Next Article in Journal
The Use of Retinoids for the Prevention and Treatment of Skin Cancers: An Updated Review
Previous Article in Journal
ADAM10 and ADAM17—Novel Players in Retinoblastoma Carcinogenesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid Coordination Networks for Removal of Pollutants from Wastewater

1
Department of Biology-Chemistry, Faculty of Chemistry, Biology, Geography, West University Timisoara, 16 Pestalozzi Street, 300115 Timisoara, Romania
2
“Coriolan Dragulescu” Institute of Chemistry, Romanian Academy, 24 Mihai Viteazul Bvd., 300223 Timisoara, Romania
3
Faculty of Industrial Chemistry and Environmental Engineering, Politehnica University Timisoara, 6 Vasile Parvan Blv., 300223 Timisoara, Romania
4
Research Institute for Renewable Energy, Politehnica University of Timişoara, G. Muzicescu 138, 300501 Timişoara, Romania
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(20), 12611; https://doi.org/10.3390/ijms232012611
Submission received: 28 September 2022 / Revised: 16 October 2022 / Accepted: 18 October 2022 / Published: 20 October 2022
(This article belongs to the Section Materials Science)

Abstract

:
The adsorption properties of two coordination polymers, resulting from the reaction of divalent metal (Ca2+ or Co2+) salts with (2-carboxyethyl)(phenyl)phosphinic acid, are presented in this paper. The structural and textural characterization before and after adsorption experiments is presented. The adsorbent materials were prepared using the hydrothermal procedure. The compound Ca[O2P(CH2CH2COOH)(C6H5)]2 (CaCEPPA) has a layered topology, with the phenyl groups oriented into the interlayer space and crystallizes in the monoclinic system. Compound Co2[(O2P(CH2CH2COO)(C6H5)(H2O)]2·2H2O (CoCEPPA) has a 1D structure composed of zig-zag chains. The adsorption performances of CaCEPPA and CoCEPPA materials were tested in the removal of cadmium and lead from aqueous solutions. The optimum pH of ions adsorption was found to be five for both adsorbent materials. Pseudo-first and second-order kinetic models were used for fitting kinetic experimental data, and Langmuir and Freundlich isotherms were used for modeling the equilibrium experimental data. The pseudo-second-order kinetic model and Langmuir isotherm best described the adsorption of Cd and Pb ions onto the studied materials, judging from the results of the error function (correlation coefficient, sum of square error, chi-square test, and average relative error) analysis. The studied materials present a higher affinity for Cd ions compared with the adsorption capacity developed for the removal of Pb ions from aqueous solutions. CoCEPPA showed the highest adsorption performance in the removal process of metal ions from aqueous solutions compared with CaCEPPA (qm = 54.9 mg Cd2+/g of CoCEPPA, qm = 36.5 mg Cd2+/g of CaCEPPA).

1. Introduction

The danger posed to humans and the environment by extremely toxic heavy metal lead and cadmium ions in the environment remains a serious issue. Lead and cadmium ions are discharged into the environment from various industries such as leather, cosmetics, electronics, and battery manufacturing. Increased amounts of Pb(II) in the human body are very toxic because Pb(II) is non-biodegradable. It disturbs nearly all body systems and produces irremediable diseases, including heart disease, hematological problems, renal dysfunction, brain harm, cancer, and even death [1,2,3]. Cadmium ions Cd(II) can cause lesions in many major organs, including the kidney, and also cause bronchiolitis, pulmonary disease, fibrosis, and skeletal damage [4,5,6]. Various methods have been used for the elimination of Pb(II) and Cd(II), such as ion exchange, adsorption, coagulation-flocculation, chemical precipitation, electro-dialysis, and biological methods [7,8].
Among these, adsorption is the most used technique due to its low cost and easy working conditions [9,10].
Porous coordination polymers, particularly the subclass metal–organic frameworks (MOFs), are obtained by linking organic and inorganic units with solid bonds. More than 20,000 MOFs have been described and studied since they first appeared in the literature. They represent a promising new class of porous crystalline solids materials with amazing properties such as high surface capacities, large pore sizes, stable porosity, high thermal stability, and open channels with tunable sizes and topology. MOFs are rigid frameworks that can coordinate in 1D, 2D, or three dimensions and form highly porous networks. The organic units are usually di or polytopic carboxylates, interconnected with metallic centers, and produce supramolecular crystalline structures. The area of the MOFs surface range from 1000 to 10,000 m2/g [11,12,13]. Several synthesis routes have been used for MOFs synthesis. The slow evaporation method is a common method to prepare MOFs in which a solution of raw materials is concentrated by slow evaporation of the solvent at a fixed temperature, most often at room temperature.
The most used method is solvothermal synthesis, but other methods are used, including microwave-assisted synthesis, which allows the faster synthesis of crystals compared to conventional heating; electrochemical synthesis, when the metal ions are continuously inserted through an anodic dissolution to the reaction mixture that contains a conducting salt and linker molecules; the mechanochemical method, where mechanical breakage of intramolecular bonds takes place followed by a chemical transformation; the sonochemical technique, where well-known compounds are reproduced by these alternative routes [14,15,16,17,18,19,20].
MOFs are promising materials for applications in separation, adsorption catalysis [21,22,23,24], drug delivery agents, gas storage, corrosion, molecular recognition [25], antibacterial agents [26], sensors, and so on [27,28] because of their designable framework structures; the size, dimensions, shape, and surface functionality of the nanochannels can be tuned by changing the combination of metal ions and organic ligands.
In addition to di-, tri-, or tetradentate carboxylate ligands, phosphonates and sulfonates have been used as organic linkers to obtain MOFs. This field is relevant to the core concept of the metal–organic framework [29,30,31,32].
Diphosphinic acids have also been used for the synthesis of MOFs. They are similar to di-carboxylic acids. A porous tubular 1D-MOF, consisting of Cu(II), 1,2-bis(4-pyridyl)ethane, and P,P′-diphenyl-diphosphinate, was obtained in an easy and direct self-assembly process in either needle microcrystal or nanorod forms, depending on the reaction conditions [33].
1D structural arrays are obtained when P,P′-diphenylmethylenediphosphinic acid (H2pcp) reacts with alkaline metal ions such as Mg(SO4) or CaCl2·6H2O [34]. Using auxiliary ligands such as 2,2′-bipyridine, a new tetranuclear complex connected through water–hydrogen-bonding of zinc(II) with P,P′-diphenylmethylenediphosphinate was synthesized [35]. P,P′-diphenylmethylenediphosphinic (H2pcp), P,P′-diphenylethylenediphosphinic acids, and copper(II) salts were synthesized in the presence of 2,2′-bipyridine, and the effect of carbon chain length connecting the two phosphinate species, both in the free acids and in CuII/2,2′-bipyridine derivatives, were investigated [36].
The properties of heterocyclic phosphorus compounds and their complexations are being extensively studied for a wide variety of applications in numerous fields [37].
Phosphorus-bearing calixarenes have received significant attention for their ionophoric receptor characteristics and for their ability to localize transition metal reactive centers present close to the cavity as host–guest and supramolecular chemistry as “building blocks” three-dimensionally. The intention is to use them as selective cation receptors and delivery systems [38,39].
(2-Carboxyethyl)(phenyl)phosphinic acid represents another type of ligand used for obtaining metal–organic frameworks because it features polar phosphinic acid and carboxy moieties at the ends of a flexible –(CH2)n– spacer (n = 2), which is an efficient asymmetric ligand for inducing the formation of noncentrosymmetric structures for metal complexes [40,41]. 2-Carboxyethyl (phenyl)phosphinic acid has also been used as a flame retardant prepared via the reaction with melamine [42] or with dimethyl terephthalate and ethylene glycol [43] or with 2,5-furandicarboxylic acid and 1,6-hexanediol [44].
Only a few papers deal with the synthesis of metal networks in which 2-carboxyethyl(phenyl)phosphinic acid was involved, and this is the first one that describes this type of material as an adsorbent for heavy metals. Furthermore, the adsorption performance of CaCEPPA and CoCEPPA materials was investigated in the removal process of cadmium and lead ions from aqueous solutions. The optimum pH, concentration, and stirring time were investigated to reach the maximum adsorption capacities for both adsorbent materials used. Pseudo-first and second-order kinetic models were performed for fitting kinetic experimental data and Langmuir and Freundlich isotherms for modeling the equilibrium experimental data.

