Next Article in Journal
Profitability Analysis of Selected Low Impact Development Methods for Decentralised Rainwater Management: A Case Study from Lublin Region, Poland
Next Article in Special Issue
Simple Preparation of Lignin-Based Phenolic Resin Carbon and Its Efficient Adsorption of Congo Red
Previous Article in Journal
A New Method for Selecting the Geometry of Systems for Surface Infiltration of Stormwater with Retention
Previous Article in Special Issue
Preparation of Porous Carbon Materials as Adsorbent Materials from Phosphorus-Doped Watermelon Rind
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of a Chemically Modified Biosorbent Based on the Invasive Plant Leucaena leucocephala and Its Application in Metformin Removal

by
Luís Fernando Cusioli
1,
Letícia Nishi
1,
Laiza Bergamasco Beltran
1,
Anna Carla Ribeiro
1,
Rosângela Bergamasco
1,*,
Milena Keller Bulla
1,
Rhana Keterly Facina
2 and
Gustavo Affonso Pisano Mateus
2
1
Department of Chemical Engineering, State University of Maringá, Paraná 87020-900, Brazil
2
Centro Universitário Cesumar, Paraná 87050-900, Brazil
*
Author to whom correspondence should be addressed.
Water 2023, 15(14), 2600; https://doi.org/10.3390/w15142600
Submission received: 25 May 2023 / Revised: 7 July 2023 / Accepted: 11 July 2023 / Published: 18 July 2023

Abstract

:
The present study investigated the use of a biosorbent produced from Leucaena leucocephala pods for the removal of metformin from aqueous solutions. The pods were subjected to chemical and thermal treatments and were referred to as L. leucocephala modified, which was characterized using scanning electron microscopy (SEM). The parameters investigated in the sorption process were temperature, contact time, adsorbent dosage, pH, and initial metformin concentration. The experimental data were in accordance with the Langmuir isothermal model. The maximum adsorption capacity reached was 56.18 mg g−1 at 313 K. In the kinetic study, stability was achieved in 300 min, with 53.24% removal, and the pseudo-first-order model agreed well with the experimental data. The thermodynamic parameters indicated a spontaneous, favorable, and exothermic reaction. The presence of NaCl, CaCl2, and MgCl2 negatively affected metformin adsorption. Thus, the importance of the study was that a developed material showed promising results in the removal of metformin, particularly because it is an innovative material, and there are no studies in the literature on drug removal using L. leucocephala.

1. Introduction

The technological development of contemporary society has advanced in different areas of everyday life, such as communication, locomotion, and the use of increasingly specialized therapeutic treatments for various pathologies. Paradoxically, the modern lifestyle is a great driver of scientific progress while also making society increasingly dependent on the indiscriminate use of various chemical substances. In this sense, the detection of synthetic compounds originating from anthropogenic activities in surface and groundwater is already a reality in several locations around the world [1].
The scientific community has signaled its constant concern about the subject since the so-called emerging compounds (EC) are capable of compromising the quality of the aquatic ecosystem, even at low concentrations [2]. Such compounds have diverse natures and include pesticides, dyes, sweetening substances, plasticizing agents, hormones, personal care products, and drugs. Pharmaceutical compounds, in particular, are a class of substances considered environmentally recalcitrant in aqueous matrices, given their high degree of stability, and, therefore, their toxic effects must be strongly taken into account [3].
The access routes of pharmaceutical metabolites to the environment are diverse and include wastewater from livestock properties as well as urban, hospital, and industrial wastewater [4]. It is estimated that, ultimately, the pharmaceutical sector has a consumption rate of about 22% of fresh water in its processes compared to other sources, thus being a very extensive source of release of such compounds into the environment through effluents not treated or treated inefficiently [5].
Among many commercialized compounds, metformin (MET) is the most prescribed oral hypoglycemic factor in the world for the treatment of type 2 diabetes mellitus (DM2) [6]. According to the International Diabetes Federation (2013), the number of people diagnosed with DM2 in 2011 was 382 million worldwide, and an estimated 592 million will be carriers of the disease by the year 2035. In 2019, metformin was ranked fourth among the most prescribed drugs in the United States (Agency for Healthcare Research and Quality 2022).
As MET is not metabolized by the human body and is excreted almost completely unchanged in the urine and feces [7], the high level of consumption of the drug by the population has led to its frequent detection in the wastewater pathway system, contaminating the hydric resources that are intended for potable water supplies [8]. Moreover, its discharge into water can expose aquatic organisms to unwanted genetic and physiologic disturbances, such as malformation, feminization, and infertility [9]. This is a relevant issue, as it is an emerging contaminant with high levels in the environment [10,11].
As already reported in many papers, conventional wastewater treatment plants are still not prepared to remove EC from water [12]. For such purposes, several methodologies have been used to ensure drug removal from contaminated water. Adsorption is a technique that has stood out because it is simple, economical, and easy to perform [13]. A great advantage of this process is the wide variety of materials that can be employed as adsorbents, including agricultural wastes from different origins, geomaterials, and minerals. In order to improve the contaminants’ adsorption capacity, different approaches to physical or chemical modifications have been applied to these materials, such as acidic, basic, thermal, or steam activation [14]. Regarding MET wastewater removal through plant-based adsorbents, there are reports in the literature of the use of Moringa oleifera seed husks functionalized with iron oxide nanoparticles [15], activated carbon from Zea mays tassels [16], orange peel [17], and chichá-do-cerrado (Sterculia striata St. Hil. et Naud) fruit shells [18].
In this context, Leucaena leucocephala is a leguminous species tree originating from Central America that has been explored for many uses, such as a cellulose source for the paper industry, firewood, and protein biomass for animal feed enrichment [19]. However, in some locations, as in Brazil, it is considered an invasive plant due to toxins that inhibit the development of endemic species [20].
In the last decade, L. leucocephala has received considerable attention from researchers. The total number of articles published using different parts of the shrub is approximately 928, and most of them are published mainly in agricultural, energy, environmental, medicinal, and pharmacology journals [21]. Focusing on the environmental field, L. leucocephala has been widely investigated as a coagulant agent [22] and biosorbent for contaminant removal from wastewater, such as heavy metals (Pb2+, Cd2+ and Ni2+) [23,24] and malachite green dye [25]. The L. leucocephala pod has already been used by the adsorption process to remove contaminants such as hexavalent chromium and malachite green dye [25,26].
Considering the circular economy perspective, the use of L. leucocephala as a biosorbent for pollutant removal from water can be a promising application for agricultural wastes or an option in places where the plant has become invasive [21]. Therefore, the investigation of the potential of L. leucocephala bark for pharmaceutical removal from an aqueous solution such as MET is extremely pertinent and urgent, given the actual scenario. With this objective, L. leucocephala bark was submitted to chemical and thermal treatment. The modifications were characterized through MEV and zeta potential, Fourier-transform infrared spectroscopy (FTIR), structural analysis (BET), and N2 fission process analysis. The MET adsorption performance was evaluated by the analysis of variations in solution pH, adsorbate dosage, adsorbent mass, and stirring speed. The obtained experimental data was submitted to the adjustment of mathematical models. Thus, the objective of the present work was to develop a new bioadsorbant for metformin removal.

