additives are foreign salts (MgCl2, Na2SO4 and MgSO4) to the calco-carbonic system and antiscalants (sodium polyacrylate ‘RPI’ and sodium-tripolyphosphate ‘STPP’). The results show that additives affects both crystallization kinetics and the CaCO3 microstructure. Sulfate and magnesium ions, added separately at constant ionic strength, influence the nucleation step more than the growth of the formed crystallites. Added simultaneously, their effect was accentuated on both nucleation and the growth of CaCO3. Furthermore, antiscalants RPI and STPP affect the crystallization process by greatly delaying the precipitation time and largely increasing the supersaturation coefficient. It was also shown that the calco-carbonic system with additives prefers the heterogeneous nucleation to the homogeneous one. X-ray diffraction patterns show that additives promote the formation of a new crystal polymorph of calcium carbonate as aragonite, in addition to the initial polymorphs formed as calcite and vaterite.

1. Introduction

Crystallization in the calco-carbonic system CaCO3-CO2-H2O and the parameters that significantly control the polymorph’s formation have been studied for more than a century by several researchers in different scientific disciplines ranging from industrial crystallization [1,2] to biomineralization [3,4,5].
Calcium carbonate exists in three crystalline forms, which are calcite, aragonite and vaterite, in the order of stability at earth surface conditions. These crystalline phases have different crystal structures and morphologies. Calcite, aragonite and vaterite crystals have rhombohedral, orthorhombic, and hexagonal structures, respectively. In addition, crystalline monohydrocalcite (CaCO3.H2O) and ikaite (CaCO3.6H2O) are known. The observation of an intermediated or transient amorphous phase (amorphous calcium carbonate ACC) has also been reported [6,7,8,9]. In accordance with the Ostwald’s step rule [10], the transformation of the unstable phase (ACC) to a metastable phase (vaterite or aragonite) is followed by the formation of the more stable phase (calcite) through dissolution–crystallization reactions of various degrees of complexity.
The nucleation and growth of these forms are generally related to kinetic and thermodynamic factors such as the degree of supersaturation, solvent, organic, and inorganic additives. CaCO3 crystallization is intensely affected by the presence of such additives, which delay or inhibit nucleation and growth, and change the polymorph morphology [11,12,13]. The additives’ impact is detected in the nucleation and growth steps [14,15,16,17,18]. In the current work, the effect of different types of chemical additives such as magnesium, sulfate, and antiscalants on the CaCO3 crystallization process was studied.
The presence of dissolved ions, for instance magnesium and sulfate, is supposed to produce modification in the kinetic and the driving force of calcium carbonate precipitation [19,20]. They have long been considered as additives in calcium carbonates, usually known to affect calcite growth and dissolution. Magnesium ions have a significant role in the formation and transformation phases of CaCO3 crystals. They inhibit calcite formation [18,21] by decelerating its growth rate [22] and favoring the appearance of aragonite [23] and monohydrocalcite [24], which are less stable phases. They extend the duration of the amorphous phase life [25], and determine the polymorph formation. Magnesium interacts directly with calcium carbonate crystals and affects their morphologies [26]. The presence of magnesium and sulfate, separately and together, modifies the shape of the calcite crystal [27]. Sulfate affects the nucleation even at lower concentrations, by suppressing the formation of CaCO3 crystals and modifying their morphology [28]. It reduces the crystallization rate and favors the formation of the aragonite form [29].
The addition of antiscalants, also known as chemical inhibitors, is an effective method to inhibit the formation of scale [30]. Indeed, antiscalants delay calcium carbonate crystallization through growth sites blocking [31]. They have three mechanisms of action, which are dispersion, chelation, and crystal modification, consisting of the distortion of the crystal to become irregular and less adhesive, thus inhibiting crystal growth at calcite surfaces [32]. The polymer adsorption on the crystal growing surface suppresses the growth rate and displaces the precipitation mode from the most thermodynamic stable phase (calcite) to the metastable phases (aragonite and vaterite) [33]. The addition of the inhibitor resulted in (i) an important inhibition of crystallization, and (ii) a porous or an unconsolidated consistency of precipitates [34]. Therefore, the incorporation of the chemical additives in the structure of the crystal phase form modified its thermodynamic properties.
The present paper aims to improve our understanding of the effect of chemical additives on the CaCO3 crystallization. The changes induced in the kinetic, thermodynamic, and morphology of CaCO3 phases formed after the use of diverse types of chemical additives will be evaluated to explore the complex behaviors of the crystallization described in the calco-carbonic system. Consequently, the experimental process for CaCO3 crystallization based on the fast controlled precipitation (FCP) method [35,36] will be used. The chemical additives employed are three foreign salt (MgCl2, Na2SO4, and MgSO4) and two antiscalants (sodium polyacrylate and sodium-tripolyphosphate).

2. Experimental Section

2.1. Solution Preparation

The pure calco-carbonic water (PCCW) was obtained by the dissolution of reagent grade calcium carbonate solid in distilled water under CO2 bubbling, as follows:
CaCO3 + CO2 + H2O ⇆ Ca2+ + 2HCO3
The initial pH value was fixed at 5.9, which corresponds to a supersaturation coefficient equal to 0.25 (calculated using Equation (2)) to maintain the solution undersaturated.
The degree of supersaturation (Ω) with respect to calcite, the most CaCO3 stable form, is defined as follows:
Ω CaCO 3 = [ Ca 2 + ] [ CO 3 2 ] γ Ca 2 + γ CO 3 2 K s ( calcite )
where [i], γi and Ks (calcite) are the ions’ concentrations, the activity coefficients and the solubility product of calcite, respectively.
All the additives used were introduced to the PCCW before the precipitation test. Studied foreign salts (MgCl2, Na2SO4, and MgSO4) quantities were calculated according to the solution ionic strengths (IS) of about 0.012 (IS0), 0.024 (IS1) and 0.036 mol L−1 (IS2). The antiscalants used were organic sodium salt of polyacrylate (C3H3NaO2)n and inorganic sodium tripolyphosphate (Na5P3O10), thereafter called, respectively, RPI and STPP. The corresponding quantities used are small and do not affect the ionic strength and the resistivity.

