Skip to main content
Analytical Science Advances logoLink to Analytical Science Advances
. 2020 Jun 25;1(2):109–123. doi: 10.1002/ansa.202000030

Design, synthesis, and study of Pd(II) ion‐imprinted functionalized polymer

Mohammad Nozari 1,, Mohammed Monier 2,3, Ali H Bashal 3
PMCID: PMC10989084  PMID: 38715904

Abstract

Selective metal ions’ extraction and recovery has various applications in the analytical field. Metal ions need to be extracted, detected, and quantified. For that purpose, ion‐imprinted polymers have earned a great deal of attention during the past two decades. Pd2+ ion‐imprinted hollow silica particles including an isatin Schiff base were prepared by Schiff base condensation of (3‐aminopropyl)triethoxysilane and isatin. The prepared Schiff base ligand was coordinated to the target Pd2+ cations, the polymerizable Pd‐complex was set aside to form gel in the company of tetraethoxysilane and the target Pd2+ cations were subsequently removed from the cross‐linked silica network by means of acidified thiourea solution. All materials throughout this synthesis process were investigated utilizing mass spectrometry, elemental analysis, FTIR, and 1H‐NMR. The morphological structure of both Pd2+ ion‐imprinted and non‐ion‐imprinted silica polymer were pictured by scanning electron microscopy. Several batches were studied exploiting both Pd2+ ion‐imprinted and non‐ion‐imprinted silica polymer to test their functionality for selective extraction of Pd2+ cations in multi‐ionic solution of Ni2+, Co2+, Cu2+, Mn2+, and Pd2+.

Keywords: (3‐aminopropyl)triethoxysilane, ion‐imprinting, isatin, palladium ions, selective extraction

1. INTRODUCTION

The particular chemical, physical, and biological characteristics of precious metals such as platinum‐group metals (PGMs), which include platinum (Pt), palladium (Pd), iridium (Ir), osmium (Os), rhodium (Rh), ruthenium (Ru) as well as gold (Au) and even silver (Ag), have attracted a great deal of attention in various fields of application such as catalysis, renewable energy systems, and the electronic and pharmaceuticals industries. 1 , 2 , 3 , 4 , 5 Within the next few years, the expected high demand for these metals and the depletion of them in the earth's crust will escalate their price. Therefore, the extraction and retrieval of these metal ions from industrial wastewater gains extreme economic importance. 6

Co‐precipitation, 7 , 8 solid phase extraction, 9 solvent extraction, 10 membrane ultra‐filtration, and reverse osmosis 11 , 12 are well‐known techniques, which have been commonly utilized in separation of heavy metal ions from aquatic wastes. However, the wide utilization of these methods is greatly restricted due to the high cost, energy demand, low separation efficiency and formation of harmful and sometimes toxic by‐products, which could be released, causing soil, or water pollution. 13 , 14 Adsorption is regarded as one of the most efficient known techniques, which is able to extract heavy metal cations from aqueous solutions without the drawbacks of the previously mentioned methods. 15 In previous studies, various adsorbents were prepared and utilized for removal of different metal ions such as modified grafted cellulosic cotton fibers, 16 , 17 grafted PET fibers, 18 modified chitosan, and alginate. 19 , 20 , 21 , 22 Despite the obvious competence of these adsorbents, the absence of selectivity could be a serious limitation to their use on a large scale.

Ion‐imprinting is considered a relatively modern technique utilized for improving the adsorbent selectivity toward target specific ionic species and rare earth metals in the environment mixed with other interfering ions. 23 , 24 The common method for the preparation of ion‐imprinted polymeric materials includes complex formation between the target ionic species and an active polymerizable ligand followed by polymerization in presence of an appropriate cross‐linker agent. The metal ions are then extracted from the cross‐linked polymeric network to leave recognition sites capable of selectively interacting with the same metal ions over the other interfering ones. 25 , 26 , 27 , 28 , 29

As a result of its high rigidity, which is considered a vital point in creation of active recognition sites, the imprinted matrices derived from mesoporous silica are regarded as one of the most promising materials. Moreover, these materials provide the target species with a high accessibility to the active recognition sites, due to the relatively high pore volume and nano‐scaled pore wall, particularly with denser cross‐linked structures, which limit the molecules’ mobility. Generally, the functionalized polymeric matrices based on mesoporous silica can be obtained by co‐condensation of tetraalkylorthosilicate in the presence of other organosilica precursors, depending on the desired functional group.

In the literature, there are several material used to create cross‐linked network for ion‐imprinting, for instance, Francisco et al 30 used 4‐vinylpyridine as polymerizable agent, ethylene glycol dimethacrylate as the crosslinking agent, and benzoyl peroxide for radical generator. Kumar et al 31 and da Santos Silva et al 32 utilized 1‐vinyl imidazole as ligand and methacrylic acid as functional monomer for synthesis of ion‐imprinted polymer. Biswas et al 33 used modified curcumin as functional monomer, ethylene glycol dimethacrylate as a binding agent and 2,2‐azobisisobutyronitrile as a free radical initiator. Yasizai et al 34 reported ion‐imprinted polymer based receptors based on styrene, N‐vinylpyrrolidone, and their copolymer. Table 1 shows a comparative review of different methodologies for detection and recovery of the palladium in aqueous samples. Several factors need to be considered for designing novel IIPs such as maximum absorbance capacity (qm ), pH of application, recovery percentage, ease of synthesis, and affordability. Therefore, it needs to be highlighted that based on comparative data in Table 1, we propose an easy to synthesize, accessible alternative that offers the second highest maximum absorbance capacity for palladium 249 mg/g; one should note that the previous work 38 that has higher absorbance capacity 275 mg/g has a lower recovery percentage 65% compared to our proposed method 98%, which overall makes the proposed Pd‐IIP a method that offers better recovery and quantification results.

