Next Article in Journal
Methylcellulose Bionanocomposite Films Incorporated with Zein Nanoparticles Containing Propolis and Curcumin for Functional Packaging
Previous Article in Journal
Gallic Acid Functionalization Improves the Pharmacological Profile of Fucoidan B: A Polysaccharide with Antioxidant Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of Pharmaceutical Compounds from Water on Chitosan/Glutaraldehyde Hydrogels: Theoretical and Experimental Analysis

by
Billy Alberto Ávila Camacho
1,
Miguel Andrés Rojas Pabón
1,
Norma Aurea Rangel Vázquez
1,*,
Edgar A. Márquez Brazón
2,
Hilda Elizabeth Reynel Ávila
3,
Didilia Ileana Mendoza Castillo
3 and
Yectli A. Huerta
4
1
TecNM/Instituto Tecnológico de Aguascalientes, Avenida Adolfo López Mateos 1801, Aguascalientes 20256, Mexico
2
Grupo de Investigaciones en Química y Biología, Departamento de Química y Biología, Facultad de Ciencias Básicas, Universidad del Norte, Carrera 51B, Km 5, vía Puerto Colombia, Barranquilla 081007, Colombia
3
Investigadoras e Investigadores por México, SECIHTI (Secretaria de Ciencia, Humanidades, Tecnologia e Innovacion), Ciudad de México 03940, Mexico
4
Minnesota Supercomputing Institute, University of Minnesota, Minneapolis, MN 55455, USA
*
Author to whom correspondence should be addressed.
Polysaccharides 2025, 6(4), 90; https://doi.org/10.3390/polysaccharides6040090
Submission received: 13 January 2025 / Revised: 1 April 2025 / Accepted: 30 September 2025 / Published: 9 October 2025

Abstract

Chitosan-based hydrogels are used in the adsorption of pharmaceutical compounds from water. The adsorption process of diclofenac and naproxen on chitosan hydrogels cross-linked with glutaraldehyde has been studied theoretically and experimentally. According to the thermodynamic properties, the adsorption processes were spontaneous and endothermic, due to the negative values of Gibbs free energy, and the enthalpies of formation were positive. Furthermore, the different systems were studied by electrostatic potential maps, where the functional groups (amino and hydroxyl) represented the active sites of the hydrogel. The maximum adsorption capacity obtained for diclofenac and naproxen was 108.85 and 97.22 mg/g, respectively, at a temperature of 308.15 K. On the other hand, the adsorbent was characterized by FTIR (Fourier Transform Infrared Spectroscopy) and XRD (X-ray Diffraction) before and after the adsorption of the drugs to confirm the binding of the adsorbates on the surface of the material.

1. Introduction

Water represents an important factor for human, animal, and plant life. However, factors such as industrial, agricultural, livestock, urban development, and climate change generate greater pollution in rivers, lakes, lagoons, or oceans. Around 80% of wastewater from industry and municipalities cause damage to the environment and human life [1,2,3]. There are different techniques for removing contaminants from water, but adsorption is a widely used physical method for monitoring and separating environmental contaminants, where the contaminant is adsorbed onto the porous surface of an adsorbent material [4,5]. The adsorption processes depend on factors such as the structures of the hydrogels and the nature of the contaminants [6,7]. The adsorption process allows the analysis of various contaminants such as drugs, nitrates, phosphates, plastics, and fertilizers [4,5].
Hydrogels are three-dimensional cross-linked structures insoluble in polar solvents such as water, with properties of biodegradability, biocompatibility, and responsiveness. There are two types of hydrogels: (a) physical, which are characterized by hydrogen bonding, ion exchange, and hydrophobic interactions, and (b) chemical, with covalent bonds. In addition, the functional groups present in hydrogels increase the swelling capacity up to 400 times their original weight [8]. Hydrogel synthesis uses synthetic or natural polymers. However, hydrogels based on natural polymers such as chitosan and dextran have better biological and chemical properties [9,10].
Chitosan is a natural biopolymer—cationic, biodegradable, and biocompatible [11,12]. Due to its properties, chitosan is widely used in the adsorption processes of heavy metals, pesticides, pharmaceuticals, chemical compounds, dyes, etc. [13,14] Chitosan has many functional groups that can interact with pharmaceutical compounds in an adsorption process. Therefore, its adsorption capacities are significantly higher compared to other polysaccharides such as cellulose and starch, because these two polysaccharides have fewer reactive functional groups and lower mechanical strength, resulting in lower adsorption capacities [15,16,17,18].
On the other hand, pharmaceutical compounds are structures with specific functionalities and physicochemical and biological characteristics; they are polar and have molecular weights between 200 and 1000 Da [19,20]. Pharmaceutical compounds are found in the water due to various situations. Pharmaceutical compounds are observed in wastewater because treatment plants do not have the capacity to remove them from effluents [21,22]. On the other hand, rainwater, farms, yard waste, and septic tanks contribute to the presence of pharmaceutical compounds (see Table 1) [23,24].
In recent years, different technologies related to the removal of pharmaceutical contaminants have been evaluated, including biological methods [31,32], advanced oxidation [33,34], photodegradation [35,36], ozonation [37], adsorption [38,39], and others.
Currently, the adsorption process is of greater importance in the removal of water contaminants because it is an economical process, allows on-line operation, produces no sludge, and enables reusability [40,41]. The use of various adsorbents derived from biopolymers has received great attention [42,43]. Chitosan has been used for the adsorption of pharmaceutical products (Figure 1, Figure S1 and S2) [44,45,46,47].
Computational modeling involves the use of computers to analyze molecules or macromolecules using mathematics, physics, chemistry, and computer science to obtain different energies, thermodynamic and QSAR properties, HOMO/LUMO orbitals, FTIR, and electronic distribution, using semi-empirical methods such as AM1 to obtain new materials [15]. QSAR properties are based on molecular and mathematical descriptors to predict the relationship between biological activity and chemical structure of the molecule or macromolecule. In addition, hybrid methods of molecular mechanics and quantum mechanics determine the binding affinity and capacity of a molecule [16,48].
As can be seen, recent studies have demonstrated the efficacy of chitosan-based hydrogels for removing pharmaceutical pollutants from water. However, most prior work has been limited to single-compound adsorption experiments and conventional isotherm/kinetic analyses, providing little insight into molecular-level interaction mechanisms or multi-contaminant scenarios. The study of multicomponent systems is crucial to subsequently compare with realistic wastewater systems. On the other hand, computational approaches (e.g., QSAR modeling or molecular simulations) are seldom integrated into these adsorption studies, leaving a knowledge gap in understanding how pharmaceutical molecules bind to the adsorbent and how adsorbent structure influences performance. In this context, the present study offers a novel integrated experimental–theoretical approach for diclofenac and naproxen uptake on a glutaraldehyde-cross-linked chitosan hydrogel.
This work combines advanced computational chemistry tools—including geometry optimizations, QSAR property analysis, electrostatic potential mapping (EPM), and Monte Carlo simulations—with traditional batch adsorption experiments and sample characterization. This methodology provides a novel insight into the adsorption mechanism.

