Next Article in Journal
Tung Oil Microcapsules Prepared with Different Emulsifiers and Their Effects on the Properties of Coating Film
Next Article in Special Issue
Preparation and Application of Graphene and Derived Carbon Materials in Supercapacitors: A Review
Previous Article in Journal
Changes in the Oral Cavity Mucosal Surface under the Influence of Wearing Protective Face Masks—Nitric Oxide Concentration Analysis—Preliminary Report
Previous Article in Special Issue
Humic Acid Removal in Water via UV Activated Sodium Perborate Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Electrostatic Field on Mixed Aqueous Solution of Calcium and Ferrous Ions: Insights from Molecular Dynamics Simulations

1
State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, China
2
School of Electrical Engineering, Yanshan University, Qinhuangdao 066004, China
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(8), 1165; https://doi.org/10.3390/coatings12081165
Submission received: 28 June 2022 / Revised: 9 August 2022 / Accepted: 10 August 2022 / Published: 12 August 2022

Abstract

:
In order to investigate the anti-scaling and anti-corrosion characteristics of an electrostatic anti-fouling system in the application process, the influence of an electrostatic field (EF) on the structure and dynamics of hydrated Ca2+ and hydrated Fe2+ in a mixed aqueous system was studied through the calculation and analysis of the radial distribution function (RDF), self-diffusion coefficients, viscosity, and hydrogen bond structure by using molecular dynamics simulation. The study results show that the EF can decrease the radius of the first water shell of hydrated Ca2+ but increase that of Fe2+, which will reduce the possibility of forming calcite. The EF can make water molecules and Fe2+ more active, which can hinder iron release and thus decrease iron corrosion products. In addition, the EF can enhance the hydrogen structure of water molecules in the aqueous solution.

1. Introduction

Fouling and iron corrosion are common problems in the water industry. Fouling results in degradation in the performance of heat exchangers. Iron corrosion can destroy the pipe and pollute the drinking water due to the suspension of iron particles. An electrostatic anti-fouling (EAF) system is one of the most common and effective ways to solve these problems. Water scale is almost entirely composed of calcium carbonate (CaCO3), which naturally occurs in six various forms: calcite, aragonite, vaterite, two hydrate phases, and amorphous calcium carbonate (ACC). ACC is perhaps the most vital to growth processes since it can create the precursor to the crystallization of those phases and plays a critical role in the biomineralization process for this substance [1]. Thus, the electric field’s impact on calcium carbonate has been the focus of the great majority of studies looking into the anti-fouling mechanism of EAF. It consists of two main aspects: how calcium carbonate crystals are formed and how the structure of scale crystals changes under electric fields. The impact of an electric field on calcium carbonate crystallization was studied by Qi et al. [2]. Their results show that the electric field applied to the CaCO3 solution affects the lattice structure and crystal shape of CaCO3 crystals. According to Li et al. [3], the scale sample produced by electrostatic water included more aragonite than before. Most equipment in industrial systems is made of carbon steel, and its main chemical composition is iron. Therefore, researchers who investigate the mechanism of metal corrosion tend to use iron as the study object. Lin et al. [4] found that the electric field accelerates steel corrosion, and the electric field favors the generation of plate-shaped γ-FeOOH and inhibits the formation of α-FeOOH. In addition, the scaling of calcium carbonate has a certain influence on iron corrosion. Sarin et al. [5] found that CaCO3 plays a role in reducing iron release, which can alleviate water pollution problems, and Harris et al. found that the scaling tendency in water is the main factor determining the corrosion rate of iron [6]. The different contents of metal cations (including Fe2+) in aqueous solutions also have a certain influence on the precipitation process of CaCO3 [7]. Less attention has been paid to the impact of an electrostatic field (EF) on the microcosmic particle structure in aqueous systems.
Due to the relatively poor solubility of calcium carbonate, there are nearly zero CO32− ions in hard water, but HCO3 ions are plentiful. As hydrated ions, Ca2+ and Fe2+ can be found in aqueous solutions. Additionally, these hydrated ions form hydrogen bonds with additional water molecules [8,9,10]. The formation, growth, and crystal structure of calcium carbonate precipitates are mainly determined by the structural and dynamic properties of Ca2+. Fe2+ plays an important role in the metal corrosion process. The main reactions of iron corrosion are shown below.
The iron in the material loses two electrons and forms Fe2+.
Fe Fe 2 + + 2 e
Electron acceptors such as oxygen or hydrogen ions in aqueous solutions serve to complete the reaction.
1 2 O 2 + H 2 O + 2 e 2 O H
2 H + + 2 e H 2
Ferrous solids such as ferrous hydroxide can form.
Fe 2 + + 2 O H Fe ( OH ) 2
Fe(OH)2 can be further oxidized to Fe(OH)3 attached to the wall of the tube.
4 Fe ( OH ) 2 + O 2 + 2 H 2 O 4 Fe ( OH ) 3
Therefore, by examining the impact of an EF on the structure and molecular dynamics of hydrated calcium ions and hydrated divalent iron ions, the mechanism of scale inhibition and corrosion under EAF in aqueous solutions may be more readily ascertained. This paper’s study involves a blend of CaCl2 and FeCl2. The position of the first water shell of calcium ions and divalent iron ions, the framework of hydrated calcium ions and divalent iron ions, the self-diffusion coefficients of calcium ions, divalent iron ions, and water molecules, and hydrogen bonding in the liquid solution were all investigated using MD simulations to study the structural and kinetic properties of aqueous CaCl2 solutions under an external EF. It should be noted that although the materials in this study and our previous work [11] are mixed aqueous solutions of two ions containing calcium ions, there are significant differences in the research details between the two. Our previous work mainly focused on the influence of magnesium ion impurities on the scaling process of calcium ions during the operation of the EAF system. However, in this study, we chose to pay attention to the influence of the EAF system on the corrosion process of iron water transmission pipes.