2. Materials and Methods

2.1. L. leucocephala Seed Preparation

L. leucocephala material was prepared using the modified method of Akhtar et al. (2007). The in natura pods were washed with tap water and deionized water for the removal of debris. Then, they were oven-dried at 105 °C for 24 h.
The L. leucocephala seeds were subjected to chemical treatment by being kept in contact with 1 M methanol (CH3OH) and 1 M nitric acid (HNO3). Subsequently, heat treatment of the material was performed using a muffle oven (Oven Jung 10.012) at 300 °C for 1 h. The objective of chemical and thermal treatment is to increase the surface area, pore volume, and porosity of the pods. After chemical and thermal treatments, the material was called L. leucocephala modified (LLMO).

2.2. LLMO Characterization

2.2.1. SEM

For scanning electron microscopy (SEM), LLMO were fixed on a support that was coated with gold metallization and observed with a high-resolution double-beam electron microscope. FEI SCIOS (Quanta 250 FEI, Eindhoven, The Netherlands) was used at a voltage of 15 kV.

2.2.2. Zeta Potential Analyzer

A particle analyzer, DelsaTMNanoC (Beckman Coulter, Brea, CA, USA), was used to analyze the surface load of the material, and a pH between 2 and 12 was used to verify which would be the isoelectric point and to observe the cationic and anionic charges.

2.2.3. FTIR and BET

Fourier-transform infrared spectroscopy (FTIR) was performed using a Vertex 70 v (Bruker, Billerica, MA, USA) to analyze the functional groups of the studied adsorbent, and the Brunauer–Emmett–Teller (BET) method was employed using adsorption/desorption isotherms.

2.3. Metformin Samples

The metformin solutions were prepared with distilled water and different concentrations of metformin (MET, PA > 99%, Sigma-Aldrich, St. Louis, MO, USA) (5–200 mg L−1). The solution was diluted with the aid of a magnetic stirrer, in which the masses of the metformin standard were weighed and consequently diluted in a 1 L volumetric flask using ultra-pure water.

2.3.1. Adsorption Experiments

Metformin was used in this study (MET, PA > 99%, Sigma-Aldrich) and adsorption tests were carried out in a shaker (Tecnal TE-4200) at 150 rpm and 25 °C. For that, different masses of adsorbent (0.01–0.05 g) were stirred in contact with 30 mL of MET solutions (w/v) at different concentrations (5–200 mg L−1) [18,27]. The parameters that were considered for the adsorption experiments are listed in Table 1.

2.3.2. Effect of Biosorbent Concentration and pH

Initially, preliminary tests were controlled using 0.03 g LLMO in 30 mL of 10 mg L−1 metformin solution in constant rotation of 150 rpm at 25 °C for 24 h. After this test, the effect of the biosorbent mass on 0.01, 0.02, 0.03, 0.04, and 0.05 g was determined. In addition, the effects of pH were determined using acid, neutral, and alkaline pH (4, 7, and 10).

2.3.3. Kinetic and Equilibrium Study

The procedure was performed with 0.01 g LLMO in 30 mL of 10 mg g−1 metformin solution maintained at an inspiration rate of 150 rpm, pH 7, and temperature of 25 °C. Aliquots were taken at specific intervals from 1 to 1440 min, and transferred to airtight jars. After withdrawing each sample, the solutions were filtered through a 0.45 µm cellulose acetate membrane. Final readings were submitted using a spectrophotometer at 229 nm. All tests were performed in duplicate. With the final concentration of each sample, the adsorption capacity was shown as shown in Equation (1),
q e = ( C i C f ) . V m
where qe is the capacity of adsorption (mg g−1). Ci (mg g−1) and Cf (mg g−1) are the initial and final concentrations of the pollutant, respectively; V is the volume of the pollutant solution (L); and m is the mass of the biosorbent (g). To explain the kinetic models, classic models such as pseudo-first-order (PFO), pseudo-second-order (PSO), and intraparticle diffusion were applied to the experimental data. The model PFO was described by Lagergren [28] (1898) as a kinetic operation that represents the behavior of a reactance, like the decomposition of a substance, as if it were a first order. It is based on the hypothesis that the speed of the reaction is directly proportional to the concentration of the species reacting at each moment. The PSO model was described by Ho (1937) and represents reactions that depend not only on the concentration of the reacting substance, but also on the resulting concentration of the reaction product. These models aim to represent the behavior of a reactance as if it were second order. Finally, the intraparticle diffusion model refers to mass transfer carried out by diffusion between interior components and is presented in Equation (2).
q t = q e [ 1 e k 1 t ]
Since qt and qe are the adsorption capacities with respect to time (t) at their equilibrium (mg g−1), k1 is the constant (min−1), and t (min). The PSO model was developed by Ho and McKay [29], as in Equation (3).
q t = k 2 q e 2 t 1 + k 2 q e t
Let k2 be a constant PSO adsorption rate and the intraparticle difference to those proposed by Weber and Morris in which it was performed in the present study. With the help of Equation (4), the effective transmission rates for the liveries were determined [30],
q t = K D i f t + C
where KDif is the constant intraparticle diffusion rate (mg g−1 min1/2), and C is a constant that is related to the thickness of the boundary layer (mg g−1). Thus, considering that the equation is valid for all solutes that tend to move by intraparticle diffusion, we can state that this equation describes the dependence of the soil solute transport rate on the thickness of the boundary layer.

