Next Article in Journal
Reducing Marine Ecotoxicity and Carbon Burden: A Life Cycle Assessment Study of Antifouling Systems
Previous Article in Journal
Mechanistic Insights into Brine Domain Assembly Regulated by Natural Potential Field: A Molecular Dynamics Exploration in Porous Media
Previous Article in Special Issue
Active Sites in Low-Loaded Copper-Exchanged Mordenite: Spectroscopic and Stability Study for Methane Oxidation Using Mild Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of β-C3N4 and Its Novel MnTeO3 Nanohybrids for Remediating Water Contaminated by Pharmaceuticals

1
Chemistry Department, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), P.O. Box 90950, Riyadh 11623, Saudi Arabia
2
Basic Science Research Center, Deanship of Scientific Research, Imam Mohammad Ibn Saud Islamic University (IMSIU), P.O. Box 90950, Riyadh 11623, Saudi Arabia
3
Department of Chemistry, College of Science, Qassim University, Buraydah 51452, Saudi Arabia
4
Central Research Laboratory, Female Students Campus, King Saud University, Riyadh 11623, Saudi Arabia
*
Author to whom correspondence should be addressed.
Processes 2025, 13(8), 2357; https://doi.org/10.3390/pr13082357
Submission received: 24 May 2025 / Revised: 15 July 2025 / Accepted: 17 July 2025 / Published: 24 July 2025

Abstract

A facile method was adopted to fabricate β-C3N4, and it was then doped with manganese and tellurium to obtain novel 10%MnTeO3@β-C3N4 (10%MnTe@β) and 20%MnTeO3@β-C3N4 (20%MnTe@β) nanohybrids. The β-C3N4, 10%MnTe@β, and 20%MnTe@β showed surface areas of 85.86, 97.40, and 109.54 m2 g−1, respectively. Using ciprofloxacin (CIP) as a pollutant example, 10%MnTe@β and 20%MnTe@β attained equilibrium at 60 and 45 min with qt values of 48.88 and 77.41 mg g−1, respectively, and both performed better at pH = 6.0. The kinetic studies revealed a better agreement with the pseudo-second-order model for CIP sorption on 10%MnTe@β and 20%MnTe@β, indicating that the sorption was controlled by a liquid film mechanism, which suggests a high affinity of CIP toward 10%MnTe@β and 20%MnTe@β. The sorption equilibria outputs indicated better alignment with the Freundlich and Langmuir models for CIP removal by 10%MnTe@β and 20%MnTe@β, respectively. The thermodynamic analysis revealed that CIP removal by 10%MnTe@β and 20%MnTe@β was exothermic, which turned more spontaneous as the temperature decreased. Applying 20%MnTe@β as the best sorbent to groundwater and seawater spiked with CIP resulted in average efficiencies of 94.8% and 91.08%, respectively. The 20%MnTe@β regeneration–reusability average efficiency was 95.14% within four cycles, which might nominate 20%MnTe@β as an efficient and economically viable sorbent for remediating CIP-contaminated water.

1. Introduction

Rapid industrialization and population growth have led to significant water contamination issues. Vast amounts of toxic elements, plastics, and pharmaceutical and dye effluents are released into rivers and other natural water bodies by various sectors, including pharmaceutical, metallurgy, mining, textile, paint, leather good, paper, and cosmetics industries [1,2]. These contaminants are known to be detrimental to humans, animals, aquatic life, and the entire environment [3]. Since both industrial intermediates and hazardous substances are physiologically carcinogenic, they are the targets of laws. Hence, it is imperative to ensure both environmental sustainability and human health by implementing laws and regulations regarding the disposal of waste effluent [4]. Some of these hazardous chemicals are found in personal care products (PCPs) and essential medications, which is why they are targeted by pollution control authorities [5]. Antibiotics (ANBs) are the most commonly used PCPPs, and a study of 76 nations showed that drug consumption increased by 65% within 15 years, with a 39% increase in ANB consumption. A 200% increase in ANB usage was anticipated for the upcoming half-decade [6]. Large amounts of ANBs and their by-products eventually make their way into the sewage system unaltered, as the majority of the doses are discharged unmetabolized from the body via excretion [7]. These ANBs endanger the ecosystem’s flora and fauna unless they are efficiently broken down or removed from wastewater. Ciprofloxacin (CIP) is a frequently prescribed antibiotic with a broad spectrum of antibacterial activity that was initially introduced to the market in 1987 [8]. It is a second-generation fluoroquinolone (FQ) that functions by inhibiting the cellular replication of bacteria and ultimately hindering their ability to proliferate. CIP is extensively utilized in the fields of medicine, aquaculture, and agriculture; moreover, in 2013, CIP was the second-most-used FQ antibiotic in China (5340 t production), which makes the case of it being chosen as a representative model pollutant ANB in many research studies concerning water remediation [8]. Due to its persistent and ongoing use, CIP has been detected in widespread concentrations in surface water and wastewater. CIP has been identified in effluents from wastewater treatment facilities at a concentration of 5.6 μg L−1 [9]. The CIP contamination levels in 25 wastewater samples from German hospital effluents ranged from 0.7 to 124.5 μg L−1 [10]. The literature indicates that wastewater from drug production can potentially be a significant source of considerably higher CIP concentrations. A survey of adjoining effluent treatment plants in Hyderabad, India, revealed a CIP concentration range from 1.0 to 14.0 mg L−1, whereas it was approximately 6.5 mg L−1 in two nearby lakes [11]. The accumulation of CIP in an organism’s body presents a considerable health risk; therefore, the effective removal of CIP is essential due to its resistance to degradation and potential ecotoxicity [12]. Unless removed by an efficient treatment procedure, CIP is predicted to persist in the environment for a considerable period, with grave adverse effects [13,14]. Numerous strategies have been proposed for the remediation of wastewater contaminated by CIP, including ozonation, advanced oxidation processes, and bioremediation [15,16,17]. However, these methods have concomitant drawbacks, including ineffective removal, intricate processes, and/or significant energy requirements [18]. Because of its low cost, high performance, and versatility, adsorption is favored [19]. Various adsorbents have been investigated for antibiotic adsorption, including carbon-based materials, polymers, resins, metal–organic frameworks, clays, and minerals [20,21]. The use of the adsorption approach in the uptake of CIP from an aqueous medium is a reasonable protocol that requires developing an appropriate sorbent and adjusting the optimum conditions in order to gain the highest benefits it offers. CIP removal from water has been studied using carbon nitride (C3N4) and its nanocomposites [22]. β-C3N4, a family member of layered materials, is among the oldest synthetic polymers known to scientists [23]. It is a special metal-free organic semiconductor photocatalyst that is active in the visible light spectrum and has adjustable electrical characteristics [24]. Despite the outstanding characteristics of β-C3N4, including its layered structure stability, tunable properties, and strong tetrahedral bonding, relatively few studies have investigated it as an adsorbent [25]. Graphene and boron nitride ribbon fibers were combined to create graphene nanoplatelet/Boron nitride, which composed an adsorbent for removing CIP from aqueous solutions [26]. β-C3N4 is inexpensive, hard, lightweight, and has a relatively high surface area, exhibiting outstanding stability under ambient conditions [27,28]. Its applications extend to various fields, including energy conversion, water purification, NOx decomposition, and the standardization of materials for identifying sites in oxidation reactions, as well as the development of H2 and fuel cells [29,30]. Numerous studies have recommended doping with transition metal oxides to enhance the properties and adsorption capabilities of base materials [31,32]. Doping can narrow the band gap, extend visible light absorption, reduce electron–hole recombination, and, most importantly for sorbents, expand the surface area of base substrates [33,34]. Although some studies have investigated manganese compounds, such as Mn/Te and Mn@NiO, in the fields of photocatalysis and antibacterial applications, these materials have not been tested as adsorbent-doping substrates [35,36,37]. Manganese tellurite is characterized by its hexagonal channel crystal structure, which features high remaining electron densities within the channels [37], resulting in partially positive and negative sites on the sorbent surface. This creates the potential for electrostatic attraction between the sorbents and sorbates, thereby enhancing the sorption process. Nevertheless, there is an investigation deficiency regarding MnTeO3 that is likely attributable to the challenge of synthesizing single-phase MnTeO3 compounds [38].
This study aimed to develop a simplified one-pot route employing guanidine hydrochloride as an unconventional precursor for synthesizing nitrogen-rich β-C3N4. The β-C3N4 fabricated in this study is modified via MnTeO3 doping at 10% and 20% ratios. β-C3N4 and both nanohybrids are characterized and examined for adsorbing CIP as a contaminant example due to pre-identified justifications. Then, the solution conditions are adjusted to achieve the best possible CIP sorption capacities onto β-C3N4 nanohybrids. The CIP sorption order, mechanism, isotherm, and thermodynamics are tested. Moreover, the most effective β-C3N4/MnTeO3 nanohybrid sorbent is utilized to remediate synthetic CIP-contaminated groundwater (GW) and seawater (SW). Furthermore, as a cost-effective part of the survey, the reusability of the selected β-C3N4/MnTeO3 nanohybrid is investigated.