2. Model and Simulation Details

An equilibrium MD simulation was performed by using the Gromacs 5.0.2 software [12]. NPT ensembles were used for all simulations. The thermostat was adjusted to 300 K. The pressure was established at 1.01 × 105 Pa. As stated in Table 1, four sets of aqueous solution systems were constructed. The CaCl2 aqueous solution had a concentration of 4 mol/kg, and there was only a 0.5 percent difference in concentration between models and tests [13]. In our study, the water molecule was represented by the SPCE model. In this simulation, ion–water interactions were simulated using the GROMOS force field [14,15,16]. A cutoff of 1.2 nm was used for Lennard-Jones interactions, and a real-space cutoff of 1.2 nm was used for electrostatic interactions when using the particle mesh Ewald (PME) [15] technique. Fourier spacing has a value of 0.12 nm. The Lennard-Jones potential parameters for the particles utilized in the GROMOS force field are shown in Table 2. The Lennard-Jones parameters of Fe2+ were taken from Fernanda et al. [17]. All of the simulations employed a 2.0 fs time step, and a 2.5 × 104 ps simulation time was used to confirm the stability of the system setup. Following a 1 × 104 ps period of system equilibration, the structures and characteristics of the sample (apart from the self-diffusion coefficient data) were determined throughout a 1.5 × 104 ps period (the final 1.5 × 104 ps of each simulation). By computing the average result of 1.5 × 104 ps, which was separated into ten segments, the average values of the parameters were acquired. The leapfrog method is a type of integration algorithm used for motion equations. The simulation box measures 4 nm, 4 nm, and 3 nm in length. The EF was 1 V/μm in intensity. The EF was used in a certain way (corresponding to the y-axis). The simulations were run as described in the step above. Previous research has confirmed the accuracy of our MD simulations and error analyses [18].

3. Results and Discussion

3.1. Structural Parameters

The hydration shell regulates the mobility of ion pairs in the solution and the interaction between ion pairs [19]. A change in the hydration shell can affect calcium carbonate precipitation [20], and the layers of hydration around these ions define the route(s) for crystal formation in aquatic settings [21]. In this segment, the average radii of the first water shells of Ca2+ (RCa-O), Fe2+ (RFe-O), Ca2+ (nCa-O), Fe2+ (nFe-O), Ca2+ (nCa-O), and Fe2+ (nFe-O) under an external EF and under no EF are provided.
It is possible to determine RCa-O and RFe-O from the RDF [g(r)]; [g(r)] [22] is calculated as shown in Equation (6):
g A B ( r ) = ρ B ( r ) ρ B l o c a l = 1 ρ B l o c a l 1 N A i A N A j B N B δ ( r i j r ) 4 π r 2
where ‹ρB(r)› is the particle density of particle B on all spheres surrounding particle A with radius rmax, ‹ρBlocal is the particle density of particle B at distance r around particle A, NA denotes the number of particles A, and NB denotes the number of particles B. Typically, rmax has a value equal to half of the box’s minimum length.
The values of nCa-O and nFe-O are calculated using Equation (7) [14]:
n Ca O ( r s ) = 4 π ρ 0 0 r s g CaO ( r ) r 2 d r
For nCa-O, rs is the location of the first local minimum of gCaO(r), which is also the location of RCa-O; for nFe-O, rs is the location of the first local minimum of gFeO(r), which is also the location of RFe-O.
Figure 1 shows RDFs and the corresponding running integration number of the CaCl2 solution containing different numbers of Fe2+. The coordinates of the first peak in gCaO(r), the coordinates of the first peak in gFeO(r), RCa-O, RFe-O, nFe-O and nCa-O are presented.
According to Figure 1a,c,e, it can be seen that the height of the first peak in Fe2+-O g(r) is higher than its Ca2+-O g(r) counterpart. The first peak position of Fe2+-O g(r) is less than the first peak position of Ca2+-O g(r), which implies that Fe2+ has a smaller size than Ca2+, and the hydrated structure of Fe2+ is more stable than Ca2+ [23]. Furthermore, the higher the content of Fe2+ ions, the lower the heights of the first peak in Ca2+-O g(r). Fe2+ was added, but it had no discernible impact on RCa-O. With an increase in the amount of Fe2+ supplied, the nCa-O drops. The decrease in nCa-O could be due to the more stable hydrated structure of Fe2+ compared to Ca2+, which means that fewer water molecules remain in the water shell of Ca2+. In the CaCl2 solution containing Fe2+ under an EF, Figure 1b,d,f demonstrate that the values of RCa-O decrease while the values of nCa-O increase, which is consistent with our earlier research on “pure” CaCl2 aqueous solutions [24]. This is because the average binding ability of water molecules increases monotonically as the hydration shell becomes smaller [25,26]. In addition, the larger height and sharpness of the ferrous peak suggest a more stable structure under the EF, which means that the EF can make the Fe2+ radius smaller. The enthalpy of formation of ACC clusters decreases as the coordination number increases in the first hydration shell of Ca2+, and the stability of ACC correlates with water content. Experimentally, it has been proposed that “transient” ACC is largely anhydrous, while “stable” ACC contains some water [20]. Therefore, the EF increases the coordination number in the first hydration shell of Ca2+ and decreases the enthalpy of ACC, which can enhance the stability of ACC. To a certain extent, the EF reduces the possibility of ACC converting to calcite by changing the structure of the first hydration shell of Ca2+, which will play a role in scale inhibition (according to our previous study [27]). The creation of carbonate ions close to Ca2+ is made easier by the reduced radius of the initial water shell of Ca2+, which promotes the development of calcium carbonate crystals [24].
Figure 1 shows that the values of RFe-O range between 0.230 and 0.240 nm, and the values of nFe-O are distributed between 4.800 and 5.000, which is consistent with the results reported in Reference [28]. In CaCl2 solution under an EF, the values of RFe-O are larger, and the values of nFe-O are lower. According to experimental findings [29], this suggests that the EF decreases hydration between Fe2+ and water molecules.
In addition, we also calculated the RDF of chloride ions around Ca2+ and Fe2+ under an EF (not shown here). The calculation results show that although the influence of the EF on the distribution of chloride ions around cations is small, it enhances the interaction between cations and anions (the number of anions around cations increases slightly). This is consistent with the conclusion of our previous paper. Therefore, the distribution of chloride ions is not analyzed in detail in this section.