TABLE 1.

Different methodologies for selective palladium detection in water samples

Reference Methodology Detection pH qm * (mg/g) Recovery (%)
Fujiwara et al 2007 20 l‐Lysine modified chitosan IIP ICP‐AES a 2 109 98
Godlewska‐Żyłkiewicz et al 2010 35 4‐vinylpyridin & styrene IIP FAAS b 5‐6 13.3 92‐100
Liu et al 2012 15 chitosan & graphene oxide composite FAAS 3‐4 216 95
Jiang et al 2013 36 2‐aminobenzonitrile & 4‐vinylpyridin IIP AAS c 0.5 38.9
Cataldo et al 2014 37 modified alginate beads ICP‐OES d 3 127
Lin et al 2015 2 chitosan ion‐imprinted fiber ICP‐AES 2 174 96
Monier et al 2016 38 modified chitosan resin ICP‐AES 6 275 65
Yu et al 2018 39 2‐hydroxyethyl methacylate IIP AAS 3‐4 46
Present Work silane & isatin IIP ICP‐AES 5‐6 249 98
*

qm maximum absorbance capacity.

a

ICP‐AES (Inductively coupled plasma ‐ atomic emission spectrometer).

b

Flame atomic absorption spectrometry).

c

Atomic absorption spectrum).

d

ICP‐OES (Inductively coupled plasma ‐ optical emission spectrometry).

Herein, we report the synthesis and full characterization of Pd2+ ion‐imprinted hollow silicone particles (Pd‐Si‐IS). The proposed Schiff base precursor ligand AT‐IS was initially produced by reaction of (3‐aminopropyl)triethoxysilane (AT) with isatin (IS). Then the template Pd(II) ions were complexed to the AT‐IS ligand and the polymerizable Pd‐complex obtained was set aside to form gel in with tetraethoxysilane (TEOS). Afterward, the Pd2+ ion‐imprinted silica network was prepared through mechanically smashing the resulting solid matter followed by Pd2+ ion desorption. All materials prepared throughout this synthetic process were investigated using elemental analysis, FTIR, and scanning electron microscopy (SEM). In addition, both the ion‐imprinted and the non‐ion‐imprinted silica particles were tested for their functionality in selective extraction of Pd2+ ions under various interfering effects of accompanying ions in order to evaluate the optimal factors that affect the selective extraction.

2. MATERIALS AND METHODS

2.1. Chemicals

(3‐Aminopropyl)triethoxysilane (AT), isatin (IS), tetraethoxysilane (TEOS), and PdCl2 were purchased from Alfa Aesar Chemicals and were utilized without any additional treatment. All utilized solvents and chemicals were obtained from Sigma–Aldrich and utilized as received.

2.2. Synthesis of HATIS Schiff base ligand

Initially (35 mmol, 5 g) of isatin was dissolved in 100 mL 200 proof anhydrous ethyl alcohol to give a clear solution, to which AT (35 mmole, 7.53 g) was added. The system was then refluxed for 4 h, after which the reaction mixture was cooled and AT‐IS Schiff base precipitated by addition of ether, the ligand was filtered off and washed numerous times with 200 proof anhydrous ethyl alcohol.

2.3. Preparation of the monomer‐template Pd2+ complex

AT‐IS Schiff base (40 mmol, 14 g) was dissolved in 100 mL hot anhydrous ethyl alcohol to which (20 mmol, 3.57 g) PdCl2 was then added. The reaction container was connected to a reflux condenser and the temperature raised to 80°C for 6 h with continuous magnetic stirring. The mixture was then cooled and the monomer‐template Pd‐complex was filtered off then washed with anhydrous ethyl alcohol.

2.4. Preparation of Pd2+ ion‐imprinted silica particles (Pd‐Si‐IS)

Typically, (7.5 mmol, 6.5 g) of Pd‐complex monomer template was dissolved in 15 mL TEOS. Then the NH4OH (1 M, 3 mL) mixed with 5 mL ethanol was added dropwise over in 15 min with continuous mixing. The mixture obtained was set aside in a desiccator for 48 h at ambient temperature and then dried in oven at 40°C for 24 h. The resulting particles were collected, mechanically ground to approximately 200 μm particle size, then washed continuously with ethyl alcohol, diluted HCl, and distilled water. Afterward, the Pd2+ cations were taken out of the cross‐linked network by stirring the silica particles in 300 mL 0.5 M acidified thiourea solution until all Pd2+ ions were entirely removed. Then Pd‐Si‐IS particles were washed with distilled water until remaining acid was removed, and then dried in the oven for 12 h at 40°C. For selectivity comparison, a blank non‐ion‐imprinted silica network was prepared by following the same procedure by AT‐IS Schiff base ligand instead of the Pd‐complex. The synthetic procedures are displayed in Figure 1.

FIGURE 1.

FIGURE 1

Preparation of Pd2+ ion‐imprinted silica particles

2.5. Instrumentation

A Perkin–Elmer 240C Elemental Analytical Instrument was utilized in performing C, H, and N elemental analyses. Palladium and chloride were estimated using the digestion method. 40 Infrared spectroscopy was completed on a Perkin–Elmer spectrometer. NMR spectra were obtained on Oxford 500 MHz instrument with samples in DMSO‐d6. Morphological structures were studied by scanning electron microscope (FEI Co.). The mass spectra were obtained by Agilent‐7800 ICP‐MS. Surface area measurement was performed by BET technique using a Micromeritics ASAP 2010 instrument. The amounts of cations were measured the using Inductively Coupled Plasma‐Atomic Emission Spectrometer (ICP‐AES; ICPS‐7500, Shimadzu, Japan).