2.4. Adsorption Isotherms

The Freundlich experiment was maintained at three different temperatures: 25, 35, and 45 °C, using 30 mL of metformin solution with concentrations distributed from 5 to 200 mg L−1 in contact with 0.01 g of LLMO at pH 7 under 150 rpm for 480 min. From the adsorption capacity calculations, the most classic models of adsorption isotherms were evaluated, namely, Langmuir and Freundlich.
The Langmuir isotherm is based on the idea that catalyst molecules have individual places where they chemically bond to the reactive material. It describes the relationship between the amount of reactive material bound to the catalyst per unit surface area (or surface occupation) through Equation (5) [31]:
q e q = q m b L C e 1 + b L C e ,
where bL is the Langmuir isotherm (L mg−1) constant. The Langmuir isotherm describes how a particle adsorbs onto a specific surface or interface. It is used in separation and decanting processes, as well as to describe the adsorption of gases onto solid surfaces. The Freundlich model becomes useful in modeling adsorption/desorption phenomena, especially when the presence of two or more components favors their absorption. The Freundlich model can be applied in several systems, such as liquid mixtures of surfactants, in which a fraction of the species is adsorbed on a solid surface material, in addition to other applications in gas–solid, liquid–solid, and others. Furthermore, the Freundlich model can also be applied to electrical processing problems, when a circuit is formed by a conductor and 2 or more components. In this case, the Freundlich model must be adjusted according to the specific conditions of the circuit, given that the temperature and density conditions directly influence the response of the circuit components [32], as presented in Equation (6):
q e q = k F C e 1 / n ,
where kF is the Freundlich isotherm constant [(mg/g)/(mg/L)1/n]. After applying the adsorption isotherm models, the equilibrium data were used to calculate thermodynamic parameters such as enthalpy (ΔH), entropy (ΔS), equilibrium constant Kc, and Gibbs free energy (ΔG). The Freundlich isotherm constant (kF), expressed in mg L−1 (L g−1)1/n, was also considered in the calculations.

2.5. Effect of Ionic Strength

An analysis of the effect of ionic strength on the adsorption of metformin in LLMO was carried out, considering the presence of common salt in surface water and its dissociated ions. For this, different salts (NaCl, CaCl2, and MgCl2) were added with concentrations of 0.1 and 0.3 M of free ions and a contact time of 24 h at 25 °C. The objective was to determine the influence of this salt and its ions on the adsorption of metformin.

3. Results and Discussion

3.1. Biosorbent Characterization

SEM images of LLMO are presented in Figure 1, revealing the formation of pores in the form of small cavities on the surface of the sample, which indicates a good possibility of metformin being adsorbed. The appearance of pores can be explained by the evaporation of volatile components during the chemical and thermal treatment processes [33,34].
The influences of particle diameter, mass, and pH were evaluated via adsorption experiments (Figure 2). The highest sieve yield and adsorption capacity of LLMO were obtained with the 600 μm sieve. It was then observed that the granulometry of 600 μm, in this particular case due to the higher amount of coal of heterogeneous size, presented a better removal rate. The adsorption capacities for adsorbent masses of 0.01, 0.02, 0.03, 0.04, and 0.05 g L−1 were 16.07, 9.25, 8.69, 7.64, and 7.19 mg g−1, respectively. Therefore, an adsorbent mass of 0.05 g L−1 was selected in subsequent experiments.
At pH 4, 7, and 10, the adsorption capacities were 5.84, 7.21, and 3.82 mg g−1, respectively. The adsorption capacity of metformin is higher at pH 7, which is in line with the results obtained for other adsorbents. Similar results were found by [35]. In that study, metformin adsorption from aqueous solutions was performed using Moringa oleifera Lam. seed husks and it was verified that pH 7 favored the interaction between the bioadsorbant and metformin. Metformin has a positive charge at pH 7, which is lower than its pKa value, while LLMO has a slightly negative surface charge at pH 7 owing to its low isoelectric point (6.2) [36]. This indicates favorable electrostatic interactions between metformin and LLMO at pH 7.
FTIR analysis was used to determine the functional groups present, as shown in Figure 3.
Cellulose nitrate is a modified cellulose polymer. It is characterized by wide bands in its FTIR spectrum, notably at 3707–3633 cm−1, indicating hydrogen bonding. If methylation has occurred, a peak at 2920 cm−1 indicates the asymmetric stretching of the C–H bond of the CH2 group. A peak at 2831 cm−1 refers to the symmetrical stretching of the C–H bond of the CH3 in a β linkage [37]. The band at 1697 cm−1 corresponds to the conformational release vibration of the bond between carbon and nitrogen (C–N). The increase in relative area and the displacement of the peak suggest that this bond was distorted, which happened to the carbon-to-carbon (C–C) bond as well, as it shifted to 1680 cm−1. These changes are related to the breakdown of the angular structure of the molecule, which occurs when it is heated. Thus, this indicates that the temperature increased and there was a change in the molecular structure.
This alteration is probably related to the increase in the length of the COO bonds and the partial distortion of the N–H secondary amine groups [38]. The peak observed at 927 cm−1 is typical of the C–O stretching vibration, which indicates the potential presence of lignin, cellulose, and hemicellulose. Possibly, the adsorption mechanism of metformin in LLMO occurred due to the assumption of electrostatic interactions between the carboxylate groups of the biosorbent particle surface along with metformin’s protonated amine and hydrogen bonds. After carrying out the adsorption, a BET analysis was performed to evaluate the porosity of the investigated material. The large specific area contributes to the adsorption of metformin in the available pores. The results obtained are shown in Table 2.
According to IUPAC, there are three categories of pore diameters: micropores, mesopores, and macropores. The developed material presents a predominance of micropores (20 Å), which indicates good interaction between the material and the contaminant. The more micropores present, the better the adsorption process [39]. This fact can be observed in [40], in which a study was conducted to evaluate the use of biosorbents in metformin removal, but it was observed that the pore volume available in the biosorbents was relatively small.