3.2. Self-Diffusion Coefficients of Ca2+ and Fe2+ and Water Molecule

This section examines Ca2+, Fe2+, and water molecule self-diffusion coefficients in the presence of an exogenous EF in the CaCl2 aqueous solution system. An important factor that may affect the course of chemical reactions in aqueous systems is the self-diffusion coefficient of ions. In addition, the self-diffusion coefficient may change significantly when the structure of the cation or anion is changed or another ion is involved [30]. It is worth noting that, in the simulation model in this paper, the solution seems to be strongly concentrated since there is 4 mol/kg CaCl2, which means 4 mol of Ca2+ and 8 mol of Cl- dissolved in 55.6 mol of water that solvates them (see Table 1). It may be considered that there are no free water molecules, because all water molecules seem to become ionic coordination water. This may make water molecules lack independent diffusivity. In fact, as the ions move in the water, the coordination water around the ions is not fixed, and they will be constantly replaced by the surrounding water molecules [13]. Therefore, the water molecules in this paper can be considered to have independent diffusivity, but this diffusivity will be affected by ions.
In this research, the “Einstein relation” [31] was used to calculate the self-diffusion coefficients of Ca2+(Dca) and water molecules (Do):
D = lim t 1 6 t [ r i ( t ) r i ( 0 ) ] 2
where ri(0) is the starting position, the parentheses represent the set mean, and ri(t) is the numerator position i at time t. Fitting the slope of the linear component of the mean square displacement (MSD) and dividing by six are the particular steps in computing the self-diffusion coefficient of a particle. The RANSAC algorithm determines the linear component of the MSD. RANSAC is a technique for robust model fitting in the presence of outliers in the data [32].
The impact of Fe2+ on the values of Dca and Do with and without an EF is shown in Table 3. With regard to the self-diffusion coefficient, the standard deviation of the simulation results presented in Table 3 is conservatively estimated to be in the region of 0.41%–5.11%. Table 3 shows that whether or not there is an EF, the values of Dca and Do both drop as the amount of Fe2+ increases. Moreover, the higher the Fe2+ ion content, the lower the values of Dca and Do. In addition, Table 3 shows that the external EF can increase the values of Dca and Do no matter how much Fe2+ exists in the CaCl2 aqueous solution during all simulations.
The valence and structural order of cations in aquatic environments impact the rate at which water molecules diffuse [33]. According to Table 3, the values of Do are 0.6626 × 10−5 cm2·s−1 and 0.7276 × 10−5 cm2·s−1 without and with an EF, respectively, when Fe2+ is not in the solutions, in contrast to the values of Do in the range of 0.12 × 10−5 cm2·s−1–0.18 × 10−5 cm2·s−1 when different numbers of Fe2+ ions are present. It is worth noting that when iron ions exist, the existence of an electric field can only slightly increase the diffusion coefficient of water molecules (with an increase of about 5%), which may be due to the binding of iron ions to water molecules, weakening the influence of the electric field. Water molecules that are coupled to metal ions are successfully immobilized in relation to the bulk water by being contained in a hydration layer [21]. As a result, when Fe2+ is available in the CaCl2 solution, its significant capacity to bind to water molecules restricts the activity of the water molecules. The Do values decline. The presence of Fe2+ is responsible for the decrease in the self-diffusion coefficient of Ca2+. The CaCl2 aqueous solution becomes more compact due to the electrostatic contact between Fe2+ and the water molecule. Ca2+ activity is constrained, and its self-diffusion coefficient is decreased.
The EF lowers the water molecule’s self-diffusion coefficient while increasing the hydrogen bond network’s binding capacity in pure water solutions [34]. However, the findings in Table 3 demonstrate that for the four example CaCl2 aqueous solutions, the EF may increase the water molecule’s self-diffusion coefficient. Therefore, the rise in the diffusion coefficient of water molecules in aqueous CaCl2 solutions can be attributed to the activity of metal ions in the solution. When Ca2+ and Fe2+ diffuse, they either displace their solvated shells along their trajectories or break the strong connections with the initial water shell [35]. Thus, the EF increases the activity of Ca2+ and Fe2+ and influences the relatively stable water molecule hydrogen bond network.
As can be seen in Table 3, the self-diffusion coefficients of water molecules and Fe2+ in aqueous CaCl2 solutions grow in the presence of EF. The speed at which oxidants are transported to the scale’s surface may be increased by increasing the activity of water molecules, which can reduce iron release and the amounts of iron corrosion products that dissolve in water [36,37]. Water contamination issues may be resolved. Additionally, the rise in Do and DFe may indicate that the EF has the ability to accelerate the oxidation of Fe2+.