2. Results and Discussion

2.1. Characterization of Materials

Figure 1 presents the morphology of the synthesized materials. Compared with CoCEPPA, whose surface is denser with particle conglomerates of several sizes and shapes, CaCEPPA based on Ca ions presents a well-ordered structure with particles of well-defined sizes and shapes.
To investigate the crystalline structure of CoCEPPA and CaCEPPA before and after adsorption of lead and cadmium, PXRD measurements were performed, Figure 2.
Judging from the PXRD spectra of both adsorbent materials, the crystallinity is highly affected after the adsorbance of metal ions on the studied materials, in concordance with IR spectra of adsorbents before and after metal ions adsorption. The XRD analysis identified the formation of CdCO3 or Cd3(PO4)2, especially in the case of CoCEPPA adsorbent materials (Figure 2b), with two more peaks appearing at 22.67° and 24.19°. For CaCEPPA, these peaks are covered by the peaks specific to the based materials [45]. In the case of lead adsorption, it is also highlighted that some complexes are formed at the adsorbent surface due to the appearance of new peaks at 2θ = 11.1° and 16.6° [46] in the XRD spectrum of both studied materials. To gain a deeper understanding of the mechanism of adsorption, in addition to the isothermal and kinetic studies, FTIR spectra of Pb(II)-impregnated and Cd(II)-impregnated CoCEPPA and CaCEPPA were recorded and compared to that of pristine CoCEPPA and CaCEPPA (Figure 3a,b). The results indicate that chemisorption may be the main adsorption mechanism, with the carboxylic oxygen in both CoCEPPA and CaCEPPA potentially coordinating with the Pb2+ and Cd2+.
FTIR spectra of CaCEPPA compounds present a band at 1684 cm−1, which is absent in the spectra of CoCEPPA, revealing the presence of a carboxylic group in calcium-based adsorbent material. CoCEPPA bands in the region of 3500–3300 cm−1 (broad), together with the vibration band of H–O–H at 1638 cm−1, reveal the presence of H-bonded water in concordance with the crystallographic data [41]. While in the IR spectrum of CoCEPPA, no peak near 1700 cm−1 for the –COOH group is observed, two strong peaks around 1500–1345 cm−1 are observed, which are characteristic of absorption of asymmetric and symmetric vibrations of coordinated carboxylate groups present in the case of CaCEPPA. Strong characteristic bands of P=O in the region from 1390 to 1000 cm−1 and of P–O around 1050–850 cm−1 are present [46,47]. An increase in the peak strength or slight broadening of the existing peaks in the region 1700–1000 cm−1 are characteristic bands for P=O and of P–O, together with the appearance of new peaks created in the range of 400–600 cm−1 is related to Me2+-O peaks. All of these are evidence of coordination interactions between Me2+ and O atoms from the carboxylic group and indicate the involvement of the Cd2+ and Pb2+ on the surface of the adsorbent during the adsorption process. Results of FT-IR showed that the adsorption of Me2+ is accomplished by two processes, chemisorption and physisorption. The thermal stability studies of the CaCEPPA and CoCEPPA compounds were previously described [41].

2.2. Adsorption Studies

2.2.1. Influence of Initial pH of Aqueous Solutions upon the Adsorption Properties of CaCEPPA and CoCEPPA Materials

The initial pH of the solution influences both phases involved in the adsorption processes, influencing the charge of the adsorbent surface and the form of the adsorbate.
In addition to the physical adsorption in the pores of the adsorbent, it is always much better to favor the conditions for achieving strong electrostatic attractions. The experimental results regarding the influence of the initial pH of aqueous solutions containing 30 mg/L of cadmium and respective lead ions on the adsorption capacity of CaCEPPA and CoCEPPA are presented in Figure 4.
It can be observed that at lower pH values, due to the competition between the Cd2+/Pb2+ ions and the H+, the studied materials developed the lowest adsorption capacities [48]. Increasing the initial pH of aqueous solutions from 2 to 5 also increases the adsorption capacities of the studied materials. In the case of Cd ions removal, a significant increase at pH = 8 and in the case of Pb ions at pH > 6 could be observed, but in these cases, the removal could not be attributed to the adsorption process. In these cases, the removal is due to the precipitation of metal ions as hydroxides, according to the distribution diagram of cadmium and lead species function on pH [49,50]. To avoid the precipitation processes and to obtain the maximum adsorption capacities developed by the studied compounds, further adsorption studies were conducted using aqueous solutions containing metal ions with an initial pH of 5.

2.2.2. Stirring Time Influence on Adsorption Properties of CaCEPPA and CoCEPPA Materials

Figure 5 and Figure 6 present the influence of stirring time on the adsorption capacities of CaCEPPA and CoCEPPA in the removal process of Cd; respective Pb ions from aqueous solutions, and the non-linear pseudo-first-order and pseudo-second-order kinetic models were used for fitting the experimental data. The equilibrium between Cd, respective Pb ions, and CaCEPPA and CoCEPPA adsorbent materials was achieved in 60 min.
The pseudo-first-order and pseudo-second-order kinetic parameters, together with the error analysis results, are presented in Table 1.
From the experimental data presented in Table 1, it can be observed that for the pseudo-second-order kinetic model, values of the correlation coefficient close to unity and the lowest values of the chi-square test χ2 are obtained. The error functions ERRSQ and ARE obtained for the pseudo-second-order kinetic models present values of order of units; instead, for the pseudo-first-order kinetic model, these errors present values of order of hundreds. The results suggest that the adsorption of metal ions (Cd2+, respective Pb2+) onto CaCEPPA and CoCEPPA is best described by the pseudo-second-order kinetic model, implying a chemical sorption mechanism.