3.2. Adsorption Study

The pseudo-first-order (PFO) and pseudo-second-order (PSO) kinetic models are shown in Figure 4, and the parameters are listed in Table 3.
The initial adsorption process was rapid, and saturation was reached approximately 300 min after the onset of agitation. After stabilization, the kinetic studies revealed a maximum adsorption capacity of 6.70 mg g−1 and a removal percentage of 53.24%. The PFO model calculated a correlation coefficient value (R2) of 0.980 and qand 7.825 mg g−1, while the PSO model calculated an R2 value of 0.974 and qand of 11.422 mg g−1. The chi-squared value (χ2 = 0.097) was also lower for the PSO model than that of the PFO model. Thus, the PSO model fit the data the best, and the calculated qand was close to the experimental value (6.70 mg g−1).
Adsorption data were analyzed according to the Langmuir and Freundlich models; the calculated parameters are presented in Table 4 and the adsorption isotherms are shown in Figure 5.
The amount of metformin adsorbed decreased with increasing temperature, suggesting an exothermic process. The Langmuir model satisfactorily fitted the experimental data. This means that adsorption occurs in a monolayer without interactions between contaminant molecules, and that each active site adsorbs only a single molecule [41]. The correlation coefficients (R2) obtained using the Langmuir model (0.920, 0.964, and 0.975 at 25, 35, and 45 °C, respectively) were higher than those obtained using the Freundlich model (0.905, 0926, and 0.913). A higher value of KL was obtained at 25 °C, indicating that interactions were stronger at this temperature, resulting in a higher adsorption capacity. The maximum adsorption capacity (qmax) (56.18 mg g−1) was obtained at a temperature of 313 K. Another biosorbent, containing chitosan and activated charcoal, reached the maximum adsorption capacity of 60.9 mg g−1 at 303 K for the removal of thionine, and thermodynamic results indicated a spontaneous and exothermic adsorption process [42].
In Table 5, the negative variation in free energy (ΔG) indicates a favorable and spontaneous process for metformin removal. The negative adsorption enthalpy (−30.39 kJ mol−1) indicates a chemical process and confirms the exothermic nature of the adsorption process [43]. The positive entropy (ΔS) value (0.027 J K−1 mol−1) indicates a favorable process at the solid–liquid interface [44].
The results indicate that, for the three different temperatures (293 K, 303 K, and 313 K), there was a favorable and spontaneous adsorption since the ΔG values were negative (−22.36, −21.31, and −20.79 KJ mol−1). This means that the system does not require external energy input during the adsorption process and that the process is favorable due to negative Gibbs free energy values [45]. The confirmation of the negative value of ΔH indicates that the process is exothermic, that is, it releases energy in the form of heat. This means that as the process takes place, the temperature of the system increases. The value of −20.39 KJ mol−1 suggests that the process occurs through a chemisorption mechanism, which means that a chemical reaction occurs between the adsorbent and the adsorbate during adsorption. This value is a measure of the amount of energy released per mole of adsorbate and can help to understand the adsorption mechanism involved in the process [43]. The results indicate that the interaction between metformin and LLMO occurred randomly at the solid–liquid interface, since the values obtained for ΔS are low and positive (0.027 KJ mol−1) [46]. A similar behavior was observed by [47], who confirmed that the adsorption of solutions containing metformin on activated carbon occurred spontaneously and exothermically.
In general, lower metformin adsorption was observed in the presence of different salts (NaCl, CaCl2, and MgCl2), as shown in Figure 6. The salts influenced contaminant adsorption, resulting in the final concentration of metformin in the solution. In their metformin studies, ref. [10] verified the metformin adsorption using graphene oxide and presented similar results, indicating significant inhibiting effects on metformin adsorption in the presence of salts.
The optimum parameters of metformin adsorption to LLMO are presented in Table 6, in which we present others’ studies in which metformin was removed with biosorbents.
One can see that the adsorption capacity of LLMO is similar to that of other bioadsorbants; thus, we can consider L. leucocephala pods a potential material for the adsorption of metformin from aqueous solutions.

4. Conclusions

Although the use of L. leucocephala pods is an innovative application, there are no studies in the literature that use L. leucocephala for the removal of contaminants from aqueous solutions. In addition, the material is a sustainable and economically viable solution because it is an invasive plant that is present in abundance in the environment, and its application as an adsorbent has the advantage of low cost and reduces its harmful presence in the environment.
In this study, the optimal parameters for the maximum removal percentage were pH 7, an adsorbent dose of 0.05 g L−1, a contact time of 300 min, and 25 °C. Surface characterization was performed using SEM, which verified the formation of pores. For adsorption and kinetic isothermal analysis, the Langmuir and PSO models satisfactorily fit the experimental data. The thermodynamic parameters confirmed a spontaneous, favorable, and exothermic process. The adsorption capacity decreased in the presence of salts. It is concluded that the new adsorbent presents promising results in the removal of metformin, demonstrating the potential of using an invasive plant as an adsorbent following chemical and thermal treatments, and more studies are being developed aimed at increasing the removal performance of this and other drugs and for the development of a biocomposite. Future studies may report the use of this discovered material on an industrial scale, which will greatly contribute to the environment.