3.3. Viscosity of the CaCl2 Aqueous Solution

Generally, the formation of calcium carbonate precipitation involves three stages: nucleation, grain growth, and aggregation of CaCO3 crystals [38]. The viscosity of the solution affects the precipitation of calcium carbonate and the corrosion behavior of iron in the solution [39,40,41,42,43]. According to experimental findings, the crystallinity of scale-forming ions mixed with water is inversely related to the viscosity of a liquid [44]. A liquid’s dynamic viscosity is another kinetic characteristic that affects how quickly solutes in the liquid alter their conformation. It is susceptible to an EF [44].
The following equation can be used to determine the viscosity (η):
η = A V ρ k 2
where ρ and V stand for density and velocity, respectively; A is a constant [44], and we can deduce the following from a cosine that meets both requirements:
a x = A cos ( k z ) , k = 2 π / l z
where lz is the box’s height. In the simulation, the instantaneous V is defined as a Fourier coefficient:
V ( t ) = 2 i = 1 N m i v i , x ( t ) cos [ k r i , z ( t ) ] i = 1 N m i
where mi is the mass of an atom, ri,z is the z coordinate, and vi,x is the x component of the velocity. The amplitude of the velocity profile can be completely formed before measuring the average for V.
Four sample CaCl2 aqueous solutions’ viscosities under an EF are shown in Figure 2. For comparison, the values in the absence of the EF are shown. As shown in Figure 2, the presence of Fe2+ increases the viscosity values in the CaCl2 system, and four typical aqueous CaCl2 solutions become viscous in the presence of Fe2+. The EF also reduces the viscosity of aqueous CaCl2 solutions.
The molecular weight of the particles inside the system is inversely proportional to the dynamic viscosity, which is a reflection of a sort of friction brought on by fluid motion in the solution system. Bakker [45] found that viscosity is a distinct feature that describes the typical behavior of a large number of water molecules in aqueous solutions and that specific ions have a significant effect on the viscosity of water because of the extended lifetime of their initial hydration shells. Viscosity is substantially influenced by the water molecules’ extended residence periods in ion hydration shells. The aqueous solution of CaCl2 also has no carbon–hydrogen bonds, making it an inorganic salt. Therefore, changes in the viscosity of CaCl2 solutions can be attributed to changes in the structural and kinetic properties of the water molecules.
Figure 2 demonstrates that without the EF, the viscosities of the four example CaCl2 aqueous solutions are 9.49 × 10−3 Pa·s, 16.31 × 10−3 Pa·s, 17.24 × 10−3 Pa·s, and 17.81 × 10−3 Pa·s. This means that the addition of Fe2+ can cause a relatively high viscosity in CaCl2 aqueous solutions. Dian et al. [46] studied Ca2+ solvation using polarizable atomic multipole potential, and their results show that the lifetime of water molecules in the first solvation shell of Ca2+ is 18 picoseconds (ps) when the temperature is 298 K. The experimental [47] values of the lifetime of water molecules in the first solvation shell of Fe2+ were determined in the range of 105–106 ps when the temperature is 298.15 K, in contrast to only a few picoseconds for Ca2+. Therefore, the addition of Fe2+ forms a water shell, which can lead to a longer residence time for water molecules in Fe2+ hydration shells compared to Ca2+. Therefore, the viscosity of the solution is greatly increased by Fe2+. Furthermore, the mobility of water molecules is related to the viscosity of the solution. According to Table 3, the activity of water molecules is more limited when the Fe2+ ion content increases, and this may result in a rise in the solution’s viscosity.
Figure 2 demonstrates that when Fe2+ ions are available in the solution, the EF can reduce the viscosity of CaCl2 aqueous solutions. This is due to the fact that the EF reduces the mean dynamic residence time of water at Ca2+ and increases the activity of water molecules, both of which lead to a drop in dynamic viscosity. The viscosity of the salt solution can be reduced by the EF [48]. Electrostatic interaction between the water molecule and Fe2+ might raise the viscosity of the solution. This indicates that the dynamic viscosity of the CaCl2 solution can be reduced by the presence of Fe2+ and an external EF. In conclusion, the aqueous CaCl2 solution can become more viscous when Fe2+ ions are present, and the viscosity can be reduced when an EF is applied to an aqueous CaCl2 solution that contains Fe2+. The decrease in viscosity of CaCl2 aqueous solutions, on the one hand, means that the scale-forming ions in the solution are reduced [40] and the fluidity of the solution is enhanced [49]. On the other hand, the decrease in the viscosity of CaCl2 aqueous solutions reduces the aggregation of calcium carbonate precipitates [41,42,43], which plays an effective role in anti-fouling. The decrease in solution viscosity reduces the adhesion of the electrode surface to ions and increases the rate of transport of oxidants, which can influence the capacitance values of the double layer and reflect changes in the surface state of the corrosion metal [36,37].

3.4. Effect of Fe2+ and Electric Field on Hydrogen Bonds

The hydrogen bond network that is generated by water molecules has a significant impact on the precipitation of calcium carbonate [20,21] and is strongly connected to the chemical and physical characteristics of aqueous solutions [50,51]. The dissolution of calcium carbonate and corrosion products is influenced by the nature of the hydrogen bond network. According to Gromacs software, a hydrogen bond occurs when a donor and acceptor are separated by a van der Waals (VDW) length less than 3.5Å(r) and when the hydrogen donor–acceptor angle is less than 30° [52]. The geometrical hydrogen bond requirement is displayed in Figure 3. The oxygen atoms in water molecules are the study’s donors and acceptors. This work estimated the hydrogen bond angle as ∠OD-HD- OA(θ), which is a suitable angle that is frequently used in the literature and offers helpful insight into the flexibility of donor hydrogen bonds [53]. Calculating the number of hydrogen bonds allows one to assess the overall importance of a hydrogen bond. The number of hydrogen bonds and the strength of the hydrogen bonds are positively correlated [52].
The current study looked at both the number of hydrogen bonds (nHB) and the percentage of water molecules with n or more hydrogen bonds (fn). Without the addition of Fe2+ (0F) and with the addition of 10 Fe2+ (10F), 20 Fe2+ (20F), and 30 Fe2+ (30F), the values of fn and nHB in CaCl2 aqueous solutions are shown in Table 4. (30F). Additionally, the effect of the EF is demonstrated.
Table 4 demonstrates how the EF may alter the percentage of molecules without hydrogen bonds while increasing the percentage of molecules with one to three hydrogen bonds and the average number of hydrogen bonds in a water molecule, regardless of the presence or absence of Fe2+. The proportion of water molecules with n or more hydrogen bonds might be used to define the dimension of the water cluster. For instance, the size of the water cluster increases with the proportion of molecules with two or three hydrogen bonds. The decrease in f0 and the rise in nHB, f1, f2, and f3 indicate that water molecule activity is constrained and that more water molecules are shifting from the free state to the hydrogen-bonded state. This indicates that the EF improves the hydrogen structure of the water molecules to some extent. The findings in Table 4 further demonstrate that, regardless of whether the EF is administered or not, Fe2+ can enhance f0 and reduce nHB, f1, f2, and f3. Additionally, in the CaCl2 systems, nHB, f1, f2, and f3 decrease as the amount of Fe2+ rises, whereas f0 increases. This is the result of the Fe2+ ion’s ability to attract ion–dipole interactions, which retains water molecules in the ion hydration shell and reduces the number of water molecules bound by hydrogen bonds [54]. This results in fewer hydrogen bonds being created, an increase in the percentage of molecules without hydrogen bonds, a drop in the percentage of molecules with one to three hydrogen bonds, and a rise in the mean number of hydrogen bonds per water molecule.