2.2.3. Influence of Equilibrium Concentration of Metal Ions upon the Adsorption Properties of CaCEPPA and CoCEPPA Materials

The experimental results regarding the equilibrium studies and their fitting with the Langmuir, Freundlich, Temkin, and Dubinin–Raduschkevich isotherms are presented in Figure 7 and Figure 8.
It can be observed that both studied materials present a higher affinity for Cd2+ ions compared with Pb2+ ions. CoCEPPA developed a higher adsorption capacity for both metal ions compared with the adsorption capacities developed by CaCEPPA. The maximum adsorption capacity determined experimentally for CoCEPPA is 51.6 mg/g for Cd2+ and 32.7 mg/g for Pb2+, respectively. CaCEPPA developed a maximum adsorption capacity of 32.6 mg/g for Cd2+ ions removal and 21.4 mg/g for Pb2+, respectively.
The Langmuir, Freundlich, Temkin, and Dubinin–Raduschkevich isotherm parameters, together with the error analysis results, are presented in Table 2.
From the results presented in Table 2, the maximum adsorption capacities calculated from the Langmuir isotherm are very close to those determined experimentally for both studied materials, and both metal ions adsorbed could be observed. Additionally, in this case, the lowest values of the error functions and the highest values for the correlation coefficient are obtained, compared with the values obtained for the other studied isotherm. The Temkin isotherm model gave the best fit after the Langmuir isotherm, with an obtained correlation coefficient higher than 0.96, but the error values are higher than the Langmuir isotherm. The Temkin isotherm model, whose adsorption is characterized by a uniform distribution of the binding energies up to a notable maximum binding energy, was verified. The positive value of the variation of adsorption energy parameter b suggests that the adsorption of Cd2+/Pb2+ onto the studied materials is exothermic in nature (b = −ΔH). The values of theoretical monolayer saturation capacity in the Dubinin–Radushkevich model obtained using non-linear regression are all lower than the experimental amounts corresponding to the adsorption isotherm plateau, and the lower values obtained for the correlation coefficient and the highest value obtained for the studied errors indicate that the modeling of Dubinin–Radushkevich for the adsorption system of the Cd2+/Pb2+ is unacceptable. The obtained results clearly show that the values predicted from the Langmuir isotherm are close to the experimental data, suggesting that the Cd2+ and Pb2+ ions adsorption onto CaCEPPA and CoCEPPA materials take place under a homogeneous mechanism as a monolayer on the adsorbent surface. Therefore, by comparison, the order of the isotherm best fits the four sets of experimental data in this study is Langmuir > Temikn > Freundlich > Dubinin–Radushkevich.
To better understand the mechanism of the adsorption, we analyzed the packing of the CoCEPPA and CaCEPPA and measured the distances between the Me2+ situated in the adjacent layer. We found that the distance between metallic centers of two adjacent layers Co2+⋅⋅⋅Co2+ is 11.769Å, while for Ca2+⋅⋅⋅Ca2+, the distance is 11.566 Å, as can be observed in Figure 9. The 1D orientation of CoCEPPA compared with the layer structure of CaCEPPA allows the metal ions to more easily access the surface of the adsorbent material, and this explains the higher adsorption capacity of CoCEPPA [41].
Additionally, the properties of Pb2+ and Cd2+ are also important. The lowest Gibbs free energy of hydration was observed for Pb2+ (1425 kJ mol−1) in comparison with Cd2+ (1755 kJ mol−1). The radius of Cd2+ is 0.095 nm, while the radius of Pb2+ is 0.119 nm. Correlating the results of the adsorption studies with the XRD and FTIR analysis and the structure of the adsorbent materials, the complex mechanism involved in the metal ion removal is clear.

2.2.4. Comparison with Other Adsorbents

The maximum adsorption capacity obtained from the Langmuir isotherm for the removal of Cd2+/Pb2+ ions onto the studied materials was compared with other findings in the literature on adsorption capacities and are presented in Table 3 and Table 4. It can be observed that the studied materials present a good adsorption efficiency in the removal of Cd2+/Pb2+ ions compared with similar materials presented in the specialty literature.

3. Materials and Methods

3.1. Materials

2-Carboxyethylphenylphosphinic acid (CEPPA) was purchased from Carbosynth Limited. Cobalt acetate tetrahydrate and calcium acetate monohydrate were purchased from Sigma-Aldrich. All chemicals were obtained from commercial sources and used without purification.

3.2. Instrumentation

Thermal analysis (TG-DTA) was performed on an SDT-Q600 analyzer from TA Instruments New Castle, DE, USA. A thermogravimetric analyzer (Perkin-Elmer, New York, NY, USA) was used at a 30–680 °C temperature range, at a heating rate of 10 °C/min, under an N2 flow. A FEG 250 microscope (Quanta, Field Electron and Iron Company (FEI), Hillsboro, OR, USA), equipped with an EDAX/ZAF quantifier, was used for obtaining SEM images. Lead and cadmium ions adsorption were measured using a SpectrAA 280 FS atomic absorption spectrophotometer (Varian, Melbourne, Australia). The adsorption studies were performed in batch mode using an SW23 shaker bath (Julabo Labortechnik GmbH, Seelbach, Germany).

3.3. Synthesis of Materials

Co2[(O2P(CH2CH2COO)(C6H5)(H2O)]2·2H2O (CoCEPPA) and Ca[O2P(CH2CH2COOH)(C6H5)]2 (CaCEPPA)

These compounds were synthesized similarly to our previous papers. A solution of Co(CH3COO)2·4H2O (50.0 mmol) or Ca(CH3COO)2·H2O (50.0 mmol) and bidistilled water (50 mL) was stirred at a constant speed of 1000 rpm until a clear (violet or white) solution was obtained. Secondly, 2-carboxyethylphenylphosphinic acid (CEPPA) (50.0 mmol) and bidistilled water (50 mL) were mixed at 60 °C until a colorless and clear solution was obtained. Both solutions were mixed, and the pH was adjusted to 2.8 with an aqueous solution of NaOH (0.1 M). Then, the clear white or violet solution was heated in an oil bath at 80 °C for 75 h, unperturbed. After 75 h heating, crystalline white (CaCEPPA) and black crystals (CoCEPPA) were isolated by filtration and, finally, air dried (yield: 62–70%) [55,56,57].