2. Materials and Methods

2.1. Computational Section

2.1.1. Geometry Optimization

Using HyperChem Professional 8.0v software, each molecular structure was drawn. Initially, the atoms of each molecule were drawn, and in the “Build” menu, the “Add H and Model Build” option was selected to complete the hydrogen atoms and spatially organize each molecule. In this study, a hybrid method between molecular mechanics (AMBER) and quantum mechanics (AM1) was used to geometrically optimize the analyzed molecular structures (naproxen and diclofenac showed in Figure 2) and to describe the potential energy of each system. The main objective of using this hybridization was to reduce the computational time of each molecule to calculate its minimum energy. For the AMBER method, the Polak–Ribiere algorithm was used with 450 iterations for each drug. For chitosan, the iterations were 3015. Then, the calculation was performed with the semi-empirical AM1 method without restrictions, using the conjugate gradient method and 450 iterations for the drugs and 3015 for chitosan.

2.1.2. Determination of QSAR and Thermodynamic Properties

In the “Compute” menu, the “QSAR Properties” option was selected, and the partition coefficient (LogP) was determined. Thermodynamic properties were obtained from the “Properties” option in the “Compute” tool, which was performed for each molecule.

2.1.3. Electrostatic Potential Map (EPM)

To determine the EPM, the option “Plot Molecular Graphs” was selected in the “Compute” menu.

2.1.4. Adsorption Process

Hyperchem Professional 8.0 was used to analyze the interactions between each nonsteroidal anti-inflammatory compound and glutaraldehyde cross-linked chitosan (Figure 3). Then, in the “Build” menu, the “Add H and Model Build” option was used to complete the hydrogen atoms.

2.2. Experimental Studies

2.2.1. Reagents

Chitosan (CAS-No: 9012-76-4) from shrimp shells, the pharmaceuticals naproxen and diclofenac with a purity of 98%, and the glutaraldehyde (25%) were obtained by Sigma-Aldrich (Sigma-Aldrich Química S. de RL. de C.V, Toluca, 50200, Mexico). Acetic acid (99.7%) from J. T. Baker and deionized water (DI water) were supplied by MERCK (MERCK Mexico, Naucalpan de Juárez, 53500, Mexico).

2.2.2. Hydrogel Synthesis

A total of 1.5 g of chitosan was dissolved in a 3% acetic acid solution. The chitosan colloid was magnetically stirred at 80 rpm for 180 min at a temperature of 50 °C.
Subsequently, a 10% glutaraldehyde solution was prepared and slowly added to the chitosan solution, this was mixed for another 2 h. A viscous solution with a pale-yellow color was obtained; this was rapidly centrifuged at 6000 rpm for 6 min to sediment the particles that remained suspended in the solution, and finally it was placed on Petri dishes to drain at 25 °C for 72 h to obtain chitosan films cross-linked with glutaraldehyde.

2.3. Characterization of Hydrogel

2.3.1. FTIR

The synthesized hydrogel was analyzed by Fourier transform infrared spectroscopy (FTIR) using a Nicolet iS10 infrared spectrophotometer from Thermo Scientific (Waltham, MA, USA). For the analysis, 32 scans were performed, and the spectra were collected in a wavenumber range of 4000–400 cm−1.

2.3.2. XRD

XRD analysis was carried out using an Empyrean diffractometer from Malvern-Panalytical and PIXel ID detectors (Malvern Panalytical, 66269, N.L, Mexico). The analysis was performed at 45 kV and 40 mA.

2.4. Adsorption Studies of Pharmaceuticals

A 500 mL of solution was prepared with 250 mg/L of each pharmaceutical (diclofenac and naproxen) at pH 6. The adsorbate concentration varied from 20 to 250 mg/L and the hydrogel mass remained fixed at 0.02 g. The experiments were performed in batches using a volume of 10 mL at temperatures of 298.15K and 308.15 K, with stirring at 120 rpm for 24 h. After 24 h, the adsorbent was separated from the pharmaceutical solutions using filters, and samples were subsequently quantified by HPLC. To obtain the adsorption capacities, Equation (1) was used.
q e = C o C e V m
where “ q e ” is the adsorption capacity, “Co” and “Ce” are the initial and equilibrium concentrations of the pharmaceutical solutions, respectively, “V” is the volume used, and “m” is the mass of the adsorbent.