4. Conclusions

A methodical long-time balanced molecular dynamics simulation was used to examine the structural and dynamic characteristics of mixed solutions of CaCl2 and FeCl2 under an EF. Some significant findings were made.
(1)
Compared to Ca2+, the hydrated structure of Fe2+ is more stable. The EF may raise the first water coordination number for Ca2+ and reduce the radius of hydrated Ca2+. The EF has the ability to reduce the first water coordination number for hydrated Fe2+ and expand its radius. The presence of Fe2+ has no discernible impact on the radius of hydrated Ca2+; however, it can lower the first water coordination number of Ca2+.
(2)
The viscosity of CaCl2 in an aqueous solution can be efficiently raised by Fe2+. The presence of an EF can make the mixed aqueous solution less viscous.
(3)
The self-diffusion coefficient of Ca2+ can be lowered by Fe2+. Ca2+, Fe2+, and water molecules’ self-diffusion coefficients can all be increased by an EF.
(4)
An EF can enhance the hydrogen structure of the water molecules in the aqueous solution.
To sum up, the EF generated by an EAF system can change the structure of the solvent layer of Ca2+, resulting in a reduction in the possibility of Ca2+ forming high-density calcite scale. In addition, the existence of an EF can also increase the activity of water molecules and reduce the viscosity of the solution, which will effectively hinder the release of Fe2+ in water and the formation of iron corrosion products.