3.4. Adsorption Studies

The CaCEPPA and CoCEPPA materials were used as adsorbents in the removal of cadmium (Cd2+) and lead (Pb2+) ions from synthetic aqueous solutions to determine their adsorption properties. The process was conducted in batch mode, determining the influence of the initial pH (2–8) solution, stirring time (5–120 min), and initial concentrations of pollutants (C0 = 5–200 mg/L) upon the adsorption performance of the studied compounds. For all the studies, the ratio between the adsorbent and aqueous solution was 1:1 (a solution of 0.025 g adsorbent in 25 mL of water was used). The initial pH of solutions was calibrated at the desired values by the addition of 1M NaOH or HCl solutions and was measured with a Mettler Toledo pH meter. When the pH and the stirring time were varied, an aqueous solution containing 30 mg/L metal ions was used. When the pH and the initial concentration of ions were varied, the samples were stirred for 60 min. The initial and equilibrium concentration of cadmium and lead ions were determined using a Varian SpectrAA 280 FS atomic absorption spectrophotometer by atomic absorption spectroscopy. The adsorption performance of the studied materials was evaluated by calculating the adsorption capacity according to Equation (1):
q e = ( C 0 C e ) · V m
where qe is the adsorption capacity of the adsorbent at equilibrium (mg/g), V is the volume of the aqueous solution (L), C0 (mg/L) represents the initial concentration of metal ions (Cd or Pb) in the aqueous solutions, and Ce (mg/L) represents the residual concentration (the equilibrium concentration) of metal ions in solutions, and m is the mass of adsorbent (g).
The adsorption mechanism and the equilibrium between the adsorbents, cadmium, and lead ions, could be determined by matching the experimental results concerning the influence of stirring time on the adsorption capacity of CaCEPPA and CoCEPPA materials with the known kinetic models, pseudo-first-order and pseudo-second-order kinetic models, presented in their non-linear form in Equations (2) and (3) [57,58].
q t = q e ( 1 exp k 1 · t )
q t = k 2 · q e 2 · t 1 + k 2 · q e · t
where qe and qt (mg/g) is the adsorption capacity of the adsorbent at equilibrium and at a specific time t, k1 represents the rate constant of the pseudo-second-order kinetic model (min−1), k2 represents the rate constant of the pseudo-second-order kinetic model (g∙mg−1∙min−1), and t is the stirring time (min).
The maximum adsorption capacity of the studied materials was determined by modeling the experimental data, taking into account the influence of the initial concentration of metal ions from aqueous solutions on the adsorption behavior of CaCEPPA and CoCEPPA materials, using non-linear forms of Langmuir, Freundlich, Temkin, and Dubinin–Raduschkevich isotherms (Equations (4)–(7)) [57,58].
q e = q max · K L · C e 1 + K L · C e
where qmax (mg/g) is the maximum saturated monolayer adsorption capacity under the given conditions, KL (L/mg) is the Langmuir constant related to the affinity between the adsorbent and adsorbate.
q e = K F · C e 1 n
where KF (mg/g)(L/mg)1/n is the Freundlich constant and 1/n (0 < n < 10) is a Freundlich intensity parameter.
q e = RT b ln ( K T C e )
where R (J/mol K) is the gas constant, and T (K) is the absolute temperature, KT is the adsorption potential or the equilibrium binding constant (Lg or Lmol−1) corresponding to the maximum binding energy and interaction between adsorbate and adsorbent, while adsorption energy, bT (Jmol−1) is related to heat of adsorption (ΔQ), that is; bT = ΔQ = −ΔH).
q e = q m e K DR ε 2
where qm (mg/g) is a constant in the Dubinin–Radushkevich isotherm model which is related to adsorption capacity; KDR (mol2/kJ2) is a constant related to the mean free energy of adsorption.
ε = RTln ( 1 + 1 C e )
where R (J/mol K) is the gas constant, and T (K) is the absolute temperature.
To predict the best kinetic and isotherm model which describes better the adsorption of cadmium and lead ions onto CaCEPPA and CoCEPPA materials, four error functions were applied: correlation coefficient R2; sum of squares of the errors ERRSQ (Equation (9)); chi-square test χ2 (Equation (10)); average relative error ARE (Equation (11)) [59,60,61,62].
ERRSQ = i = 1 n ( q e   calc q e   exp ) i 2
χ 2 = i = 1 n ( q e   exp q e   calc   ) 2 q e   calc
ARE = 100 n i = 1 n | q e   calc q e   exp q e   calc | i

4. Conclusions

CaCEPPA and CoCEPPA showed acceptable performance in the removal of Cd2+ and Pb2+ from an aqueous solution. The metal ions’ adsorption onto the studied materials is better described by the pseudo-second-order kinetic model and by the Langmuir isotherm, suggesting that the adsorption mechanism could be chemical sorption in the case of Me2+ ions. CoCEPPA developed the highest adsorption capacity compared to CaCEPPA, and both materials presented a higher affinity for Cd ions compared to Pb2+ ions. In addition to physical adsorption, the phenyl functional groups of the adsorbent materials, which attract the metal ions through electrostatic interactions or Me-O bonds, have a great contribution. This explains the better adsorption performance developed by the CoCEPPA, which is a zig-zag chain 1D structure with available functional groups for adsorption on the entire length of the chain. In the case of CaCEPPA, a monoclinic system, due to the fact that phenyl groups are orientated into the interlayer space, the metal ions do not easily reach them to form a Me-O bond; therefore, less adsorption performance is developed by this material.

Author Contributions

Conceptualization, A.V., G.I., B.M. and L.L.; Methodology, M.M., B.M., V.C. and L.L.; Software, A.V. and L.L.; Validation, A.V., G.I., B.M. and L.L.; Formal Analysis, I.H. and V.C.; Investigation, A.V., G.I., B.M. and L.L.; Resources, A.V., G.I., B.M. and L.L.; Data Curation, A.V. and L.L.; Writing—Original Draft Preparation, A.V., G.I., B.M., V.C. and L.L.; Writing—Review and Editing, A.V., G.I., B.M. and L.L.; Supervision G.I.; Project Administration, G.I.; Funding Acquisition, G.I. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Ministry of Research, Innovation, and Digitization, CNCS-UEFISCDI, project number PN-III-P4-PCE-2021-0089, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data analyzed during this study are included in this published article.