2.2. Fast Controlled Precipitation Set Up

The experimental set up of the fast controlled precipitation (FCP) method was presented in Figure 1. Thermostatic water bath was used to maintain the temperature of the FCP test at 30 °C. A 0.5 L volume of pure calco-carbonic water (PCCW) was filled in a polytetrafluoroethylene (PTFE) cell and was stirred at 800 rpm. The solution pH and resistivity values were constantly measured each 5 min, using a pH meter (Hanna HI 110, Hanna Instruments, Woonsocket, RI, USA) and a conductivity meter (Meter Lab CDM210, Radiometer Analytical’s, Villeurbanne, France). The measuring electrode positions in the round bottom cell were well controlled. The calcium ion concentration was measured using EDTA complexometric titration.
In the FCP vessel, CaCO3 can precipitate on the cell wall and in the bulk solution. After the completion of the experiment, homogeneously formed precipitate mb can be recuperated by filtration using cellulose nitrate membrane with 0.45 μm porosity. The amount of Ca2+ residual in the solution was measured, and then the total precipitated calcium carbonate mt was calculated. The heterogeneous mass mw deposited on the cell wall can be deduced (mw = mt − mb), and its rate is calculated as follows:
% h e t e = m w m t × 100
Figure 2 represents a model of an FCP test reporting the evolution of pH and resistivity as a function of time of PCCW throughout the crystallization procedure. The figure emphasizes the limit between three different zones of precipitation. During zone 1, the resistivity is roughly constant and pH values increase up to reach the prenucleation pH “pHprenuc” at tprenuc. In zone 2, the pH continues rising up with a small change in the resistivity variation speed before precipitation time “tprec”. The medium is considered to be supersaturated, and calcium carbonate precipitation is thermodynamically possible. In zone 3, the slope of the resistivity temporal evolution diverges significantly, which highlights the precipitation beginning. After attaining the precipitation threshold at “pHprec”, the pH decreases considerably. Therefore, the precipitation is rapid and followed by a crystal growth stage.
The reproducibility of the experimental measurements was verified for three FCP tests using 400 mg L−1 of PCCW, as seen in Figure 3.

2.3. X-ray Diffraction Method for Solid Precipitate Characterization

The X-ray Diffraction (XRD) method is a very efficient and universal tool for determining the structure of crystals. At the end of each FCP test, the precipitate was recuperated by solution filtration through cellulose nitrate membrane with 0.45 μm porosity. The collected samples were dried at ambient conditions before XRD analysis. The analyses were performed using a diffractometer Philips X’PERT PRO (PANalytical, Philips, Amsterdam, The Netherlands) in step-scanning mode using Cu Kα (1.54 Å) radiation. The XRD patterns were chronicled in the angular range 2θ = 10–60°, with a slight step size of 2θ = 0.017° and a fixed count time of 4 s. The software ‘X-Pert HighScore Plus’ was used to determine the XRD reflection positions. The XRD patterns of the formed precipitates were compared to the joint committee on powder diffraction standards data.

3. Results and Discussion

3.1. Effect of Initial Calcium Carbonate Concentration

The calcium carbonate concentrations vary in natural waters in a range of 100 to 600 mg L−1. Examples include geothermal water (100 mg L−1), tap water (200 to 400 mg L−1), and saline water (600 mg L−1) [37]. Therefore, three solutions with different dissolved CaCO3 content levels (200, 400, 600 mg L−1) were prepared to study the effect of initial calcium carbonate amount on the nucleation threshold. Figure 4 reports the temporal evolution of pH and ΔResistivity curves during the FCP tests by varying the initial concentration of calcium carbonate. The main results are presented in Table 1.
As seen in Figure 4, the prenucleation (tprenuc) and the precipitation (tprec) time decrease as the CaCO3 amount increases. Indeed, tprenuc and tprec are roughly divided by two and by five, respectively, as the initial concentration increases from 200 to 600 mg L−1 (Table 1). Ben Amor et al. [38] proved that the precipitation time was delayed from 90 to 7 min after increasing the water hardness from 20 to 50 °F by using a CO2 degasification test. Additionally, Fathi et al. [39] found that the increase in Ca(HCO3)2 content initially dissolved from 30 to 50 °F accelerates the nucleation time from 20 to 7 min.
Thus, the nucleation step (∆t = tprec − tprenuc) is shorter as the initial calcium carbonate concentration increases. According to Table 1, this nucleation phase lasts 120 min for 200 mg L−1 and only 20 min for 600 mg L−1. This can be explained by the fact that the medium saturation state is affected by the initial concentration. According to Raffaella et al. [8], the lifetime of a cluster, before the growth phase, becomes longer when the concentration decreases as a result of the reduction in collision frequency. Indeed, the variation in the solution concentration modifies the frequency of collisions between the different species present (ions, ions pairs, and clusters). Thus, the maximum size that a germ can reach varies according to the different equilibriums between the association and dissociation processes.
From a thermodynamic point of view, the prenucleation threshold is attained at low Ω values less than two despite the difference in time to reach it (8 min for 600 mg L−1 and 18 min for 200 mg L−1). However, at the precipitation threshold, the supersaturation coefficients increase from 24 to 51 and the time decreases from 138 to 28 min as the initial concentration of calcium carbonate increases from 200 mg L−1 to 600 mg L−1, respectively.
Table 1 also reveals that the initial concentration of calcium carbonate [CaCO3]i influences the heterogeneous precipitation rate (%hete) of CaCO3. Indeed, %hete decreases from 56 to 34% as [CaCO3]i increases from 200 to 600 mg L−1. This is in agreement with Fathi et al. [39], who found that the heterogeneous precipitation rate, which was around 90% for 30 and 40 °F CaCO3 solutions, reached 63% for 50 °F solution. The same observation can be made for Ben Amor et al. [38], who proved that when the water hardness increases from 20 to 50 °F, the %hete decreases from 98 to 73%. Consequently, the adhesion of tartar on the cell walls is favored for low levels of calcium carbonate initially dissolved, and therefore for low nucleation thresholds. This can be due to the slow aggregation of ion pairs and clusters in low concentrated solutions, so the pre-nucleus remain at a small size for a longer time. Thus, it is more likely that they reach the surface and cling before they become large enough to not adhere to the cell wall.

