Next Article in Journal
Crystallization Kinetics and Structure Evolution during Annealing of Ni-Co-Mn-In Powders Obtained by Mechanical Alloying
Previous Article in Journal
Bio-Stimulated Surface Healing of Historical and Compatible Conservation Mortars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Determination of the Ni(II) Ions Sorption Mechanism on Dowex PSR2 and Dowex PSR3 Ion Exchangers Based on Spectroscopic Studies

by
Justyna Bąk
1,*,
Weronika Sofińska-Chmiel
2,
Maria Gajewska
2,
Paulina Malinowska
2 and
Dorota Kołodyńska
1
1
Department of Inorganic Chemistry, Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie-Skłodowska University, Maria Curie-Skłodowska Sq. 2, 20-031 Lublin, Poland
2
Analytical Laboratory, Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie Skłodowska University, Maria Curie Skłodowska Sq. 3, 20-031 Lublin, Poland
*
Author to whom correspondence should be addressed.
Materials 2023, 16(2), 644; https://doi.org/10.3390/ma16020644
Submission received: 13 December 2022 / Revised: 4 January 2023 / Accepted: 5 January 2023 / Published: 9 January 2023

Abstract

:
This paper estimates the suitability of the strongly basic anion exchangers, Dowex PSR2 and Dowex PSR3, as sorbents of nickel ions in aqueous solutions. These actions are aimed at searching for new solutions due to the growing discharge of nickel into wastewaters, primarily due to its addition to steel. The nickel sorption experiments were conducted under static conditions and resulted in the optimization of pH, phase contact time, initial solution concentration, and temperature. The next step was to calculate the kinetic, isothermal, and thermodynamic parameters. Moreover, the ion exchangers were characterized by means of Fourier transform infrared spectroscopy, X-ray photoelectron spectroscopy, and CHN elemental analysis. It was found that the sorption process was most effective at pH 6 after 240 min and at the temperature of 293 K. The values of the thermodynamic parameters revealed that the adsorption was exothermic and spontaneous. The physicochemical analyses combined with the experimental research enabled determination of the sorption mechanism of Ni(II) ions.

1. Introduction

Economic growth in the world involves the use of numerous materials that are dangerous to the natural environment. As a result, the risk of air, water, and soil pollution is increasing [1,2]. The most dangerous pollutants for humans are heavy metals, poorly decomposing organic substances, and insoluble mineral compounds. An example of a heavy metal with allergic and carcinogenic effects is nickel [3,4,5]. The WHO recommends less than 0.1 mg/L of Ni(II) as the maximum permissible content in drinking water [6]. In terms of its distribution on earth, nickel ranks seventh. The average nickel content in the Earth’s crust is 100 mg/kg, and in the seawater it is about 5.4 g/L. Despite its wide distribution in nature, pure nickel itself is rare, most often in the form of sulfide, oxide, and silicate minerals [7]. Due to its properties, this element has many industrial applications. Nickel as a metal is malleable and suitable for welding, forging, or rolling. In addition, it forms alloys that are easily resistant to high temperatures and corrosion [8]. In addition, under the normal conditions, the metal is stable in water and air. Due to these properties, nickel is used in the electroplating industry to protect the surfaces of other metals [9]. This element is also used as a catalyst due to the fact that its crystal lattice can absorb hydrogen atoms [10]. The electrochemical properties of nickel are also used in the production of batteries, cadmium–nickel, and nickel–metal hydride cells [11,12]. This element is widely applied in the production of stainless steel. In the jewelry industry, despite its allergenic properties, nickel is used to make jewelry or eyeglass frames [13]. The basic information about nickel properties and applications is presented in Figure 1.
Ecology as a whole and the groundwater could be seriously endangered by nickel contamination from industrial and mining activities. This is dangerous due to the bioaccumulative nature of this element [14]. There are many ways to purify water, and one of the most appreciated is the ion exchange with the use of ion exchangers [15,16]. Compared with purification by precipitation or solvent extraction, the usage of ion exchange resins can offer a more efficient and straightforward purification [17]. Moreover, the ion exchange has also been successfully applied in a variety of applications, including the recovery of heavy metals from industrial wastes and the separation of rare earths or gas combinations to enhance the treatment of industrial effluents, as well as in the food sector [18]. There are many studies examining the efficiency of removing nickel ions from solutions with the use of various types of ion exchangers. Kołodyńska [19] described the results of research on the removal of Cu(II), Zn(II), and Ni(II) ion complexes with 1-hydroxyethylene-1,1-diphosphonic acid (HEDP) using Amberlite IRA 458, Amberlite IRA 67, Amberlite IRA 958, Purolite S-920, and Purolite S-930. Heavy metal complexes with HEDP within 20 min and at pH 11.5 are very well sorbed (even above 99%) on the highly basic anion exchangers Amberlite IRA 458 and Amberlite IRA 958, as well as the chelating ion exchangers Purolite S-920 and Purolite S-930, with the exception of the weakly basic anion exchanger Amberlite IRA 67, where the optimum pH was determined to be 7.0 [19]. Stefan and Meghea [20] investigated the simultaneous removal of Ca(II), Ni(II), Pb(II), and Al(III) ions using Purolite S-930. It was found that the ion exchange capacity of Ca(II), Ni(II), and Pb(II) ions increases relatively linearly in the pH range from 2.0 to 6.5. The greatest selectivity of the resin occurs for Ni(II), then Pb(II), while for the Al(III) ions, the maximum value of qt is observed at pH 4.45. At higher pH values, the sorption capacity decreases due to the microprecipitation of aluminum on the resin grid [20]. Dizge et al. [21] conducted studies on the sorption of nickel(II) ions from the aqueous solution using Lewatit MonoPlus SP112 (strongly acidic, macroporous cation exchange resin). The adsorption process was relatively fast, and it took about 90 min to reach equilibrium. However, approximately 3 h of the contact time was required to achieve the process efficiency of 95–100%. The maximum adsorption capacity of the resin was 170.94 mg/g at 298 K according to the Langmuir isotherm [21].
The purpose of this study was to evaluate the effectiveness of Ni(II) ion removal from aqueous solutions using the ion exchangers manufactured by DOW Chemical Company (Midland, MI, USA). In order to design technologies for the purification of water from heavy metals, it is necessary to research the structure and properties of materials, especially since there are very few literature reports on them. Understanding the structure of the exchangers combined with the results of research on the reaction rate and factors influencing the sorption processes (pH, phase contact time, initial concentration of the solution, and temperature) is undoubtedly a novel aspect of this paper that will allow us to fully understand the reaction mechanisms, as well as to determine the potential of the tested materials. Figure 2 depicts the general scheme of ion exchange, as well as the purpose of the research.

2. Materials and Methods

2.1. Sorbate and Sorbents Characteristics and Calculations

The appropriate amount of Ni(NO3)2∙6H2O (analytically pure) (Avator Performance Materials, Gliwice, Poland) was dissolved in distilled water and a 1000 mg/L nickel ions stock solution was prepared. Working solutions for the studies of the kinetics and sorption isotherms with the concentrations in the range of 10–500 mg/L were prepared by diluting the stock solution. The dropwise additions of 1 mol/L HNO3 and/or 1 mol/L NaOH to the solution promoted obtaining the required pH. A laboratory shaker (Elpin + 358A, Katowice, Poland) was used to shake the 100 mL Erlenmeyer flasks containing 0.1 g of ion exchanger and 50 mL of solution continuously at 180 rpm and the amplitude 8 during the batch adsorption studies. The impacts of pH (2–7), interaction time (1–240 min), initial adsorbate concentrations (10–500 mg/L), and temperature (293–333 K) were tested. After shaking, the sorbent was removed from the mixture by filtration, and the pH of the filtrate was assessed using the pHmeter pHM82 (Radiometer, Copenhagen, Denmark). After that, the atomic absorption analyzer SpectrAA240 FS (Varian, Palo Alto, CA, USA) at the wavelength of 232.0 nm was applied to determine the final nickel ions concentration. The basic adsorption parameters were calculated according to the equations:
q t = ( C 0 C t ) V m
q e = ( C 0 C e ) V m
where qt and qe are the adsorption capacities at time t (mg/g) and equilibrium (mg/g), respectively. C0, Ct, and Ce are the initial, after time t, and at the equilibrium concentrations (mg/L), respectively, V is the solution volume (mL), and m is the sorbent mass (g).
The physicochemical properties of Dowex PSR2 and Dowex PSR3 are listed in Table 1.
The kinetic models, such as the pseudo-first-order (PFO), pseudo-second-order (PSO), Elovich kinetic equations (EKE), and intraparticle diffusion (IPD) ones, were applied to evaluate the results. To analyze the equilibrium data, the Langmuir, Freundlich, Temkin, and Dubinin–Raduszkiewicz, as well as Sips, isotherm models were used. The formulae and parameters of the kinetic and isotherm models are presented in Table 2.
Using the formulae provided in [22], the thermodynamic parameters’ entropy change ΔS° (kJ/mol), free energy change ΔG° (kJ/mol), and enthalpy change ΔH° (kJ/mol) were calculated and interpreted.