Acknowledgments

We acknowledge the Romanian Academy, Program 2, “Coriolan Dragulescu” Institute of Chemistry. This work was supported by a grant from the Ministry of Research, Innovation, and Digitization, CNCS-UEFISCDI, project number PN-III-P4-PCE-2021-0089, within PNCDI III.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dhahri, R.; Yılmaz, M.; Mechi, L.; Alsukaibi, A.K.D.; Alimi, F.; ben Salem, R.; Moussaoui, Y. Optimization of the Preparation of Activated Carbon from Prickly Pear Seed Cake for the Removal of Lead and Cadmium Ions from Aqueous Solution. Sustainability 2022, 14, 3245. [Google Scholar] [CrossRef]
  2. Flora, G.; Gupta, D.; Tiwari, A. Toxicity of lead: A review with recent updates. Interdiscip. Toxicol. 2012, 5, 47–58. [Google Scholar] [CrossRef] [PubMed]
  3. Maranescu, B.; Lupa, L.; Visa, A. Synthesis, characterizations and Pb(II) sorption properties of cobalt phosphonate materials. Pure Appl. Chem. 2016, 88, 979–992. [Google Scholar] [CrossRef] [Green Version]
  4. Skipper, A.; Sims, J.N.; Yedjou, C.G.; Tchounwou, P.B. Cadmium chloride induces DNA damage and apoptosis of human liver carcinoma cells via oxidative stress. Int. J. Environ. Res. Public Health 2016, 13, 88. [Google Scholar] [CrossRef] [Green Version]
  5. Xu, M.Y.; Wang, P.; Sun, Y.J.; Yi-Jun Wu, Y.J. Disruption of kidney metabolism in rats after subchronic combined exposure to low-dose cadmium and chlorpyrifos. Chem. Res. Toxicol. 2019, 32, 122–129. [Google Scholar] [CrossRef]
  6. Abbas, A.; Al-Amer, A.M.; Laoui, T.; Al-Marri, M.J.; Nasser, M.S.; Khraisheh, M.; Atieh, M.A. Heavy metal removal from aqueous solution by advanced carbon nanotubes: Critical review of adsorption applications. Sep. Purif. Technol. 2016, 157, 141–161. [Google Scholar]
  7. Marques, J.P.; Vaz, C.M.P.; Sígolo, J.B.; Rodrigues, V.G.S. Soils of the Ribeira Valley (Brazil) as Environmental Protection Barriers: Characterization and Adsorption of Lead and Cadmium. Sustainability 2022, 14, 5135. [Google Scholar] [CrossRef]
  8. Kim, H.; Lee, B.I.; Byeon, S.H. The inner filter effect of Cr (VI) on Tb–doped layered rare earth hydroxychlorides: New fluorescent adsorbents for the simple detection of Cr (VI). Chem. Commun. 2015, 51, 725–728. [Google Scholar] [CrossRef]
  9. Lin, R.B.; Li, T.Y.; Zhou, H.L.; He, C.T.; Zhang, J.P.; Chen, X.M. Tuning fluorocarbon adsorption in new isoreticular porous coordination frameworks for heat transformation applications. Chem. Sci. 2015, 6, 2516–2521. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, K.; Gu, J.; Yin, N. Efficient removal of Pb(II) and Cd(II) Using NH2-Functionalized Zr-MOFs via rapid microwave-promoted synthesis. Ind. Eng. Chem. Res. 2017, 56, 1880–1887. [Google Scholar] [CrossRef]
  11. Wang, Z.; Cohen, S.M. Postsynthetic modification of metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1315–1329. [Google Scholar] [CrossRef] [PubMed]
  12. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hu, M. Design, Synthesis and Applications of Metal Organic Frameworks. Ph.D. Thesis, Department of Chemistry and Biochemistry, Worcester Polytechnic Institute, Worcester, MA, USA, 2 September 2011. Available online: https://digitalcommons.wpi.edu/etd-theses/964/ (accessed on 8 March 2022).
  14. Sun, Y.; Zhou, H.C. Recent progress in the synthesis of metal–organic frameworks. Sci. Technol. Adv. Mater. 2015, 16, 054202. [Google Scholar] [CrossRef] [PubMed]
  15. Stock, N.; Biswas, S. Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites. Chem. Rev. 2012, 112, 933–969. [Google Scholar] [CrossRef]
  16. Macarie, L.; Simulescu, V.; Ilia, G. Ultrasonic irradiation used in synthesis of aminophosphonates. Monatsh. Chem. 2019, 150, 163–171. [Google Scholar] [CrossRef]
  17. Khan, M.A.; Jhung, S.H. Synthesis of metal-organic frameworks (MOFs) with microwave or ultrasound: Rapid reaction, phase-selectivity, and size reduction. Coord. Chem. Rev. 2015, 285, 11–23. [Google Scholar] [CrossRef]
  18. Iliescu, S.; Ilia, G.; Pascariu, A.; Popa, A.; Plesu, N. Novel synthesis of phosphorus containing polymers under inverse phase transfer catalysis. Polymer 2006, 47, 6509–6512. [Google Scholar] [CrossRef]
  19. Visa, A.; Mracec, M.; Maranescu, B.; Maranescu, V.; Ilia, G.; Popa, A.; Mracec, M. Structure simulation into a lamellar supramolecular network and calculation of the metal ions/ligands ratio. Chem. Cent. J. 2012, 6, 91. [Google Scholar] [CrossRef] [Green Version]
  20. Popa, A.; Ilia, G.; Pascariu, A.; Iliescu, S.; Plesu, N. Grafted styrene-divinylbenzene copolymers containing benzaldehyde and their Wittig reactions with various phosphonium salts. Chin. J. Polym. Sci. 2005, 23, 651–656. [Google Scholar] [CrossRef]
  21. Maranescu, B.; Lupa, L.; Visa, A. Heavy metal removal from waste waters by phosphonate metal organic frameworks. Pure Appl. Chem. 2018, 90, 35–47. [Google Scholar] [CrossRef]
  22. Nistor, M.A.; Muntean, S.G.; Maranescu, B.; Visa, A. Phosphonate metal organic frameworks used as dyes removal materials from wastewaters. Appl. Organomet. Chem. 2020, 34, e5939. [Google Scholar] [CrossRef]