Author Contributions

Methodology, L.F.C.; Formal analysis, L.B.B.; Investigation, R.K.F.; Resources, A.C.R.; Data curation, M.K.B.; Writing—original draft, L.N.; Visualization, G.A.P.M.; Supervision, R.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors declare that no funds, grants, or other support were received during the preparation of this manuscript.

Data Availability Statement

The data that support the findings of this study are available on request from the corresponding author.

Acknowledgments

The authors thank the Department of Chemical Engineering (State University of Maringá) and Cesumar Institute of Science, Technology, and Innovation (UniCesumar) for assistance with characterization and analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahammad, N.A.; Ahmad, M.A.; Hameed, B.H.; Mohd Din, A.T. A Mini Review of Recent Progress in the Removal of Emerging Contaminants from Pharmaceutical Waste Using Various Adsorbents. Environ. Sci. Pollut. Res. 2022. [Google Scholar] [CrossRef]
  2. Samal, K.; Mahapatra, S.; Hibzur Ali, M. Pharmaceutical Wastewater as Emerging Contaminants (EC): Treatment Technologies, Impact on Environment and Human Health. Energy Nexus 2022, 6, 100076. [Google Scholar] [CrossRef]
  3. O’Flynn, D.; Lawler, J.; Yusuf, A.; Parle-Mcdermott, A.; Harold, D.; Mc Cloughlin, T.; Holland, L.; Regan, F.; White, B. A Review of Pharmaceutical Occurrence and Pathways in the Aquatic Environment in the Context of a Changing Climate and the COVID-19 Pandemic. Anal. Methods 2021, 13, 575–594. [Google Scholar] [CrossRef] [PubMed]
  4. Masanabo, N.; Orimolade, B.; Idris, A.O.; Nkambule, T.T.I.; Mamba, B.B.; Feleni, U. Advances in Polymer-Based Detection of Environmental Ibuprofen in Wastewater. Environ. Sci. Pollut. Res. 2023, 30, 14062–14090. [Google Scholar] [CrossRef] [PubMed]
  5. Eniola, J.O.; Kumar, R.; Barakat, M.A.; Rashid, J. A Review on Conventional and Advanced Hybrid Technologies for Pharmaceutical Wastewater Treatment. J. Clean. Prod. 2022, 356, 131826. [Google Scholar] [CrossRef]
  6. Foretz, M.; Viollet, B. New Promises for Metformin: Advances in the Understanding of Its Mechanisms of Action. Med. Sci. 2014, 30, 82–92. [Google Scholar] [CrossRef] [Green Version]
  7. Scheurer, M.; Michel, A.; Brauch, H.J.; Ruck, W.; Sacher, F. Occurrence and Fate of the Antidiabetic Drug Metformin and Its Metabolite Guanylurea in the Environment and during Drinking Water Treatment. Water Res. 2012, 46, 4790–4802. [Google Scholar] [CrossRef]
  8. Yan, J.H.; Xiao, Y.; Tan, D.Q.; Shao, X.T.; Wang, Z.; Wang, D.G. Wastewater Analysis Reveals Spatial Pattern in Consumption of Anti-Diabetes Drug Metformin in China. Chemosphere 2019, 222, 688–695. [Google Scholar] [CrossRef]
  9. Ambrosio-Albuquerque, E.P.; Cusioli, L.F.; Bergamasco, R.; Sinópolis Gigliolli, A.A.; Lupepsa, L.; Paupitz, B.R.; Barbieri, P.A.; Borin-Carvalho, L.A.; de Brito Portela-Castro, A.L. Metformin Environmental Exposure: A Systematic Review. Environ. Toxicol. Pharmacol. 2021, 83, 103588. [Google Scholar] [CrossRef]
  10. Zhu, S.; Liu, Y.G.; Liu, S.B.; Zeng, G.M.; Jiang, L.H.; Tan, X.F.; Zhou, L.; Zeng, W.; Li, T.T.; Yang, C.P. Adsorption of Emerging Contaminant Metformin Using Graphene Oxide. Chemosphere 2017, 179, 20–28. [Google Scholar] [CrossRef]
  11. Ussery, E.; Bridges, K.N.; Pandelides, Z.; Kirkwood, A.E.; Bonetta, D.; Venables, B.J.; Guchardi, J.; Holdway, D. Effects of Environmentally Relevant Metformin Exposure on Japanese Medaka (Oryzias latipes). Aquat. Toxicol. 2018, 205, 58–65. [Google Scholar] [CrossRef] [PubMed]
  12. Sahay, P.; Mohite, D.; Arya, S.; Dalmia, K.; Khan, Z.; Kumar, A. Removal of the Emergent Pollutants (Hormones and Antibiotics) from Wastewater Using Different Kinds of Biosorbent—A Review. Emergent Mater. 2023, 6, 373–404. [Google Scholar] [CrossRef]
  13. de Souza Leite, L.; Hoffmann, M.T.; de Vicente, F.S.; dos Santos, D.V.; Daniel, L.A. Adsorption of Algal Organic Matter on Activated Carbons from Alternative Sources: Influence of Physico-Chemical Parameters. J. Water Process Eng. 2021, 44, 102435. [Google Scholar] [CrossRef]
  14. Zaman, D.; Tiwari, M.K.; Mishra, S. Low-cost adsorptive removal techniques for pharmaceuticals and personal care products. In Energy, Environment, and Sustainability; Springer: Singapore, 2020. [Google Scholar]
  15. Cusioli, L.F.; Quesada, H.B.; de Brito Portela Castro, A.L.; Gomes, R.G.; Bergamasco, R. Development of a New Low-Cost Adsorbent Functionalized with Iron Nanoparticles for Removal of Metformin from Contaminated Water. Chemosphere 2020, 247, 125852. [Google Scholar] [CrossRef] [PubMed]
  16. Kalumpha, M.; Guyo, U.; Zinyama, N.P.; Vakira, F.M.; Nyamunda, B.C. Adsorptive Potential of Zea Mays Tassel Activated Carbon towards the Removal of Metformin Hydrochloride from Pharmaceutical Effluent. Int. J. Phytoremediat. 2020, 22, 148–156. [Google Scholar] [CrossRef] [PubMed]
  17. Jimoh, O.S.; Ibrahim, A.O.; Bello, O.S. Metformin Adsorption onto Activated Carbon Prepared by Acid Activation and Carbonization of Orange Peel. Int. J. Phytoremediat. 2023, 25, 125–136. [Google Scholar] [CrossRef]
  18. Quesada, H.B.; Baptista, A.T.A.; Cusioli, L.F.; Seibert, D.; de Oliveira Bezerra, C.; Bergamasco, R. Surface Water Pollution by Pharmaceuticals and an Alternative of Removal by Low-Cost Adsorbents: A Review. Chemosphere 2019, 222, 766–780. [Google Scholar] [CrossRef]
  19. Wardatun, S.; Harahap, Y.; Mun’im, A.; Saputri, F.C.; Sutandyo, N. Leucaena leucocephala (Lam.) de Wit Seeds: A New Potential Source of Sulfhydryl Compounds. Pharmacogn. J. 2020, 12, 298–302. [Google Scholar] [CrossRef] [Green Version]
  20. Magalhães-Ghiotto, G.A.V.; de Oliveira, A.M.; Natal, J.P.S.; Bergamasco, R.; Gomes, R.G. Green Nanoparticles in Water Treatment: A Review of Research Trends, Applications, Environmental Aspects and Large-Scale Production. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100526. [Google Scholar] [CrossRef]