2.6. Batch adsorption studies

For preparing the batches, 50 mg of either Pd‐Si‐IS or NII‐Si‐IS particles were immersed in 50 mL Pd2+ cation solution with known concentrations. The solution pH was altered to the determined values by means of the proper buffer solutions. The containers were located on a thermostated shaker adjusted to 150 rpm at the desired temperatures for appropriate periods of time (10‐120 min). After the suspensions reached equilibrium, the containers were taken out of the shaker, the particles were filtered off and the amounts of adsorbed Pd2+ cations were determined by measuring the residual amount in the filtrate using inductively coupled plasma‐atomic emission spectrometer (ICP‐AES) and using the Equation (1).

qe=CiCeV/W (1)

where qe (mg/g) is the adsorbed Pd2+ at equilibrium; Ci (mg/L) and C e (mg/L) are the initial and final Pd2+ cation concentrations, respectively, V (L) is solution volume, and W(g) is the particles’ weight.

2.7. Initial pH effect

The Pd2+ adsorption using Pd‐Si‐IS and NII‐Si‐IS was studied in the pH ranging 1‐6 along with 50 mg of the target adsorbent particles in 50 mL of 200 ppm aqueous Pd2+ cations solution at 30°C. The batches were equilibrated on a shaker at 150 rpm for 3 h then filtered to calculate the residual metal ion amount. The removal percentage was calculated by applying Equation (2).

Percentremoval(%)=CiCeCi×100 (2)

2.8. Effect of temperature

A series of batches containing 50 mg of the target adsorbent particles immersed in 50 mL of 50 ppm aqueous Pd2+ cation solution was equilibrated in the temperature ranging from 20 to 40°C. After the reaction was complete the residual Pd2+ contents were estimated as previously described.

2.9. Kinetics

Two sets were made ready, each by soaking 500 mg of either Pd‐Si‐IS or NII‐Si‐IS in 500 mL of 200 ppm Pd2+ cation solution. The containers were equilibrated at 30°C and 150 rpm and pH 5, an aliquot of 1 mL was sampled every 10 min to evaluate the level of Pd2+ cation residue in the solution.

2.10. Isotherm experiments

Several batch adsorption tests were completed using initial adsorbate Pd2+ cations for which the concentration was changed from 50 to 500 ppm. In each instance, either Pd2+ ion‐imprinted or non‐ion‐imprinted silica particles (50 mg) were equilibrated in 50 mL aqueous Pd2+ cation solution at 150 rpm and 30°C for 3 h.

2.11. Ion selectivity

For evaluating the effect of the ion‐imprinting process on improving the selectivity of the silica particle for the target Pd2+ cations, the extraction process was completed in a solution containing 40 ppm of several metal ions of the interfering ions Co2+, Cu2+, Mn2+, and Ni2+ together with the target Pd2+ cations using either Pd‐Si‐IS or NII‐Si‐IS silica particles. After agitating at 150 rpm, pH 5, and 30°C for 3 h, each set was filtered in order to measure the residual amounts of Pd2+ cations and each co‐existing interfering metal cation. The following relationship was used to evaluate the imprinting effect on the selective extraction of Pd2+ cations. 41

D=CiCfCf×VW (3)

where D, Ci , and Cf (mg/L) are the distribution coefficient and initial and final Pd2+ concentrations, respectively.

The affinity of each of the investigated silica particle types toward Pd2+ was contrasted to interfering metal ions was using:

βPd2+/Mn+=DPd2+DMn+ (4)

While the impact of the ion‐imprinting could be estimated by using Equation (5):

βr=βimprintedβnonimprint (5)

2.12. Regeneration

A mixture solution of 1:1 aqueous 1 M thiourea and 1 M acid (HCl) was used for desorption and regeneration. The Pd2+ cation‐loaded silica particles were agitated with this regeneration solution for 2 h at 150 rpm at ambient temperature. The particles were then extracted and treated with dilute NaOH solution for neutralization and washed with distilled water before being reused in a subsequent extraction process. Desorption ratio and regeneration efficiency were calculated by:

Desorption%=AmountofPdIIdesorbedAmountofPdIIadsorbed×100 (6)
Regenerationefficiency%=PdIIadsorbedinthesecondtimePdIIadsorbedinthefirsttime×100 (7)

3. RESULTS AND DISCUSSION

3.1. Characterization

The percentage amount of C, H, and N were acquired from elemental analysis for both Pd and Cl, together with the values that were estimated by standard digestion method 40 are collected in Table 2. The mass spectra of the Schiff base ligand HATIS presented a major signal at m/z 351.21 (M+1)+ without the characteristic signals at m/z 147.13 or 222.13 due to isatin or AT, respectively. These obtained results endorse the HATIS Schiff base synthesis. The mass spectra of the Pd‐complex revealed the major signal at m/z 806.22, as a solid evidence for complex synthesis based on the proposed composition [Pd(ATIS)2] (calculated m/z 806.22).

TABLE 2.

Elemental analysis of HATIS and [Pd(ATIS)2]

Found (calculated) (%)
Compound C H N Metal
HATIS 58.5 (58.3) 7.4 (7.5) 7.9 (8.0)
[Pd(ATIS)2] 50.4 (50.7) 6.1 (6.3) 6.8 (7.0) 13.0 (13.2)

The Pd2+’s coordination to the HATIS ligand was further evidenced through comparison of the FTIR spectra of the free ligand and corresponding Pd‐complex. Table 3 summarizes the basic characteristic absorptions, which undergo the main shifts or changes upon coordination. It has been reported that isatin motifs demonstrate amido‐oximido tautomerism in solution. 42 The synthesized ligand exhibited characteristic bands at 3244, 1720, and 1640 cm−1 related to N‐H, amidic C=O, and C=N, respectively, confirming that in the solid state, the ligand exists mainly in amido‐form. After coordinating the ligand to the Pd2+ cations, the characteristic C=N absorption exhibited a noticeable shift to 1620 cm−1. On the other hand, the absence of both amidic C=O and N‐H absorptions with the simultaneous emergence of a novel C=N absorption at 1650 cm−1 is clear evidence for the amido‐oximido transformation associated with Pd2+ ion‐binding and coordination via the oxygen of the deprotonated hydroxyl group and the nitrogen of the C=N species, in a square‐planar five‐membered ring system as displayed in Figure 1.