Author Contributions

Conceptualization, methodology, software, and validation, Y.H.; data curation and writing—original draft preparation, B.W.; visualization and investigation, X.G.; supervision and writing—reviewing and editing, T.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 51408525), University Science and Technology Research Projects of Hebei Province (No. ZD2020324), The Central Guidance on Local Science and Technology Development Fund of Hebei Province (No. 216Z3601G), and the Natural Science Foundation of Hebei Province (No. E2022203061).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Weiner, S.; Sagi, I.; Addadi, L. Choosing the crystallization path less traveled. Science 2005, 309, 1027–1028. [Google Scholar] [CrossRef] [PubMed]
  2. Qi, J.Q.; Guo, R.; Wang, Y.; Liu, X.W.; Chan, H.L.W. Electric field-controlled crystallizing CaCO3 nanostructures from solution. Nanoscale Res. Lett. 2016, 11, 120. [Google Scholar] [CrossRef] [PubMed]
  3. Li, H.H.; Liu, Z.F.; Gao, Y.H.; Zhang, L.H.; Li, X.H. Influence of electrostatic water treatment on crystallization behavior of CaCO3 and synergistic scale inhibition with a green scale inhibitor. CIESC J. Chem. Ind. Eng. 2013, 64, 1736–1742. [Google Scholar] [CrossRef]
  4. Lin, D.; Dai, N.; Chen, Y.; Ni, Q.; Zhang, X.; Zhang, J. Effect of direct current electric field on initial corrosion of steel exposed to simulated marine environment. Corros. Sci. Protetion Technol. 2017, 29, 63–67. [Google Scholar] [CrossRef]
  5. Sarin, P.; Snoeyink, V.L.; Lytle, D.A.; Kriven, W.M. Iron corrosion scales: Model for scale growth, iron release, and colored water formation. J. Environ. Eng. 2004, 130, 364–373. [Google Scholar] [CrossRef]
  6. Finan, M.A.; Harris, A.; Marshall, A. Ecologically acceptable method of preventing corrosion/scale problems by control of water conditions. Natl. Assoc. Corros. Eng. 1980, 204, 24–29. [Google Scholar]
  7. de Leeuw, N.H. Molecular dynamics simulations of the growth inhibiting effect of Fe2+, Mg2+, Cd2+, and Sr2+ on calcite crystal growth. J. Phys. Chem. B 2002, 106, 5241–5249. [Google Scholar] [CrossRef]
  8. Zhao, G.-J.; Liu, J.-Y.; Zhou, L.-C.; Han, K.-L. Site-selective photoinduced electron transfer from alcoholic solvents to the chromophore facilitated by hydrogen bonding: A new fluorescence quenching mechanism. J. Phys. Chem. B 2007, 111, 8940–8945. [Google Scholar] [CrossRef]
  9. Zhao, G.J.; Han, K.L. Hydrogen bonding in the electronic excited state. Accounts. Chem. Res. 2012, 45, 404. [Google Scholar] [CrossRef]
  10. Chen, J.S.; Yuan, M.H.; Wang, J.P.; Yang, Y.; Chu, T.S. Sensing mechanism for biothiols chemosensor DCO: Roles of excited-state hydrogen-bonding and intramolecular charge transfer. J. Phys. Chem. A 2014, 118, 8986–8995. [Google Scholar] [CrossRef]
  11. Zhu, L.; Han, Y.; Zhang, C.; Zhao, R.; Tang, S. Molecular dynamics simulation for the impact of an electrostatic field and impurity Mg2+ ions on hard water. RSC Adv. 2017, 7, 47583. [Google Scholar] [CrossRef]
  12. Nedyalkova, M.; Madurga, S.; Pisov, S.; Pastor, I.; Vilaseca, E.; Mas, F. Molecular dynamics simulation of the spherical electrical double layer of a soft nanoparticle: Effect of the surface charge and counterion valence. J. Chem. Phys. 2012, 137, 174701. [Google Scholar] [CrossRef] [PubMed]
  13. Li, M.; Duan, Z.; Zhang, Z.; Zhang, C.; Weare, J. The structure, dynamics and solvation mechanisms of ions in water from long time molecular dynamics simulations: A case study of CaCl2 (Aq) aqueous solutions. Mol. Phys. 2008, 106, 2685–2697. [Google Scholar] [CrossRef]
  14. Chialvo, A.A.; Simonson, J.M. The structure of CaCl2 aqueous solutions over a wide range of concentration. Interpretation of diffraction experiments via molecular simulation. J. Chem. Phys. 2003, 119, 8052–8061. [Google Scholar] [CrossRef]
  15. Isele-Holder, R.E.; Mitchell, W.; Ismail, A.E. Development and application of a particle-particle particle-mesh ewald method for dispersion interactions. J. Chem. Phys. 2012, 137, 174107. [Google Scholar] [CrossRef] [PubMed]
  16. Gunsteren, W.F.V.; Billeter, S.R.; Eising, A.A.; Hünenberger, P.H.; Krüger, P.K.H.C.; Mark, A.E.; Scott, W.R.P.; Tironi, I.G. Biomolecular Simulation: The GROMOS96 Manual and User Guide; Vdf Hochschulverlag AG an der ETH Zürich: Zürich, Switzerland, 1996; Volume 86, pp. 1–1044. [Google Scholar]
  17. Duarte, F.; Bauer, P.; Barrozo, A.; Amrein, B.A.; Purg, M.; Åqvist, J.; Kamerlin, S.C.L. Force field independent metal parameters using a nonbonded dummy model. J. Phys. Chem. B 2014, 118, 4351–4362. [Google Scholar] [CrossRef] [PubMed]
  18. Han, Y.; Zhao, Y. Influence to the structure and dynamics of hydrated Ca2+ and water molecules in CaCl2 aqueous solution caused by external AC electric field: A mechanism study by molecular dynamics simulation. Int. J. Electrochem. Sci. 2012, 7, 10008–10026. [Google Scholar]
  19. Lopez-Berganza, J.A.; Diao, Y.; Pamidighantam, S.; Espinosa-Marzal, R.M. Ab initio studies of calcium carbonate hydration. J. Phys. Chem. A 2015, 119, 11591–11600. [Google Scholar] [CrossRef]
  20. Raiteri, P.; Gale, J.D. Water is the key to nonclassical nucleation of amorphous calcium carbonate. J. Am. Chem. Soc. 2010, 132, 17623–17634. [Google Scholar] [CrossRef]
  21. Dorvee, J.R.; Veis, A. Water in the formation of biogenic minerals: Peeling away the hydration layers. J. Struct. Biol. 2013, 183, 278–303. [Google Scholar] [CrossRef]
  22. York, D.M.; Darden, T.A.; Pedersen, L.G.; Anderson, M.W. Molecular dynamics simulation of HIV-1 protease in a crystalline environment and in solution. Biochemistry 1993, 32, 3196. [Google Scholar] [CrossRef]