2.2. FTIR Spectroscopy

The FTIR-ATR spectroscopic experiments were conducted to study the chemical composition and clarify the mechanism of Ni(II) ions sorption on Dowex PSR2 and PSR3. The FTIR Thermo Nicolet 8700 (Thermo Scientific, Waltham, MA, USA) spectrometer with the Smart OrbitTM attachment diamond ATR and the DTGS (deuterated triclycine sulfate) detector was used to test the Dowex PSR2 and Dowex PSR3 ion exchanger samples before and after the Ni(II) ions sorption process at 10, 100, 200, and 500 mg/L. Before the measurements, the samples of ion exchangers were ground in the agate mortar. The tests were carried out at room temperature in the wavenumber range of 4000–400 cm−1, with the spectral resolution of 4 cm−1. A total of 64 scans were made for each spectrum. Using the Omnic SpectaTM software(number 833-036200), the collected spectra were subjected to ATR correction, automated baseline correction, and normalizing.

2.3. XPS Spectroscopy

In order to examine the chemical structure and explain the mechanism of Ni(II) ions sorption on Dowex PSR2 and Dowex PSR3, the XPS tests were made using the Ultra-High Vacuum multichamber analytical system (Prevac, Rogów, Polska). The tests were carried out using Dowex PSR2 and Dowex PSR3 before and after the Ni(II) ions sorption. The samples were degassed at ambient temperature to a high continuous vacuum of approximately 5 × 10−8 mbar in the UHV system sluice after being fixed on a molybdenum carrier. Following their entrance into the system’s analytical chamber, the proper analysis was carried out using the XPS spectroscopy method. Spectra were collected using the hemispherical Scienta R4000 electron analyzer (Scienta Scientific AB, Uppsala, Sweden) The Scienta SAX-100 X-ray source (Al Kα, 1486.6 eV, 0.8 eV band) equipped with the XM 650 X-ray Monochromator (Inrad Optics Inc., Northvale, NJ, USA) (0.2 eV band) was used as a complementary.
The pass energy of the analyzer was set to 200 eV for the survey spectra (with 500 meV step) and 50 eV for the C1s region (high-resolution spectra with the 50 meV step). The base pressure in the analysis chamber was 5 × 10−9 mbar. During the spectra collection, it was not higher than 3 × 10−8 mbar.

2.4. CHN Analysis

The samples of Dowex PSR2 and Dowex PSR3 ion exchangers before the sorption process were tested using the CHN/CHNS EuroEA3000 analyzer (EuroVector, Milan, Italy) which enables automatic and simultaneous determination of the percentage content of carbon, hydrogen, and nitrogen. Prior to the measurements, the samples were weighed using the analytical microbalance M2P by Sartorius, with the accuracy of 0.001 mg. The test was based on the Dumas dynamic combustion method, followed by the chromatographic separation of gaseous fractions resulting from this combustion (N2, CO2, and H2O), and then their analysis using the catarometer. The test results were compiled using the Callidus package. Two combustion processes were performed for each sample.

3. Results and Discussion

3.1. Optimization of the Sorption Process

Based on the nickel speciation plot, it is concluded that pH plays an important role in the sorption process. Figure 3 indicates the prevalent Ni species at various pH values, for example, Ni2+ at pH < 8.0, Ni(OH)+ at pH 9.0–10.0, Ni(OH)2 at pH 10.0–11.0, and Ni(OH)3 at pH > 11.0. At pH > 6.7, the majority of nickel exists in the form of solid hydroxides, but pH 6.5 is found to make all nickel complexes generally soluble [23].
Figure 4 presents the effects of Ni(III) ions sorption on Dowex PSR2 and PSR3. The research was carried out in the pH range of 2.0–7.0, because at pH above 7.0 nickel hydroxide precipitated. The amount of adsorbed ions increases gradually as pH rises from 2.0 to 6.0, and the maximum value is obtained at pH 6.0. As a result, this amount was chosen for further investigations. The amount of adsorbed Ni(II) ions was 5.46 mg/g for PSR2 and 8.65 mg/g for PSR3. Furthermore, the final pH value increases after the sorption process.
The next step in optimizing the sorption process was determining the effects of phase contact time and initial concentration of solution (Figure 5). The studies were carried out in the contact time range of 0–240 min and the initial solution concentration from 10 to 100 mg/L.
The value of adsorbed Ni(II) ions increases with time, regardless of the initial solution concentration, which improved adsorption efficiency. After 240 min, the qt values were equal to: 2.24, 3.30, 4.92, and 10.30 mg/g for the concentrations: 10, 25, 50, and 100 mg/L, respectively, for Dowex PSR2. These values for Dowex PSR3 for the same concentrations are: 2.51, 6.76, 8.59, and 19.10 mg/g. The much greater amounts of the adsorbed ions, especially for the concentration of 100 mg/L for the PSR3 ion exchanger, indicate a greater affinity of nickel ions for PSR3. The same results were obtained by Wołowicz [24], who investigated the zinc(II) ions removal from the model chloride and chloride–nitrate(V) solutions using different ion exchangers, among others, PSR2 and PSR3. In the zinc(II) sorption from the chloride solution at 0.1 mol/L concentration, the qt equals 2.4 mg/g for PSR2 and 4.1 mg/g for PSR3 [24]. Additionally, it was found that the initial solution concentrations affected the time required to reach the equilibrium of sorption. A total of 30 min of sorption time is sufficient to reach the equilibrium for the concentration of 10 mg/L. For 25 and 50 mg/L, the time is extended to 60 min. The slowest equilibration occurs at the highest concentration of 100 mg/L.
Table 3 shows the comparison of the equilibrium capacities and sorption conditions of the other ion exchangers in relation to the nickel ions. Many authors have undertaken research on the removal of nickel ions from aqueous, chloride, and nitrate solutions under various conditions. On the basis of the equilibrium capacities listed in the table, it can be concluded that Dowex PSR2 and PSR3 have high qe values. Wołowicz and Hubicki [25] carried out research under similar conditions (C0 100 mg/L, t 240 min, and T 298 K), and the only difference was the removal of nickel ions from the chloride solutions at different concentrations of HCl, from 0.1 to 6.0 mol/L. Regardless of the hydrochloric acid concentration, smaller amounts of adsorbed Ni(II) ions were obtained. Therefore, it can be stated that the removal of nickel ions from the aqueous solutions using the Dowex PSR2 and PSR3 ion exchangers is very effective and can be successfully transferred to an industrial scale.
The sorption process was optimized studying the impact of temperature in the next step. In this case, the investigations were carried out at 293, 313, 323, and 333 K for the initial nickel ions concentrations from 10 to 500 mg/L (Figure 6). These tests will also be used to calculate the thermodynamic parameters of the process and to indicate its endo- or exothermic character.
The equilibrium capacity increases with the increasing initial concentration, reaching 16.93 mg/g at the highest concentration for Dowex PSR2 and a temperature of 293 K. As the temperature increases, it can be seen that the qe values decrease and are equal to 15.51, 14.14, and 13.09 mg/g for 313, 323, and 333 K, respectively (for the concentration of 500 mg/L and PSR2). For Dowex PSR3, the qe values changed as follows: 29.40, 26.59, 24.09, and 22.48 mg/g for the temperatures of 293, 313, 323, and 333 K. The decrease in the nickel ion diffusion toward the outer surface and into the pores of ion exchangers caused by the rise in temperature can be utilized to explain this. Thus, the optimal temperature for nickel ions sorption on Dowex PSR2 and PSR3 is 293 K. These investigations also demonstrate the greater efficiency of removing nickel ions from the aqueous solutions using Dowex PSR3.