TABLE 3.

IR absorptions (cm 1 ) of the HATIS and [Pd(ATIS)2]

Compound v(N‐H) v(C = O) v(C = N)1 v(C = N)2
HATIS 3244 1720 1640
[Pd(ATIS)2] 1620 1650

The electronic spectra of the synthesized Pd‐complex displayed two bands at 17 300 and 19 200 cm−1 assigned to the 1A1g1A2g and 1A1g1B1g transitions, respectively, in square planar geometry. 43 , 44

The 1H‐NMR spectrum of HATIS metal‐free ligand is shown in Figure 2A, similar to previous literature, 45 the diagnostic AT resonances were evidently noticed between 0.56 and 3.83 ppm. Another determining factor for Schiff base synthesis was the amide N‐H peak presence at 10.03 ppm, along with aromatic doublets at 7.73 and 7.93 ppm, and triplets at 7.34‐7.60 ppm, which are typical for the inserted isatin moieties. 46 The Pd‐complex exhibited a similar spectrum (Figure 2B). However, the absence of the amidic N‐H peak indicates the Pd2+ coordination through the previously mentioned deprotonated oximido‐form.

FIGURE 2.

FIGURE 2

1H NMR spectra of (A) HATIS, (B) [Pd(ATIS)2]

The 13C‐NMR spectra of HATIS and related Pd‐complex are shown in Figure 3; the peaks for both C=N and C–O showed noticeable shifts after bonding to the Pd2+ cations, ratifying the involvement of these functional groups in the process of complex synthesis.

FIGURE 3.

FIGURE 3

13C NMR spectra of (A) HATIS, (B) [Pd(ATIS)2]

Conformational analysis along with geometry optimization for both HATIS ligand and Pd‐complex were performed by HyperChem 8.0 software and utilizing the MM+ and semi‐empirical PM3 force‐field methods. Molecular models of the ligand in both amido and oximido tautomeric forms are presented in Figure 4. The calculated theoretical total energies were 20.768 and 27.287 kcal/mol for amido and oximido forms, respectively. These obtained results are in line with the experimental results from the FTIR spectra, which confirm the existence of the amido rather than oximido form in solid state.

FIGURE 4.

FIGURE 4

Proposed structures of (A) amido form and (B) oximido form

The proposed geometry‐optimized structures of the Pd‐complex are presented in Figure 5, the square planar geometry of the Pd‐complex provides proper organization of the EtO groups with the least steric interruption for an effective sol‐gel polymerization.

FIGURE 5.

FIGURE 5

Molecular model of Pd‐complex

The surface morphology of both ion‐imprinted and non‐ion‐imprinted silica particles was pictured by SEM (Figure 6). The ion‐imprinted Pd‐Si‐IS exhibited a rough surface compared to the smooth one observed for non‐ion‐imprinted NII‐Si‐IS. The resultant rough hollow morphology may be attributed to the Pd2+ extraction from the cross‐linked matrix of the silica particles. Moreover, the surface area obtained by BET measurements exposed a rough irregular morphology as a result of the ion‐imprinting. Pd‐Si‐IS and NII‐Si‐IS exhibited areas of 1220.25 and 895.45 m2/g, respectively.

FIGURE 6.

FIGURE 6

SEM images of (A) Pd‐Si‐IS and (B) NII‐Si‐IS

The FTIR spectra for both Pd2+ ion‐free and Pd‐loaded Pd‐Si‐IS along with the NII‐Si‐IS are shown in Figure 7; both the Pd2+ ion‐free Pd‐Si‐IS and NII‐Si‐IS particles gave approximately identical spectra with characteristic IR absorptions at 3244, 1720, and 1640 cm−1 related to N‐H, amidic C=O, and C=N, respectively. These observations endorse that the active isatin Schiff base species are still included in the cross‐linked silica network after the Pd2+ cation leaching treatment. On the other hand, the Pd2+ ion‐loaded Pd‐Si‐IS particles demonstrated spectra close to the one observed for the Pd‐complex, which indicates that bonding to Pd2+ ions induces the amido‐oximido transformation and coordination via the oxygen of the deprotonated hydroxyl group and the nitrogen of the C=N species in a square‐planar five‐membered ring system.

FIGURE 7.

FIGURE 7

FTIR spectra of (A) NII‐Si‐IS, (B) Pd‐Si‐IS before Pd(II) ion removal, and (C) Pd‐Si‐IS after Pd2+ ion removal

3.2. Adsorption studies

3.2.1. Solution pH effect

The adsorption of ionic species in aqueous solutions is typically distressed by the solution pH, which determines the charge and surface ionization. The percentage removal of Pd2+ cations by Pd‐Si‐IS and NII‐Si‐IS were determined for the range 1 ≤ pH ≤ 6 (Figure 8). At pH 1, the percentage removal was moderately elevated that can be attributed to the ionic interactions between the dominant [PdCl4] anions and the protonated functional groups on the adsorbent surface. Conversely, the adsorption reduced at pH 2. By increasing pH, both adsorbents exhibited greater removal efficacy. In low acidic media, palladium ions exist generally as divalent Pd2+ cations, conversely the chelating sites still have positive charges. This phenomenon can in turn limit the target Pd2+ metal ions’ approaching to the active sites.

FIGURE 8.