  23. Ye, Q.; Chen, X.; Zhou., J.; Zhao, H.F.; Chu, W.S.; Zheng, X.S.; Marcelli, A.; Wu, Z.Y. Local hydrated structure of an Fe2+/Fe3+ aqueous solution: An investigation using a combination of molecular dynamics and X-ray absorption fine structure methods. Chin. Phys. C 2013, 37, 118–123. [Google Scholar] [CrossRef]
  24. Han, Y.; Zhu, L.; Zhang, Y. Molecular dynamics simulation for the impact of external electric fields on CaCl2 aqueous solution. Chem. Res. Chin. Univ. 2016, 32, 641–646. [Google Scholar] [CrossRef]
  25. Cunningham, A.B.; Phillips, A.J.; Troyer, E.; Lauchnor, E.; Hiebert, R.; Gerlach, R.; Spangler, L. Wellbore leakage mitigation using engineered biomineralization. Energy Procedia 2014, 63, 4612–4619. [Google Scholar] [CrossRef]
  26. Yang, Q.; Liu, Y.; Gu, A.; Ding, J.; Shen, Z. Investigation of calcium carbonate scaling inhibition and scale morphology by AFM. J. Colloid. Interf. Sci. 2001, 240, 608–621. [Google Scholar] [CrossRef] [PubMed]
  27. Han, Y.; Zhang, C.; Wu, L.; Zhang, Q.; Zhu, L.; Zhao, R. Influence of alternating electromagnetic field and ultrasonic on calcium carbonate crystallization in the presence of magnesium ions. J. Cryst. Growth 2018, 499, 67–76. [Google Scholar] [CrossRef]
  28. Floris, F.; Persico, M.; Tani, A.; Tomasi, J. Ab Initio effective pair potentials for simulations of the liquid state, based on the polarizable continuum model of the solvent. Chem. Phys. Lett. 1992, 199, 518–524. [Google Scholar] [CrossRef]
  29. Busch, K.W.; Busch, M.A.; Parker, D.H.; Darling, R.E.; McAtee, J.L. Studies of a water treatment device that uses magnetic fields. Corrosion 1986, 42, 211–221. [Google Scholar] [CrossRef]
  30. Tsuzuki, S.; Shinoda, W.; Saito, H.; Mikami, M.; Tokuda, H.; Watanabe, M. Molecular dynamics simulations of ionic liquids: Cation and anion dependence of self-diffusion coefficients of ions. J. Phys. Chem. B 2009, 113, 10641–10649. [Google Scholar] [CrossRef]
  31. Henritzi, P.; Bormuth, A.; Klameth, F.; Vogel, M. A molecular dynamics simulations study on the relations between dynamical heterogeneity, structural relaxation, and self-diffusion in viscous liquids. J. Chem. Phys. 2015, 143, 164502. [Google Scholar] [CrossRef]
  32. Fischler, M.A.; Bolles, R.C. Random sample consensus: A paradigm for model fitting with applications to image analysis and automated cartography. Commun. Assoc. Comput. 1981, 24, 381. [Google Scholar] [CrossRef]
  33. Guo, B.; Han, H.; Chai, F. Influence of magnetic field on microstructural and dynamic properties of sodium, magnesium and calcium ions. Trans. Nonferrous Met. Soc. China 2011, 21, s494–s498. [Google Scholar] [CrossRef]
  34. Wei, S.; Zhong, C.; Su-Yi, H. Molecular dynamics simulation of liquid water under the influence of an external electric field. Mol. Simulat. 2005, 31, 555–559. [Google Scholar] [CrossRef]
  35. Bruneval, F.; Donadio, D.; Parrinello, M. Molecular dynamics study of the solvation of calcium carbonate in water. J. Phys. Chem. B 2007, 111, 12219–12227. [Google Scholar] [CrossRef]
  36. Ying, T.Y.; Yang, K.L.; Yiacoumi, S.; Tsouris, C. Electrosorption of ions from aqueous solutions by nanostructured carbon aerogel. J. Colloid. Interf. Sci. 2002, 250, 18–27. [Google Scholar] [CrossRef]
  37. Morimoto, T.; Hiratsuka, K.; Sanada, Y. Development and current status of electric double-layer capacitors. MRS Proc. 1995, 393, 397. [Google Scholar] [CrossRef]
  38. Chen, L.; Zhang, H.; Qi, Z.-P.; Yao, W.; Wan, J.-J.; Shao, M.; Zhang, C.-C. Synthesis of calcium carbonate crystals by using bacteria. Synth. React. Inorg. Met. Nano-Metal Chem. 2012, 42, 972–975. [Google Scholar] [CrossRef]
  39. Pach, L.; Hrabe, Z.; Komarneni, S.; Roy, R. Controlled crystallization of vaterite from viscous solutions of organic colloids. J. Mater. Res. 1990, 5, 2928–2932. [Google Scholar] [CrossRef]
  40. Ricciardiello, F.; Roitti, S. The corrosion of Fe and Ag in S liquid at lowtemperature. Effect of S viscosity. Corros. Sci. 1972, 12, 651–659. [Google Scholar] [CrossRef]
  41. Jada, A.; Verraes, A. Preparation and microelectrophoresis characterisation of calcium carbonate particles in the presence of anionic polyelectrolyte. Colloids Surf. A Physicochem. Eng. Asp. 2003, 219, 7–15. [Google Scholar] [CrossRef]
  42. Domingo, C.; García-Carmona, J.; Loste, E.; Fanovich, A.; Fraile, J.; Gómez-Morales, J. Control of calcium carbonate morphology by precipitation in compressed and supercritical carbon dioxide media. J. Cryst. Growth 2004, 271, 268–273. [Google Scholar] [CrossRef]
  43. Xie, A.J.; Zhang, C.Y.; Shen, Y.H.; Qiu, L.-G.; Xiao, P.P.; Hu, Z.Y. Morphologies of calcium carbonate crystallites grown from aqueous solutions containing polyethylene glycol. Cryst. Res. Technol. 2006, 41, 967–971. [Google Scholar] [CrossRef]
  44. Hess, B. Determining the shear viscosity of model liquids from molecular dynamics simulations. J. Chem. Phys. 2002, 116, 209. [Google Scholar] [CrossRef]
  45. Bakker, H.J. Structural dynamics of aqueous salt solutions. Chem. Rev. 2008, 108, 1456–1473. [Google Scholar] [CrossRef] [PubMed]
  46. Jiao, D.; King, C.; Grossfield, A.; Darden, T.A.; Ren, P. Simulation of Ca2+ and Mg2+ solvation using polarizable atomic multipole potential. J. Phys. Chem. B 2006, 110, 18553–18559. [Google Scholar] [CrossRef]
  47. Ohtaki, H.; Radnai, T. Structure and dynamics of hydrated ions. Chem. Rev. 1993, 93, 1157–1204. [Google Scholar] [CrossRef]
  48. Han, G.; Fang, Z.; Chen, M. Modified Eyring viscosity equation and calculation of activation energy based on the liquid quasi-lattice model. Sci. China Phys. Mech. Astron. 2010, 53, 1853–1860. [Google Scholar] [CrossRef]