3.2. Kinetic, Isotherm, and Thermodynamic Parameters

As can be seen in Table 4 and Table S1, the parameters of the two kinetic models showed the better fit of the PSO model than that of the PFO one with higher R2 values. Additionally, the qe values predicted by the PSO model were in excellent agreement with the experimental values. This suggests that this model was more appropriate to describe the sorption behavior of Ni(II) ions. This can reflect the fact that the nickel removal is controlled by both surface diffusion and chemical reactions at the liquid–solid interface [34].
The PSO model, which does not take into account the reaction mechanism, exhibits a fit for the whole time period under study. Due to the fact that the PFO and PSO models do not describe the adsorptive diffusion mechanism, the intraparticle diffusion (IPD) model is used for this purpose. This is to define the steps that control the rate of the reaction [35]. The first stage is related to the diffusion of heavy metal ions through the solution to the outer sorbent surface. In this stage, the adsorption process is fast, and the curve is steeply sloped. The intramolecular diffusion process slows down the reaction rate in the second step, flattening the curve as a result. In the third stage, when the concentration of ions in the solution decreases due to reaching the equilibrium, the diffusion process slows down [36]. The values of the coefficients ki, varying in the range ki1 > ki2 > ki3, suggest a reduction in the reaction rate for the subsequent stages of the diffusion process.
The Elovich equation (EKE) is used to describe chemisorption. The increase in the initial Ni(II) ions concentration from 10 to 100 mg/L resulted only in a slight decrease in the desorption constant (β) for Dowex PSR2 from 2.190 to 0.562 g/mg and from 3.885 to 0.735 g/mg for Dowex PSR3. A small β value (<1) suggests an irreversible adsorption process. Interestingly, for the concentrations in the range of 10–50 mg/L, the β values are above unity, suggesting reversibility of the process, and for the concentrations of 100 mg/L, they are below unity, which may suggest an irreversible process [37]. For the EKE model, obtained correlation coefficients are greater than 0.909. It should be noted that the other kinetic models provide a superior fit with the experimental data, notwithstanding the Elovich equation utility in estimating reaction kinetics.
Table 5 displays the findings of the analysis of the equilibrium data using the Langmuir, Freundlich, Temkin, and Dubinin–Raduszkiewicz isotherms. The assumption that the uniform adsorption takes place on the sorbent active sites forms the basis of the Langmuir isotherm, which was used to characterize the adsorption phenomena. Furthermore, once an adsorbate occupies an active site, no more sorption can take place there. On the other hand, the nonideal sorption on the heterogeneous surfaces was described by means of the Freundlich isotherm model [38,39]. Compared with the determination coefficients of the Langmuir and Freundlich models, higher values were obtained for the Langmuir one. The maximum removal of Ni(II) ions predicted by the Langmuir model was 19.16 mg/g for Dowex PSR2 and 31.72 mg/g for PSR3. These values are slightly overestimated, because the experimental data equal 16.93 mg/g for PSR2 and 29.40 mg/g for PSR3.
According to the Temkin model, the heat of adsorption for all particles decreases linearly with an increase in the surface area covered by the adsorbent. Additionally, the spread of binding energy up to the topmost binding energy describes the adsorption process precisely [40]. The change in the adsorption energy is associated with the bT constant of the Temkin isotherm. An endothermic adsorption reaction is indicated by a negative value of the bT constant. However, in this research, the exothermic nature of the adsorption reaction is suggested by the found positive values, which is supported by the thermodynamic parameters [41]. The calculated determination coefficient ≥0.915 indicates the participation of electrostatic interactions in the sorption process.
The Dubinin–Raduszkiewicz isotherm model defines the heterogeneity of ion adsorption on the surface and makes a distinction between physical and chemical adsorption processes [40]. The evaluation of the sorption character is based on the estimated parameter Ea value, which is defined as the free energy transfer of 1 mole of the solute from the infinite of the sorbent surface [42]. When Ea is in the range of 1–8 kJ/mol, it indicates physisorption. The activation energy from 8 to 16 kJ/mol indicates the ion exchange. On the other hand, the values in the range of 16–40 kJ/mol confirm chemisorption [42,43]. The activation energies: 10.074 kJ/mol for Dowex PSR2 and 9.753 kJ/mol for PSR3 typify ion exchange as the main mechanism for removing nickel ions.
The Sips isotherm combines the Freundlich and Langmuir isotherms and is useful in heterogeneous adsorption, in which the adsorbed molecule has several adsorption sites. Adsorbate–adsorbate synergy is not taken into account by the model. The estimated adsorption capacity for PSR2 is overpriced and equals 27.74, but for PSR3, it is understated and equals 25.53 mg/g. For Dowex PSR2, the correlation coefficient results were best fitted with the Sips isotherm model with R2 ≥0.982.
Table 6 lists the thermodynamic parameters of Ni(II) ions sorption on two ion exchangers. The negative value of ΔH° suggests that the process is exothermic, which is confirmed by the values of bT parameters determined on the basis of the Temkin isotherm. The adsorption process is spontaneous, as indicated by the positive value of ΔS° and the negative value of Gibbs free energy (ΔG°). Additionally, the values of the ΔG° decrease with rising temperature, which proves that the sorption process is more effective at lower temperatures [44,45].

3.3. Physicochemical Analyses

3.3.1. FTIR Spectroscopy

In order to confirm the chemical structure of the tested Dowex PSR2 and Dowex PSR3 ion exchangers and to observe the changes occurring under the influence of Ni(II) ions sorption, FTIR spectroscopic studies were carried out using the ATR technique. FTIR-ATR spectra of the pure ion exchanger were obtained before the sorption process and after the Ni(II) ions sorption process, with the concentrations of 10, 100, 200, and 500 mg/L. The test results are presented in Figure 7, Figure 8, Figure 9 and Figure 10.
The FTIR-ATR tests of the Dowex PSR2 ion exchanger before the sorption process showed a band of great intensity in the range of 3600–3200 cm−1. The stretching vibrations of the O-H groups and the N-H groups are represented by this band [46]. In the wavenumber range of 3090–3000 cm−1, the presence of bands corresponding to the stretching vibrations of the C-H groups was found [47]. The FTIR-ATR study also presented the presence of a band corresponding to the symmetric and asymmetric stretching vibrations originating from the aliphatic CH2 groups in the wavenumber range of 3000–2850 cm−1 [46,47,48]. An intense peak at 1705 cm−1 was also found, corresponding to the stretching vibrations of the C=O groups [49]. In the wavenumber range of 1610–1500 cm−1, the bands’ characteristic of symmetric C=C stretching vibrations of aromatic rings, characteristic of the ion exchange matrix, at 1454, 1378, and 900–703 cm−1, were observed [29,50,51]. The presence of numerous, low-intensity bands, including deformation vibrations of CH2 groups, as well as stretching vibrations of C-C in the range of 1215–1150 cm−1, were found. The tests also showed the presence of delocalized skeletal vibrations of the ion exchanger matrix in the position of 552 cm−1.
The FTIR spectroscopic studies carried out on the Dowex PSR2 ion exchanger after the Ni(II) ions sorption process showed the presence of the same FTIR-ATR spectral bands as found before the sorption process. However, some differences in signal intensity across the spectrum were observed. The intensity of the signals corresponding to the stretching vibrations of the O-H and the N-H groups (in the range of 3600–3200 cm−1) decreases as the concentration of the Ni(II) ions solution rises. With the increasing concentration of Ni(II) ions in the FTIR spectra, changes in the peak intensity in the range of 1530–1250 cm−1 were observed. In this range, there are peaks in positions 1512 cm−1, 1454, 1378, and 900–703 cm−1, corresponding to the stretching vibrations of aliphatic amino groups present in the ion exchanger functional groups [29,50,51]. Changes in the intensity of the peaks in this range may indicate the ongoing sorption process involving the nitrogen atoms present in the tri-n-butylammonium functional groups.
In the FTIR-ATR spectra of the Dowex PSR3 ion exchanger, a band of high intensity in the range of 3600–3200 cm−1 was observed before the sorption process. This band corresponds to the stretching vibrations of the O-H groups and the N-H groups [52]. At the wavenumbers in the range of 3090–3000 cm−1, stretching vibrations of the C-H groups were observed [47]. In the range of 3000–2850 cm−1, the bands corresponding to the symmetric and asymmetric stretching vibrations originating from the aliphatic CH2 groups were observed [46,47,48]. In the wavenumber range of 1610–1500 cm−1, the bands characteristic of the symmetric C=C stretching vibrations of the aromatic rings of the ion exchanger matrix were observed. The tests showed the presence of stretching vibrations originating from the aliphatic amino groups at the following signal positions: 1512 cm−1, 1454, 1378, and 900–703 cm−1 [29,50,51]. The FTIR spectra also showed the presence of numerous, low-intensity bands covering the deformation vibrations of the CH2 groups, and stretching C-C in the range of 1215–1150 cm−1. In the position of 552 cm−1, the presence of delocalized skeletal vibrations of the ion exchanger matrix was found.
Comparison of the FTIR spectra of the Dowex PSR3 ion exchanger before and after the process of Ni(II) ions sorption also showed significant differences in the signal positions in the range of 1454–1200 cm−1. In this range, the signals corresponding to the stretching vibrations of the aliphatic amino groups present in the functional group of the ion exchanger were located. This may indicate the ongoing sorption process with the participation of nitrogen atoms present in the tri-n butylammonium functional groups [29,50,51].