FIGURE 8

Effect of pH on the removal of Pd2+ ions by Pd‐Si‐IS and NII‐Si‐IS

3.2.2. Thermodynamics

To understand the adsorption systems’ spontaneity, it is useful to determine the thermodynamic parameters. The adsorption of Pd2+ cations by both Pd‐Si‐IS and NII‐Si‐IS was thus performed at various temperatures and the results were used to calculate the standard free energy (ΔG°), enthalpy (Δ), and entropy (ΔS°) values for adsorption, employing the Van't Hoff approach. 47

lnKd=ΔG0RT=ΔS0RΔH0RT (8)

Estimates for ΔH° and ΔS° were obtained as shown in Figure 9.

FIGURE 9.

FIGURE 9

LnKC plot for the Pd2+ cations’ uptake by Pd‐Si‐IS and NII‐Si‐IS chelating silica particles

ΔG°ads , ΔH°ads and ΔS°ads values for Pd2+ extraction for Pd‐Si‐IS and NII‐Si‐IS are displayed in Table 4. In all of the tested temperatures, the adsorption process was spontaneous. Also, Pd2+ ion adsorption is enthalpically disfavored but entropically driven, which can be associated with the release of H+ and hydration water after coordination of Pd2+ cations.

TABLE 4.

Thermodynamic parameters for the adsorption of Pd2+ cations on Pd‐Si‐IS and NII‐Si‐IS silica particles

K c −ΔGo ads (kJ/mol)
System 293 K 303 K 313 K 293 K 303 K 313 K ΔHo ads (kJ/mol) ΔSo ads (J/molK)
Pd‐Si‐IS 19999 24999 45453 24.12 25.51 27.91 32.21 191.63
NI‐Si‐IS 205.6 278.3 426.3 12.97 14.18 15.76 28.57 141.50

3.2.3. Adsorption kinetics

Figure 10 shows the contact time effect on the adsorption of Pd2+ cations onto Pd‐Si‐IS and NII‐Si‐IS. The rapid initial rate of extraction from the solution may have been because of the coordination sites’ availability on the adsorbent surface. After increased contact time, the most accessible sites become saturated, which in turn can considerably decrease the adsorption rate. Both adsorbent types reach equilibrium within 1 h. To interpret the kinetic results, both pseudo‐first‐order and second‐order models were tested. 48 , 49

FIGURE 10.

FIGURE 10

Contact time effect on removal of the Pd2+ cations by Pd‐Si‐IS and NII‐Si‐IS particles

The pseudo‐first‐order model is represented by:

lnqeqt=lnqe1k1t (9)

where qe and qt (mg/g) are the extracted Pd2+ cations by Pd‐Si‐IS or NII‐Si‐IS at equilibrium and any given time (t), respectively. k1 (min−1) is the apparent first‐order rate constant. The first‐order parameters k1 and qe1 were estimated using the slope and intercept of ln(qe‐qt) against t plot, respectively.

The pseudo‐second‐order model is represented by:

tqt=1k2qe22+tqe (10)

where k2 (g mg−1 min−1) is the pseudo‐second‐order rate constant. The first‐order model parameters k2 and qe2 were calculated using the slope and intercept of t/qt against t plot.

The estimated kinetic parameters using the mathematical models above are tabulated in Table 5. The relatively higher R 2 and lower SD as long with the estimated qe values highlights that the adsorption of Pd2+ cations onto Pd‐Si‐IS and NII‐Si‐IS exhibited the most appropriate results using the second‐order equation, which may indicate that the metal ion coordination is the main mechanism governing the overall solid‐phase extraction procedure.

TABLE 5.

Kinetic parameters for Pd2+ ions extraction by Pd‐Si‐IS and NII‐Si‐IS silica particles

Adsorbent First‐order model
k1 (min−1) qe1 (mg/g) R2
Pd‐Si‐IS 0.112 220 ± 3 0.8765
NII‐Si‐IS 0.087 140 ± 3 0.7894
Adsorbent Second‐order model
k2 (g/(mg min)) qe2 (mg/g) R2
Pd‐Si‐IS 2.8 × 10−4 200 ± 2 0.9998
NII‐Si‐IS 1.9 × 10−4 136 ± 2 0.9987

3.2.4. Adsorption isotherms

For investigating the adsorption capacity, several concentration experiments have been performed with variations of Pd‐Si‐IS and NII‐Si‐IS as a function of the Pd2+ cations’ concentration and the isotherms are illustrated in Figure 11. The capacity of Pd2+ ion‐imprinted and non‐ion‐imprinted silica particles amplified considerably by increasing the initial Pd2+ cation concentration until reaching a maximum level that is the saturation point of the active coordination sites. For more understanding and a better description of the adsorption equilibrium isotherms in aqueous solution, both Langmuir and the Freundlich models are used.

FIGURE 11.

FIGURE 11

Pd2+ cation isotherms by Pd‐Si‐IS and NII‐Si‐IS silica particles

Based on the Langmuir model (Equation 11), the Pd2+ cations are adsorbed in form of a single layer onto energetically equivalent coordination active sites, while no interaction among adsorbed Pd2+ cations. The Langmuir model is presented in Equation 11.

Ceqe=1qmKL+Ceqm (11)

where qe is the adsorption capacity (mg/g), Ce is the equilibrium concentration of Pd2+ cations (mg/L), qm is the maximum adsorption capacity for Pd2+ cations (mg/g), and KL is the Langmuir adsorption constant (L/mg). The slope and intercept of the linear plot of Ce/qe against Ce were used to calculate the parameters qm and KL .

On the other hand, for a Freundlich model, the formation of an adsorbate multilayer onto energetically heterogeneous adsorption sites is expressed by:

lnqe=lnKF+1nlnCe (12)

where KF and n are the Freundlich model parameters, which can be acquired from the slope and intercept of the linear plot of lnqe against lnCe .