3. Results

3.1. Theoretical Studies

The energetic properties of diclofenac, naproxen, and the adsorption systems yielded negative results for the hydrogel–drug systems (see Table 2), that is, the adsorption systems are spontaneous and endothermic.
The dipole moment is a measure of the separation of opposite electrical charges, so it is possible to find the nature (polar, nonpolar, or ionic) of any bond. In this case, glutaraldehyde is considered a polar molecule, meaning that the charge distribution in both pharmaceuticals and adsorption systems is asymmetric. Therefore, the electronic distribution will be oriented toward the most electronegative atom. This is because the value of this parameter was greater than 0 in all cases. The dipole moment is very important since it predicts an improvement in polarity and affinity in the union of the bonds of a molecule. A higher dipole moment reveals a greater hydrogen bond interaction and higher binding affinity [49].
The partition coefficient (Log P) is a parameter that determines whether a molecule has affinity with water (hydrophilic) or, conversely, repels water (hydrophobic). In addition, it indicates whether the studied molecule can be dissolved in polar solvents. All the studied systems presented negative Log P partition coefficient values (−0.21, −0.56, −14.34, and −12.86 for diclofenac, naproxen, hydrogel–diclofenac, and hydrogel–naproxen, respectively) (see Table 2), indicating a hydrophilic character. This negative value of the partition coefficient also shows that diclofenac and naproxen can be dissolved in polar solvents.
For diclofenac, the EPM obtained ranged from −0.085 to +0.328, with a negative electron-rich region marked in red (see Figure 4a), where the oxygens are located, which has the possibility of receiving an electrophilic attack due to its high electronegativity and the sp2 hybridization in one of the oxygens of the carboxyl group. The green color showed a region of medium reactivity where the chlorines are attached to their respective carbon. The hydrogens are found in the positive region represented by blue, indicating the possibility of a nucleophilic attack due to the lower presence of electrons.
The electrostatic potential map (EPM) allows determination of the reactive zones that may exhibit electrophilic and nucleophilic charge in the calculated geometry. For the naproxen molecule, the EPM has a range from −0.138 to +0.472 (see Figure 4b). The regions with negative partial charge (red color) are located on the oxygen atoms of the carboxyl and ether groups, these are areas that can receive electrophilic attacks, while the region of positive partial charge was located on the hydrogen atom of the carboxyl group. Some hydrogen atoms from the methyl and aromatic groups were observed for nucleophilic attacks.
The electrostatic potential map of the interaction between the hydrogel and diclofenac had a range from −0.064 to +1.041 (see Figure 5). The amino and hydroxyl groups of the hydrogel are negatively charged (red areas) and can be exposed to electrophilic attacks. The neutral region (green area) was in the hydrogen atoms bonded to carbon along the chain. Finally, positive regions (blue and translucent areas) were observed in the carbon–hydrogen and nitrogen–hydrogen bonds, where nucleophilic attacks can occur. The diclofenac molecule presented a low electron density prone to electrophilic attacks (red areas) in the carbons and hydrogens of the aromatic group and the oxygens of the carboxyl group. Furthermore, neutral charge (green area) was observed on chlorine atoms, and a positive charge (blue areas) was found on hydrogen–carbon bonds, which are prone to nucleophilic attacks. The electrostatic potential map determined the possible reaction site for adsorption, where the carboxyl group of the diclofenac form hydrogen bonds with an amino group of the hydrogel.
Figure 6 shows the potential map of the interaction between hydrogel and naproxen. A high electron density was generally observed in the hydrogel in the oxygens and nitrogen of the amines (red regions); these areas are susceptible to electrophilic attacks. Likewise, a slightly positive zone was shown in the external hydrogens (blue region) with the possibility of receiving nucleophilic attacks. For naproxen, there is a negatively charged zone (red region) along its carbons and oxygens prone to receiving electrophilic attacks. Positive charges were found in its hydrogens (blue regions), capable of receiving nucleophilic attacks.
A hydrogen bond can be formed with the sp2-hybridizied oxygen of the carboxyl group of naproxen and a hydrogen bonded to an oxygen atom (OH) of the hydrogel.

3.2. Monte Carlo Simulation at 308.15 K

3.2.1. Hydrogel–Diclofenac

Table 3 shows the energetics and QSAR properties obtained by the Monte Carlo method. The Gibbs free energy was constant with the change in temperature, which means that the adsorption occurred stably and spontaneously. Moreover, the increase in temperature favors the dipole moment and promotes adsorption by increasing the intermolecular forces. The dipole moment is a measure of the distribution of electrical charges in a molecule and is related to polarity. The increase in the dipole moment can be related to the influence of temperature on the distribution of electrical charges. At higher temperatures, molecules tend to have greater thermal agitation, which can lead to changes in conformations and charge distribution. That is, at 308.15 K, a conformational change in the molecule or a redistribution of charges occurred, resulting in this notable increase in the dipole moment. This change was influenced by interactions between the different parts of the molecules, favoring their attraction, as well as by the intermolecular forces present in the environment.
Volume is a physical property related to the space occupied by a molecule, and surface area is related to the amount of space a molecule occupies on a surface. Both properties are affected by several factors, such as intermolecular interactions and temperature. The observed changes in the system’s volume and surface area could be due to interactions between diclofenac and hydrogel. On the other hand, OH stretching was observed between 3637 and 3382 cm−1 (see Table 4).
In addition, the carboxyl (diclofenac) and the carbonyl group (hydrogel) presented close stretching vibrations at the analyzed temperatures, suggesting the possibility that these groups are favorable sites for obtaining hydrogen bonds during the adsorption process. However, a shift was observed in the vibrations of the carbonyl group of the hydrogel due to the interaction of the adsorbed compounds and the hydrogel, which increased with temperature.

3.2.2. Hydrogel–Naproxen

The energetic characteristics at several temperatures for the removal of naproxen by the hydrogel (see Table 5) showed stability in Gibbs free energy, indicating a spontaneous and stable process that can occur in nature. Naproxen presented higher dipole moment values compared to diclofenac, due to its interaction and affinity with the hydrogel during the adsorption process.
Table 6 shows the main FTIR signals of the hydrogel/naproxen adsorption at 308.15 K, where peaks between 3489 and 3288 cm−1 corresponded to the OH stretching. Carboxyl and carbonyl groups were observed between 2164 and 2077 cm−1 and between 1593 and 1583 cm−1, which allowed physical adsorption by hydrogen bonding. The shift in the carboxyl group bands at 308.15 K was attributed to the charge distribution of the molecule, which generated higher energy and increased its vibration.