3.2. Effect of Foreign Salts

3.2.1. Influence on the Nucleation Threshold

The most abundant salts found in natural waters are MgCl2, Na2SO4 and MgSO4. Their effect on the nucleation process of CaCO3 will be investigated. The results of FCP tests obtained in PCCW-MgCl2, PCCW-Na2SO4 and PCCW-MgSO4 solutions are compared to the results of PCCW solution. The pH and resistivity time curves are represented in Figure 5 and Figure 6.
As shown in Table 2, the prenucleation time tprenuc varies from 12 to 60 min as a function of the solution composition. During this time, the resistivity remains invariable, corresponding to the prenucleation threshold frontier. As expected, the increase of Mg2+ and SO42− ions concentrations highly delays the nucleation time for both ionic strengths (Table 2). Any precipitation time obtained in the presence of foreign ions was more than the 46 min precipitation time of CaCO3 in a PCCW solution. The MgSO4 ion greatly delayed the precipitation of CaCO3 compared to Mg2+ and SO42− ions. Indeed, tprec increased from 46 to 214 min when MgSO4 concentration increased from 0 (IS0) to 828 (IS2) mg L−1. As to Mg2+ and SO42−, they increased the precipitation time from 46 min (PCCW) to only 102 and 64 min, respectively. Moreover, the ion pairs’ formation (tprenuc) and aggregation (tprec) times were remarkably affected by the addition of each foreign ion. Indeed, the duration of the stable nuclei formation (∆t = tprec − tprenuc) was extended after the addition of these ions to a PCCW solution. At the same ionic strength (IS2), magnesium had more of an influence on the formation of CaCO3 stable nuclei than sulfate since ∆t was equivalent to 74 and 40 min for Mg2+ and SO42− ions, respectively.
Furthermore, Table 2 shows that the supersaturation coefficient values of the prenucleation threshold Ωprenuc were very low and varied from 1 to 9. By contrast, the precipitation threshold was reached at high Ωprec, especially after adding Na2SO4 to the PCCW solution. In fact, at the same ionic strength IS1, SO42− ion created a large supersaturation coefficient, which reached up to 1.55 times higher than the PCCW test (reference), while Mg2+ ion created a small supersaturation coefficient, which was 0.9 times lesser than that of PCCW. By combining these two ions to obtain MgSO4, Ωprec went down to 0.7 times less than that of PCCW. Consequently, sulfate ion had more of an impact on the pH of prenucleation and precipitation. As for magnesium ion, it influenced the nucleation time. Thus, the presence of magnesium and sulfate ions affected the nucleation and growth in different ways. Indeed, according to Waly et al. [40], two mechanisms can occur in the presence of magnesium and sulfate ions individually or together. The first mechanism involved is complexation, affecting the activity coefficients of calcium and carbonate ions by decreasing the total present ions for precipitation. The second implicated mechanism is inhibition, where the formed nuclei are inhibited from growing more by blocking the active growth sites. Therefore, both ions can attach to the recently formed calcium carbonate nuclei, causing a decrease in the nuclei growth rate due to the reduction in the available sites for growth by the blockage of the growth steps [41]. The growth can be suppressed entirely if the presence of foreign ions is dominant compared to the presence of calcium and carbonate ions [42].