3.3.2. XPS Spectroscopy

To identify the elemental composition and chemical bonds characteristic of the tested ion exchanger, XPS spectroscopic tests were carried out. The analyses were made for the pure Dowex PSR2 and Dowex PSR3 ion exchangers and after the Ni(II) ions sorption with the concentration of 100 mg/L. The test results are given in Figure 11 and Figure 12 and Table 7, Table 8, Table 9 and Table 10.
The XPS tests of Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process showed the following elemental composition: Dowex PSR 2: 83.3% carbon, 2.2% nitrogen, 11.5% oxygen, and 1.6% chlorine; Dowex PSR3: 87.7% carbon, 1.6% nitrogen, 8.1% oxygen, and 2.7% chlorine. In addition, in the sample of the Dowex PSR 2 ion exchanger, the presence of aluminum, silicon, and sulfur was detected in an amount not exceeding 1 atomic %.
The XPS studies of Dowex PSR2 and Dowex PSR3 ion exchangers carried out in a narrow range of carbon binding energy showed the presence of bonds characteristic of both the ion exchange matrix (DVB cross-linked polystyrene) and tri-n-butylamine functional groups. The presence of C=C, C-H, C-C, and CN bonds and carbon bound to the aromatic groups was confirmed. The XPS tests carried out for oxygen showed the presence of nitrates and C-O bonds. The presence of chlorine in the form of Cl and chlorates ions was also observed. The presence of chlorates was confirmed in the Dowex PSR3 ion exchanger sample. The analyses of the Dowex PSR2 and Dowex PSR3 ion exchangers made in a narrow range of binding energy for nitrogen showed the presence of nitrogen in the form of a quaternary amine, which agrees with the information provided by the ion exchanger manufacturer. The presence of a protonated amine and nitro group was also observed in the Dowex PSR2 ion exchanger.
The XPS analysis of the Dowex PSR2 and Dowex PSR3 ion exchangers after the Ni(II) ions sorption process showed the following elemental compositions: Dowex PSR2: 80.10% carbon, 3.0% nitrogen, 14.4% oxygen, and 0.3% chlorine; Dowex PSR3: 82.3% carbon, 3.1% nitrogen, 14.2% oxygen, and 0.5% chlorine. In addition, the Dowex PSR2 ion exchanger sample contained aluminum, silicon, and sulfur in an amount not exceeding 1%. Due to the too low sensitivity of the method, the XPS tests for both ion exchangers after the Ni(II) ions sorption process did not show the presence of this element. The nickel content in the tested samples was determined by the XRF (X-ray fluorescence spectroscopy). The test showed a nickel content of 4.0 mg/L in the Dowex PSR2 ion exchanger and 6.4 mg/L in the Dowex PSR3 ion exchanger. The tests of ion exchangers carried out in a narrow range of binding energies for carbon showed the presence of the following bonds: C=C, C-H, C-C, CN, -COOH, and C-O and carbon linked to the aromatic groups. Particular attention in the XPS analysis was paid to the spectra of nitrogen after the Ni(II) ions sorption process. The XPS studies of the Dowex PSR2 ion exchanger after the sorption process showed a smaller number of groups in the form of quaternary amine from 31.5% to 21.1% and the protonated amine from 61.9% to 50.7%, which suggests the participation of nitrogen from the functional groups of the ion exchanger in the Ni(II) ions sorption process. In the case of the Dowex PSR3 ion exchanger, changes in the nitrogen spectra were also observed. A decrease in the amount of nitrogen was found in the form of quaternary amine from 100% to 66.8%, as well as the appearance of a nitro group which, as in the case of the Dowex PSR2 ion exchanger, indicates the participation of nitrogen in the Ni(II) ions sorption. The atomic percent is used to express the test results. In addition, the spectra of PSR2 and PSR3 before and after the sorption process in a narrow range of binding energies for C, O, N and Cl are presented in Supplementary Materials in Figures S1–S4.

3.3.3. Elemental Analysis of CHN

Due to the fact that the hydrogen content is not determined by the XPS electron spectroscopy, the CHN analysis was performed for the pure ion exchangers. The test results are presented in Table 11.
To estimate the number of functional groups per one cross-linked polystyrene DVB, the theoretical calculations were made of the % content of elements included in the ion exchanger (carbon, nitrogen, and hydrogen). The calculation results are given in the mass %. Calculations were made for the following matrix:functional group configurations: 1:1, 2:1, and 1:2, and they are presented in Table 12.
The ratio of nitrogen present in the functional groups to carbon and hydrogen indicates the internal structure of the ion exchanger, in which one functional group occurs per two units of DVB cross-linked polystyrene.

3.4. Sorption Mechanism

The physicochemical properties of the Dowex PSR2 and Dowex PSR3 ion exchangers were examined before and after the sorption process to indicate the possible mechanism of Ni(II) ions sorption. The comparison of the FTIR spectra of the ion exchangers before and after sorption process revealed differences in the positions and intensities of spectral bands, which corresponded to the stretching vibrations of the aliphatic amino groups present in the functional groups of the ion exchanger. This shows how the boundaries of these groups change as a result of the Ni(II) ions sorption process.
The XPS spectroscopic studies on nitrogen after the sorption of Ni(II) ions revealed changes within the functional groups as a result of the sorption process. These changes imply that nitrogen from the quaternary amine is involved in the Ni(II) ions sorption process. Both ion exchangers showed a decrease in the amount of nitrogen in the form of the quaternary amine and an increase in the amount of NCl- bonds. The XPS studies showed a greater number of NCl-groups after the sorption process in the Dowex PSR3 ion exchanger compared with the Dowex PSR2 ion one. These studies corroborate the research conducted on the kinetics of the sorption process. In the case of the Dowex PSR3 ion exchanger, greater equilibrium capacities were obtained compared with the Dowex PSR2 ion exchanger. The results of XRF tests also confirm the greater efficiency of the Ni(II) ions sorption process on the Dowex PSR3 ion exchanger. The test showed a nickel content of 4.0 mg/L in the Dowex PSR2 ion exchanger and 6.4 mg/L in the Dowex PSR3 one.
Based on the parameters determined from the Temkin model and the thermodynamic parameters, it was found that the process is exothermic. In addition, the calculated activation energy from the Dubinin–Raduszkiewicz model confirms that ion exchange is the main mechanism of sorption.
As follows from the results of XPS tests for nitrogen, it can be assumed that, in the case of the Dowex PSR2 ion exchanger, one in four functional groups participates in the Ni(II) ions sorption process, and in the case of the Dowex PSR3 ion exchanger, one in three functional groups participates in the Ni(II) ions sorption process.
The sorption studies, combined with the ion exchangers instrumental studies, allow us to propose the following mechanism for the Ni(II) ions sorption reaction on the Dowex PSR2 and Dowex PSR3 ion exchangers (Figure 13).
The figure shows that the sorption process takes place through the exchange of chloride ions present in the quaternary amine for nickel ions in stoichiometric terms—two NCl groups for one nickel ion.

4. Conclusions

The focus of this paper was an evaluation of the suitability of the strongly basic anion exchangers, Dowex PSR2 and Dowex PSR3, as potential sorbents in the removal of nickel ions from aqueous solutions. It was reported that pH 6 and 293 K are the ideal conditions for the sorption process. The phase contact time of 240 min is appropriate for establishing equilibrium, and the amount of adsorbed Ni(II) ions after this time is 10.30 mg/g for Dowex PSR2 and 19.10 mg/g for Dowex PSR3 (for the initial solution concentration of 100 mg/L). The PSO kinetic model showed the best fit of the experimental data to the theoretical ones and coefficients of determination close to unity. Based on the identified isotherms, it can be stated that in the case of the ion–exchangers–nickel ions solution systems, the best fit was observed on the Sips adsorption isotherm, where the R2 equals 0.982 for PSR2 and for PSR3, and the Langmuir isotherm, where R2 is equal 0.995. On the other hand, the parameters calculated from the Temkin isotherm model prove the exothermic nature of the process (which is confirmed by the enthalpy value) and indicate a share of electrostatic interactions in the process of removing nickel ions. The FTIR spectroscopic studies of the Dowex PSR2 and Dowex PSR3 ion exchangers showed the presence of bands characteristic of the ion exchanger matrix and tri-n-butylamine functional groups. The XPS spectroscopic studies allowed to determine the elemental composition of the tested ion exchangers and confirmed the presence of bonds in the functional groups and the ion exchanger matrix. The elemental composition tests carried out by means of the CHN elemental analysis method suggest the structure of the ion exchangers, in which there is one tri-n-butylammonium functional group for two polymers constituting the ion exchange matrix. The differences in the position and intensity of the spectral bands associated with the stretching vibrations of the aliphatic amino groups present in the functional groups of the ion exchanger indicate that chloride ions are exchanged for nickel ions within these groups. The dominance of ion exchange in the sorption process was confirmed by the activation energy calculated from the Dubinin–Raduszkiewicz model.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16020644/s1, Table S1: Parameters of adsorption kinetics of nickel ions on Dowex PSR2 and PSR3 (C0 10–50 mg/L). Figure S1: XPS spectra of carbon, oxygen, nitrogen, and chlorine made for the Dowex PSR 2 ion exchanger before the sorption process. Figure S2: XPS spectra of carbon, oxygen, nitrogen, and chlorine made for the Dowex PSR 2 ion exchanger after the sorption process. Figure S3: XPS spectra of carbon, oxygen, nitrogen, and chlorine made for the Dowex PSR 3 ion exchanger before the sorption process. Figure S4: XPS spectra of carbon, oxygen, nitrogen, and chlorine made for the Dowex PSR 3 ion exchanger after the sorption process.