3.3. Experimental Analysis

Prior to adsorption experiments, glutaraldehyde cross-linked chitosan hydrogel was analyzed by FTIR and XRD. Figure 7 shows the IR of hydrogel before the drug adsorption experiments. The broad band located between 3450 and 3200 cm−1 was assigned to the O-H bond; the low intensity peaks located at 2950 and 2850 correspond to C-H (chitosan chain), and the peak at 1550 cm−1 corresponded to the amino groups (NH). The band at 1400 cm−1 is attributed to C-N bond of the hydrogel, and that at 1100–1000 cm−1 corresponded to the glycosidic zones (chitosan) [50,51].
The XRD analysis before the adsorption experiments showed the following. The hydrogel presents two important zones, at 10°, which were assigned to the amorphous zone of the adsorbent, and at 20°, which corresponded to the crystalline zone, assigned to the inter- and intramolecular of the hydroxyl and amine bonds of the polymer chain (see Figure 8).

3.3.1. Adsorption Isotherms in a Single System

The isotherms of the individual experiments (Figure 9) indicated a maximum adsorption of 108.85 and 94.37 mg/g for diclofenac, and 97.22 and 82.90 mg/g for naproxen at 308.15 and 298.15 K, respectively. The results determined an endothermic process, that is, the removal of naproxen and diclofenac depends on temperature [52,53].

3.3.2. Adsorption Isotherms of Naproxen and Diclofenac in a Binary System

The isotherms indicated that the maximum adsorption was 96.29 and 83.78 mg/g for diclofenac and naproxen, respectively, at 308.15 K. On the other hand, the maximum capacities obtained at 298.15 K were 78.78 and 72.11 mg/g for diclofenac and naproxen, respectively (Figure 10). The pharmaceutical removal followed the trend diclofenac > naproxen under the indicated operating conditions. The adsorption capacities decreased slightly with respect to the single systems, which was attributed to competition for the amine and carboxyl groups.

3.3.3. FTIR After Adsorption Experiments

The spectra of the adsorbent after adsorption are observed in Figure 11b–d for naproxen, diclofenac, and the combination of both, respectively. Specifically, the area of 3450–3200 cm−1 was assigned to the addition of naproxen/diclofenac to the hydrogel surface. Similarly, the peaks at 2950–2850 cm−1 increased in intensity due to the conjugated double bonds of both pharmaceuticals. The bands between 1650 and 1750 cm−1 corresponded to both diclofenac and naproxen, respectively [54].

3.3.4. XRD After Adsorption Experiments

Figure 12 shows a decrease in the signals of both regions of the hydrogel, which may be associated with the ionic interaction between the hydrogel groups and the COOH groups of the drugs [54].

4. Conclusions

Based on the results obtained, the hydrogel represents an excellent option for the adsorption of drugs. The adsorption capacities are significant, considering that in a real effluent, the concentration of contaminants can be lower, which ensures the complete removal of drugs from water. The characterization results confirm the presence of adsorbates on the hydrogel. In addition, different functional groups, mainly hydroxyl (OH) and amine (NH), contribute to the removal of drugs. The competition of drugs over the hydrogel in binary solutions is mainly due to the similarity between diclofenac and naproxen molecules, since they both contain a carboxyl group (COOH) and conjugated double-bond rings, which also interact with the hydrogel. Finally, these results demonstrate the potential of this hydrogel not only for drug removal but also in the removal of other types of contaminants.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polysaccharides6040090/s1, Figure S1: Adsorption process of diclofenac and Figure S2: Adsorption process of naproxen.

Author Contributions

Conceptualization—B.A.Á.C., M.A.R.P. and N.A.R.V.; Methodology—B.A.Á.C., M.A.R.P. and N.A.R.V.; Software—B.A.Á.C. and M.A.R.P.; Validation—B.A.Á.C., M.A.R.P. and N.A.R.V.; Formal Analysis—B.A.Á.C., M.A.R.P., N.A.R.V., E.A.M.B., H.E.R.Á. and D.I.M.C.; Investigation—B.A.Á.C., M.A.R.P. and N.A.R.V.; Resources—B.A.Á.C., M.A.R.P. and N.A.R.V.; Data Curation—B.A.Á.C., M.A.R.P. and N.A.R.V.; Writing—Original Draft—B.A.Á.C., M.A.R.P., H.E.R.Á. and D.I.M.C.; Writing—Review and Editing—E.A.M.B. and Y.A.H.; Visualization—N.A.R.V., E.A.M.B. and Y.A.H.; Supervision—N.A.R.V., H.E.R.Á. and D.I.M.C.; Project Administration—N.A.R.V., E.A.M.B. and Y.A.H.; Funding Acquisition—B.A.Á.C. and N.A.R.V. All authors have read and agreed to the published version of the manuscript.

Funding

To TecNM for the funding granted for the development of project 19315.24-P, of the 2024 Call: SCIENTIFIC RESEARCH, TECHNOLOGICAL DEVELOPMENT AND INNOVATION PROJECTS.

Data Availability Statement

The original contributions presented in this study are included in the article/supplementary material. Further inquiries can be directed to the corresponding author.