3.2.2. Influence on the Surface Scaling

The effect of foreign ions on CaCO3 formation has been investigated systematically in this paper, including aspects of kinetics, thermodynamics and behaviors. However, their influence on scale surface deposition has not yet been studied. For this, the total precipitation rate and the heterogeneous precipitation percentage are calculated using the weight method [38] in the absence and presence of foreign ions. As illustrated by Table 2, the total precipitation rate τprec decreases as the ionic strength increases. Indeed, τprec decreases for IS1 by 5% for magnesium, 18% for sulfate and 9% for the two combined ions (MgSO4) compared to the PCCW solution test. Moreover, τprec is practically invariant by increasing the ionic strength for each salt. Furthermore, in the presence of magnesium or sulfate ions or combined magnesium sulfate ions, more than 70% of the CaCO3 precipitates are formed on the different surfaces in contact with water (cell-wall and probes) for both ionic strengths, whereas this heterogeneous percentage does not exceed 45% for pure calco-carbonic water. Subsequently, the introduction of one of these two ions, or both at once, leads to an increase in the precipitation on the walls for both ionic strengths studied. At the macroscopic scale, it would seem that both magnesium and sulfate ions, although presenting opposite charges, have the same effect on the orientation of CaCO3 precipitation to the heterogeneous precipitation, regardless of their modes of action in solution. However, in the study of Chen et al. [43], it was proved that Mg2+ ions apparently inhibit both CaCO3 bulk precipitation, regarded as a homogenous process, and CaCO3 surface deposition, regarded as a heterogeneous process. It is known that Mg will bond with water and form a complex at room temperature. This complex prohibits the nucleation and growth of calcite structure carbonate [44].

3.2.3. Influence on the Solid Precipitate Microstructure

The effect of foreign salts addition on the solid precipitate was studied. At the end of each FCP experiment, the precipitate obtained was characterized using the X-ray diffraction analyses. Figure 7, Figure 8 and Figure 9 present, respectively, the XRD patterns of the precipitate obtained in the absence and presence of MgCl2, Na2SO4 and MgSO4 for IS1 and IS2. CaCO3 precipitated in the PCCW is mainly vaterite, with a little fraction of calcite. The formation of these calcite crystals is inhibited when Mg2+ ions (from IS1) and SO42− ions are separately added (Figure 7 and Figure 8). Nevertheless, the presence of Mg2+ promotes the aragonite (Figure 7), while SO42− has no effect and vaterite remains the main CaCO3 polymorphous (Figure 8). The action mode of magnesium ions can be by (i) the substitution of calcium ions, or (ii) the insertion in or (iii) adsorption on the CaCO3 lattice [45]. If substitution or insertion leads to magnesian calcite, the adsorption at the earlier stage of nucleation and growth affects the CaCO3 crystalline structure, which is the most probable action mode for our case. Indeed, the interactions between Mg and CaCO3 nuclei are strong because they are improved by electrostatic interactions [46]. Moreover, Mg2+ competes with Ca2+ due to having the same charge, which favors the magnesium adsorption on the surface of calcite and vaterite [47]. The amount of Mg2+ being adsorbed onto CaCO3 crystal active growth sites can be influenced by the precipitation rate and the local solution chemistry [22]. In the presence of MgSO4 at IS1, the vaterite is completely transformed into aragonite (Figure 9) as for the precipitate obtained in MgCl2-solution (Figure 7). This confirms the role of magnesium ions in favoring the aragonite instead of vaterite. However, at such IS and in the presence of SO42−, Mg2+ ions could not inhibit the crystallization of the calcite form (Figure 9). By increasing the ionic strength (IS2) in the case of MgSO4, calcite is gradually inhibited.

3.3. Effect of Antiscalant

3.3.1. Influence on the Nucleation Threshold

Calcium carbonate scale deposits in natural water installations are a stimulating issue, mainly by obstructing the water flow. The formation of such undesirable deposits can be inhibited by the addition of antiscalants such as organic sodium salt of polyacrylate (RPI) and inorganic sodium tripolyphosphate (STPP). The results of FCP tests obtained in PCCW-RPI and PCCW-STPP solutions were compared to the results of the PCCW solution.
Figure 10 and Figure 11 report the temporal evolution of pH and ΔResistivity as a function of RPI and STPP antiscalants amount. They show that the presence of both antiscalants RPI and STPP significantly influences the kinetics of the nucleation process. Indeed, tprenuc increased from 12 min for PCCW to 20 min in the presence of 4 ppm RPI and 0.8 ppm STPP, as seen in Table 3. During this time, the resistivity remained invariable corresponding to the prenucleation threshold limit. In addition, the increase of the antiscalant amount added to PCCW led to a higher precipitation time. In addition, STPP greatly delayed the precipitation of CaCO3 compared to RPI. In fact, tprec increased from 46 to 210 min and to 365 min for 4 ppm RPI and 0.8 ppm STPP, respectively. Furthermore, the duration of the stable nuclei formation (Δt = tprec − tprenuc) was longer as the antiscalant amount was larger. Therefore, the distinction between the prenucleation and precipitation thresholds became easier. Moreover, the prenucleation threshold was always attained at Ωprenuc values less than 7. Conversely, the precipitation threshold was reached at very high Ωprec, especially after adding STPP to PCCW solution, which went up to 353 at 0.8 ppm. Consequently, both antiscalants affected the kinetics and thermodynamics of calcium carbonate prenucleation and precipitation with different manners.

3.3.2. Influence on the Surface Scaling

The influence of the two antiscalants on scale surface deposition was studied by calculating the total precipitation rate and the heterogeneous precipitation percentage using the weight method. As seen in Table 3, after the addition of each antiscalant, the total precipitation rate τprec decreased remarkably from 45% for PCCW to 29% and 33% in the presence of RPI (4 ppm) and STPP (0.8 ppm), respectively. Moreover, for the same antiscalant, τprec decreased by increasing its amount added to PCCW. Furthermore, both antiscalants greatly affected the scale adherence to the cell wall and probes. Indeed, %hete increased from 45% for PCCW to 79% and to 70% after adding 4 ppm of RPI and 0.8 ppm of STPP, respectively. Despite RPI being organic and STPP being inorganic antiscalants, they both favorably promote precipitation on the cell wall detrimentally to the bulk solution scaling.