The experimental isotherms of Pd2+ cation adsorption onto both Pd‐Si‐IS and NII‐Si‐IS are shown in Figure 11; the corresponding parameters calculated by fitting the results with the aforesaid two mathematical models are listed in Table 6. As can be noticed, for both of adsorbents, Langmuir models displayed the better fit with the equilibrium isotherm data with R 2 > 0.99, which indicated the adsorption of Pd2+ cations in form of single layer onto the homogeneous surface of both of the modified silica particles. The qm value corresponding to the Pd‐Si‐IS was considerably higher than the value related to NII‐Si‐IS, which can be attributed to the comparatively larger surface area by the ion‐imprinted NII‐Si‐IS as well as the formation of Pd2+ cations recognition hollows by imprinting procedure.

TABLE 6.

Pd2+ cations adsorption parameters for Pd‐Si‐IS and NII‐Si‐IS silica particles based on different equilibrium models

Langmuir isotherm constants
Adsorbent KL (L/g) qm (mg/g) R2
Pd‐Si‐IS 2.43 249.6 0.9987
NII‐Si‐IS 2.12 141.3 0.9897
Adsorbent Freundlich isotherm constants
KF n R2
Pd‐Si‐IS 125.20 6.34 0.7689
NI‐Si‐IS 95.31 7.52 0.6897

3.2.5. Adsorptive selectivity

The efficiency of selective Pd2+ cations recognition by both Pd‐Si‐IS and NII‐Si‐IS was evaluated. Calculated selectivity parameters are summarized in Table 7. The higher selectivity of Pd‐Si‐IS toward the target Pd2+ cations is obvious by the relatively higher D values compared to the other interfering ionic species. Moreover, the calculated selectivity coefficients of the target Pd2+ ions corresponding to Pd‐Si‐IS all have values larger than 1. Alternatively in case of NII‐Si‐IS, the projected selectivity coefficient amounts were approximately equal to or less than 1. These selectivity parameters’ amounts indicate the greater ability of the Pd2+ imprint sites within the network of Pd‐Si‐IS‐modified silica particles to selectively extract the target Pd2+ cations from multi‐ionic solutions including competing metal ions.

TABLE 7.

Selective adsorption of target Pd2+ cations from multi‐ionic solutions by Pd‐Si‐IS and NII‐Si‐IS silica particles

Distribution ratio (L/g) Selectivity coefficient βPd 2+ /M n+‏
Metal Pd‐Si‐IS NII‐Si‐IS Pd‐Si‐IS NII‐Si‐IS Relative selectivity coefficient βr
Pd2+ 523.67 15.76
Co2+ 12.65 12.85 41.4 1.22 33.93
Cu2+ 20.54 18.45 25.5 0.85 30.00
Mn2+ 13.89 14.67 37.7 1.07 35.23
Ni2+ 9.45 10.88 55.4 1.45 38.21

3.2.6. Regeneration

The metal ions’ desorption from the adsorbent particles without adsorption efficiency loss is important for functionality of ion‐imprinted particles. In this regard, the adsorbed Pd2+ cations were extracted from the Pd‐Si‐IS particles via acidified thiourea solution as a desorption medium. The silica particles were then reutilized and desorbed five times in a row and each individual time, both the desorption efficiency and regeneration efficiency were calculated using Equations (8) and (9). It was noticed that the adsorption capacity marginally reduced by 5% after the fifth experiment. This might be pertaining to the partial hydrolysis of the C=N bond of the Schiff base that occurs due to continuous regeneration in acidic media.

3.2.7. Conclusions

Modified Pd2+ ion‐imprinted hollow chelating silica particles were prepared by the reaction of AT with isatin to synthesize the polymerizable Schiff base ligand that was consequently coordinated to the template Pd2+. The Pd2+ complex solution was mixed with tetraethoxysilane and set aside to form gel under slightly basic conditions. The template Pd2+ cations were leached from the resultant cross‐linked matrix. The molecules and substances synthesized throughout the synthetic route were studied using appropriate physical methods to distinguish the chemical structure of the obtained chelating silica particles. Furthermore, the synthesized ion‐imprinted as well as non‐ion‐imprinted silica particles were employed in several experiments in order to optimize and examine the selective Pd2+ removal process.

CONFLICT OF INTEREST

The authors declare no conflict of interest.

Nozari M, Monier M, Bashal AH. Design, synthesis, and study of Pd(II) ion‐imprinted functionalized polymer. Anal Sci Adv. 2020;1:109–123. 10.1002/ansa.202000030

DATA AVAILABILITY STATEMENT

Authors confirm that the data supporting this study are available within the article. The synthetic route is explained in full details in experimental sections in order to resynthesize the ligand and the complex. All the adsorption experiments were performed three times and the points were taken as an average. Any further information is available from the corresponding author, M.N., upon reasonable request.