Acknowledgments

Billy Alberto Ávila-Camacho would like to thank the SECIHTI for the Doctorate in Engineering Science 834799 scholarship.To TecNM for the funding granted for the development of project 19315.24-P, of the 2024 Call: SCIENTIFIC RESEARCH, TECHNOLOGICAL DEVELOPMENT AND INNOVATION PROJECTS.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lin, L.; Yang, H.; Xu, X. Effects of water pollution on human health and disease heterogeneity: A Review. Front. Environ. Sci. 2022, 10, 880246. [Google Scholar] [CrossRef]
  2. De Souza, S.A.; de Souza, A. Invisible pollutants: Environmental, economic and social impacts as threats to water quality. Rev. Juríd. 2019, 16, 95–107. [Google Scholar] [CrossRef]
  3. Ariho, A.; Aja, L.; Muhammad, T.; Mohammad, L. Water pollution: Causes, impacts and current efforts to address the issues of water pollution along river Meizimera-kihihi, Kanugu District, Uganda. F1000Reserch 2024, 13, 1298. [Google Scholar] [CrossRef]
  4. Singh, N.B.; Nagpal, G.; Rachna, S.A. Water purification by using Adsorbents: A review. Environ. Technol. Innov. 2018, 11, 187–240. [Google Scholar] [CrossRef]
  5. Rathi, B.S.; Kumar, P.S. Application of adsorption process for effective removal of emerging contaminants from water and wastewater. Environ. Pollut. 2021, 280, 116995. [Google Scholar] [CrossRef]
  6. Sabzehmeidani, M.M.; Mahnaee, S.; Ghaedi, M.; Heidari, H.; Roy, V.A.L. Carbon based materials: A review of adsorbents for inorganic and organic compounds. Mater. Adv. 2021, 2, 598–627. [Google Scholar] [CrossRef]
  7. Ghosh, S.; Falyouna, O.; Malloum, A.; Othmani, A.; Bedair, H.; Onyeaka, H.; Al-sharify, Z.T.; Oluwaseun, A.; Miri, T.; Osagie, C.; et al. A general review on the use of advance oxidation and adsorption processes for the removal of furfural from industrial effluents. Micropor. Mesopor. Mater. 2022, 331, 2021. [Google Scholar] [CrossRef]
  8. Far, F.; Omrani, M.; Jamal, M.N.; Javanshir, S. Multi-responsive chitosan-based hydrogels for controlled release of vincristine. Commun. Chem. 2023, 6, 28. [Google Scholar]
  9. Hidaka, M.; Kojima, M.; Sakai, S.; Delattre, C. Characterization of chitosan hydrogels obtained through phenol and tripolyphosphate anionic crosslinking. Polymers 2024, 16, 1274. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, F.; Wang, L.; Zhai, X.; Ji, S.; Ye, J.; Zhu, Z.; Teng, C.; Dong, W.; Wei, W. A multi-functional double cross-linked chitosan hydrogel with tunable mechanical and antibacterial properties for skin wound dressing. Carbohyd. Polym. 2023, 322, 121344. [Google Scholar] [CrossRef] [PubMed]
  11. Nilsen-Nygaard, J.; Strand, S.P.; Vårum, K.M.; Draget, K.I.; Nordgård, C.T. Chitosan: Gels and interfacial properties. Polymers 2015, 7, 552–579. [Google Scholar] [CrossRef]
  12. Ahmadi, F.; Oveisi, Z.; Samani, S.M.; Amoozgar, Z. Chitosan based hydrogels: Characteristics and pharmaceutical applications. Res. Pharmaceut. Sci. 2015, 10, 1–16. [Google Scholar]
  13. Ávila-Camacho, B.A.; Rangel-Vázquez, N.A. Analysis of the adsorption of Hg2+, Ni2+ and Cu2+ on chitosan hydrogels. Polimeros 2024, 34, e20240035. [Google Scholar] [CrossRef]
  14. Oliveira-Gonçalves, J.; Strieder, M.M.; Oliveira-Silva, L.F.; dos Reis, G.S.; Dotto, G.L. Advanced technologies in water treatment: Chitosan and its modifications as effective agents in the adsorption of contaminants. Int. J. Biol. Macromol. 2024, 270, 132307. [Google Scholar] [CrossRef]
  15. Grisales-Cifuentes, C.M.; Galvis, E.A.S.; Porras, J.; Flórez, E.; Torres-Palma, R.A.; Acelas, N. Kinetics, isotherms, effect of structure, and computational analysis during the removal of three representative pharmaceuticals from water by adsorption using a biochar obtained from oil palm fiber. Bioresour. Technol. 2021, 326, 124753. [Google Scholar] [CrossRef] [PubMed]
  16. Can, C.; Jianlong, W. Correlating metal ionic characteristics with biosorption capacity using QSAR model. Chemosphere 2007, 69, 1610–1616. [Google Scholar] [CrossRef]
  17. Beppu, M.M.; Vieira, R.S.; Aimoli, C.G.; Santana, C.C. Crosslinking of chitosan membranes using glutaraldehyde: Effect on ion permeability and water absorption. J. Membr. Sci. 2007, 301, 126–130. [Google Scholar] [CrossRef]
  18. Arianita, A.; Amalia, B.; Pudjiastuti, W.; Melanie, S.; Fauzia, V.; Imawan, C. Effect of glutaraldehyde to the mechanical properties of chitosan/nanocellulose. J. Phys. Conf. Ser. 2019, 1317, 012045. [Google Scholar] [CrossRef]
  19. Kümmerer, K. The presence of pharmaceuticals in the environment due to human use–present knowledge and future challenges. J. Environ. Manag. 2009, 90, 2354–2366. [Google Scholar] [CrossRef]
  20. Boxall, A.B. Pharmaceuticals in the environment and human health. In Health Care Environ Contamination, 1st ed.; Boxall, A.B.A., Kookana, R.S., Eds.; Elsevier: London, UK, 2018; Volume 1, pp. 123–136. [Google Scholar]
  21. Phillips, P.; Chalmers, A. Wastewater effluent, combined sewer overflows, and other sources of organic compounds to Lake Champlain 1. J. Am. Water Resour. Assoc. 2009, 45, 45–57. [Google Scholar] [CrossRef]
  22. Schwarzenbach, R.P.; Escher, B.I.; Fenner, K.; Hofstetter, T.B.; Johnson, C.A.; Von Gunten, U.; Wehrli, B. The challenge of micropollutants in aquatic systems. Science 2006, 313, 1072–1077. [Google Scholar] [CrossRef]
  23. Bavumiragira, J.P.; Yin, H. Fate and transport of pharmaceuticals in water systems: A processes review. Sci. Total Environ. 2022, 823, 153635. [Google Scholar] [CrossRef]
  24. Watkinson, A.J.; Murby, E.J.; Kolpin, D.W.; Costanzo, S.D. The occurrence of antibiotics in an urban watershed: From wastewater to drinking water. Sci. Total Environ. 2009, 407, 2711–2723. [Google Scholar] [CrossRef]
  25. Lin, A.Y.C.; Yu, T.H.; Lin, C.F. Pharmaceutical contamination in residential, industrial, and agricultural waste streams: Risk to aqueous environments in Taiwan. Chemosphere 2008, 74, 131–141. [Google Scholar] [CrossRef]
  26. Kim, S.D.; Cho, J.; Kim, I.S.; Vanderford, B.J.; Snyder, S.A. Occurrence and removal of pharmaceuticals and endocrine disruptors in South Korean surface, drinking, and waste waters. Water Res. 2007, 41, 1013–1021. [Google Scholar] [CrossRef] [PubMed]