  21. Hermes Ulises, R.S.; Montiel Aida Lucia, F.; Alma Delia, O.B.; Villaseñor Odila, D.L.T. Impacts of Climate Change on the Water Sector in Mexico. Asian J. Environ. Ecol. 2022, 17, 37–57. [Google Scholar] [CrossRef]
  22. Ismawati, R. Zeolite: Structure and potential in agriculture. J. Pena Sains 2018, 5, 57–64. [Google Scholar] [CrossRef] [Green Version]
  23. Cimá-Mukul, C.A.; Olguín, M.T.; Abatal, M.; Vargas, J.; Barrón-Zambrano, J.A.; Ávila-Ortega, A.; Santiago, A.A. Assessment of Leucaena leucocephala as Bio-Based Adsorbent for the Removal of Pb2+, Cd2+ and Ni2+ from Water. Desalin. Water Treat. 2020, 173, 331–342. [Google Scholar] [CrossRef]
  24. Wan Ibrahim, W.M.H.; Mohamad Amini, M.H.; Sulaiman, N.S.; Kadir, W.R.A. Powdered Activated Carbon Prepared from Leucaena leucocephala Biomass for Cadmium Removal in Water Purification Process. Arab J. Basic Appl. Sci. 2019, 26, 30–40. [Google Scholar] [CrossRef] [Green Version]
  25. Lee, Y.C.; Amini, M.H.M.; Sulaiman, N.S.; Mazlan, M.; Boon, J.G. Batch Adsorption and Isothermic Studies of Malachite Green Dye Adsorption Using Leucaena leucocephala Biomass as Potential Adsorbent in Water Treatment. Songklanakarin J. Sci. Technol. 2018, 40, 563–569. [Google Scholar] [CrossRef]
  26. Yusuff, A.S. Adsorption of Hexavalent Chromium from Aqueous Solution by Leucaena leucocephala Seed Pod Activated Carbon: Equilibrium, Kinetic and Thermodynamic Studies. Arab J. Basic Appl. Sci. 2019, 26, 89–102. [Google Scholar] [CrossRef] [Green Version]
  27. Alnajjar, M.; Hethnawi, A.; Nafie, G.; Hassan, A.; Vitale, G.; Nassar, N.N. Silica-Alumina Composite as an Effective Adsorbent for the Removal of Metformin from Water. J. Environ. Chem. Eng. 2019, 7, 102994. [Google Scholar] [CrossRef]
  28. Lagergren, S. About the Theory of So-Called Adsorption of Solid Substance. Handlinger 1898, 24, 1–39. [Google Scholar]
  29. Ho, Y.S.; McKay, G. Pseudo-Second Order Model for Sorption Processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  30. Weber, W.J.; Morris, J.C. Kinetics of Adsorption on Carbon from Solution. J. Sanit. Eng. Div. 1963, 89, 31–60. [Google Scholar] [CrossRef]
  31. Langmuir, I. The Constitution and Fundamental Properties of Solids and Liquids. Part I. Solids. J. Am. Chem. Soc. 1916, 38, 2221–2295. [Google Scholar] [CrossRef] [Green Version]
  32. Freundlich, H.M. Over the Adsorption in Solution. J. Phys. Chem. 1906, 57, 385–470. [Google Scholar]
  33. Ameen, F.; AlNadhari, S.; Al-Homaidan, A.A. Marine Microorganisms as an Untapped Source of Bioactive Compounds. Saudi J. Biol. Sci. 2021, 28, 224–231. [Google Scholar]
  34. Cusioli, L.F.; Quesada, H.B.; Baptista, A.T.A.; Gomes, R.G.; Bergamasco, R. Soybean Hulls as A Low-Cost Biosorbent for Removal of Methylene Blue Contaminant. Environ. Prog. Sustain. Energy 2019, 39, e13328. [Google Scholar] [CrossRef]
  35. Cusioli, L.F.; Quesada, H.B.; Barbosa de Andrade, M.; Gomes, R.G.; Bergamasco, R. Application of a Novel Low-Cost Adsorbent Functioned with Iron Oxide Nanoparticles for the Removal of Triclosan Present in Contaminated Water. Microporous Mesoporous Mater. 2021, 325, 111328. [Google Scholar] [CrossRef]
  36. Rivera-Utrilla, J.; Sánchez-Polo, M.; Ferro-García, M.Á.; Prados-Joya, G.; Ocampo-Pérez, R. Pharmaceuticals as Emerging Contaminants and Their Removal from Water. A Review. Chemosphere 2013, 93, 1268–1287. [Google Scholar] [PubMed]
  37. Araújo, C.S.T.; Alves, V.N.; Rezende, H.C.; Almeida, I.L.S.; De Assunção, R.M.N.; Tarley, C.R.T.; Segatelli, M.G.; Coelho, N.M.M. Characterization and Use of Moringa Oleifera Seeds as Biosorbent for Removing Metal Ions from Aqueous Effluents. Water Sci. Technol. 2010, 62, 2198–2203. [Google Scholar] [CrossRef]
  38. Pradhan, B.K.; Sandle, N.K. Effect of Different Oxidizing Agent Treatments on the Surface Properties of Activated Carbons. Carbon N. Y. 1999, 37, 1323–1332. [Google Scholar] [CrossRef]
  39. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, L.A.; Rouquerol, J.; Siemieniewska, T. International Union of Pure and Applied Chemistry Physical Chemistry Division Reporting Physisorption Data for Gas/Soils Systems with Special Reference to the Determination of Surface Area and Porosity. Pure Appl. Chem. 1985, 31, 579–638. [Google Scholar] [CrossRef]
  40. Adewuyi, A. Chemically Modified Biosorbents and Their Role in the Removal of Emerging Pharmaceuticalwaste in the Water System. Water 2020, 12, 1551. [Google Scholar] [CrossRef]
  41. Gupta, V.K.; Mittal, A.; Jain, R.; Mathur, M.; Sikarwar, S. Adsorption of Safranin-T from Wastewater Using Waste Materials—Activated Carbon and Activated Rice Husks. J. Colloid Interface Sci. 2006, 303, 80–86. [Google Scholar] [CrossRef] [PubMed]
  42. Jawad, A.H.; Abdulhameed, A.S. Mesoporous Iraqi Red Kaolin Clay as an Efficient Adsorbent for Methylene Blue Dye: Adsorption Kinetic, Isotherm and Mechanism Study. Surf. Interfaces 2020, 18, 100422. [Google Scholar] [CrossRef]
  43. Hayward, D.O.; Trapnell, B.M.W. Chemisorption, 2nd ed.; Butterworths: Washington, DC, USA, 1964. [Google Scholar]
  44. Ali, A.M.; Rønning, H.T.; Al Arif, W.M.; Kallenborn, R.; Kallenborn, R. Occurrence of Pharmaceuticals and Personal Care Products in Effluent-Dominated Saudi Arabian Coastal Waters of the Red Sea. Chemosphere 2017, 175, 505–513. [Google Scholar] [CrossRef]
  45. Elliott, J.R.; Lira, C.T. Introductory Chemical Engineering Thermodynamics; Prentice Hall: Upper Saddle River, NJ, USA, 2012. [Google Scholar]
  46. Crini, G. Non-Conventional Low-Cost Adsorbents for Dye Removal: A Review. Bioresour. Technol. 2006, 97, 1061–1085. [Google Scholar] [CrossRef] [PubMed]