REFERENCES

  • 1. Ho KY, Yeung KL. Effects of ozone pre‐treatment on the performance of Au/TiO2 catalyst for CO oxidation reaction. J Catal. 2006;242:131‐141. [Google Scholar]
  • 2. Lin S, Wei W, Wu X, et al. Selective recovery of Pd(II) from extremely acidic solution using ion‐imprinted chitosan fiber: adsorption performance and mechanisms. J Hazard Mater. 2015;299:10‐17. [DOI] [PubMed] [Google Scholar]
  • 3. Das N. Recovery of precious metals through biosorption ‐a review. Hydrometallurgy. 2010;103:180‐189. [Google Scholar]
  • 4. Park SI, Kwak IS, Bae MA, et al. Recovery of gold as a type of porous fiber by using biosorption followed by incineration. Bioresour Technol. 2012;104:208‐214. [DOI] [PubMed] [Google Scholar]
  • 5. Won SW, Kim S, Kotte P, et al. Cationic polymer‐immobilized polysulfone‐based fibers as high performance sorbents for Pt(IV) recovery from acidic solutions. J Hazard Mater. 2013;263:391‐397. [DOI] [PubMed] [Google Scholar]
  • 6. Zheng H, Hu D, Zhang L, et al. Thiol functionalized mesoporous silicas for selective adsorption of precious metals. Miner Eng. 2012;35:20‐26. [Google Scholar]
  • 7. Soylak M, Tuzen M. Coprecipitation of gold(III), palladium(II) and lead(II) for their flame atomic absorption spectrometric determinations. J Hazard Mater. 2008;152:651‐656. [DOI] [PubMed] [Google Scholar]
  • 8. Soylak M, Saracoglu S, Divrikli U, et al. Coprecipitation of heavy metals with erbium hydroxide for their flame atomic absorption spectrometric determinations in environmental samples. Talanta. 2005;66:1098‐1102. [DOI] [PubMed] [Google Scholar]
  • 9. Maheswari MA, Subramanian MS. Selective enrichment of U(VI), Th(IV) and La(III) from high acidic streams using a new chelating ion‐exchange polymeric matrix. Talanta. 2004;64:202‐209. [DOI] [PubMed] [Google Scholar]
  • 10. Mao J, Lee SY, Won SW, et al. Surface modified bacterial biosorbent with poly(allylamine hydrochloride): development using response surface methodology and use for recovery of hexachloroplatinate(IV) from aqueous solution. Water Res. 2010;44:5919‐5928. [DOI] [PubMed] [Google Scholar]
  • 11. Ihara K, Hasegawa SI, Naito K. The separation of aluminum(III) ions from the aqueous solution on membrane filter using Alizarin Yellow E. Talanta. 2008;75:944‐949. [DOI] [PubMed] [Google Scholar]
  • 12. Feng L, Zhang Y, Wen L, et al. Colorimetric determination of copper(II) ions by filtration on sol‐gel membrane doped with diphenylcarbazide. Talanta. 2011;84:913‐917. [DOI] [PubMed] [Google Scholar]
  • 13. Cho DH, Kim EY. Characterization of Pb2+ biosorption from aqueous solution by Rhodotorula glutinis . Bioprocess Biosyst Eng. 2003;25:271‐277. [DOI] [PubMed] [Google Scholar]
  • 14. Goksungur Y, Uren S, Guvenc U. Biosorption of cadmium and lead ions byethanol treated waste baker's yeast biomass. Bioresour Technol. 2005;96:103‐109. [DOI] [PubMed] [Google Scholar]
  • 15. Liu L, Li C, Bao C, et al. Preparation and characterization of chitosan/graphene oxide composites for the adsorption of Au(III) and Pd(II). Talanta. 2012;93:350‐357. [DOI] [PubMed] [Google Scholar]
  • 16. Monier M, Ibrahim AA, Metwally MM, et al. Surface ion‐imprinted amino‐functionalized cellulosic cotton fibers for selective extraction of Cu(II) ions. Int J Biol Macromolec. 2015;81:736‐746. [DOI] [PubMed] [Google Scholar]
  • 17. Monier M, Akl MA, Ali WM. Modification and characterization of cellulose cotton fibers for fast extraction of some precious metal ions. Int J Biol. 2014;66:125‐134. [DOI] [PubMed] [Google Scholar]
  • 18. Monier M, Abdel‐Latif DA. Modification and characterization of PET fibers for fast removal of Hg(II), Cu(II) and Co(II) metal ions from aqueous solutions. J Hazard Mater. 2013;250‐251:122‐130. [DOI] [PubMed] [Google Scholar]
  • 19. Liu F, Liu Y, Xu Y, et al. Efficient static and dynamic removal of Sr(II) from aqueos solution using chitosan ion‐impronted polymer functionalized with dithiocarbamate. J Environ Chem Eng. 2015;3:1061‐1071. [Google Scholar]
  • 20. Fujiwara K, Ramesh A, Maki T, et al. Adsorption of platinum(IV), palladium(II) and gold(III) from aqueous solutions onto l‐lysine modified crosslinked chitosan resin. J Hazard Mater. 2007;146:39‐50. [DOI] [PubMed] [Google Scholar]
  • 21. Monier M, Abdel‐Latif DA, Mohammed HA. Synthesis and characterization of uranyl ion‐imprinted microspheres based on amidoximated modified alginate. Int J Biol Macromol. 2015;75:354‐363. [DOI] [PubMed] [Google Scholar]
  • 22. Zhang Y, Bai Z, Luo W, et al. Ion imprinted adsorbent for the removal of Ni(II) from waste water: preparation, characterization, and adsorption. J Dispers Sci Technol. 2019;40:1‐10. [Google Scholar]
  • 23. Fu J, Chen L, Li J, et al. Current status and challenges of ion imprinting. J Mater Chem A. 2015;3:13598‐13627. [Google Scholar]
  • 24. Yusoff MM, Mostapa NRN, Sarkar MS, et al. Synthesis of ion imprinted polymers for selective recognition and separation of rare earth metals. J Rare Earth. 2017;35:177‐186. [Google Scholar]
  • 25. Monier M, Youssef I, El‐Mekabaty A. Preparation of functionalized ion‐imprinted phenolic polymer for efficient removal of copper ions. Polym Int. 2020;69:31‐40. [Google Scholar]
  • 26. Loy DA, Shea KJ. Bridged polysilsesquioxanes. Highly porous hybrid organic‐inorganic materials. Chem Rev. 1995;95:1431‐1442. [Google Scholar]