2. Materials and Methods

2.1. Materials

Guanidine hydrochloride (GU-HCl) was obtained from Loba-Chem, Mumbai, India; potassium tellurite and manganese chloride (tetrahydrate) were purchased from Sigma Aldrich, Saint Louis, MO, USA. All the reagents were of analytical grade and used without any further purification.

2.2. Synthesis of β-C3N4 Nanoparticles

β-C3N4 was prepared by placing 20 g of GU-HCl in a 200 mL covered crucible, and it was then pyrolyzed in a muffle furnace at 500 °C for 2.0 h. The product was then cooled, and the obtained powdered β-C3N4 was collected.

2.3. Fabrication of β-C3N4MnTeO3 Nanocomposites

An accurate 4.5 g of β-C3N4 was dispersed in 200 mL of distilled water (D-W) in a 400 mL beaker. Totals of 0.4356 g of MnCl2 · 4H2O and 0.5579 g of K2TeO3 were dissolved separately in 50 mL of deionized water (D-W). The last two beakers were poured simultaneously into the β-C3N4 beaker with vigorous stirring. The formed suspensions were sonicated for 30 min, filtered under suction, washed with D-W, and then dried at 110 °C for 1.0 h in an oven. They were then calcined at 400 °C in a tubular furnace under a nitrogen flow of 60 mL/min and labeled as (10%MnTe@β). The 20%MnTe@β-C3N4 was synthesized similarly, except for using 4.0 g of β-C3N4, 0.8712 g of MnCl2.4H2O, and 1.158 g of K2TeO3, and it was then labeled as (20%MnTe@β).

2.4. Characterization Techneaques

Using standard physical techniques, β-C3N4, 10%MnTe@β, and 20%MnTe@β sorbents were characterized. The Rigaku Smart-Lab X-ray diffractometer, equipped with PDXL2 software-2015 (Tokyo, Japan), was employed to match the XRD patterns and compute the lattice parameters. The surface morphologies of β-C3N4, 10%MnTe@β, and 20%MnTe@β were surveyed via JSM-IT300-scanning (JEOL Ltd., Tokyo, Japan) electron-energy dispersive X-ray technique (SEM–EDX). The functional groups of β-C3N4, 10%MnTe@β, and 20%MnTe@β were explored via a Bruker TENSOR-FTIR spectrophotometer (Berlin, Germany). The β-C3N4, 10%MnTe@β, and 20%MnTe@β surface and porosity properties were analyzed by N2 adsorption–desorption isotherms (ASAP 2020 Micromeritics analyzer, Miami, FL, USA).

2.5. Adsorption Experiments

The contact time study involved preparing a stock solution of 600 mL of 50.0 mg L−1 CIP in distilled water. An amount of 120 mL of CIP solution was stirred separately with 50 mg of each sorbent, and 5 mL of each mixture was withdrawn at serial intervals. Following this, each solution was filtered using a 0.22 μm nylon syringe filter and measured via a UV-Vis spectrophotometer, from which the adsorption capacity (qt, (mg g−1) was computed via the following Equation (1):
Q t = v C 0 C t / m
where m is the mass of the sorbent (g), v denotes the solution’s volume (L), C0 is the initial supplied concentration, and Ct is the unabsorbed concentration. The nonlinear pseudo-first-order (NPFO, Equation (2)) and second-order (NPSO, Equation (3)) models were used to investigate CIP sorption rates onto 10%MnTe@β and 20%MnTe@β nanocomposites. Additionally, Equations (4) and (5) were utilized to elucidate the adsorption control mechanisms via the liquid film and intraparticle diffusion models (LFM and IPM). These equations are shown as follows:
q t = q e ( 1 e x p k 1 . t )
q t = k 2 . q e 2 . t 1 + k 2 . q e . t
q t = K I P t 1 2 + C i
l n ( 1 F ) = K L F t
where Ci represents the boundary layer factor, while k1 (min−1), k2 (g mg−1 min−1), KIP (mg g−1 min−1/2), and KLF (min−1) denote the constants for NPFO, NPSO, LFM, and IPM, respectively.
The initial CIP feed concentration impact was tested utilizing 10, 20, 30, and 50 mg L−1 solutions, and temperature impact was examined in a range of 293–323 K.
To enhance the understanding of CIP sorption onto 10%MnTe@β and 20%MnTe@β, the concentration study outputs were used to investigate the isotherms. The Langmuir (LIM, Equation (6)) and Freundlich (FIM, Equation (7)) isotherm models were utilized to examine the hypothesis of single-layered CIP sorption and the potential for multilayer sorption, as follows:
q e =   K l q m C e 1 + q m C e  
q e = K F . C e 1 n
where Ce (mg L−1) represents the CIP concentration; qm denotes the maximum qt; KL (L mg−1) is the LIM constant; KF (L g−1) and 1/n are the FIM constant and favorability factor, respectively [39].
For a better understanding of CIP removal by 10%MnTe@β and 20%MnTe@β, the thermodynamics were examined utilizing the concentration–temperature study outputs. The entropy (ΔS°) and enthalpy (ΔH°) were computed via Equation (8). The Gibbs free energy (ΔG°) was then derived by introducing their values into Equation (9).
l n   K c =   Δ H o R T   +   Δ S o R
Δ G o = Δ H o T Δ S o
where R stands for the ideal gas constant (0.008314 kJ mol−1), T stands for temperature (K), and Kc stands for the CIP sorption distribution coefficient.
Furthermore, the pH influence was surveyed between pH 3.0 and 10.0 to optimize the parameters for CIP sorption on 10%MnTe@β and 20%MnTe@β. Portions of 50 mg L−1 CIP solution were adjusted at the desired value (pH = 3 to 10) via adding HCl and/or NaOH. In order to prevent any potential errors arising from the alteration of CIP chromophore by the specific pH, an adequate quantity of 50 mg L−1 CIP solution was adjusted to utilize each extra solution as a standard for its typical pH sample. Moreover, the optimized parameters were applied to GW and SW as natural samples after being spiked to achieve 5 and 10 mg CIP pollution levels. Furthermore, the reusability of the two sorbents was examined through four consecutive cycles.