  47. Bernal, V.; Giraldo, L.; Moreno-Piraján, J.C. Thermodynamic Study of Triclosan Adsorption from Aqueous Solutions on Activated Carbon. J. Therm. Anal. Calorim. 2020, 139, 913–921. [Google Scholar] [CrossRef]
  48. Sanchez-Silva, J.M.; Collins-Martínez, V.H.; Padilla-Ortega, E.; Aguilar-Aguilar, A.; Labrada-Delgado, G.J.; Gonzalez-Ortega, O.; Palestino-Escobedo, G.; Ocampo-Pérez, R. Characterization and Transformation of Nanche Stone (Byrsonima crassifolia) in an Activated Hydrochar with High Adsorption Capacity towards Metformin in Aqueous Solution. Chem. Eng. Res. Des. 2022, 183, 580–594. [Google Scholar] [CrossRef]
  49. Balasubramani, K.; Sivarajasekar, N.; Sarojini, G.; Naushad, M. Removal of Antidiabetic Pharmaceutical (Metformin) Using Graphene Oxide Microcrystalline Cellulose (GOMCC): Insights to Process Optimization, Equilibrium, Kinetics, And Machine Learning. Ind. Eng. Chem. Res. 2022, 62, 4713–4728. [Google Scholar] [CrossRef]
  50. Mohammad, A.H.; Radovic, I.; Ivanović, M.; Kijevčanin, M. Adsorption of Metformin on Activated Carbon Produced from the Water Hyacinth Biowaste Using H3PO4 as a Chemical Activator. Sustainability 2022, 14, 11144. [Google Scholar] [CrossRef]
  51. Hellmann, L.; Schmitz, A.P.D.O.; Módenes, A.N.; Hinterholz, C.L.; Antoniolli, C.D.A. Peach Pit Chemically Treated Biomass as a Biosorbent for Metformin Hydrochloride Removal: Modeling and Sorption Mechanisms. Eng. Agric. 2021, 41, 181–195. [Google Scholar] [CrossRef]
  52. Mustafa, S.; Bhatti, H.N.; Maqbool, M.; Iqbal, M. Microalgae Biosorption, Bioaccumulation and Biodegradation Efficiency for the Remediation of Wastewater and Carbon Dioxide Mitigation: Prospects, Challenges and Opportunities. J. Water Process Eng. 2021, 41, 102009. [Google Scholar]
  53. Mahmoud, M.E.; El-Ghanam, A.M.; Saad, S.R.; Mohamed, R.H.A. Promoted Removal of Metformin Hydrochloride Anti-Diabetic Drug from Water by Fabricated and Modified Nanobiochar from Artichoke Leaves. Sustain. Chem. Pharm. 2020, 18, 100336. [Google Scholar] [CrossRef]
  54. Akpomie, K.G.; Conradie, J. Banana Peel as a Biosorbent for the Decontamination of Water Pollutants. A Review. Environ. Chem. Lett. 2020, 18, 1085–1112. [Google Scholar]
  55. Shaikhiev, I.G.; Kraysman, N.V.; Sverguzova, S.V. Review of Peach (Prúnus pérsica) Shell Use to Remove Pollutants from Aquatic Environments. Biointerface Res. Appl. Chem. 2023, 13, 459. [Google Scholar]
Figure 1. (a,b) Scanning electron microscopy images of Leucaena leucocephala modified.
Figure 1. (a,b) Scanning electron microscopy images of Leucaena leucocephala modified.
Water 15 02600 g001
Figure 2. Effect of Leucaena leucocephala modified (a) mass and (b) pH on metformin adsorption.
Figure 2. Effect of Leucaena leucocephala modified (a) mass and (b) pH on metformin adsorption.
Water 15 02600 g002
Figure 3. FTIR spectra LLMO.
Figure 3. FTIR spectra LLMO.
Water 15 02600 g003
Figure 4. Kinetics of metformin adsorption in Leucaena leucocephala modified.
Figure 4. Kinetics of metformin adsorption in Leucaena leucocephala modified.
Water 15 02600 g004
Figure 5. Metformin adsorption isotherms of Leucaena leucocephala modified at (A) 25, (B) 35, and (C) 45 °C.
Figure 5. Metformin adsorption isotherms of Leucaena leucocephala modified at (A) 25, (B) 35, and (C) 45 °C.
Water 15 02600 g005
Figure 6. Metformin removal capacity in the presence of NaCl, CaCl2, and MgCl2 in the solution.
Figure 6. Metformin removal capacity in the presence of NaCl, CaCl2, and MgCl2 in the solution.
Water 15 02600 g006
Table 1. Adsorption experiments evaluated parameters.
Table 1. Adsorption experiments evaluated parameters.
pHTemperature
(°C)
Metformin Concentration
(mg L−1)
Adsorbent Mass
(g)
Stirring Speed
(rpm)
725100.01, 0.02, 0.03, 0.04 and 0.05150
4, 7, and 1025100.03150
725100.03150
725, 35, and 455–2000.03150
Table 2. Texture properties of LLMO.
Table 2. Texture properties of LLMO.
ParametersLLMO
BET Specific surface area (m² g−1)5.32
Average pore diameter (Å)14.79
Total pore volume (cm³ g−1)0.0892
Micropore volume (cm³ g−1)0.0874
Mesopore volume (cm³ g−1)0.0018
Table 3. Kinetic parameters for metformin biosorption by Leucaena leucocephala modified (LLMO).
Table 3. Kinetic parameters for metformin biosorption by Leucaena leucocephala modified (LLMO).
ModelsParametersLLMO
Pseudo-first-order (PFO)qe (mg g−1)7.825
k1 (min−1)0.008
R20.980
χ20.097
Pseudo-second-order (PSO)qe (mg g−1)11.422
k2 (g mg−1 min−1)0.003
R20.974
χ20.173
Table 4. Model of Langmuir and Freundlich at 293, 303 and 313 K.
Table 4. Model of Langmuir and Freundlich at 293, 303 and 313 K.
ModelsParameters293 K303 K313 K
Langmuirqmáx (mg g−1)14.6439.7956.18
KL (L mg−1)0.0240.0570.073
R20.9200.9640.975
FreundlichKF [(mg/g)/(mg/L)1/n]19.6014.8211.67
nF1.1601.0120.945
R20.9050.9260.913
Table 5. Thermodynamic parameters obtained from the adsorption of metformin at temperatures of 25 °C, 35 °C, and 45 °C.
Table 5. Thermodynamic parameters obtained from the adsorption of metformin at temperatures of 25 °C, 35 °C, and 45 °C.
T (°C)T (K)(ΔG) (KJ mol−1)(ΔH) (KJ mol−1)(ΔS) (KJ mol−1)
25298−22.36−30.390.027
35308−21.31
45318−20.79
Table 6. Studies of metformin removal with biosorbents.
Table 6. Studies of metformin removal with biosorbents.
Adsorbentqe
(mg g−1)
Mass (g)pHTemperature (K)Reference
Acid modification of orange peels (OPAC)50.99 0.17323[17]
Byrsonima crassifolia activated hydrochar113.62.07293[48]
Graphene oxide microcrystalline cellulose (GOMCC)132.100.58.531.8[49]
Activated Carbon chemical activation122.471.08.5318[50]
Peach pit chemically treated biomass82.540.58323[51]
Moringa seed husks are functionalized with nanoparticles65.010.037298[15]
Microalgae45.6716.3298[52]
Artichoke Leaves17.10.012–12333[53]
Banana peel43.281.54298[54]
Peach (Prúnus pérsica)38.970.58303[55]
LLMO56.180.037298This study
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