  • 27. Tani T, Mizoshita N, Inagaki S. Functionalized periodic mesoporous organosilicas for catalysis. J Mater Chem. 2009;19:4451‐4456. [Google Scholar]
  • 28. Jung BM, Kim MS, Kim WJ, et al. Molecularly imprinted mesoporous silica particles showing a rapid kinetic binding. Chem Commun. 2010;46:3699‐3701. [DOI] [PubMed] [Google Scholar]
  • 29. Alotaibi MR, Monier M, Elsayed N. Fabrication and investigation of gold (III) ion‐imprinted functionalized silica particles. J Mol Recognit. 2019;e2813:1‐14. [DOI] [PubMed] [Google Scholar]
  • 30. Francisco JE, Feiteira FN, da Silva WA, et al. Synthesis and application of ion‐imprinted polymer for the determination of mercury II in water sample. Environ Sci Pollut Res. 2019;26:19588‐19597. [DOI] [PubMed] [Google Scholar]
  • 31. Kumar A, Balouch A, Pathan AA, et al. Novel chromium imprinted polymer: synthesis, characterization and analytical applicability for the selective remediation of Cr(VI) from an aqueous system. Int J Environ Anal Chem. 2019;99:454‐473. [Google Scholar]
  • 32. da Santos Silva RC, Pires BC, Borges KB. Double‐imprinted polymer based on cross‐linked poly(vinylimidazole–trimethylolpropane trimethacrylate) in solid phase extraction for determination of lead from wastewater samples by UV–vis spectrophotometry. Int J Environ Anal Chem. 2019;99:949‐967. [Google Scholar]
  • 33. Biswas TK, Yusoff MM, Sarjadi MS, et al. Ion‐imprinted polymer for selective separation of cobalt, cadmium and lead ions from aqueous media. Sep Sci Technol. 2019:1‐10. 10.1080/01496395.2019.1575418. [DOI] [Google Scholar]
  • 34. Yasinzai M, Mustafa G, Asghar N, et al. Ion‐imprinted polymer‐based receptors for sensitive and selective detection of mercury ions in aqueous environment. J Sens. 2018;2018:1‐6. [Google Scholar]
  • 35. Godlewska‐Żyłkiewicz B, Leśniewska B, Wawreniuk I. Assessment of ion imprinted polymers based on Pd(II) chelate complexes for preconcentration and FAAS determination of palladium. Talanta. 2010;83:596‐604. [DOI] [PubMed] [Google Scholar]
  • 36. Jiang Y, Kim D. Synthesis and selective adsorption behavior of Pd(II)‐imprinted porous polymer particles. Chem Eng J. 2013;232:503‐509. [Google Scholar]
  • 37. Cataldo S, Gianguzza A, Pettignano A. Sorption of Pd(II) ion by calcium alginate gel beads at different chloride concentrations and pH. A kinetic and equilibrium study. Arab J Chem. 2016;9:656‐667. [Google Scholar]
  • 38. Monier M, Abdel‐Latif DA, Abou El‐Reash YG. Ion‐imprinted modified chitosan resin for selective removal of Pd(II) ions. J Colloid Interface Sci. 2016;469:344‐354. [DOI] [PubMed] [Google Scholar]
  • 39. Yu H, Shao P, Fang L, et al. Palladium ion‐imprinted polymers with PHEMA polymer brushes: role of grafting polymerization degree in anti‐interference. Chem Eng J. 2018. 10.1016/j.cej.2018.11.149. [DOI] [Google Scholar]
  • 40. Vogel. Quantitative Inorganic Analysis. 7th ed. London, UK: Longman; 2010. [Google Scholar]
  • 41. Monier M, Elsayed NH, Abdel‐Latif DA. Synthesis and application of ion‐imprinted resin based on modified melamine–thiourea for selective removal of Hg(II). Polym Int. 2015;64:1465‐1474. [Google Scholar]
  • 42. Cerchiaro G, da Costa Ferreira AM. Oxindoles and copper complexes with oxindole‐derivatives as potential pharmacological agents. J Braz Chem Soc. 2006;17:1473‐1485. [Google Scholar]
  • 43. Rabar VJ, Shinde VM. Metal complexes of phenanthraquinone monothiosemicarbazone & 2‐hydroxy‐1,4‐naphthaquinone thiosemicarbazone. Indian J Chem. 1983;22A:477‐490. [Google Scholar]
  • 44. Bhave NS, Kharat RB. Magnetic and spectral properties of Fe(II), Fe(III), Co(II), Ni(II), Cu(II) and Pd(II) chelates of 2’‐hydroxy‐3’‐bromo‐4‐methoxy‐5’‐methylchalcone oxime (HBMMCO). J Inorg Nucl Chem. 1980;42:977‐983. [Google Scholar]
  • 45. Kidsaneepoiboon P, Wanichwecharungruang SP, Chooppawa T, et al. Organic–inorganic hybrid polysilsesquioxane nanospheres as UVA/UVB absorber and fragrance carrier. J Mater Chem. 2011;21:7922‐7930. [Google Scholar]
  • 46. Monier M, Ayad DM, Wei Y, et al. Adsorption of Cu(II), Co(II), and Ni(II) ions by modified magnetic chitosan chelating resin. J Hazard Mater. 2010;177:962‐970. [DOI] [PubMed] [Google Scholar]
  • 47. Tan I, Ahmad A, Hameed B. Adsorption isotherms, kinetics, thermodynamics and desorption studies of 2,4,6‐trichlorophenol on oil palm empty fruit bunch‐based activated carbon. J Hazard Mater. 2009;164:473‐482. [DOI] [PubMed] [Google Scholar]
  • 48. Ho YS, McKay G. Sorption of dye from aqueous solution by peat. Chem Eng J. 1998;70:115‐124. [Google Scholar]
  • 49. Ho YS, Wase DAJ, Forster CF. Kinetic studies of competitive heavy metal adsorption by sphagnum moss peat. Environ Technol. 1996;17:71‐77. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Authors confirm that the data supporting this study are available within the article. The synthetic route is explained in full details in experimental sections in order to resynthesize the ligand and the complex. All the adsorption experiments were performed three times and the points were taken as an average. Any further information is available from the corresponding author, M.N., upon reasonable request.


Articles from Analytical Science Advances are provided here courtesy of Wiley

RESOURCES