3. Results and Discussions

3.1. Yield Percentage of β-C3N4

Calcining 20.0 g of GU-HCl at 500 °C produced a pale yellow product mass of 12.3 g β-C3N4, indicating a 61.5% yield of utilizing this precursor via the suggested method parameters. Such a relatively high yield suggests that GU-HCl is a suitable precursor for a facile one-pot synthesis method compared to the other studied precursors.

3.2. Characterization

Figure 1a illustrates the stacked plot of β-C3N4, 10%MnTe@β, and 20%MnTe@β diffraction patterns. The Rigaku Smart-Lab PDXL2 (2015) software was employed to match the obtained patterns with reference JCPDS cards. The two distinctive β-C3N4 peaks of 2θ°= 13.79° and 27.56° allocated to (100) and (110) β-C3N4 crystal planes [40], in addition to other weak and broad peaks at 44.65 (300) and 57.08 (220) JCPDS -00-087-1526), confirming the successful fabrication of β-C3N4 from GU-HCl [41,42]. The presence of pre-identified peaks in the 10%MnTe@β and 20%MnTe@β patterns indicated the existence of β-C3N4 without chemical alteration. Moreover, the 10%MnTe@β and 20%MnTe@β peaks at 2θ° of 18.25 (112), 21.22 (101), 25.73 (020), 32.75 (002), (37.95 (220), 42.10 (022), 43.21 (202), 52.10 (103), and 56.57 (400) align with the centered cubic MnTeO3 pattern JCPDS 00-078-1713) [43]. Hence, these data confirmed the incorporation of MnTeO3 within the β-C3N4 nanosheets and further indicated the excellent purity of the synthesized nanohybrids, as no foreign peak was observed. The XRD lattice parameters β-C3N4, 10%MnTe@β, and 20%MnTe@β are illustrated in Table 1.
The FT-IR analysis was employed to elucidate the functional groups of β-C3N4, 10%MnTe@β, and 20%MnTe@β (Figure 1b and Table 2). Stretching vibrations of the moisture O-H and the primary and secondary amine bonds appeared at 3000–3500 cm−1, while the weak band of 2933 cm−1 may be allocated to traces of C-H bond stretching. In 10%MnTe@β and 20%MnTe@β spectra, it is possible to attribute the two shallow peaks at 2374 and 2364 cm−1 to the C≡N stretching, while the stretching mode of heterocycles of C–N is attributed to the peaks at 1200 and 1600 cm−1 [44]. The out-of-plane bending vibration of the triazine ring is responsible for the steep peak at 809 cm−1, confirming the presence of β-C3N4 in the composite framework. Additionally, the 592 and 589 cm−1 bands are ascribed to the Mn–O and Te–O bonds, showing that MnTeO3 was incorporated efficiently [45]. These findings support the XRD indications of the successful synthesis of β-C3N4 and its MnTeO3 nanocomposites via the simple procedure adopted in this work.
The size and morphology of β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials were obtained using SEM imaging (Figure 2a–c). This revealed the appearance of the pristine β-C3N4 as irregularly folded sheets, particles, and plates, as illustrated in Figure 2a. Both 10%MnTe@β and 20%MnTe@β revealed that the prepared materials are composed of fine MnTeO3 nanoparticles dispersed on the surface and/or impeded within the β-C3N4 plates. The nanoparticles of manganese tellurite were dispersed on the surface of the base material as clusters of spherical particles. The 10%MnTe@β and 20%MnTe@β average particle sizes obtained from SEM images were 32.08 ± 5.8 and 40.40 ± 6.6 nm, respectively.
The EDX spectrum of β-C3N4, 10%MnTe@β, and 20%MnTe@β, shown in Figure 2d–f, confirms the presence of the expected elements in each sorbent. The homogeneity of elements constituting β-C3N4, 10%MnTe@β, and 20%MnTe@β was tested via EDX mappings shown in Figure 3a, b, and c, respectively. The contrast between the dark and light colors ensures an even distribution of C, N, Mn, Te, and O atoms among the entire tested area of the nanohybrids.
The surface properties of β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials were examined using the N2 adsorption–desorption technique. Figure 4a–c illustrates that N2 adsorption was negligible at low pressures in type IV isotherms, which increased significantly with increasing gas pressure. The figures show that the three products exhibit type IV isotherm curves with H3 hysteresis loops, which are characteristic of mesoporous constructions with more likely slit-shaped pores [45,46]. Table 3 gathers the BET surface area (SA), average pore diameter (PD), and pore volume (PV). The 20%MnTe@β demonstrates the largest surface area (109.54 m2 g−1), suggesting that it is the most effective adsorbent among the three sorbents. The elevated pore volume of 0.365 cm3 g−1 of this sorbent supports this hypothesis.