Cusioli, L.F.; Nishi, L.; Beltran, L.B.; Ribeiro, A.C.; Bergamasco, R.; Bulla, M.K.; Facina, R.K.; Mateus, G.A.P. Synthesis of a Chemically Modified Biosorbent Based on the Invasive Plant Leucaena leucocephala and Its Application in Metformin Removal. Water 2023, 15, 2600. https://doi.org/10.3390/w15142600

AMA Style

Cusioli LF, Nishi L, Beltran LB, Ribeiro AC, Bergamasco R, Bulla MK, Facina RK, Mateus GAP. Synthesis of a Chemically Modified Biosorbent Based on the Invasive Plant Leucaena leucocephala and Its Application in Metformin Removal. Water. 2023; 15(14):2600. https://doi.org/10.3390/w15142600

Chicago/Turabian Style

Cusioli, Luís Fernando, Letícia Nishi, Laiza Bergamasco Beltran, Anna Carla Ribeiro, Rosângela Bergamasco, Milena Keller Bulla, Rhana Keterly Facina, and Gustavo Affonso Pisano Mateus. 2023. "Synthesis of a Chemically Modified Biosorbent Based on the Invasive Plant Leucaena leucocephala and Its Application in Metformin Removal" Water 15, no. 14: 2600. https://doi.org/10.3390/w15142600

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