3.3.3. Influence on the Solid Precipitate Microstructure

The X-ray diffraction patterns of the precipitate formed in the presence of RPI and STPP are presented in Figure 12 and Figure 13. The patterns of deposit scale obtained in PCCW reveal the formation of calcite with majority vaterite. After the addition of 4 ppm RPI, the crystallization was orientated towards the formation of calcite and to the complete inhibition of vaterite. In the presence of 0.8 ppm STPP, the phases formed remained vaterite and calcite, with the appearance of a new aragonite phase. Consequently, the two antiscalants acted on the CaCO3 precipitation in different ways. XRD characterizations proved that the presence of antiscalant can affect the morphology and the crystal shape of calcium carbonate [48].

4. Conclusions

Chemical additives play a significant role in several stages of the formation and transformation of CaCO3 crystals. In the present paper, the fast controlled precipitation method was used to study the influence of chemical additives on the CaCO3 crystallization process. The additives employed were three foreign salts (MgCl2, Na2SO4 and MgSO4) and two antiscalants (sodium polyacrylate and sodium-tripolyphosphate). The results showed that the nucleation time was retarded after the addition of each foreign salt, regardless of its ionic strength value. The MgSO4 ion greatly delayed the precipitation of CaCO3 compared to Mg2+ and SO42− ions. Moreover, the sulfate ion had more impact on the pH of prenucleation and precipitation. As for magnesium ion, it influenced the nucleation time. Thus, the presence of magnesium and sulfate ions affected nucleation and growth through different ways. At the macroscopic scale, it would seem that both magnesium and sulfate ions, although presenting opposite charges, had the same effect on the orientation of CaCO3 precipitation to the heterogeneous precipitation, regardless of their modes of action in solution. Furthermore, both antiscalants RPI and STPP had a great effect on the crystallization kinetic by greatly delaying the precipitation time. Indeed, the prenucleation stage duration (Δt = tprec − tprenuc) was longer when the antiscalant concentration was larger. Therefore, the distinction between the prenucleation and precipitation thresholds became easier. In addition, the presence of each antiscalant affected the precipitation threshold remarkably, leading to a large variation in the supersaturation coefficient. In addition, the chemical inhibition favorably promoted precipitation on the cell wall, detrimental to the bulk solution scaling. Finally, the X-ray diffraction patterns of deposit scale formed in the absence and presence of each chemical additive revealed the formation of different crystal forms of calcium carbonate, which were calcite, vaterite and aragonite.