Author Contributions

Conceptualization, J.B. and W.S.-C.; methodology, J.B. and W.S.-C.; validation W.S.-C.; formal analysis, J.B. and W.S.-C.; investigation, J.B., W.S.-C., M.G. and P.M.; resources, J.B. and W.S.-C.; writing—original draft preparation, J.B. and W.S.-C.; writing—review and editing, J.B., W.S.-C. and D.K.; visualization, J.B., W.S.-C. and M.G.; supervision, D.K.; project administration, J.B., W.S.-C. and D.K.; provided ion exchangers for research, D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained in the paper and Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anthony, F.; Godfrey, K. Simultaneous Adsorption of Ni(II) and Mn(II) Ions from Aqueous Solution unto a Nigerian Kaolinite Clay. J. Mater. Res. Technol. 2014, 3, 129–141. [Google Scholar] [CrossRef] [Green Version]
  2. Flieger, J.; Kawka, J.; Płazinski, W.; Panek, R.; Madej, J. Sorption of Heavy Metal Ions of Chromium, Manganese, Selenium, Nickel, Cobalt, Iron from Aqueous Acidic Solutions in Batch and Dynamic Conditions on Natural and Synthetic Aluminosilicate Sorbents. Materials 2020, 13, 5271. [Google Scholar] [CrossRef] [PubMed]
  3. Šuránek, M.; Melichová, Z.; Kureková, V.; Kljajevic, L.; Nenadovic, S. Removal of Nickel from Aqueous Solutions by Natural Bentonites from Slovakia. Materials 2021, 14, 282. [Google Scholar] [CrossRef] [PubMed]
  4. Charazinska, S.; Burszta-Adamiak, E.; Lochynski, P. Recent Trends in Ni(II) Sorption from Aqueous Solutions Using Natural Materials. Rev. Environ. Sci. Biotechnol. 2022, 5, 105–138. [Google Scholar] [CrossRef]
  5. Ebisike, K.; Okoronkwo, A.E.; Alaneme, K.K. Nickel Sorption onto Chitosan-Silica Hybrid Aerogel from Aqueous Solution. Walailak J. Sci. Technol. 2021, 18, 9454. [Google Scholar] [CrossRef]
  6. Mende, M.; Schwarz, D.; Steinbach, C.; Boldt, R.; Schwarz, S. The Influence of Salt Anions on Heavy Metal Ion Adsorption on the Example of Nickel. Materials 2018, 11, 373. [Google Scholar] [CrossRef] [Green Version]
  7. Buxton, S.; Garman, E.; Heim, K.E.; Lyons-Darden, T.; Schlekat, C.E.; Taylor, M.D.; Oller, A.R. Concise Review of Nickel Human Health Toxicology and Ecotoxicology. Inorganics 2019, 7, 89. [Google Scholar] [CrossRef] [Green Version]
  8. Henckens, M.L.C.M.; Worrell, E. Reviewing the Availability of Copper and Nickel for Future Generations. The Balance between Production Growth, Sustainability and Recycling Rates. J. Clean. Prod. 2020, 264, 121460. [Google Scholar] [CrossRef]
  9. Farooq, A.; Ahmad, S.; Hamad, K.; Deen, K.M. Effect of Ni Concentration on the Surface Morphology and Corrosion Behavior of Zn-Ni Alloy Coatings. Metals 2022, 12, 96. [Google Scholar] [CrossRef]
  10. Aramini, M.; Magnani, G.; Pontiroli, D.; Milanese, C.; Girella, A.; Bertoni, G.; Gaboardi, M.; Zacchini, S.; Marini, A.; Riccò, M. Nickel Addition to Optimize the Hydrogen Storage Performance of Lithium Intercalated Fullerides. Mater. Res. Bull. 2020, 126, 110848. [Google Scholar] [CrossRef]
  11. Blumbergs, E.; Serga, V.; Platacis, E.; Maiorov, M.; Shishkin, A. Cadmium Recovery from Spent Ni-Cd Batteries: A Brief Review. Metals 2021, 11, 1714. [Google Scholar] [CrossRef]
  12. Genchi, G.; Carocci, A.; Lauria, G.; Sinicropi, M.S.; Catalano, A. Nickel: Human Health and Environmental Toxicology. Int. J. Environ. Res. Public Health 2020, 17, 679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Saradesh, K.M.; Vinodkumar, G.S. Metallurgical Processes for Hardening of 22Karat Gold for Light Weight and High Strength Jewelry Manufacturing. J. Mater. Res. Technol. 2020, 9, 2009–2020. [Google Scholar] [CrossRef]
  14. Veneu, D.M.; Yokoyama, L.; Cunha, O.G.C.; Schneider, C.L.; Monte, M.B.D.M. Nickel Sorption Using Bioclastic Granules as a Sorbent Material: Equilibrium, Kinetic and Characterization Studies. J. Mater. Res. Technol. 2019, 8, 840–852. [Google Scholar] [CrossRef]
  15. Abou-Mesalam, M.M. Sorption Kinetics of Copper, Zinc, Cadmium and Nickel Ions on Synthesized Silico-Antimonate Ion Exchanger. Colloids Surf. A Physicochem. Eng. Asp. 2003, 225, 85–94. [Google Scholar] [CrossRef]
  16. Jumadilov, T.; Yskak, L.; Imangazy, A.; Suberlyak, O. Ion Exchange Dynamics in Cerium Nitrate Solution Regulated by Remotely Activated Industrial Ion Exchangers. Materials 2021, 14, 3491. [Google Scholar] [CrossRef]
  17. Mendes, F.D.; Martins, A.H. Selective Sorption of Nickel and Cobalt from Sulphate Solutions Using Chelating Resins. Int. J. Miner. Process. 2004, 74, 359–371. [Google Scholar] [CrossRef]
  18. Yang, B.; Li, C.; Wang, J.; Wei, W.; Hao, B.; Yang, L.; Tang, J.; Zhang, C. Removal of Nickel Ions from Automobile Industry Wastewater Using Ion Exchange Resin: Characterization and Parameter Optimization. IOP Conf. Ser. Earth Environ. Sci. 2020, 467, 012182. [Google Scholar] [CrossRef]
  19. Kołodyńska, D. Cu(II), Zn(II), Ni(II), and Cd(II) Complexes with HEDP Removal from Industrial Effluents on Different Ion Exchangers. Ind. Eng. Chem. Res. 2010, 49, 2388–2400. [Google Scholar] [CrossRef]
  20. Stefan, D.S.; Meghea, I. Mechanism of Simultaneous Removal of Ca2+, Ni2+, Pb2+ and Al3+ Ions from Aqueous Solutions Using Purolite® S930 Ion Exchange Resin. Comptes Rendus Chim. 2014, 17, 496–502. [Google Scholar] [CrossRef]
  21. Dizge, N.; Keskinler, B.; Barlas, H. Sorption of Ni(II) Ions from Aqueous Solution by Lewatit Cation-Exchange Resin. J. Hazard. Mater. 2009, 167, 915–926. [Google Scholar] [CrossRef] [PubMed]
  22. Bąk, J.; Thomas, P.; Kołodyńska, D. Chitosan-Modified Biochars to Advance Research on Heavy Metal Ion Removal: Roles, Mechanism and Perspectives. Materials 2022, 15, 6108. [Google Scholar] [CrossRef] [PubMed]
  23. Das, K.K.; Reddy, R.C.; Bagoji, I.B.; Das, S.; Bagali, S.; Mullur, L.; Khodnapur, J.P.; Biradar, M.S. Primary Concept of Nickel Toxicity—An Overview. J. Basic Clin. Physiol. Pharmacol. 2019, 30, 141–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wołowicz, A. Zinc(II) Removal from Model Chloride and Chloride-Nitrate(V) Solutions Using Various Sorbents. Physicochem. Probl. Miner. Process. 2019, 55, 1517–1534. [Google Scholar] [CrossRef]
  25. Wołowicz, A.; Hubicki, Z. Sorption Behavior of Dowex PSR-2 and Dowex PSR-3 Resins of Different Structures for Metal(II) Removal. Solvent Extr. Ion Exch. 2016, 34, 375–397. [Google Scholar] [CrossRef]
  26. Kołodyńska, D. Polyacrylate Anion Exchangers in Sorption of Heavy Metal Ions with the Biodegradable Complexing Agent. Chem. Eng. J. 2009, 150, 280–288. [Google Scholar] [CrossRef]