  27. Spongberg, A.L.; Witter, J.D. Pharmaceutical compounds in the wastewater process stream in northwest Ohio. Sci. Total Environ. 2008, 397, 148–157. [Google Scholar] [CrossRef]
  28. Yoon, Y.; Ryu, J.; Oh, J.; Choi, B.G.; Snyder, S.A. Occurrence of endocrine disrupting compounds, pharmaceuticals, and personal care products in the Han River (Seoul, South Korea). Sci. Total Environ. 2010, 408, 636–643. [Google Scholar] [CrossRef] [PubMed]
  29. Tyumina, E.; Subbotina, M.; Polygalov, M.; Tyan, S.; Ivshina, I. Ketoprofen as an emerging contaminant: Occurrence, ecotoxicity and (bio) removal. Front. Microbiol. 2023, 14, 1200108. [Google Scholar] [CrossRef] [PubMed]
  30. Chaves, M.J.S.; Barbosa, S.C.; Primel, E.G. Emerging contaminants in Brazilian aquatic environment: Identifying targets of potential concern based on occurrence and ecological risk. Environ. Sci. Pollut. Res. Int. 2021, 28, 67528–67543. [Google Scholar] [CrossRef]
  31. Tiwari, B.; Sellamuthu, B.; Ouarda, Y.; Drogui, P.; Tyagi, R.D.; Buelna, G. Review on fate and mechanism of removal of pharmaceutical pollutants from wastewater using biological approach. Bioresour. Technol. 2017, 224, 1–12. [Google Scholar] [CrossRef]
  32. Cecconet, D.; Molognoni, D.; Callegari, A.; Capodaglio, A.G. Biological combination processes for efficient removal of pharmaceutically active compounds from wastewater: A review and future perspectives. J. Environ. Chem. Eng. 2017, 5, 3590–3603. [Google Scholar] [CrossRef]
  33. Hofman-Caris, C.H.M.; Siegers, W.G.; van de Merlen, K.; De Man, A.W.A.; Hofman, J.A.M.H. Removal of pharmaceuticals from WWTP effluent: Removal of EfOM followed by advanced oxidation. Chem. Eng. J. 2017, 327, 514–521. [Google Scholar] [CrossRef]
  34. Kanakaraju, D.; Glass, B.D.; Oelgemöller, M. Advanced oxidation process-mediated removal of pharmaceuticals from water: A review. J. Environ. Manag. 2018, 219, 189–207. [Google Scholar] [CrossRef]
  35. Bastami, T.R.; Ahmadpour, A.; Hekmatikar, F.A. Synthesis of Fe3O4/Bi2WO6 nanohybrid for the photocatalytic degradation of pharmaceutical ibuprofen under solar light. J. Ind. Eng. Chem. 2017, 51, 244–254. [Google Scholar] [CrossRef]
  36. Beni, F.A.; Gholami, A.; Ayati, A.; Shahrak, M.N.; Sillanpää, M. UV-switchable phosphotungstic acid sandwiched between ZIF-8 and Au nanoparticles to improve simultaneous adsorption and UV light photocatalysis toward tetracycline degradation. Micropor. Mesopor. Mater. 2020, 303, 110275. [Google Scholar] [CrossRef]
  37. Gomes, J.; Costa, R.; Quinta-Ferreira, R.M.; Martins, R.C. Application of ozonation for pharmaceuticals and personal care products removal from water. Sci. Total Environ. 2017, 586, 265–283. [Google Scholar] [CrossRef]
  38. Kyzas, G.Z.; Fu, J.; Lazaridis, N.K.; Bikiaris, D.N.; Matis, K.A. New approaches on the removal of pharmaceuticals from wastewaters with adsorbent materials. J. Mol. Liq. 2015, 209, 87–93. [Google Scholar] [CrossRef]
  39. Mlunguza, N.Y.; Ncube, S.; Mahlambi, P.N.; Chimuka, L.; Madikizela, L.M. Adsorbents and removal strategies of non-steroidal anti-inflammatory drugs from contaminated water bodies. J. Environ. Chem. Eng. 2019, 7, 103142. [Google Scholar] [CrossRef]
  40. Ayati, A.; Ranjbari, S.; Tanhaei, B.; Sillanpää, M. Ionic liquid-modified composites for the adsorptive removal of emerging water contaminants: A review. J. Mol. Liq. 2019, 275, 71–83. [Google Scholar] [CrossRef]
  41. Ranjbari, S.; Tanhaei, B.; Ayati, A.; Sillanpää, M. Novel Aliquat-336 impregnated chitosan beads for the adsorptive removal of anionic azo dyes. Int. J. Biol. Macromol. 2019, 125, 989–998. [Google Scholar] [CrossRef]
  42. Arya, V.; Philip, L. Adsorption of pharmaceuticals in water using Fe3O4 coated polymer clay composite. Micropor. Mesopor. Mater. 2016, 232, 273–280. [Google Scholar]
  43. Ravi, S.; Choi, Y.; Choe, J.K. Novel phenyl-phosphate-based porous organic polymers for removal of pharmaceutical contaminants in water. Chem. Eng. J. 2020, 379, 122290. [Google Scholar]
  44. Abdolmaleki, A.Y.; Zilouei, H.; Khorasani, S.N.; Zargoosh, K. Adsorption of tetracycline from water using glutaraldehyde-crosslinked electrospun nanofibers of chitosan/poly (vinyl alcohol). Water Sci. Technol. 2018, 77, 1324–1335. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, Y.; Liu, R.; Li, M.; Yu, F.; He, C. Removal of pharmaceuticals by novel magnetic genipin-crosslinked chitosan/graphene oxide-SO3H composite. Carbohyd. Polym. 2019, 220, 141–148. [Google Scholar]
  46. Mojiri, A.; Andasht Kazeroon, R.; Gholami, A. Cross-linked magnetic chitosan/activated biochar for removal of emerging micropollutants from water: Optimization by the artificial neural network. Water 2019, 11, 551. [Google Scholar]
  47. Pereira, A.K.D.S.; Reis, D.T.; Barbosa, K.M.; Scheidt, G.N.; da Costa, L.S.; Santos, L.S.S. Antibacterial effects and ibuprofen release potential using chitosan microspheres loaded with silver nanoparticles. Carbohyd. Res. 2020, 488, 107891. [Google Scholar]
  48. Li, J.; Zhao, T.; Yang, Q.; Du, S.; Xu, L. A review of quantitative structure-activity relationship: The development and current status of data sets, molecular descriptors and mathematical models. Chemometr. Intell. Lab. 2025, 256, 105278. [Google Scholar]
  49. Uzzaman, M.; Hasan, M.K.; Mahmud, S.; Yousuf, A.; Islam, S.; Uddin, M.N.; Barua, A. Physicochemical, spectral, molecular docking and ADMET studies of bisphenol analogues: A computational approach. Inform. Med. Unlocked 2021, 25, 100706. [Google Scholar]
  50. Alberti, G.; Amendola, V.; Pesavento, M.; Biesuz, R. Beyond the synthesis of novel solid phases: Review on modelling of sorption phenomena. Coordin. Chem. Rev. 2012, 256, 28–45. [Google Scholar]
  51. Rao, G.P.; Lu, C.; Su, F. Sorption of divalent metal ions from aqueous solution by carbon nanotubes: A review. Sep. Purif. Technol. 2007, 58, 224–231. [Google Scholar] [CrossRef]
  52. Quintero-Álvarez, F.G.; Rojas-Mayorga, C.K.; Mendoza-Castillo, D.I.; Aguayo-Villarreal, I.A.; Bonilla-Petriciolet, A. Physicochemical modeling of the adsorption of pharmaceuticals on mil-100-Fe and mil-101-Fe MOFs. Adsorpt. Sci. Technol. 2022, 2022, 4482263. [Google Scholar]