Author Contributions

Conceptualization, R.H. and M.M.T.; methodology, R.H. and M.M.T.; software, R.H.; validation, R.H. and M.M.T.; formal analysis, R.H. and M.M.T.; investigation, R.H.; resources, R.H. and M.M.T.; data curation, R.H.; writing—original draft preparation, R.H.; writing—review and editing, R.H. and M.M.T.; visualization, R.H.; supervision, R.H. and M.M.T.; project administration, R.H. and M.M.T.; funding acquisition, R.H. and M.M.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia, Project number (IF2/PSAU/2022/01/21605).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on reasonable request.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia for funding this research work through the Project number (IF2/PSAU/2022/01/21605).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, J.; Ruan, S.; Liu, Y.; Wang, T.; Zeng, Q.; Yan, D. Morphological characteristics of calcium carbonate crystallization in CO2 pre-cured aerated concrete. RSC Adv. 2022, 12, 14610–14620. [Google Scholar] [CrossRef] [PubMed]
  2. Feng, Z.; Yang, T.; Dong, S.; Wu, T.; Jin, W.; Wu, Z.; Wang, B.; Liang, T.; Cao, L.; Yu, L. Industrially synthesized biosafe vaterite hollow CaCO3 for controllable delivery of anticancer drugs. Mater. Today Chem. 2022, 24, 100917. [Google Scholar] [CrossRef]
  3. Lu, H.; Huang, Y.-C.; Hunger, J.; Gebauer, D.; Cölfen, H.; Bonn, M. Role of Water in CaCO3 Biomineralization. J. Am. Chem. Soc. 2021, 143, 1758–1762. [Google Scholar] [CrossRef] [PubMed]
  4. Kontrec, J.; Tomašić, N.; Mlinarić, N.M.; Kralj, D.; Džakula, B.N. Effect of pH and Type of Stirring on the Spontaneous Precipitation of CaCO3 at Identical Initial Supersaturation, Ionic Strength and a(Ca2+)/a(CO32−) Ratio. Crystals 2021, 11, 1075. [Google Scholar] [CrossRef]
  5. Dobberschütz, S.; Nielsen, M.R.; Sand, K.K.; Civioc, R.; Bovet, N.; Stipp, S.L.S.; Andersson, M.P. The mechanisms of crystal growth inhibition by organic and inorganic inhibitors. Nat. Commun. 2018, 9, 1578. [Google Scholar] [CrossRef] [Green Version]
  6. Lázár, A.; Molnár, Z.; Demény, A.; Kótai, L.; Trif, L.; Béres, K.A.; Bódis, E.; Bortel, G.; Aradi, L.E.; Karlik, M.; et al. Insights into the amorphous calcium carbonate (ACC) → ikaite → calcite transformations. CrystEngComm 2023, 25, 738–750. [Google Scholar] [CrossRef]
  7. Dong, L.; Xu, Y.-J.; Sui, C.; Zhao, Y.; Mao, L.-B.; Gebauer, D.; Rosenberg, R.; Avaro, J.; Wu, Y.-D.; Gao, H.-L.; et al. Highly hydrated paramagnetic amorphous calcium carbonate nanoclusters as an MRI contrast agent. Nat. Commun. 2022, 13, 5088. [Google Scholar] [CrossRef]
  8. Raffaella, D.; Paolo, R.; Julian, D.G.; David, Q.; Denis, G. Stable prenucleation mineral clusters are liquid-like ionic polymers. Nat. Commun. 2011, 2, 590. [Google Scholar]
  9. Gebauer, D.; Völkel, A.; Cölfen, H. Stable prenucleation calcium carbonate clusters. Science 2008, 322, 1819–1822. [Google Scholar] [CrossRef] [Green Version]
  10. Ostwald, W. Studies on formation and transformation of solid materials (Studien uber die bildung und umwandlung fester korper). Z. Phys. Chem. 1897, 22, 289–330. [Google Scholar] [CrossRef]
  11. Hamdi, R.; Tlili, M.M. Influence of foreign salts on the CaCO3 pre-nucleation stage: Application of the conductometric method. CrystEngComm 2022, 24, 3256–3267. [Google Scholar] [CrossRef]
  12. Wolf, S.L.P.; Jähme, K.; Gebauer, D. Synergy of Mg2+ and poly(aspartic acid) in additive-controlled calcium carbonate precipitation. CrystEngComm 2015, 17, 6857–6862. [Google Scholar] [CrossRef] [Green Version]
  13. Song, R.-Q.; Cölfen, H. Additive controlled crystallization. CrystEngComm 2011, 13, 1249–1276. [Google Scholar] [CrossRef]
  14. Raj, K.S.; Devi, M.N.; Palanisamy, K.; Subramanian, V.K. Individual and synergetic effect of EDTA and NTA on polymorphism and morphology of CaCO3 crystallization process in presence of barium. J. Solid State Chem. 2021, 302, 122026. [Google Scholar]
  15. Sancho-Tomás, M.; Fermani, S.; Durán-Olivencia, M.A.; Otálora, F.; Gómez-Morales, J.; Falini, G.; García-Ruiz, J.M. Influence of Charged Polypeptides on Nucleation and Growth of CaCO3 Evaluated by Counterdiffusion Experiments. Cryst. Growth Des. 2013, 13, 3884–3891. [Google Scholar] [CrossRef]
  16. Ihli, J.; Kim, Y.-Y.; Noel, E.H.; Meldrum, F.C. The Effect of Additives on Amorphous Calcium Carbonate (ACC): Janus Behavior in Solution and the Solid State. Adv. Funct. Mater. 2013, 23, 1575–1585. [Google Scholar] [CrossRef]
  17. Xiang, L.; Xiang, Y.; Wang, Z.G.; Jin, Y. Influence of chemical additives on the formation of super-fine calcium carbonate. Powder Technol. 2002, 126, 129–133. [Google Scholar] [CrossRef]
  18. Davis, K.J.; Dove, P.M.; De Yoreo, J.J. The Role of Mg2+ as an Impurity in Calcite Growth. Science 2000, 290, 1134–1137. [Google Scholar] [CrossRef]
  19. Loste, E.; Wilson, R.M.; Seshadri, R.; Meldrum, F.C. The role of magnesium in stabilising amorphous calcium carbonate and controlling calcite morphologies. J. Cryst. Growth 2003, 254, 206–218. [Google Scholar] [CrossRef]
  20. Chong, T.H.; Sheikholeslami, R. Thermodynamics and kinetics for mixed calcium carbonate and calcium sulfate precipitation. Chem. Eng. Sci. 2001, 56, 5391–5400. [Google Scholar] [CrossRef]