  27. Wołowicz, A.; Hubicki, Z. Polyacrylate Ion Exchangers in Sorption of Noble and Base Metal Ions from Single and Tertiary Component Solutions. Solvent Extr. Ion Exch. 2014, 32, 189–205. [Google Scholar] [CrossRef]
  28. Wołowicz, A.; Wawrzkiewicz, M. Screening of Ion Exchange Resins for Hazardous Ni(II) Removal from Aqueous Solutions: Kinetic and Equilibrium Batch Adsorption Method. Processes 2021, 9, 285. [Google Scholar] [CrossRef]
  29. Wołowicz, A.; Hubicki, Z. The Use of the Chelating Resin of a New Generation Lewatit MonoPlus TP-220 with the Bis-Picolylamine Functional Groups in the Removal of Selected Metal Ions from Acidic Solutions. Chem. Eng. J. 2012, 197, 493–508. [Google Scholar] [CrossRef]
  30. Wołowicz, A.; Hubicki, Z. Carbon-Based Adsorber Resin Lewatit AF 5 Applicability in Metal Ion Recovery. Microporous Mesoporous Mater. 2016, 224, 400–414. [Google Scholar] [CrossRef]
  31. Fila, D.; Hubicki, Z.; Kołodyńska, D. Recovery of Metals from Waste Nickel-Metal Hydride Batteries Using Multifunctional Diphonix Resin. Adsorption 2019, 25, 367–382. [Google Scholar] [CrossRef] [Green Version]
  32. Keränen, A.; Leiviskä, T.; Salakka, A.; Tanskanen, J. Removal of Nickel and Vanadium from Ammoniacal Industrial Wastewater by Ion Exchange and Adsorption on Activated Carbon. Desalin. Water Treat. 2015, 53, 2645–2654. [Google Scholar] [CrossRef]
  33. Kiefer, R.; Höll, W.H. Sorption of Heavy Metals onto Selective Ion-Exchange Resins with Aminophosphonate Functional Groups. Ind. Eng. Chem. Res. 2001, 40, 4570–4576. [Google Scholar] [CrossRef]
  34. Guo, B.; Wang, Y.; Shen, X.; Qiao, X.; Jia, L.; Xiang, J.; Jin, Y. Study on CO2 Capture Characteristics and Kinetics of Modified Potassium-Based Adsorbents. Materials 2020, 13, 877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Karunanithi, R.; Ok, Y.S.; Dharmarajan, R.; Ahmad, M.; Seshadri, B.; Bolan, N.; Naidu, R. Sorption, Kinetics and Thermodynamics of Phosphate Sorption onto Soybean Stover Derived Biochar. Environ. Technol. Innov. 2017, 8, 113–125. [Google Scholar] [CrossRef]
  36. Cheung, W.H.; Szeto, Y.S.; McKay, G. Intraparticle Diffusion Processes during Acid Dye Adsorption onto Chitosan. Bioresour. Technol. 2007, 98, 2897–2904. [Google Scholar] [CrossRef]
  37. Aniagor, C.O.; Afifi, M.A.; Hashem, A. Modelling of Basic Blue-9 Dye Sorption onto Hydrolyzed Polyacrylonitrile Grafted Starch Composite. Carbohydr. Polym. Technol. Appl. 2021, 2, 100141. [Google Scholar] [CrossRef]
  38. Osemeahon, S.A.; Dimas, B.J. Removal of Crude Oil from Aqueous Medium by Sorption on Sterculis Setigera. Asian J. Appl. Chem. Res. 2020, 10, 1–12. [Google Scholar] [CrossRef]
  39. Al-Zawahreh, K.; Al-Degs, Y.; Barral, M.T.; Paradelo, R. Optimization of Direct Blue 71 Sorption by Organic Rich-Compost Following Multilevel Multifactor Experimental Design. Arab. J. Chem. 2022, 15, 103468. [Google Scholar] [CrossRef]
  40. Sakr, A.K.; Cheira, M.F.; Hassanin, M.A.; Mira, H.I.; Mohamed, S.A.; Khandaker, M.U.; Osman, H.; Eed, E.M.; Sayyed, M.I.; Hanfi, M.Y. Adsorption of Yttrium Ions on 3-Amino-5-Hydroxypyrazole Impregnated Bleaching Clay, a Novel Sorbent Material. Appl. Sci. 2021, 11, 10320. [Google Scholar] [CrossRef]
  41. Menkiti, M.C.; Aniagor, C.O.; Agu, C.M.; Ugonabo, V.I. Effective Adsorption of Crystal Violet Dye from an Aqueous Solution Using Lignin-Rich Isolate from Elephant Grass. Water Conserv. Sci. Eng. 2018, 3, 33–46. [Google Scholar] [CrossRef]
  42. Kamaraj, R.; Pandiarajan, A.; Jayakiruba, S.; Naushad, M.; Vasudevan, S. Kinetics, Thermodynamics and Isotherm Modeling for Removal of Nitrate from Liquids by Facile One-Pot Electrosynthesized Nano Zinc Hydroxide. J. Mol. Liq. 2016, 215, 204–211. [Google Scholar] [CrossRef]
  43. Tahir, S.S.; Rauf, N. Removal of a Cationic Dye from Aqueous Solutions by Adsorption onto Bentonite Clay. Chemosphere 2006, 63, 1842–1848. [Google Scholar] [CrossRef] [PubMed]
  44. Salvestrini, S.; Fenti, A.; Chianese, S.; Iovino, P.; Musmarra, D. Diclofenac Sorption from Synthetic Water: Kinetic and Thermodynamic Analysis. J. Environ. Chem. Eng. 2020, 8, 104105. [Google Scholar] [CrossRef]
  45. Song, Y.; Wang, K.; Zhao, F.; Du, Z.; Zhong, B.; An, G. Preparation of Powdered Activated Carbon Composite Material and Its Adsorption Performance and Mechanisms for Removing RhB. Water 2022, 14, 3048. [Google Scholar] [CrossRef]
  46. Shi, J.; Yi, S.; He, H.; Long, C.; Li, A. Preparation of Nanoscale Zero-Valent Iron Supported on Chelating Resin with Nitrogen Donor Atoms for Simultaneous Reduction of Pb2+ and NO3-. Chem. Eng. J. 2013, 230, 166–171. [Google Scholar] [CrossRef]
  47. Silverstein, F.X.M.; Webster, D.J.K. Spektroskopowe Metody Identyfikacji Związków Organicznych; Wydawnictwo Naukowe PWN: Warszawa, Poland, 2007. [Google Scholar]
  48. Zieliński, W. Metody Spektroskopowe i Ich Zastosowanie Do Identyfikacji Związków Organicznych; WNT: Warszawa, Poland, 1996. [Google Scholar]
  49. Gabrienko, A.A.; Ewing, A.V.; Chibiryaev, A.M.; Agafontsev, A.M.; Dubkov, K.A.; Kazarian, S.G. New Insights into the Mechanism of Interaction between CO2 and Polymers from Thermodynamic Parameters Obtained by in Situ ATR-FTIR Spectroscopy. Phys. Chem. Chem. Phys. 2016, 18, 6465–6475. [Google Scholar] [CrossRef]
  50. Zong, L.; Liu, F.; Chen, D.; Zhang, X.; Ling, C.; Li, A. A Novel Pyridine Based Polymer for Highly Efficient Separation of Nickel from High-Acidity and High-Concentration Cobalt Solutions. Chem. Eng. J. 2018, 334, 995–1005. [Google Scholar] [CrossRef]
  51. Zagorodni, A.A.; Kotova, D.L.; Selemenev, V.F. Infrared Spectroscopy of Ion Exchange Resins: Chemical Deterioration of the Resins. React. Funct. Polym. 2002, 53, 157–171. [Google Scholar] [CrossRef]
  52. Payne, B.P.; Biesinger, M.C.; McIntyre, N.S. X-Ray Photoelectron Spectroscopy Studies of Reactions on Chromium Metal and Chromium Oxide Surfaces. J. Electron Spectros. Relat. Phenom. 2011, 184, 29–37. [Google Scholar] [CrossRef]
  53. Gobbo, P.; Novoa, S.; Biesinger, M.C.; Workentin, M.S. Interfacial Strain-Promoted Alkyne-Azide Cycloaddition (I-SPAAC) for the Synthesis of Nanomaterial Hybrids. Chem. Commun. 2013, 49, 3982–3984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Raicopol, M.; Andronescu, C.; Atasiei, R.; Hanganu, A.; Pilan, L. Post-Polymerization Electrochemical Functionalization of a Conducting Polymer: Diazonium Salt Electroreduction at Polypyrrole Electrodes. J. Electrochem. Soc. 2014, 161, G103–G113. [Google Scholar] [CrossRef]
  55. Wanger, W.C.; Riggs, L.; Davis, J.; Moulder, G.M. Handbook of X-Ray Photoelectron Spectroscopy; Physical Electronics Division: Eden Prairie, MN, USA, 1979. [Google Scholar]