  53. Varma, R.; Vasudevan, S.; Chelladurai, S.; Narayanasamy, A. Synthesis and physicochemical characteristics of chitosan extracted from Pinna deltoides. Lett. Appl. NanoBioSci. 2021, 11, 4061–4070. [Google Scholar]
  54. Medeiros, R.S.; Ferreira, A.P.; Venâncio, T.; Cavalheiro, É.T. Preparation, characterization and study of the dissociation of naproxen from its chitosan salt. Molecules 2022, 27, 5801. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of adsorption process by chitosan hydrogels. Color blue: water, color yellow: hydrogel, color red: naproxen and color green: diclofenac, respectively.
Figure 1. Schematic illustration of adsorption process by chitosan hydrogels. Color blue: water, color yellow: hydrogel, color red: naproxen and color green: diclofenac, respectively.
Polysaccharides 06 00090 g001
Figure 2. Pharmaceutical structures.
Figure 2. Pharmaceutical structures.
Polysaccharides 06 00090 g002
Figure 3. Cross-linking process: (a) chitosan and (b) chitosan/glutaraldehyde.
Figure 3. Cross-linking process: (a) chitosan and (b) chitosan/glutaraldehyde.
Polysaccharides 06 00090 g003
Figure 4. Electrostatic potential map of (a) diclofenac and (b) naproxen.
Figure 4. Electrostatic potential map of (a) diclofenac and (b) naproxen.
Polysaccharides 06 00090 g004
Figure 5. EPM of the hydrogel–diclofenac system at 298.15 K.
Figure 5. EPM of the hydrogel–diclofenac system at 298.15 K.
Polysaccharides 06 00090 g005
Figure 6. EPM of the hydrogel–naproxen system at 298.15 K.
Figure 6. EPM of the hydrogel–naproxen system at 298.15 K.
Polysaccharides 06 00090 g006
Figure 7. IR spectrum for glutaraldehyde cross-linked chitosan hydrogel.
Figure 7. IR spectrum for glutaraldehyde cross-linked chitosan hydrogel.
Polysaccharides 06 00090 g007
Figure 8. X-ray diffractogram of the hydrogel.
Figure 8. X-ray diffractogram of the hydrogel.
Polysaccharides 06 00090 g008
Figure 9. Single adsorption isotherms of diclofenac at (a) 308.15 K and (b) 298.15 K, and of naproxen at (c) 308.15 K and (d) 298.15 K.
Figure 9. Single adsorption isotherms of diclofenac at (a) 308.15 K and (b) 298.15 K, and of naproxen at (c) 308.15 K and (d) 298.15 K.
Polysaccharides 06 00090 g009
Figure 10. Binary adsorption isotherms of diclofenac at (a) 308.15 K and (b) 298.15 K, and of naproxen at (c) 308.15 K and (d) 298.15 K.
Figure 10. Binary adsorption isotherms of diclofenac at (a) 308.15 K and (b) 298.15 K, and of naproxen at (c) 308.15 K and (d) 298.15 K.
Polysaccharides 06 00090 g010
Figure 11. FTIR spectra of (a) hydrogel, (b) hydrogel–naproxen, (c) hydrogel–diclofenac, and (d) hydrogel–naproxen–diclofenac.
Figure 11. FTIR spectra of (a) hydrogel, (b) hydrogel–naproxen, (c) hydrogel–diclofenac, and (d) hydrogel–naproxen–diclofenac.
Polysaccharides 06 00090 g011
Figure 12. XRD patterns of hydrogel before adsorption and loaded with naproxen and diclofenac.
Figure 12. XRD patterns of hydrogel before adsorption and loaded with naproxen and diclofenac.
Polysaccharides 06 00090 g012
Table 1. Pharmaceutical compounds are found in wastewater.
Table 1. Pharmaceutical compounds are found in wastewater.
PharmaceuticalWater SourceConcentration
(ng/L)
Reference
AmpicillinIndustrial effluents5.8 × 103[25]
CaffeineUrban effluents23–776[26]
Surface water2.9–194[26]
WWTP influents2448–4865[27]
DiclofenacUrban effluents8.8–127[26]
Surface water1.1–6.8[26]
WWTP influents9–13[27]
River21–98[28]
IbuprofenUrban effluents10–137[26]
Surface water11–38[26]
River35–270[28]
KetoprofenHospital effluents1034.5[29]
WWTP influents24–177[29]
River17–620[30]
NaproxenUrban effluents20–483[26]
Surface water1.8–18[26]
River81–360[28]
Salicylic acidWWTP influents433–8036[27]
TetracyclineIndustrial effluents1.5 × 103[25]
Table 2. Energetic properties and QSAR at 298.15 K.
Table 2. Energetic properties and QSAR at 298.15 K.
Energetic PropertiesDiclofenacNaproxenHydrogel–DiclofenacHydrogel–Naproxen
Gibbs free energy (kcal/mol)−81,614−67,954−586,105−572,472
Enthalpy of formation (kcal/mol)−48−10016831721
Dipolar moment (Debyes)1.792.219.7216.38
QSAR PropertiesDiclofenacNaproxenHydrogel–DiclofenacHydrogel–Naproxen
Volume (Å3)76871040183969
Surface area (Å2)46043820272006
Log P−0.21−0.56−14.34−12.86
Polarizability (Å3)29.6925.32160.68156.31
Table 3. Energetic properties and QSAR of hydrogel–diclofenac adsorption at several temperatures.
Table 3. Energetic properties and QSAR of hydrogel–diclofenac adsorption at several temperatures.
Energetic Properties298.15 K308.15 K
Gibbs free energy (kcal/mol)−585,943−585,950
Dipolar moment (Debyes)11.8912.3
QSAR Properties298.15 K308.15 K
Volume   ( Å 3 ) 39893987
Surface   area   ( Å 2 ) 19221928
Log P−14.34−14.34
Table 4. FTIR vibrations of hydrogel–diclofenac at different temperatures.
Table 4. FTIR vibrations of hydrogel–diclofenac at different temperatures.
Functional Group298.15 K308.15 KVibrational Mode
OH *3637, 3577, 33753634, 3535, 3382Stretching
C=O *20572042Stretching
C=O Δ19791976Stretching
C=N *19612042Stretching
NH2 *17091727Scissoring
* hydrogel, Δ diclofenac.
Table 5. Energetic properties and QSAR of hydrogel–naproxen adsorption at different temperatures.
Table 5. Energetic properties and QSAR of hydrogel–naproxen adsorption at different temperatures.
Energetic Properties298.15 K308.15 K
Gibbs free energy (kcal/mol)−572,290−572,251
Dipolar moment (Debyes)16.3516.02
QSAR Properties298.15 K308.15 K
Volume   ( Å 3 ) 39703978
Surface   area   ( Å 2 ) 20022015
Log P−12.86−12.86
Table 6. FTIR vibrations of hydrogel–naproxen.
Table 6. FTIR vibrations of hydrogel–naproxen.
Functional Group298.15 K308.15 KVibrational Mode
OH *3489, 3418, 32923470, 3407, 3288Stretching
C=O Δ20772164Stretching
C=O *1996, 19781964Stretching
C=N *19712036, 1876Stretching
NH2 *17131731Scissoring
C-O *15831593Stretching
* hydrogel, Δ naproxen.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ávila Camacho, B.A.; Rojas Pabón, M.A.; Rangel Vázquez, N.A.; Márquez Brazón, E.A.; Reynel Ávila, H.E.; Mendoza Castillo, D.I.; Huerta, Y.A. Adsorption of Pharmaceutical Compounds from Water on Chitosan/Glutaraldehyde Hydrogels: Theoretical and Experimental Analysis. Polysaccharides 2025, 6, 90. https://doi.org/10.3390/polysaccharides6040090