  21. Chen, T.; Neville, A.; Yuan, M. Assessing the effect of Mg2+ on CaCO3 scale formation–bulk precipitation and surface deposition. J. Cryst. Growth 2005, 275, e1341–e1347. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Dawe, R.A. Influence of Mg2+ on the kinetics of calcite precipitation and calcite crystal morphology. Chem. Geol. 2000, 163, 129–138. [Google Scholar] [CrossRef]
  23. Niedermayr, A.; Köhler, S.J.; Dietzel, M. Impacts of aqueous carbonate accumulation rate, magnesium and polyaspartic acid on calcium carbonate formation (6–40 °C). Chem. Geol. 2013, 340, 105–120. [Google Scholar] [CrossRef]
  24. Blue, C.R.; Giuffre, A.; Mergelsberg, S.; Han, N.; De Yoreo, J.J.; Dove, P.M. Chemical and physical controls on the transformation of amorphous calcium carbonate into crystalline CaCO3 polymorphs. Geochim. Cosmochim. Acta 2017, 196, 179–196. [Google Scholar] [CrossRef] [Green Version]
  25. Rodriguez-Blanco, J.D.; Shaw, S.; Bots, P.; Roncal-Herrero, T.; Benning, L.G. The role of Mg in the crystallization of monohydrocalcite. Geochim. Cosmochim. Acta 2014, 127, 204–220. [Google Scholar] [CrossRef]
  26. Zhang, A.; Xie, H.; Liu, N.; Chen, B.-L.; Ping, H.; Fu, Z.-Y.; Su, B.-L. Crystallization of calcium carbonate under the influences of casein and magnesium ions. RSC Adv. 2016, 6, 110362–110366. [Google Scholar] [CrossRef]
  27. Nielsen, M.R.; Sand, K.K.; Rodriguez-Blanco, J.D.; Bovet, N.; Generosi, J.; Dalby, K.N.; Stipp, S.L.S. Inhibition of Calcite Growth: Combined Effects of Mg2+ and SO42–. Cryst. Growth Des. 2016, 16, 6199–6207. [Google Scholar] [CrossRef]
  28. Sheikholeslami, R.; Ong, H.W.K. Kinetics and thermodynamics of calcium carbonate and calcium sulfate at salinities up to 1.5 M. Desalination 2003, 157, 217–234. [Google Scholar] [CrossRef]
  29. Tang, Y.; Zhang, F.; Cao, Z.; Jing, W.; Chen, Y. Crystallization of CaCO3 in the presence of sulfate and additives: Experimental and molecular dynamics simulation studies. J. Colloid Interface Sci. 2012, 377, 430–437. [Google Scholar] [CrossRef]
  30. Hamdi, R.; Tlili, M.M. Investigation of scale inhibitors effect on calcium carbonate nucleation process. Desalination Water Treat. 2019, 160, 14–22. [Google Scholar] [CrossRef] [Green Version]
  31. Jain, T.; Sanchez, E.; Owens-Bennett, E.; Trussell, R.; Walker, S.; Liu, H. Impacts of antiscalants on the formation of calcium solids: Implication on scaling potential of desalination concentrate. Environ. Sci. Water Res. Technol. 2019, 5, 1285–1294. [Google Scholar] [CrossRef]
  32. Bai, S.; Naren, G.; Nakano, M.; Okaue, Y.; Yokoyama, T. Effect of polysilicic acid on the precipitation of calcium carbonate. Colloids Surf. A Physicochem. Eng. Asp. 2014, 445, 54–58. [Google Scholar] [CrossRef]
  33. Chhim, N.; Haddad, E.; Neveux, T.; Bouteleux, C.; Teychené, S.; Biscans, B. Performance of green antiscalants and their mixtures in controlled calcium carbonate precipitation conditions reproducing industrial cooling circuits. Water Res. 2020, 186, 116334. [Google Scholar] [CrossRef] [PubMed]
  34. Eichinger, S.; Boch, R.; Leis, A.; Baldermann, A.; Domberger, G.; Schwab, C.; Dietzel, M. Green inhibitors reduce unwanted calcium carbonate precipitation: Implications for technical settings. Water Res. 2022, 208, 117850. [Google Scholar] [CrossRef]
  35. Gauthier, G.; Chao, Y.; Horner, O.; Alos-Ramos, O.; Hui, F.; Lédion, J.; Perrot, H. Application of the Fast Controlled Precipitation method to assess the scale-forming ability of raw river waters. Desalination 2012, 299, 89–95. [Google Scholar] [CrossRef]
  36. Lédion, J.; François, B.; Vienne, J. Characterization of the scaling properties of water by fast controlled precipitation test. Eur. J. Water Qual. 1997, 28, 15–35. [Google Scholar] [CrossRef]
  37. Rosset, R.; Sok, P.; Poindessous, G.; Amor, M.B.; Walha, K. Caractérisation de la compacité des dépôts de carbonate de calcium d’eaux géothermales du Sud tunisien par impédancemétrie. Comptes Rendus L’académie Sci. Ser. IIC Chem. 1998, 1, 751–759. [Google Scholar] [CrossRef]
  38. Amor, M.B.; Zgolli, D.; Tlili, M.M.; Manzola, A.S. Influence of water hardness, substrate nature and temperature on heterogeneous calcium carbonate nucleation. Desalination 2004, 166, 79–84. [Google Scholar] [CrossRef]
  39. Fathi, A.; Mohamed, T.; Claude, G.; Maurin, G.; Mohamed, B.A. Effect of a magnetic water treatment on homogeneous and heterogeneous precipitation of calcium carbonate. Water Res. 2006, 40, 1941–1950. [Google Scholar] [CrossRef]
  40. Waly, T.; Kennedy, M.D.; Witkamp, G.-J.; Amy, G.; Schippers, J.C. The role of inorganic ions in the calcium carbonate scaling of seawater reverse osmosis systems. Desalination 2012, 284, 279–287. [Google Scholar] [CrossRef]
  41. Mullin, J.W. Crystallization; Butterworth-Heinemann: Oxford, UK, 2001. [Google Scholar]
  42. Söhnel, O.; Garside, J.H. Precipitation: Basic Principles and Industrial Applications; Butterworth-Heinemann: Oxford, UK, 1992. [Google Scholar]
  43. Chen, T.; Neville, A.; Yuan, M. Influence of on formation—Bulk precipitation and surface deposition. Chem. Eng. Sci. 2006, 61, 5318–5327. [Google Scholar] [CrossRef]