  23. Maranescu, B.; Popa, A.; Lupa, L.; Maranescu, V.; Visa, A. Use of chitosan complex with aminophosphonic groups and cobalt for the removal of Sr2+ ions. Sep. Sci. Technol. 2018, 53, 1058–1064. [Google Scholar] [CrossRef]
  24. Lupa, L.; Maranescu, B.; Visa, A. Equilibrium and kinetic studies of chromium ions adsorption on Co(II) based phosphonate metal organic frameworks. Sep. Sci. Technol. 2018, 53, 1017–1026. [Google Scholar] [CrossRef]
  25. Maranescu, B.; Lupa, L.; Tara Lunga Mihali, M.; Plesu, N.; Maranescu, V.; Visa, A. The Corrosion Inhibitor Behavior of Iron in Saline Solution by the Action of Magnesium Carboxyphosphonate. Pure Appl. Chem. 2018, 90, 1713–1722. [Google Scholar] [CrossRef]
  26. Popa, A.; Ilia, G.; Iliescu, S.; Dehelean, G.; Pascariu, A.; Bora, A.; Pacureanu, L. Mixed quaternary ammonium and phosphonium salts bound to macromolecular supports for removal bacteria from water. Mol. Cryst. Liq. Cryst. 2004, 418, 195–203. [Google Scholar] [CrossRef]
  27. Uemura, T.; Yanai, N.; Kitagawa, S. Polymerization reactions in porous coordination polymers. Chem. Soc. Rev. 2009, 38, 1228–1236. [Google Scholar] [CrossRef] [Green Version]
  28. Sharmin, E.; Zafar, F. (Eds.) Introductory Chapter: Metal Organic Frameworks (MOFs). In Metal-Organic Frameworks; IntechOpen Limited: London, UK, 2016; pp. 3–16. Available online: https://www.intechopen.com/books/metal-organic-frameworks (accessed on 12 October 2016).
  29. Shimizu, G.K.H.; Vaidhyanathan, R.; Taylor, J.M. Phosphonate and sulfonate metal organic frameworks. Chem. Soc. Rev. 2009, 38, 1430–1449. [Google Scholar] [CrossRef] [PubMed]
  30. Clearfield, A.; Demadis, K.D. Metal Phosphonate Chemistry: From Synthesis to Applications; Royal Society of Chemistry: London, UK, 2012. [Google Scholar]
  31. Maranescu, B.; Visa, A.; Mracec, M.; Ilia, G.; Maranescu, V.; Simon, Z.; Mracec, M. Lamellar Co2+ vinylphosphonate metal organic framework.PM3 semi-empirical analysis of structural properties. Rev. Roum. Chim. 2011, 56, 473–482. [Google Scholar]
  32. Maranescu, B.; Visa, A.; Ilia, G.; Simon, Z.; Demadis, K.; Colodrero, R.M.P.A.; Vallcorba, O.; Rius, J.; Choquesillo-Lazarte, D. Synthesis and characterization of styryl phosphonic acid and its use as new ligand for phosphonate metal organic framework. J. Coord. Chem. 2014, 67, 1562–1572. [Google Scholar] [CrossRef]
  33. Bataille, T.; Bracco, S.; Comotti, A.; Costantino, F.; Guerri, A.; Ienco, A.; Marmottini, F. Solvent dependent synthesis of micro- and nano- crystalline phosphinate based 1D tubular MOF: Structure and CO2 adsorption selectivity. Cryst. Eng. Comm. 2012, 14, 7170–7173. [Google Scholar] [CrossRef]
  34. Midollini, S.; Lorenzo-Luis, P.; Orlandini, A. Inorganic–organic hybrid materials of p,p0-diphenylmethylenediphosphinic acid (H2pcp) with magnesium and calcium ions: Synthesis and characterization of [Mg(Hpcp)2], [Mg(Hpcp)2(H2O)4], [Mg(pcp)(H2O)3](H2O), [Ca(Hpcp)2] and [Ca(pcp)(H2O)] complexes. Inorg. Chim. Acta 2006, 359, 3275–3282. [Google Scholar] [CrossRef]
  35. Ienco, A.; Midollini, S.; Orlandini, A.; Costantino, F. Synthesis and Structural Characterization of a Tetranuclear Zinc(II) Complex with P,P′ diphenylmethylenediphosphinate (pcp) and2,2′-Bipyridine (2,2′-bipy) Ligands. Z. Naturforsch. 2007, 62b, 1476–1480. [Google Scholar] [CrossRef]
  36. Costantino, F.; Ienco, A.; Midollini, S.; Orlandini, A.; Sorace, L.; Vacca, A. Copper(II) complexes with bridging diphosphinates—The effect of the elongation of the aliphatic chain on the structural arrangements around the metal centres. Eur. J. Inorg. Chem. 2008, 2008, 3046–3055. [Google Scholar] [CrossRef]
  37. Plinta, H.J.; Neda, I.; Schmutzler, R. 1.3-Dimethyl-l,3-diaza-2-R-5,6-benzo-2 λ3-phosphorinan-4-ones (R = F, Me2N, 2-Methylpiperidino, MeC(:O)NH-) as Ligands in Transition-Metal Complexes; Synthesis and Structure of DichloroPlatinum(II)- and Tetracarbonyl-Metal(0) Coordination Compounds (Metal = Cr, Mo and W). Z. Naturforsch. 1994, 49b, 100–110. [Google Scholar]
  38. Dieleman, C.B.; Matt, D.; Neda, I.; Schmutzler, R.; Harriman, A.; Yaftian, R. Hexahomotrioxacalix[3]arene: A scaffold for a C3-symmetric phosphine ligand that traps a hydridorhodium fragment inside a molecular funnel. Chem. Commun. 1999, 18, 1911–1912. [Google Scholar] [CrossRef]
  39. Vollbrecht, A.; Neda, I.; Thonnessen, H.; Jones, P.G.; Harris, R.K.; Crowe, L.A.; Schmutzler, R. Chemische Synthesis, structure, and reactivity of tetrakis(o,o-phosphorus)-bridged calix[4]resorcinols and their derivatives. Berichte 1997, 130, 1715–1720. [Google Scholar] [CrossRef]
  40. Hu, Q.S.; Zhang, X.Z.; Luo, S.F.; Sun, Y.H.; Du, Z.-Y. Two polymorphs of (2-carboxyethyl)-(phenyl)phosphinic acid. Acta Crystallogr. 2011, C67, o195–o197. [Google Scholar] [CrossRef]
  41. Bazaga-García, M.; Vílchez-Cózar, Á.; Maranescu, B.; Olivera-Pastor, P.; Marganovici, M.; Ilia, G.; Cabeza Díaz, A.; Visa, A.; Colodrero, R.M.P. Synthesis and electrochemical properties of metal(II)-carboxyethylphenylphosphinates. Dalton Trans. 2021, 50, 6539–6548. [Google Scholar] [CrossRef]
  42. You, L.; Hui, Y.; Shi, X.; Peng, Z. Study on the synthesis and characterization of a novel phosphorus-nitrogen containing intumescent flame retardant. Adv. Mater. Res. 2012, 399–401, 1376–1380. [Google Scholar] [CrossRef]