3.3. Contact Time and Effect of pH

The produced β-C3N4, 10%MnTe@β, and 20%MnTe@β nanocomposites were tested for adsorbing CIP from aqueous solutions. A diagrammatic representation of the time impact on the adsorption of CIP is presented in Figure 5a. After sixty minutes, the β-C3N4 and 10%MnTe@β sorbent systems achieved a state of equilibrium, displaying qt values of 27.30 and 48.88 mg g−1, respectively; on the other hand, 20%MnTe@β attained equilibrium at min 45 with a qt of 77.41 mg g−1, which positioned 20%MnTe@β among the effective CIP adsorbing materials (Table 4). Notably, within the first two minutes, 20%MnTe@β exhibited a rapid adsorption rate, collecting 75% of its total qt value, which can be attributed to the availability of binding sites resulting from the exterior surface and wider pores of 20%MnTe@β. The observed increase in qt in the 20%MnTe@β composite is most likely attributed to the substantial presence of MnTeO3 nanoparticles, which have the potential to enhance the adsorption of CIP on the material. Additionally, the impediment of ionically bonded Mn, Te, and O (MnTeO3) within and between the β-C3N4 layers created electrostatic diversity via the partially positive Mn and Te sites and the partially negative O sites. Such polarization might lead to the electrostatic attraction force becoming more involved in the sorption process onto 20%MnTe@β, especially given the even distribution revealed by the EDX mapping. The proportional increase in the qt value as the MnTeO3 concentration increased indicated the possibility of achieving a higher qt by further increasing the MnTeO3 doping dose.
Figure 5b illustrates that the adsorption of CIP onto 10%MnTe@β and 20%MnTe@β increased from a pH value of 3 to 6, followed by a slight decrease from pH 7 to pH 8, while notable reductions in qt are observed with further increases in pH. The highest qt value was obtained at pH 5.0 for β-C3N4, while pH 6 was most favorable for the 10%MnTe@β and 20%MnTe@β sorbents. The adsorption behavior of CIP is attributable to the electrostatic interaction between the CIP form in solution and the surface charge of 10%MnTe@β and 20%MnTe@β, with CIP present in the mono-cationic form at pH levels below 6. CIP exists as a zwitterion within the pH range of 6 to 9, while, above pH 9, CIP may adopt an anionic form due to carboxylic group ionization and/or repulsion with the surrounding -OH groups on the sorbents [47,48,49].
Table 4. Comparing the performance of sorbents found in the literature with the prepared 10%MnTe@β and 20%MnTe@β in removing ciprofloxacine from water.
Table 4. Comparing the performance of sorbents found in the literature with the prepared 10%MnTe@β and 20%MnTe@β in removing ciprofloxacine from water.
Sorbentqt (mg g−1)Reference
10%MnTe@β48.88This study
20%MnTe@β77.41This study
Carbon nanoparticles (CNPs) derived from sunflower seed waste103.6[50]
Carbon Nanoparticles prepared from coffee skin waste142.6[51]
Carbon nanotubes synthesized from commercial gasoline95.5[39]
Aluminous oxide13.6[52]
Magnetite12.73[53]
Activated carbon derived from pomegranate peel waste2.353[54]
Zeolites5.79[55]
Activated carbon1.86[56]
Modified montmorillonite1.67[56]
Alumina1.15[56]

3.4. Adsorption Kinetics

The k1 and k2 values shown in Table 5 were computed from the slope values of the NPFO and NPSO graphs depicted in Figure 6a,b. The correlation coefficient (R2), in conjunction with the residual sum of squares and the reduced Chi-square, served as the primary criterion for selecting the model that provided the optimal fit [57]. The CIP sorption by 10%MnTe@β and 20%MnTe@β fitted NPSO for CIP, yielding R2 values of 0.992 and 0.994, respectively. The qt values derived from Equation 3 were closely aligned with the experimental values, the k2 values were similarly comparable, and the lower RSS and X2 values with the higher R2 values verified fitting the NPSO, indicating that CIP uptakes by 10%MnTe@β and 20%MnTe@β are contingent upon the availability of CIP and the active sites on the 10%MnTe@β and 20%MnTe@β surfaces [58,59]. Figure 6c,d, along with the results presented in Table 4, demonstrate the outcomes of the CIP sorption control mechanism. CIP diffusion toward 10%MnTe@β and penetration through the interior layers of 10%MnTe@β both contributed to controlling CIP sorption, as indicated by the R2 values for LFD and IPD, which are nearly typical. The results gathered in Table 4 demonstrate that LFM regulates CIP sorption onto the 20%MnTe@β nanocomposite. This indicates that the diffusion of CIP toward the 20%MnTe@β activate sites was slower compared to CIP penetration into 20%MnTe@β interior shells, demonstrating the high affinity of CIP toward the 20%MnTe@β nanocomposite [60].

3.5. Sorption Equilibria

The impact of CIP concentration on its sorption by 10%MnTe@β and 20%MnTe@β was investigated (Figure 7a,b). The qt increased proportionally as the concentration rose from 10 to 50 mg L−1, where the line showed significant inflation, indicating the sorbent’s unsuitability to treat more than 30 mg L−1 with such a solution-sorbent ratio (12:5). Raising the CIP solution temperature from 293 to 323K decreased the sorption efficiency significantly, demonstrating that CIP adsorption by 10%MnTe@β and 20%MnTe@β was exothermic.
Figure 7c,d illustrate the collective nonlinear plots of LIM and FIM regarding CIP adsorption on 10%MnTe@β and 20%MnTe@β, and their results are presented in Table 6. The outputs from the fitting indicated that the sorption of CIP onto 10%MnTe@β was more in line with FIM; hence, the process of CIP adsorption onto this sorbent was characterized by multilayer adsorption. Additionally, adsorption was favorable, as the 1/n values fell within the range from 0.0 to 1.0. Conversely, the LIM model was shown to be the most appropriate for CIP adsorption onto 20%MnTe@β, and it predicted that 1.0 g of it could uptake 136.0 mg CIP. In addition, the fact that 1/n was above one suggested the unfavorability of multilayered CIP sorption onto 20%MnTe@β [61].
Figure 7e,f present the thermodynamic plot for 10%MnTe@β and 20%MnTe@β, and the computed parameters are shown in Table 6. The negative ΔH° and rising ΔG° values as the temperature increased indicated exothermic CIP sorption on 10%MnTe@β and 20%MnTe@β, which tended towards spontaneity as the temperature decreases. The increase in ΔG° values with a decreasing starting concentration is encouraging for the application of both 10%MnTe@β and 20%MnTe@β in water treatment, particularly in scenarios with low pollutant concentrations, with a preference for 20%MnTe@β. Moreover, the ΔG° values of less than 40.0 kJ mol−1 indicated that the CIP was eliminated by 10%MnTe@β and 20%MnTe@β via physisorption process.