  49. Ravi, K.R.; Pillai, R.M.; Amaranathan, K.R.; Pai, B.C.; Chakraborty, M. Fluidity of aluminum alloys and composites: A review. J. Alloy. Compd. 2008, 456, 201–210. [Google Scholar] [CrossRef]
  50. Lee, W.J.; Chang, J.G.; Ju, S.P. Hydrogen-Bond structure at the interfaces between Water/Poly(Methyl Methacrylate), Water/Poly(Methacrylic Acid), and Water/Poly(2-Aminoethylmethacrylamide). Langmuir 2010, 26, 12640–12647. [Google Scholar] [CrossRef]
  51. Pan, Z.; Chen, J.; Lü, G.; Geng, Y.Z.; Zhang, H.; Ji, Q. An Ab initio molecular dynamics study on hydrogen bonds between water molecules. J. Chem. Phys. 2012, 136, 164313. [Google Scholar] [CrossRef]
  52. Li, Y.X.; Wei, Z.D.; Zhao, Q.L.; Ding, W.; Zhang, Q.; Chen, S.G. Preparation of Pt/Graphene catalyst and its catalytic performance for oxygen reduction. Acta. Phys-Chim. Sin. 2011, 27, 858–862. [Google Scholar] [CrossRef]
  53. Lee, H.-S.; Tuckerman, M.E. Structure of liquid water at ambient temperature from ab initio molecular dynamics performed in the complete basis set limit. J. Chem. Phys. 2006, 125, 154507. [Google Scholar] [CrossRef] [PubMed]
  54. Karmakar, A.; Choudhuri, J.R.; Yadav, V.K.; Mallik, B.S.; Chandra, A. A first principles simulation study of vibrational spectral diffusion in aqueous NaBr solutions: Dynamics of water in ion hydration shells. Chem. Phys. 2013, 412, 13–21. [Google Scholar] [CrossRef]
Figure 1. RDFs (curves with one obvious peak) and corresponding running integration numbers (single increment curves) of CaCl2 solution containing different numbers of Fe2+ with an electric field applied (b,d,f) or without an electric field (a,c,e): (a,b) 10 Fe2+ added, (c,d) 20 Fe2+ added, and (e,f) 30 Fe2+ added.
Figure 1. RDFs (curves with one obvious peak) and corresponding running integration numbers (single increment curves) of CaCl2 solution containing different numbers of Fe2+ with an electric field applied (b,d,f) or without an electric field (a,c,e): (a,b) 10 Fe2+ added, (c,d) 20 Fe2+ added, and (e,f) 30 Fe2+ added.
Coatings 12 01165 g001
Figure 2. The 103 average viscosity with no Fe2+ ions added (0F) and 10 Fe2+ added (10F), 20 Fe2+ added (20F), and 30 Fe2+ added (30F) with and without EF and EF.
Figure 2. The 103 average viscosity with no Fe2+ ions added (0F) and 10 Fe2+ added (10F), 20 Fe2+ added (20F), and 30 Fe2+ added (30F) with and without EF and EF.
Coatings 12 01165 g002
Figure 3. Geometrical hydrogen bond criterion.
Figure 3. Geometrical hydrogen bond criterion.
Coatings 12 01165 g003
Table 1. Composition of the CaCl2 solution studied by Gromacs.
Table 1. Composition of the CaCl2 solution studied by Gromacs.
Sample[H2O][Fe2+][Ca2+][Cl][Total]
0F11100801601350
10F111010801801380
20F111020802001410
30F111030802201440
NOTE: “0F” indicates no Fe2+ ions were added, “10F” indicates 10 Fe2+ were added, “20F” indicates 20 Fe2+ were added, and “30F” indicates 30 Fe2+ were added.
Table 2. The MD simulations employed the Lennard-Jones potential parameters according to the combining rules of Lorentz and Berthelot, i.e., εij = (εiiεjj)1/2, σij = (σii + σjj)/2.
Table 2. The MD simulations employed the Lennard-Jones potential parameters according to the combining rules of Lorentz and Berthelot, i.e., εij = (εiiεjj)1/2, σij = (σii + σjj)/2.
εii/(kj/mol)σii/nm
Ca2+0.50690.2813
Fe2+2.135 × 10−61.9129
Cl1.28890.3470
O1.72500.2626
Table 3. The 105 self-diffusion coefficients of Ca2+ (Dca) and water molecules (Do) with no Fe2+ ions added (0F) and ten Fe2+ added (10F), twenty Fe2+ added (20F), and thirty Fe2+ added (30F) with and without EF.
Table 3. The 105 self-diffusion coefficients of Ca2+ (Dca) and water molecules (Do) with no Fe2+ ions added (0F) and ten Fe2+ added (10F), twenty Fe2+ added (20F), and thirty Fe2+ added (30F) with and without EF.
SampleAverage
10−5 cm2·s−1
St.dev
10−5 cm2·s−1
Average
10−5 cm2·s−1
St.dev
10−5 cm2·s−1
Without Electrostatic FieldUnder Electrostatic Field
Dca
0F0.06080.00030.07070.0001
10F0.01850.00050.02390.0001
20F0.01240.00030.01370.0007
30F0.01080.00050.01280.0001
Do
0F0.66260.00040.72760.0003
10F0.18370.00040.18700.0006
20F0.15730.00050.15960.0001
30F0.12830.00050.13430.0003
DFe
10F0.01880.00090.01950.0004
20F0.01580.00040.01800.0002
30F0.01420.00020.01720.0004
Table 4. The proportion of water molecules that have n hydrogen bonds (fn), the amount of Fe2+ that was added F(n*F), and the average number of hydrogen bonds per water molecule (nHB).
Table 4. The proportion of water molecules that have n hydrogen bonds (fn), the amount of Fe2+ that was added F(n*F), and the average number of hydrogen bonds per water molecule (nHB).
Without Electrostatic FieldUnder Electrostatic Field
0F10F20F30F0F10F20F30F
f0(%)58.0961.5462.2065.0257.9960.9361.1064.08
f1(%)31.1729.7529.1628.3931.0829.8829.8029.36
f2(%)8.237.377.306.598.347.807.707.04
f3(%)1.561.291.281.181.631.341.361.27
nHB580.67536.78528.55512.19584.43549.33546.80524.60
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Han, Y.; Wei, B.; Guo, X.; Jiao, T. Influence of Electrostatic Field on Mixed Aqueous Solution of Calcium and Ferrous Ions: Insights from Molecular Dynamics Simulations. Coatings 2022, 12, 1165. https://doi.org/10.3390/coatings12081165

AMA Style

Han Y, Wei B, Guo X, Jiao T. Influence of Electrostatic Field on Mixed Aqueous Solution of Calcium and Ferrous Ions: Insights from Molecular Dynamics Simulations. Coatings. 2022; 12(8):1165. https://doi.org/10.3390/coatings12081165

Chicago/Turabian Style

Han, Yong, Bingjia Wei, Xiaoqiang Guo, and Tifeng Jiao. 2022. "Influence of Electrostatic Field on Mixed Aqueous Solution of Calcium and Ferrous Ions: Insights from Molecular Dynamics Simulations" Coatings 12, no. 8: 1165. https://doi.org/10.3390/coatings12081165

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