  43. Li, L.J.; Duan, R.T.; Zhang, J.B.; Wang, X.L.; Chen, L.; Yu-Zhong, W. Phosphorus-Containing Poly(ethylene terephthalate): Solid-State Polymerization and Its Sequential Distribution. Ind. Eng. Chem. Res. 2013, 52, 5326–5333. [Google Scholar] [CrossRef]
  44. Wang, G.; Jiang, M.; Zhang, Q.; Wang, P.; Qu, X.; Zhou, G. Poly(hexamethylene 2,5-furandicarboxylate) copolyesters containing phosphorus: Synthesis, crystallization behavior, thermal, mechanical and flame retardant properties. Polym. Degrad. Stab. 2018, 153, 272–280. [Google Scholar] [CrossRef]
  45. YiYi, D.; Huang, S.; Laird, D.A.; Wang, X.; Dong, C. Quantitative mechanisms of cadmium adsorption on rice straw- and swine manure-derived biochars. Environ. Sci. Pollut. Res. 2018, 25, 32418–32432. [Google Scholar]
  46. Sharma, R.; Sarswat, A.; Pittman, C.U.; Mohan, D. Cadmium and lead remediation using magnetic and non-magnetic sustainable biosorbents derived from Bauhinia purpurea pods. RSC Adv. 2017, 7, 8606–8624. [Google Scholar] [CrossRef] [Green Version]
  47. Xue, C.C.; Li, M.X.; Shao, M. Two novel 2D cadmium compounds with noncentrosymmetric or symmetric network dependent on different pH values. Russ. J. Coord. Chem. 2016, 42, 442–448. [Google Scholar] [CrossRef]
  48. Wang, X.; Wang, L.; Wang, Y.; Tan, R.; Ke, X.; Zhou, X.; Cheng, J.; Hou, H.; Zhou, M. Calcium sulfate hemihydrate whiskers obtained from flue gas desulfurization gypsum and used for the adsorption removal of lead. Crystals 2017, 7, 270. [Google Scholar] [CrossRef] [Green Version]
  49. Liang, Y.; Jun, M.; Liu, W. Enhanced removal of lead(II) and cadmium(II) from water in alum coagulation by ferrate(VI) pretreatment. Water Environ. Res. 2007, 79, 2420–2426. [Google Scholar] [CrossRef] [PubMed]
  50. Yang, T.; Sheng, L.; Wang, Y.; Wyckoff, K.N.; He, C.; He, Q. Characteristics of Cadmium Sorption by Heat-Activated Red Mud in Aqueous Solution. Sci. Rep. 2018, 8, 13558. [Google Scholar] [CrossRef] [Green Version]
  51. Abbasi, A.; Moradpour, T.; Van Hecke, K. A new 3D cobalt (II) metal–organic framework nanostructure for heavy metal adsorption. Inorg. Chim. Acta 2015, 430, 261–267. [Google Scholar] [CrossRef]
  52. Rahimi, E.; Mohaghegh, N. Removal of Toxic Metal Ions from Sungun Acid Rock Drainage Using Mordenite Zeolite, Graphene Nanosheets, and a Novel Metal–Organic Framework. Mine Water Environ. 2016, 35, 18–28. [Google Scholar] [CrossRef]
  53. Waritu, H.H.; Aregahegn, D.A.; Abdisa, C.M.; Minaleshewa, A. High Performance Copper Based Metal Organic Framework for Removal of Heavy Metals From Wastewater. Front. Mater. 2022, 9, 840806. [Google Scholar]
  54. Soltani, R.; Pelalak, R.; Pishnamazi, M.; Marjani, A.; Albadarin, A.B.; Sarkar, M.S.; Shirazian, S. A novel and facile green synthesis method to prepare LDH/MOF nanocomposite for removal of Cd(II) and Pb(II). Sci Rep. 2021, 11, 1609. [Google Scholar] [CrossRef] [PubMed]
  55. Visa, A.; Plesu, N.; Maranescu, B.; Ilia, G.; Borota, A.; Crisan, L. Combined Experimental and Theoretical Insights into the Corrosion Inhibition Activity on Carbon Steel Iron of Phosphonic Acids. Molecules 2021, 26, 135. [Google Scholar] [CrossRef] [PubMed]
  56. Visa, A.; Maranescu, B.; Bucur, A.; Iliescu, S.; Demadis, K. Synthesis and characterization of a novel phosphonate metal organic framework starting from copper salts. Phosphorus Sulfur Silicon Relat. Elem. 2014, 189, 630–639. [Google Scholar] [CrossRef]
  57. Visa, A.; Maranescu, B.; Lupa, L.; Crisan, L.; Borota, A. New Efficient Adsorbent Materials for the Removal of Cd(II) from Aqueous Solutions. Nanomaterials 2020, 10, 899. [Google Scholar] [CrossRef] [PubMed]
  58. Lin, J.; Wang, L. Comparison between linear and non-linear forms of pseudo-first-order and pseudo-second-order adsorption kinetic models for the removal of methylene blue by activated carbon. Front. Environ. Sci. Eng. China 2009, 3, 320–3244. [Google Scholar] [CrossRef]
  59. Kumar, K.V.; Porkodi, K.; Rocha, F. Comparison of various error functions in predicting the optimum isotherm by linear and non-linear regression analysis for the sorption of basic red 9 by activated carbon. J. Hazard. Mater. 2008, 150, 158–165. [Google Scholar] [CrossRef]
  60. Obayomi, K.S.; Bello, J.O.; Yahya, M.D.; Chukwunedum, E.; Adeoye, J.B. Statistical analyses on effective removal of cadmium and hexavalent chromium ions by multiwall carbon nanotubes (MWCNTs). Heliyon 2020, 6, e04174. [Google Scholar] [CrossRef]
  61. Tǎmaş, A.; Cozma, I.; Cocheci, L.; Lupa, L.; Rusu, G. Adsorption of orange II onto Zn2Al–layered double hydroxide prepared from zinc. Ash. Front. Chem. 2020, 8, 573535. [Google Scholar] [CrossRef]
  62. Azmi, S.N.H.; Al Lawati, W.M.; Al Hoqani, U.H.A.; Al Aufi, E.; Al Hatmi, K.; Al Zadjali, J.S.; Rahman, N.; Nasir, M.; Rahman, H.; Khan, S.A. Development of a Citric-Acid-Modified Cellulose Adsorbent Derived from Moringa peregrina Leaf for Adsorptive Removal of Citalopram HBr in Aqueous Solutions. Pharmaceuticals 2022, 15, 760. [Google Scholar] [CrossRef]
Figure 1. SEM images and EDX of the synthesized adsorbent materials (a) CaCEPPA; (b) CoCEPPA.
Figure 1. SEM images and EDX of the synthesized adsorbent materials (a) CaCEPPA; (b) CoCEPPA.