3.6. Application of 20%MnTe@β to Remediate Natural Water

As suggested by the contact time outcome, 20%MnTe@β was selected for this test. The total dissolved solids (TDS) for the natural samples (GW and SW) were 1.59 and 33.62 g L−1, with pH values of 7.8 and 8.1, respectively. Thermodynamic data indicated a preferred CIP removal as the temperature was reduced; hence, incorporating a heating phase to enhance sorption was not required. The pH of GW and SW was set at the optimal pH level (6.0). Figure 8a presents the outcomes of the 20%MnTe@β effective eradication of CIP as a model pollutant from GW and SW. The removal percentages of CIP from GW and SW were 96.93% and 92.78% for 5 mg L−1 spiked samples and 92.64% and 89.38% for 10 mg L−1 concentration, indicating that no significant alterations occurred due to the presence of other natural water constituents. The relative reduction in CIP removal percentage in SW compared to GW may be attributed to the higher salt content in SW, which potentially obstructs specific active sorption sites on 20%MnTe@β and/or hinders the diffusion of the CIP solution to the surface of 20%MnTe@β.

3.7. Regeneration and Reusability

The reusability of 20%MnTe@β was investigated using 50 mL of the 5.0 mg L−1 CIP concentration. Following the initial cycle, further adsorption experiments were conducted using the regenerated 20%MnTe@β adsorbent. The preparation of the 20%MnTe@β nanocomposite for reuse involved washing with 10 mL of 0.2 M NaOH, followed by two washes with 10 mL of ethanol each, and then drying at 80 °C for 2.0 h to be applied in the next round. Figure 8 illustrates the 20%MnTe@β reusability outcomes in removing CIP; the effectiveness of the 20%MnTe@β nanocomposite in the fourth cycle was 92.8%, which represents a slight decrease in the CIP adsorption percentage. These findings suggest that the 20%MnTe@β nanocomposites can be regenerated and subjected to multiple uses with suitable reuse, at least four times, without significant loss in performance. Although tellurite salts were known to be sparingly soluble in ethanol and/or cold water, it is possible to attribute the slight decrease in the 20%MnTe@β to a partial loss of doping material.

4. Conclusions

The present investigation employed a straightforward pyrolysis technique to produce β-C3N4 from guanidine hydrochloride, subsequently doped with manganese tellurite, resulting in the creation of the novel nanohybrids 10%MnTe@β and 20%MnTe@β. The XRD patterns confirmed the β-C3N4 phase and the success of incorporating MnTeO3 into β-C3N4. SEM pictures showed particle sizes of 27.97 nm, 32.08 nm, and 40.40 nm for β-C3N4, 10%MnTe@β, and 20%MnTe@β, respectively. The surface areas were 85.86, 97.40, and 109.54 m2 g−1, respectively. This size–surface area contradiction for 10%MnTe@β and 20%MnTe@β could be attributed to the 20%MnTe@β particles being sufficiently large enough to separate the β-C3N4 layers, exposing additional plate surfaces. In adsorbing CIP from aqueous solutions, β-C3N4 and 10%MnTe@β reached equilibrium at 60 min, while 20%MnTe@β required 45 min, with qt values of 27.30, 48.88, and 77.41 mg g−1, respectively. It is worth noting that, with a pH of 6.0, 10%MnTe@β and 20%MnTe@β worked more effectively. In the kinetic studies, more agreement was found with the NPSO for CIP sorption on 10%MnTe@β and 20%MnTe@β. Moreover, the LFM controlled CIP sorption onto 20%MnTe@β, indicating a strong affinity towards CIP; on the other hand, both LFM and IPD participated in influencing CIP sorption onto 10%MnTe@β. The sorption equilibria outputs suggested a superior alignment with the Freundlich and Langmuir models for eliminating CIP by 10%MnTe@β and 20%MnTe@β, respectively. The multilayered sorption mechanism might explain the higher qt value of 20%MnTe@β. The thermodynamic results demonstrated that CIP sorption by 10%MnTe@β and 20%MnTe@β was exothermic and became more spontaneous as the temperature in the system dropped. Applying 20%MnTe@β as the most effective sorbent to natural GW and SW spiked by 5 and 10 mg L−1 CIP, and average efficiencies of 94.8 and 91.08 were obtained. Throughout four cycles, the average efficiency of the 20%MnTe@β regeneration–reusability process was 95.14%. These outcomes of the 20%MnTe@β performance, with no need for heating, may nominate 20%MnTe@β as a practical and feasible sorbent for remedying CIP-contaminated water.

Author Contributions

Conceptualization, M.R.E. and B.Y.A.; Methodology, M.S.; Software, A.A. and R.R.; Validation, T.G.I. and R.R.; Formal analysis, N.Y.E., M.S. and R.R.; Investigation, A.A.; Writing—original draft, M.R.E. and T.G.I.; Writing—review & editing, B.Y.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2502).

Data Availability Statement

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

Conflicts of Interest

The authors declare that they have no conflicts of interest.