  44. Fang, Y.; Zhang, F.; Farfan, G.A.; Xu, H. Low-Temperature Synthesis of Disordered Dolomite and High-Magnesium Calcite in Ethanol–Water Solutions: The Solvation Effect and Implications. ACS Omega 2022, 7, 281–292. [Google Scholar] [CrossRef]
  45. Tlili, M.M.; Amor, M.B.; Gabrielli, C.; Joiret, S.; Maurin, G.; Rousseau, P. Study of Electrochemical Deposition of CaCO3 by In Situ Raman Spectroscopy: II. Influence of the Solution Composition. J. Electrochem. Soc. 2003, 150, C485–C493. [Google Scholar] [CrossRef]
  46. Zahid, A. Mineral Scale Formation and Inhibition; Plenum Press: New York, NY, USA, 1995. [Google Scholar]
  47. Compton, R.G.; Brown, C.A. The Inhibition of Calcite Dissolution/Precipitation: Mg2+ Cations. J. Colloid Interface Sci. 1994, 165, 445–449. [Google Scholar] [CrossRef]
  48. Sheng, K.; Ge, H.; Huang, X.; Zhang, Y.; Song, Y.; Ge, F.; Zhao, Y.; Meng, X. Formation and Inhibition of Calcium Carbonate Crystals under Cathodic Polarization Conditions. Crystals 2020, 10, 275. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Fast controlled precipitation experimental set up.
Figure 1. Fast controlled precipitation experimental set up.
Crystals 13 00516 g001
Figure 2. Temporal evolution of pH and resistivity curves of a PCCW 400 mg L−1 at 30 °C, 800 rpm.
Figure 2. Temporal evolution of pH and resistivity curves of a PCCW 400 mg L−1 at 30 °C, 800 rpm.
Crystals 13 00516 g002
Figure 3. Reproducibility of the temporal evolution of pH and resistivity curves for three FCP tests at 30 °C, 800 rpm and PCCW 400 mg L−1.
Figure 3. Reproducibility of the temporal evolution of pH and resistivity curves for three FCP tests at 30 °C, 800 rpm and PCCW 400 mg L−1.
Crystals 13 00516 g003
Figure 4. pH and ΔResistivity vs. time curves by varying the initial amount of calcium carbonate.
Figure 4. pH and ΔResistivity vs. time curves by varying the initial amount of calcium carbonate.
Crystals 13 00516 g004
Figure 5. pH and ΔResistivity vs. time curves as a function of the added salt for IS0 and IS1.
Figure 5. pH and ΔResistivity vs. time curves as a function of the added salt for IS0 and IS1.
Crystals 13 00516 g005
Figure 6. pH and ΔResistivity vs. time curves as a function of the added salt for IS0 and IS2.
Figure 6. pH and ΔResistivity vs. time curves as a function of the added salt for IS0 and IS2.
Crystals 13 00516 g006
Figure 7. XRD patterns of the precipitate formed in the presence of MgCl2 by varying IS.
Figure 7. XRD patterns of the precipitate formed in the presence of MgCl2 by varying IS.
Crystals 13 00516 g007
Figure 8. XRD patterns of the precipitate formed in the presence of Na2SO4 by varying IS.
Figure 8. XRD patterns of the precipitate formed in the presence of Na2SO4 by varying IS.
Crystals 13 00516 g008
Figure 9. XRD patterns of the precipitate formed in the presence of MgSO4 by varying IS.
Figure 9. XRD patterns of the precipitate formed in the presence of MgSO4 by varying IS.
Crystals 13 00516 g009
Figure 10. pH and ΔResistivity vs. time curves as a function of RPI amount.
Figure 10. pH and ΔResistivity vs. time curves as a function of RPI amount.
Crystals 13 00516 g010
Figure 11. pH and ΔResistivity vs. time curves as a function of STPP amount.
Figure 11. pH and ΔResistivity vs. time curves as a function of STPP amount.
Crystals 13 00516 g011
Figure 12. XRD patterns of the precipitate formed in the presence of RPI.
Figure 12. XRD patterns of the precipitate formed in the presence of RPI.
Crystals 13 00516 g012
Figure 13. XRD patterns of the precipitate formed in the presence of STPP.
Figure 13. XRD patterns of the precipitate formed in the presence of STPP.
Crystals 13 00516 g013
Table 1. FCP tests results by varying the initial amount of calcium carbonate.
Table 1. FCP tests results by varying the initial amount of calcium carbonate.
[CaCO3]i (mg L−1)tprenuc (min)prenuctprec (min)prec∆t (min)%hete
200181.11382412056
400121.746483445
60082.028512034
Table 2. FCP tests results by varying the added salt amount.
Table 2. FCP tests results by varying the added salt amount.
IS
(mol L−1)
Solutiontprenuc
(min)
prenuctprec
(min)
precτprec (%)%hete
IS0 = 0.012PCCW121.746484545
IS1 = 0.024PCCW-Na2SO4205.062742777
PCCW-MgCl2203.276434080
PCCW-MgSO4221.780353672
IS2 = 0.036PCCW-Na2SO4248.764552672
PCCW-MgCl2284.4102353879
PCCW-MgSO4601.6214333571
Table 3. FCP test results by varying the antiscalant amount.
Table 3. FCP test results by varying the antiscalant amount.
Solutiontprenuc
(min)
prenuctprec
(min)
precτprec%hete
PCCW121.746484545
PCCW-RPI (0.2 ppm)15355973474
PCCW-RPI (4 ppm)2052102862979
PCCW-STPP (0.1 ppm)152.51101573766
PCCW-STPP (0.8 ppm)2073653533370
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

Hamdi, R.; Tlili, M.M. Influence of Foreign Salts and Antiscalants on Calcium Carbonate Crystallization. Crystals 2023, 13, 516. https://doi.org/10.3390/cryst13030516

AMA Style

Hamdi R, Tlili MM. Influence of Foreign Salts and Antiscalants on Calcium Carbonate Crystallization. Crystals. 2023; 13(3):516. https://doi.org/10.3390/cryst13030516

Chicago/Turabian Style

Hamdi, Raghda, and Mohamed Mouldi Tlili. 2023. "Influence of Foreign Salts and Antiscalants on Calcium Carbonate Crystallization" Crystals 13, no. 3: 516. https://doi.org/10.3390/cryst13030516

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