Ijms 23 12611 g001
Figure 2. XRD patterns for (a) CaCEPPA; (b) CoCEPPA before and after adsorption of Cd2+ and Pb2+.
Figure 2. XRD patterns for (a) CaCEPPA; (b) CoCEPPA before and after adsorption of Cd2+ and Pb2+.
Ijms 23 12611 g002
Figure 3. FT-IR spectra of CaCEPPA (a) and CoCEPPA (b).
Figure 3. FT-IR spectra of CaCEPPA (a) and CoCEPPA (b).
Ijms 23 12611 g003
Figure 4. The influence of pH on the adsorption capacity of CaCEPPA and CoCEPPA in the removal process of cadmium and lead from aqueous solutions.
Figure 4. The influence of pH on the adsorption capacity of CaCEPPA and CoCEPPA in the removal process of cadmium and lead from aqueous solutions.
Ijms 23 12611 g004
Figure 5. Kinetics of Cd2+ and Pb2+ adsorption onto CaCEPPA.
Figure 5. Kinetics of Cd2+ and Pb2+ adsorption onto CaCEPPA.
Ijms 23 12611 g005
Figure 6. Kinetics of Cd2+ and Pb2+ adsorption onto CoCEPPA.
Figure 6. Kinetics of Cd2+ and Pb2+ adsorption onto CoCEPPA.
Ijms 23 12611 g006
Figure 7. Equilibrium isotherms of Cd2+ and Pb2+ adsorption onto CaCEPPA.
Figure 7. Equilibrium isotherms of Cd2+ and Pb2+ adsorption onto CaCEPPA.
Ijms 23 12611 g007
Figure 8. Equilibrium isotherms of Cd2+ and Pb2+ adsorption onto CoCEPPA.
Figure 8. Equilibrium isotherms of Cd2+ and Pb2+ adsorption onto CoCEPPA.
Ijms 23 12611 g008
Figure 9. Packing of the CaCEPPA (a) and CoCEPPA (b).
Figure 9. Packing of the CaCEPPA (a) and CoCEPPA (b).
Ijms 23 12611 g009
Table 1. Pseudo-first-order and pseudo-second-order kinetic parameters for Cd2+ and Pb2+ adsorption from aqueous solutions onto CaCEPPA and CoCEPPA materials.
Table 1. Pseudo-first-order and pseudo-second-order kinetic parameters for Cd2+ and Pb2+ adsorption from aqueous solutions onto CaCEPPA and CoCEPPA materials.
Kinetic ModelParametersCaCEPPACoCEPPA
CdPbCdPb
qe exp, mg/g15.1411.8521.0817.23
Pseudo-first-orderqe calc, mg/g7.155.5458.569.44
k1, min−10.04240.03440.0240.0342
R20.96740.89320.78400.8573
ERRSQ353221970331
χ299.584.731874.7
ARE, %228253458183
Pseudo-second-orderqe calc, mg/g15.6712.421.9218.25
k2, g mg−1 min−10.01490.01530.01040.0079
R20.99750.99590.99680.9928
ERRSQ4.213.0910.46.39
χ20.3890.450.7270.59
ARE, %5.817.317.148.26
Table 2. Langmuir, Freundlich, Temkin, and Dubini–Raduschkeich isotherm parameters for Cd2+ and Pb2+ adsorption from aqueous solutions onto CaCEPPA and CoCEPPA materials.
Table 2. Langmuir, Freundlich, Temkin, and Dubini–Raduschkeich isotherm parameters for Cd2+ and Pb2+ adsorption from aqueous solutions onto CaCEPPA and CoCEPPA materials.
Isotherm ModelParametersCaCEPPACoCEPPA
CdPbCdPb
qm exp, mg/g32.621.451.632.7
Langmuirqm calc, mg/g36.523.5254.935.8
kL, L/mg0.05560.04830.08860.0739
R20.99640.99680.99090.9968
ERRSQ10.62.2927.310.6
χ20.460.212.10.54
ARE, %5.35.9918.15.78
FreundlichkF, (mg/g) (L/mg)n3.1062.1886.653.509
1/n0.51830.48050.45900.5067
R20.94750.94480.96480.9193
ERRSQ16538.5253259
χ24.862.385.238.01
ARE, %17.315.314.622.3
Temkinb, J/mol353563275352
KT, L/g0.7680.7131.880.44
R20.97000.99310.98010.9628
ERRSQ22.91.8336.029.9
χ21.820.4227.591.85
ARE, %18.59.3642.616.3
Dubinin–Raduschkevichqm, mg/g21.614.131.022.3
KDR, mol2/J22 × 10−62 × 10−64 × 10−71 × 10−6
R20.79850.77520.77850.7630
ERRSQ26995.3746324
χ214.87.5825.916.5
ARE, %43.824.637.734.0
Table 3. Comparison of the maximum adsorption capacities given/demonstrated by similar adsorbents in the treatment process of aqueous solutions containing Cd2+ ions.
Table 3. Comparison of the maximum adsorption capacities given/demonstrated by similar adsorbents in the treatment process of aqueous solutions containing Cd2+ ions.
Adsorbentqm (mg g−1)References
Functionalized Zr-MOFs41.32[10]
UiO-66-NHC(S)NHMe49[51]
TMU-543[52]
Cu-DPA MOF1.334[53]
LDH/MOF NC415.3[54]
CaCEPPA32.6This paper
CoCEPPA51.6
Table 4. Comparison of the maximum adsorption capacities given/demonstrated by similar adsorbents in the treatment process of aqueous solutions containing Pb2+ ions.
Table 4. Comparison of the maximum adsorption capacities given/demonstrated by similar adsorbents in the treatment process of aqueous solutions containing Pb2+ ions.
Adsorbentqm (mg g−1)References
Functionalized Zr-MOFs50.51[10]
UiO-66-NHC(S)NHMe232[51]
TMU-5251[52]
Cu-DPA MOF2.19[53]
LDH/MOF NC301.4[54]
CaCEPPA21.4This paper
CoCEPPA32.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Marganovici, M.; Maranescu, B.; Visa, A.; Lupa, L.; Hulka, I.; Chiriac, V.; Ilia, G. Hybrid Coordination Networks for Removal of Pollutants from Wastewater. Int. J. Mol. Sci. 2022, 23, 12611. https://doi.org/10.3390/ijms232012611

AMA Style

Marganovici M, Maranescu B, Visa A, Lupa L, Hulka I, Chiriac V, Ilia G. Hybrid Coordination Networks for Removal of Pollutants from Wastewater. International Journal of Molecular Sciences. 2022; 23(20):12611. https://doi.org/10.3390/ijms232012611

Chicago/Turabian Style

Marganovici, Marko, Bianca Maranescu, Aurelia Visa, Lavinia Lupa, Iosif Hulka, Vlad Chiriac, and Gheorghe Ilia. 2022. "Hybrid Coordination Networks for Removal of Pollutants from Wastewater" International Journal of Molecular Sciences 23, no. 20: 12611. https://doi.org/10.3390/ijms232012611

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