References

  1. Kishor, R.; Purchase, D.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Chandra, R.; Bharagava, R.N. Ecotoxicological and health concerns of persistent coloring pollutants of textile industry wastewater and treatment approaches for environmental safety. J. Environ. Chem. Eng. 2021, 9, 105012. [Google Scholar] [CrossRef]
  2. Alsukaibi, A.K. Various approaches for the detoxification of toxic dyes in wastewater. Processes 2022, 10, 1968. [Google Scholar] [CrossRef]
  3. Stec, J.; Kosikowska, U.; Mendrycka, M.; Stępień-Pyśniak, D.; Niedźwiedzka-Rystwej, P.; Bębnowska, D.; Hrynkiewicz, R.; Ziętara-Wysocka, J.; Grywalska, E. Opportunistic pathogens of recreational waters with emphasis on antimicrobial resistance—A possible subject of human health concern. Int. J. Environ. Res. Public Health 2022, 19, 7308. [Google Scholar] [CrossRef] [PubMed]
  4. Obaideen, K.; Shehata, N.; Sayed, E.T.; Abdelkareem, M.A.; Mahmoud, M.S.; Olabi, A. The role of wastewater treatment in achieving sustainable development goals (SDGs) and sustainability guideline. Energy Nexus 2022, 7, 100112. [Google Scholar] [CrossRef]
  5. Montes-Grajales, D.; Fennix-Agudelo, M.; Miranda-Castro, W. Occurrence of personal care products as emerging chemicals of concern in water resources: A review. Sci. Total Environ. 2017, 595, 601–614. [Google Scholar] [CrossRef] [PubMed]
  6. Klein, E.Y.; van Boeckel, T.P.; Martinez, E.M.; Pant, S.; Gandra, S.; Levin, S.A.; Goossens, H.; Laxminarayan, R. Global increase and geographic convergence in antibiotic consumption between 2000 and 2015. Proc. Natl. Acad. Sci. USA 2018, 115, E3463–E3470. [Google Scholar] [CrossRef] [PubMed]
  7. Almomani, F.; Bhosale, R.; Kumar, A.; Khraisheh, M. Potential use of solar photocatalytic oxidation in removing emerging pharmaceuticals from wastewater: A pilot plant study. Sol. Energy 2018, 172, 128–140. [Google Scholar] [CrossRef]
  8. Levy, S.B. The Antibiotic Paradox: How Miracle Drugs Are Destroying the Miracle; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  9. Yan, Y.; Pengmao, Y.; Xu, X.; Zhang, L.; Wang, G.; Jin, Q.; Chen, L. Migration of antibiotic ciprofloxacin during phytoremediation of contaminated water and identification of transformation products. Aquat. Toxicol. 2020, 219, 105374. [Google Scholar] [CrossRef] [PubMed]
  10. Hartmann, A.; Golet, E.; Gartiser, S.; Alder, A.; Koller, T.; Widmer, R. Primary DNA damage but not mutagenicity correlates with ciprofloxacin concentrations in German hospital wastewaters. Arch. Environ. Contam. Toxicol. 1999, 36, 115–119. [Google Scholar] [CrossRef] [PubMed]
  11. Fick, J.; Söderström, H.; Lindberg, R.H.; Phan, C.; Tysklind, M.; Larsson, D.J. Contamination of surface, ground, and drinking water from pharmaceutical production. Environ. Toxicol. Chem. 2009, 28, 2522–2527. [Google Scholar] [CrossRef] [PubMed]
  12. Yan, Y.; Xu, X.; Shi, C.; Yan, W.; Zhang, L.; Wang, G. Ecotoxicological effects and accumulation of ciprofloxacin in Eichhornia crassipes under hydroponic conditions. Environ. Sci. Pollut. Res. 2019, 26, 30348–30355. [Google Scholar] [CrossRef] [PubMed]
  13. Chopra, S.; Kumar, D. Pharmaceuticals and personal care products (PPCPs) as emerging environmental pollutants: Toxicity and risk assessment. Adv. Anim. Biotechnol. Its Appl. 2018, 9, 337–353. [Google Scholar]
  14. Ziylan-Yavas, A.; Santos, D.; Flores, E.M.M.; Ince, N.H. Pharmaceuticals and personal care products (PPCPs): Environmental and public health risks. Environ. Prog. Sustain. Energy 2022, 41, e13821. [Google Scholar] [CrossRef]
  15. Moreira, N.F.; Sousa, J.M.; Macedo, G.; Ribeiro, A.R.; Barreiros, L.; Pedrosa, M.; Faria, J.L.; Pereira, M.F.R.; Castro-Silva, S.; Segundo, M.A. Photocatalytic ozonation of urban wastewater and surface water using immobilized TiO2 with LEDs: Micropollutants, antibiotic resistance genes and estrogenic activity. Water Res. 2016, 94, 10–22. [Google Scholar] [CrossRef] [PubMed]
  16. Polesel, F.; Andersen, H.R.; Trapp, S.; Plósz, B.G. Removal of Antibiotics in Biological Wastewater Treatment Systems—A Critical Assessment Using the Activated Sludge Modeling Framework for Xenobiotics (ASM-X). Environ. Sci. Technol. 2016, 50, 10316–10334. [Google Scholar] [CrossRef] [PubMed]
  17. Biancullo, F.; Moreira, N.F.; Ribeiro, A.R.; Manaia, C.M.; Faria, J.L.; Nunes, O.C.; Castro-Silva, S.M.; Silva, A.M. Heterogeneous photocatalysis using UVA-LEDs for the removal of antibiotics and antibiotic resistant bacteria from urban wastewater treatment plant effluents. Chem. Eng. J. 2019, 367, 304–313. [Google Scholar] [CrossRef]
  18. Tang, Z.; Lin, X.; Chen, Y.; Pan, Y.; Yang, Y.; Mondal, A.K.; Yu, M.; Wu, H. Preparation of mussel-inspired polydopamine-functionalized TEMPO-oxidized cellulose nanofiber-based composite aerogel as reusable adsorbent for water treatment. Ind. Crops Prod. 2023, 206, 117735. [Google Scholar] [CrossRef]
  19. Akhtar, M.S.; Ali, S.; Zaman, W. Innovative adsorbents for pollutant removal: Exploring the latest research and applications. Molecules 2024, 29, 4317. [Google Scholar] [CrossRef] [PubMed]
  20. Alnajrani, M.N.; Alsager, O.A. Removal of antibiotics from water by polymer of intrinsic microporosity: Isotherms, kinetics, thermodynamics, and adsorption mechanism. Sci. Rep. 2020, 10, 794. [Google Scholar] [CrossRef] [PubMed]
  21. Kong, Y.; Wang, L.; Ge, Y.; Su, H.; Li, Z. Lignin xanthate resin–bentonite clay composite as a highly effective and low-cost adsorbent for the removal of doxycycline hydrochloride antibiotic and mercury ions in water. J. Hazard. Mater. 2019, 368, 33–41. [Google Scholar] [CrossRef] [PubMed]