Figure 1. Nickel properties and application.
Figure 1. Nickel properties and application.
Materials 16 00644 g001
Figure 2. General scheme of ion exchange process and research objectives.
Figure 2. General scheme of ion exchange process and research objectives.
Materials 16 00644 g002
Figure 3. Speciation diagram of nickel.
Figure 3. Speciation diagram of nickel.
Materials 16 00644 g003
Figure 4. Impact of solution pH on the nickel ions sorption on (a) Dowex PSR2 and (b) Dowex PSR3.
Figure 4. Impact of solution pH on the nickel ions sorption on (a) Dowex PSR2 and (b) Dowex PSR3.
Materials 16 00644 g004
Figure 5. Impact of the phase contact time and the initial concentration of the solution on the nickel ions sorption on (a) Dowex PSR2 and (b) Dowex PSR3.
Figure 5. Impact of the phase contact time and the initial concentration of the solution on the nickel ions sorption on (a) Dowex PSR2 and (b) Dowex PSR3.
Materials 16 00644 g005
Figure 6. Impact of temperature on the nickel ions sorption on (a) Dowex PSR2 and (b) Dowex PSR3.
Figure 6. Impact of temperature on the nickel ions sorption on (a) Dowex PSR2 and (b) Dowex PSR3.
Materials 16 00644 g006
Figure 7. FTIR-ATR spectrum of Dowex PSR2 ion exchanger sample before the sorption process.
Figure 7. FTIR-ATR spectrum of Dowex PSR2 ion exchanger sample before the sorption process.
Materials 16 00644 g007
Figure 8. Summary of the FTIR-ATR spectra of the Dowex PSR2 ion exchanger samples before and after the Ni(II) ions sorption process at 10, 100, 200, and 500 mg/L.
Figure 8. Summary of the FTIR-ATR spectra of the Dowex PSR2 ion exchanger samples before and after the Ni(II) ions sorption process at 10, 100, 200, and 500 mg/L.
Materials 16 00644 g008
Figure 9. FTIR-ATR spectrum of Dowex PSR3 ion exchanger sample before the sorption process.
Figure 9. FTIR-ATR spectrum of Dowex PSR3 ion exchanger sample before the sorption process.
Materials 16 00644 g009
Figure 10. Summary of the FTIR-ATR spectra of the Dowex PSR3 ion exchanger samples before and after the Ni(II) ions sorption process at 10, 100, 200, and 500 mg/L.
Figure 10. Summary of the FTIR-ATR spectra of the Dowex PSR3 ion exchanger samples before and after the Ni(II) ions sorption process at 10, 100, 200, and 500 mg/L.
Materials 16 00644 g010
Figure 11. XRS spectra in a wide range of ion exchanger binding energies: (a) Dowex PSR 2 and (b) Dowex PSR3 before the process of Ni(II) ions sorption.
Figure 11. XRS spectra in a wide range of ion exchanger binding energies: (a) Dowex PSR 2 and (b) Dowex PSR3 before the process of Ni(II) ions sorption.
Materials 16 00644 g011
Figure 12. XRS spectra in a wide range of ion exchanger binding energies: (a) Dowex PSR 2 and (b) Dowex PSR3 after the Ni(II) ions sorption process.
Figure 12. XRS spectra in a wide range of ion exchanger binding energies: (a) Dowex PSR 2 and (b) Dowex PSR3 after the Ni(II) ions sorption process.
Materials 16 00644 g012
Figure 13. Proposed mechanism of the Ni(II) ions sorption process on the ion exchangers: Dowex PSR2 and PSR3.
Figure 13. Proposed mechanism of the Ni(II) ions sorption process on the ion exchangers: Dowex PSR2 and PSR3.
Materials 16 00644 g013
Table 1. Comparison of physicochemical properties of the ion exchangers.
Table 1. Comparison of physicochemical properties of the ion exchangers.
FeaturesIon Exchangers
Dowex PSR2Dowex PSR3
Resin typestrongly basic anion exchangersstrongly basic anion exchangers
Physical formbeige beadsbeige beads
Skeletonmicroporous, cross-linked polystyrene DVBmacroporous, cross-linked polystyrene DVB
Functional groups
Matrix
tri-n-butyl ammonium
Materials 16 00644 i001
tri-n-butyl ammonium
Materials 16 00644 i002
Ion exchange capacity (val/L)0.650.60
Water retention (%)40–4850–65
Mean bead size (mm)0.3–1.20.3–1.2
Working temperature (K)<373<373
ProducerDOW Chemical CompanyDOW Chemical Company
Table 2. Kinetic and isotherm models.
Table 2. Kinetic and isotherm models.
ModelEquationParameters
Kinetic
PFO log ( q 1 q t ) = log ( q 1 ) k 1 t 2.303 k1—the rate constant of PFO equation (1/min)
PSO t q t = 1 k 2 q 2 2 + t q 2 k2—the rate constant of PSO equation (g/mg∙min)
EKE q t = 1 β ln ( α β ) + 1 β ln ( t ) α—the initial adsorption rate (mg/g∙min)
β—the constant of desorption connected with the activaton energy chemisorption and the range of surface coverage (g/mg)
IPD q t = k i t 1 2 + C ki—the rate constant of intraparticle diffusion equation (mg/g/min0.5)
C—the intercept which reflects the boundary layer impact
Isotherm
Langmuir q e = q 0 K L C e 1 + K L C e q0—the Langmuir monolayer sorption capacity (mg/g)
KL—the characteristics of Langmuir equation (L/mg)
Freundlich q e = K F C e 1 n KF—the adsorption capacity of the Freundlich equation (mg/g)
n—the Freundlich constant associated with the surface heterogenity
Temkin q e = R T b T ln ( A C e ) bT—the Temkin constant related to the sorption heat (kJ/mol)
A—the equilibrium Temkin binding constant (L/g)
Dubinin-Raduszkiewicz ln q e = ln q m β ε 2
E a = 1 2 β
qm—the Dubinin–Raduszkiewicz constant
associated with the adsorption capacity (mg/g)
β—the Dubinin–Raduszkiewicz constant related to the adsorption
mean free energy (mol2/kJ2)
Ea—the activation energy (kJ/mol)
Sips q e = q m ( K C e ) n ( 1 + K C e ) n qm, K—model’s constants (mg/g) and (L/mg)
n—the heterogeneity index whose magnitude increases with heterogeneity
Table 3. Equilibrium capacities of Ni(II) sorption on different ion exchangers as follows from the literature reports.
Table 3. Equilibrium capacities of Ni(II) sorption on different ion exchangers as follows from the literature reports.
Ion ExchangersSorption ConditionsEquilibrium Capacity (mg/g)References
Amberlite IRA458C0 58.7 mg/L, t 240 min, T 293 K, in aqueous solution (Ni(II)-IDS 1-1)5.89[26]
Amberlite IRA9585.84
Amberlite IRA675.73
Purolite S984C0 100 mg/L, t 240 min, T 298 K,
in chloride solution (0.1 mol/L HCl)
4.95[27]
Purolite A8304.60[28]
Lewatit MonoPlus SR74.56
Purolite A400TL4.72
Dowex PSR23.70[25]
Dowex PSR34.73
Lewatit MonoPlus TP2206.24[29]
Lewatit AF54.89[30]
DiphonixC0 100 mg/L, t 120 min, T 293 K, in nitrate solution (0.2 mol/L HNO3)5.03[31]
AmberjetTM 1200HC0 2.44 mg/L, t 24 h, in real ammoniacal industrial wastewater4.36[32]
Purolite S940C0 2935 mg/L, t 7 days, 298 K, in chloride solution4.22[33]
Purolite S9503.31
Dowex PSR2C0 100 mg/L, t 240 min, T 293 K, in aqueous solution10.30this research
Table 4. Parameters of adsorption kinetics of nickel ions on Dowex PSR2 and PSR3 (C0 100 mg/L).