AMA Style

Ávila Camacho BA, Rojas Pabón MA, Rangel Vázquez NA, Márquez Brazón EA, Reynel Ávila HE, Mendoza Castillo DI, Huerta YA. Adsorption of Pharmaceutical Compounds from Water on Chitosan/Glutaraldehyde Hydrogels: Theoretical and Experimental Analysis. Polysaccharides. 2025; 6(4):90. https://doi.org/10.3390/polysaccharides6040090

Chicago/Turabian Style

Ávila Camacho, Billy Alberto, Miguel Andrés Rojas Pabón, Norma Aurea Rangel Vázquez, Edgar A. Márquez Brazón, Hilda Elizabeth Reynel Ávila, Didilia Ileana Mendoza Castillo, and Yectli A. Huerta. 2025. "Adsorption of Pharmaceutical Compounds from Water on Chitosan/Glutaraldehyde Hydrogels: Theoretical and Experimental Analysis" Polysaccharides 6, no. 4: 90. https://doi.org/10.3390/polysaccharides6040090

APA Style

Ávila Camacho, B. A., Rojas Pabón, M. A., Rangel Vázquez, N. A., Márquez Brazón, E. A., Reynel Ávila, H. E., Mendoza Castillo, D. I., & Huerta, Y. A. (2025). Adsorption of Pharmaceutical Compounds from Water on Chitosan/Glutaraldehyde Hydrogels: Theoretical and Experimental Analysis. Polysaccharides, 6(4), 90. https://doi.org/10.3390/polysaccharides6040090

Article Metrics

Back to TopTop