  22. Sharma, G.; Thakur, B.; Kumar, A.; Sharma, S.; Naushad, M.; Stadler, F.J. Gum acacia-cl-poly (acrylamide)@carbon nitride nanocomposite hydrogel for adsorption of ciprofloxacin and its sustained release in artificial ocular solution. Macromol. Mater. Eng. 2020, 305, 2000274. [Google Scholar] [CrossRef]
  23. Cao, S.; Low, J.; Yu, J.; Jaroniec, M. Polymeric photocatalysts based on graphitic carbon nitride. Adv. Mater. 2015, 27, 2150–2176. [Google Scholar] [CrossRef] [PubMed]
  24. Wu, K.; Chen, D.; Fang, J.; Wu, S.; Yang, F.; Zhu, X.; Fang, Z. One-step synthesis of sulfur and tungstate co-doped porous g-C3N4 microrods with remarkably enhanced visible-light photocatalytic performances. Appl. Surf. Sci. 2018, 462, 991–1001. [Google Scholar] [CrossRef]
  25. Li, K.; Xue, D. Hardness of group IVA and IVB nitrides. Phys. Scr. 2010, 2010, 014073. [Google Scholar] [CrossRef]
  26. Han, L.; Khalil, A.M.; Wang, J.; Chen, Y.; Li, F.; Chang, H.; Zhang, H.; Liu, X.; Li, G.; Jia, Q. Graphene-boron nitride composite aerogel: A high efficiency adsorbent for ciprofloxacin removal from water. Sep. Purif. Technol. 2021, 278, 119605. [Google Scholar] [CrossRef]
  27. Hayat, A.; Sohail, M.; Ali Shah Syed, J.; Al-Sehemi, A.G.; Mohammed, M.H.; Al-Ghamdi, A.A.; Taha, T.; Salem AlSalem, H.; Alenad, A.M.; Amin, M.A. Recent advancement of the current aspects of g-C3N4 for its photocatalytic applications in sustainable energy system. Chem. Rec. 2022, 22, e202100310. [Google Scholar] [CrossRef] [PubMed]
  28. Feng, D.; Cheng, Y.; He, J.; Zheng, L.; Shao, D.; Wang, W.; Wang, W.; Lu, F.; Dong, H.; Liu, H. Enhanced photocatalytic activities of g-C3N4 with large specific surface area via a facile one-step synthesis process. Carbon 2017, 125, 454–463. [Google Scholar] [CrossRef]
  29. Long, B.; Lin, J.; Wang, X. Thermally-induced desulfurization and conversion of guanidine thiocyanate into graphitic carbon nitride catalysts for hydrogen photosynthesis. J. Mater. Chem. A 2014, 2, 2942–2951. [Google Scholar] [CrossRef]
  30. Wu, P.; Wang, J.; Zhao, J.; Guo, L.; Osterloh, F.E. Structure defects in gC 3 N 4 limit visible light driven hydrogen evolution and photovoltage. J. Mater. Chem. A 2014, 2, 20338–20344. [Google Scholar] [CrossRef]
  31. Almufarij, R.S.; Abdulkhair, B.Y.; Salih, M. Fast-simplistic fabrication of MoO3@ Al2O3-MgO triple nanocomposites for efficient elimination of pharmaceutical contaminants. Results Chem. 2024, 7, 101281. [Google Scholar] [CrossRef]
  32. Khairy, M.; Algethami, F.K.; Alotaibi, A.N.; Almufarij, R.S.; Abdulkhair, B.Y. Enhancing the conductivity and dielectric characteristics of bismuth oxyiodide via activated carbon doping. Molecules 2024, 29, 2082. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, Y.; Wang, Q.; Chen, Q.; Li, X.; Li, Y.; Kang, M.; Li, Q.; Wang, J. A new boron modified carbon nitride metal-free catalyst for the cycloaddition of CO2 and bisepoxides. Appl. Catal. A Gen. 2024, 675, 119615. [Google Scholar] [CrossRef]
  34. Zhou, P.; Yang, Q.; Tang, Y.; Cai, Y. Recent trends in metal-doped carbon nitride-catalyzed heterogeneous light-driven organic transformations. Nano Trends 2023, 4, 100019. [Google Scholar] [CrossRef]
  35. Umar, M.; Ajaz, H.; Javed, M.; Mansoor, S.; Iqbal, S.; Rauf, A.; Alhujaily, A.; Awwad, N.S.; Ibrahium, H.A.; Althobiti, R.A. Design of a highly efficient heterostructure of transition metal tellurides with outstanding photocatalytic and antimicrobial potential. J. Saudi Chem. Soc. 2023, 27, 101760. [Google Scholar] [CrossRef]
  36. Srinandhini, S.; Padmanaban, A.; Sheeja, R.; Narayanan, V. Synthesis and characterization of MnTe nanoparticles and its photocatalytic activity. J. Indian Chem. Soc. 2019, 96, 147–149. [Google Scholar]
  37. Hasan, G.G.; Laouini, S.E.; Osman, A.I.; Bouafia, A.; Althamthami, M.; Meneceur, S.; Kir, I.; Mohammed, H.; Lumbers, B.; Rooney, D.W. Nanostructured Mn@NiO composite for addressing multi-pollutant challenges in petroleum-contaminated water. Environ. Sci. Pollut. Res. 2024, 31, 44254–44271. [Google Scholar] [CrossRef] [PubMed]
  38. Eder, F.; Weil, M.; Missen, O.P.; Kolitsch, U.; Libowitzky, E. The Family of MII3(TeIVO3)2(OH)2 (M = Mg, Mn, Co, Ni) Compounds—Prone to Inclusion of Foreign Components into Large Hexagonal Channels. Crystals 2022, 12, 1380. [Google Scholar] [CrossRef]
  39. Elamin, M.R.; Abdulkhair, B.Y.; Elzupir, A.O. Removal of ciprofloxacin and indigo carmine from water by carbon nanotubes fabricated from a low-cost precursor: Solution parameters and recyclability. Ain Shams Eng. J. 2023, 14, 101844. [Google Scholar] [CrossRef]
  40. Aliakbari, A.; Ghamsari, M.S.; Mozdianfard, M. β-Carbon nitride nanoflake with enhanced visible light emission. Opt. Mater. 2020, 107, 110036. [Google Scholar] [CrossRef]
  41. Yin, L.-W.; Li, M.-S.; Liu, Y.-X.; Sui, J.-L.; Wang, J.-M. Synthesis of beta carbon nitride nanosized crystal through mechanochemicalreaction. J. Phys. Condens. Matter 2003, 15, 309. [Google Scholar] [CrossRef]
  42. Chen, Y.A.; Guo, L.; Wang, E. alpha beta Experimental evidence for alpha-and beta-phases of pure crystalline C3N4 in films deposited on nickel substrates. Philos. Mag. Lett. 1997, 75, 155–162. [Google Scholar] [CrossRef]
  43. He, Y.; Guo, Y.; Yu, Y.; Luo, P.; Qiu, H.; Huang, G. Facile Synthesis of Manganese Tellurite Nanoparticles for Magnetic Resonance Imaging-Guided Photothermal Therapy. Part. Part. Syst. Charact. 2021, 38, 2100038. [Google Scholar] [CrossRef]
  44. Ganesamurthi, J.; Keerthi, M.; Chen, S.-M.; Shanmugam, R. Electrochemical detection of thiamethoxam in food samples based on Co3O4 Nanoparticle@Graphitic carbon nitride composite. Ecotoxicol. Environ. Saf. 2020, 189, 110035. [Google Scholar] [CrossRef] [PubMed]