Table 4. Parameters of adsorption kinetics of nickel ions on Dowex PSR2 and PSR3 (C0 100 mg/L).
ModelParametersUnitsIon Exchangers
PSR2PSR3
PFOqemg/g10.3019.10
k11/min0.0280.024
q1mg/g7.574.97
R2-0.9760.931
PSOk2(g/mg∙min)0.0080.023
q2(mg/g)10.7919.20
R2-0.9981.000
EKEα(mg/g∙min)3.22711,471.622
β(g/mg)0.5620.735
R2-0.9600.950
IPDfirst stepki1mg/g/min0.51.5341.988
C1-0.1689.916
R21-0.7490.873
second stepki2mg/g/min0.50.9660.410
C2-1.62714.703
R22-0.8640.821
third stepki3mg/g/min0.50.0600.061
C3-9.37418.152
R23-1.0001.000
Table 5. Parameters of the isotherms for the nickel ions sorption on Dowex PSR2 and PSR3.
Table 5. Parameters of the isotherms for the nickel ions sorption on Dowex PSR2 and PSR3.
ModelParametersUnitsIon Exchangers
PSR2PSR3
Langmuirqmmg/g19.1631.72
KLL/mg0.0120.021
R2-0.9720.995
FreundlichKFmg/g1.152.00
n-2.2572.172
R2-0.9640.937
TemkinAL/g0.2570.376
bTkJ/mol0.7620.442
R2-0.9150.968
Dubinin-Raduszkiewiczqmmg/g0.00050.0011
βmol2/kJ20.00490.0053
EakJ/mol10.0749.753
R2-0.9450.969
qmmg/g27.7425.53
SipsKL/mg0.0020.0002
n 0.6340.469
R2 0.9820.989
Table 6. Thermodynamic parameters of the nickel ions sorption on Dowex PSR2 and PSR3.
Table 6. Thermodynamic parameters of the nickel ions sorption on Dowex PSR2 and PSR3.
ParametersTemperature (K)Ion Exchangers
PSR2PSR3
Kd29337.758.7
31334.352.5
32331.147.1
33328.943.7
ΔH° (kJ/mol)-−5.37−6.01
ΔS° (kJ/mol)-11.9613.47
ΔG° (kJ/mol)293−8.87−9.96
313−9.11−10.23
323−9.23−10.36
333−9.35−10.50
Table 7. XPS results obtained in a wide range of binding energy for the Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process.
Table 7. XPS results obtained in a wide range of binding energy for the Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process.
Ion ExchangerNamePosition (eV)Raw Area% Atom Concentration
Dowex PSR2C 1s284.723,366.683.3
N 1s401.7618.92.2
O 1s532.23235.611.5
Al 2p74.7139.30.5
Si 2p101.7134.50.5
S 2p167.7106.80.4
Cl 2p197.0459.81.6
Dowex PSR3C 1s284.720,357.887.7
N 1s401.7649.01.6
O 1s531.55529.78.1
Cl 2p197.01420.32.7
Table 8. XPS results obtained in a narrow range of binding energy for Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process [53,54,55].
Table 8. XPS results obtained in a narrow range of binding energy for Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process [53,54,55].
ElementPosition (eV)Raw Area% Atom ConcentrationPhase
Dowex PSR2
C 1s284.73338.974.7C=C, C-H, C-C
285.91086.324.3-CH-, CN
288.645.31C-H–Ar
O 1s531.9404.952.1C-O
532.9372.347.9nitrates
Cl 2p196.868.350.7Cl-
198.466.549.3
N 1s399.9033.131.5quaternary amine
402.1365.061.9protonated amine
406.147.06.6nitro group
Dowex PSR3
C 1s284.77633.6686.5C=C, C-H, C-C
286.21119.6612.7-CH-, CN
288.774.73480.8C-H–Ar
O 1s532.22210.184.4C-O
533.8408.315.6nitrates
Cl 2p196.6264.874.4Cl-
197.9132.4--
198.691.225.6chlorates
199.745.6--
N 1s401.9374.6100quaternary amine
Table 9. XPS results obtained in a wide range of binding energy for Dowex PSR2 and Dowex PSR3 ion exchangers after the Ni(II) ions sorption process.
Table 9. XPS results obtained in a wide range of binding energy for Dowex PSR2 and Dowex PSR3 ion exchangers after the Ni(II) ions sorption process.
Ion ExchangerNamePosition (eV)Raw Area% Atom Concentration
Dowex PSR2C 1s284.723,842.480.1
N 1s401.7889.73.0
O 1s532.24281.014.4
Al 2p73.2277.10.9
Si 2p101.7363.11.2
S 2p167.748.80.2
Cl 2p197.081.40.3
Dowex PSR3C 1s284.718,315.382.3
N 1s401.71249.13.1
O 1s532.29236.314.2
Cl 2p197.7235.30.5
Ni 2p856.2482.5-
Table 10. XPS results obtained in a narrow range of binding energy for Dowex PSR2 and Dowex PSR3 ion exchangers after the Ni(II) ions sorption process [53,54,55].
Table 10. XPS results obtained in a narrow range of binding energy for Dowex PSR2 and Dowex PSR3 ion exchangers after the Ni(II) ions sorption process [53,54,55].
ElementPosition (eV)Raw Area% Atom ConcentrationPhase
Dowex PSR2
C 1sC 1s A284.777.5C=C, C-H, C-C
C 1s B286.021.4-CH-, CN
C 1s C288.71.1C-H–Ar
O 1sO 1s A533.95.2C-O
O 1s B532.294.8nitrates
Cl 2pCl 2p 3/211.450.7Cl-
Cl 2p 1/211.149.3
N 1sN 1s A399.821.2quaternary amine
N 1s B402.250.7protonated amine
N 1s C406.128.1nitro group
Dowex PSR3
C 1s284.75194.363.6C=C, C-H, C-C
284.91122.913.7-CH-, CN
286.11378.516.9C-H–Ar
288.7457.85.6-COOH
291.218.90.2π→π *
O 1s531.92817.663.5C-O
533.41621.236.5nitrates
Cl 2p196.8114.2100.0Cl-
199.557.1--
N 1s401.9239.766.1quaternary amine
405.9123.033.9nitro group
* transition from ground to excited state
Table 11. Percentage of elements in the Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process.
Table 11. Percentage of elements in the Dowex PSR2 and Dowex PSR3 ion exchangers before the Ni(II) ions sorption process.
Ion ExchangerElementMeasurement I [%]Measurement II
[%]
Mean (%)
Dowex PSR2C70.0870.2270.15
H8.818.798.80
N2.802.772.78
Dowex PSR3C74.8174.7374.77
H12.0712.1012.09
N2.412.412.41
Table 12. Elemental composition of the Dowex PSR2 and Dowex PSR3 ion exchangers was determined using the CHN method and the theoretical calculations prior to the Ni(II) ions sorption process.
Table 12. Elemental composition of the Dowex PSR2 and Dowex PSR3 ion exchangers was determined using the CHN method and the theoretical calculations prior to the Ni(II) ions sorption process.
Ion ExchangerElement% Content Determined by the CHN MethodTheoretical% Content (1:1)
Materials 16 00644 i003
Theoretical% Content (2:1)
Materials 16 00644 i004
Theoretical% Content (1:2)
Materials 16 00644 i005
Dowex PSR2C70.1586.1288.3483.72
H8.8010.539.5111.63
N2.783.352.154.65
Dowex PSR3C74.7786.1288.3483.72
H12.0910.539.5111.63
N2.413.352.154.65
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

Bąk, J.; Sofińska-Chmiel, W.; Gajewska, M.; Malinowska, P.; Kołodyńska, D. Determination of the Ni(II) Ions Sorption Mechanism on Dowex PSR2 and Dowex PSR3 Ion Exchangers Based on Spectroscopic Studies. Materials 2023, 16, 644. https://doi.org/10.3390/ma16020644

AMA Style

Bąk J, Sofińska-Chmiel W, Gajewska M, Malinowska P, Kołodyńska D. Determination of the Ni(II) Ions Sorption Mechanism on Dowex PSR2 and Dowex PSR3 Ion Exchangers Based on Spectroscopic Studies. Materials. 2023; 16(2):644. https://doi.org/10.3390/ma16020644

Chicago/Turabian Style

Bąk, Justyna, Weronika Sofińska-Chmiel, Maria Gajewska, Paulina Malinowska, and Dorota Kołodyńska. 2023. "Determination of the Ni(II) Ions Sorption Mechanism on Dowex PSR2 and Dowex PSR3 Ion Exchangers Based on Spectroscopic Studies" Materials 16, no. 2: 644. https://doi.org/10.3390/ma16020644

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