  45. Bekele, E.A.; Korsa, H.A.; Desalegn, Y.M. Electrolytic synthesis of γ-Al2O3 nanoparticle from aluminum scrap for enhanced methylene blue adsorption: Experimental and RSM modeling. Sci. Rep. 2024, 14, 16957. [Google Scholar] [CrossRef] [PubMed]
  46. Song, S.; Liu, Z.; Zhang, J.; Jiao, C.; Ding, L.; Yang, S. Synthesis and adsorption properties of novel bacterial cellulose/graphene oxide/attapulgite materials for Cu and Pb Ions in aqueous solutions. Materials 2020, 13, 3703. [Google Scholar] [CrossRef] [PubMed]
  47. Ajduković, M.; Stevanović, G.; Marinović, S.; Mojović, Z.; Banković, P.; Radulović, K.; Jović-Jovičić, N. Ciprofloxacin adsorption onto a smectite–chitosan-derived nanocomposite obtained by hydrothermal synthesis. Water 2023, 15, 2608. [Google Scholar] [CrossRef]
  48. Liu, T.; Liu, W.; Li, X.; Wang, H.; Lan, Y.; Zhang, S.; Wang, Y.; Liu, H. Effect of environmental factors on adsorption of ciprofloxacin from wastewater by microwave alkali modified fly ash. Sci. Rep. 2024, 14, 19831. [Google Scholar] [CrossRef] [PubMed]
  49. Rodríguez-López, L.; Santás-Miguel, V.; Cela-Dablanca, R.; Núñez-Delgado, A.; Álvarez-Rodríguez, E.; Rodríguez-Seijo, A.; Arias-Estévez, M. Sorption of Antibiotics in Agricultural Soils as a Function of pH. Span. J. Soil Sci. 2024, 14, 12402. [Google Scholar] [CrossRef]
  50. Almufarij, R.S.; Abdulkhair, B.Y.; Salih, M.; Aldosari, H.; Aldayel, N.W. Optimization, nature, and mechanism investigations for the adsorption of ciprofloxacin and malachite green onto carbon nanoparticles derived from low-cost precursor via a green route. Molecules 2022, 27, 4577. [Google Scholar] [CrossRef] [PubMed]
  51. Salih, M.; Abdulkhair, B.Y.; Alotaibi, M. Insight into the Adsorption Behavior of Carbon Nanoparticles Derived from Coffee Skin Waste for Remediating Water Contaminated with Pharmaceutical Ingredients. Chemistry 2023, 5, 1870–1881. [Google Scholar] [CrossRef]
  52. Gu, C.; Karthikeyan, K. Sorption of the antimicrobial ciprofloxacin to aluminum and iron hydrous oxides. Environ. Sci. Technol. 2005, 39, 9166–9173. [Google Scholar] [CrossRef] [PubMed]
  53. Rakshit, S.; Sarkar, D.; Elzinga, E.J.; Punamiya, P.; Datta, R. Mechanisms of ciprofloxacin removal by nano-sized magnetite. J. Hazard. Mater. 2013, 246, 221–226. [Google Scholar] [CrossRef] [PubMed]
  54. Elsaim, M.H. Removal of ciprofloxacin hydrochloride from aqueous solution by pomegranate peel grown in Alziedab agricultural scheme-River Nile State, Sudan. Biochemistry 2017, 5, 89–96. [Google Scholar]
  55. Genç, N.; Dogan, E.C. Adsorption kinetics of the antibiotic ciprofloxacin on bentonite, activated carbon, zeolite, and pumice. Desalination Water Treat. 2015, 53, 785–793. [Google Scholar] [CrossRef]
  56. Avcı, A.; İnci, İ.; Baylan, N. A comparative adsorption study with various adsorbents for the removal of ciprofloxacin hydrochloride from water. Water Air Soil Pollut. 2019, 230, 250. [Google Scholar] [CrossRef]
  57. Elamin, M.R.; Abdulkhair, B.Y.; Algethami, F.K.; Khezami, L. Linear and nonlinear investigations for the adsorption of paracetamol and metformin from water on acid-treated clay. Sci. Rep. 2021, 11, 13606. [Google Scholar] [CrossRef] [PubMed]
  58. Revellame, E.D.; Fortela, D.L.; Sharp, W.; Hernandez, R.; Zappi, M.E. Adsorption kinetic modeling using pseudo-first order and pseudo-second order rate laws: A review. Clean. Eng. Technol. 2020, 1, 100032. [Google Scholar] [CrossRef]
  59. Tran, Q.T.; Do, T.H.; Ha, X.L.; Nguyen, H.P.; Nguyen, A.T.; Ngo, T.C.Q.; Chau, H.D. Study of the ciprofloxacin adsorption of activated carbon prepared from mangosteen peel. Appl. Sci. 2022, 12, 8770. [Google Scholar] [CrossRef]
  60. An, B. Cu(II) and As(V) adsorption kinetic characteristic of the multifunctional amino groups in chitosan. Processes 2020, 8, 1194. [Google Scholar] [CrossRef]
  61. Ibrahim, T.G.; Almufarij, R.S.; Abdulkhair, B.Y.; Ramadan, R.S.; Eltoum, M.S.; Abd Elaziz, M.E. A Thorough Examination of the Solution Conditions and the Use of Carbon Nanoparticles Made from Commercial Mesquite Charcoal as a Successful Sorbent for Water Remediation. Nanomaterials 2023, 13, 1485. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The XRD spectra and (b) FTIR spectra of the fabricated β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials.
Figure 1. (a) The XRD spectra and (b) FTIR spectra of the fabricated β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials.
Processes 13 02357 g001
Figure 2. The SEM images of (a) β-C3N4, (b) 10%MnTe@β, and (c) 20%MnTe@β; the EDX spectra of (d) β-C3N4, (e) 10%MnTe@β, and (f) 20%MnTe@β.
Figure 2. The SEM images of (a) β-C3N4, (b) 10%MnTe@β, and (c) 20%MnTe@β; the EDX spectra of (d) β-C3N4, (e) 10%MnTe@β, and (f) 20%MnTe@β.
Processes 13 02357 g002aProcesses 13 02357 g002b
Figure 3. The elemental mapping of (a) carbon and nitrogen in β-C3N4; (b) and (c) carbon nitrogen, oxygen, tellurium, and manganese in 10%MnTe@β and 20%MnTe@β, respectively.
Figure 3. The elemental mapping of (a) carbon and nitrogen in β-C3N4; (b) and (c) carbon nitrogen, oxygen, tellurium, and manganese in 10%MnTe@β and 20%MnTe@β, respectively.
Processes 13 02357 g003
Figure 4. The surface characteristics of (a) β-C3N4, (b) 10%MnTe@β, and (c) 20%MnTe@β.
Figure 4. The surface characteristics of (a) β-C3N4, (b) 10%MnTe@β, and (c) 20%MnTe@β.
Processes 13 02357 g004
Figure 5. (a) the contact time results of β-C3N4, 10%MnTe@β, and 20%MnTe@β; (b) the pH effect on CIP sorption by β-C3N4, 10%MnTe@β, and 20%MnTe@β.
Figure 5. (a) the contact time results of β-C3N4, 10%MnTe@β, and 20%MnTe@β; (b) the pH effect on CIP sorption by β-C3N4, 10%MnTe@β, and 20%MnTe@β.
Processes 13 02357 g005
Figure 6. (a) NPFO, (b) NPSO, (c) LFM, and (d) IPM plots of CIP removal onto 10%MnTe@β and 20%MnTe@β from 100 mg L−1 solution.
Figure 6. (a) NPFO, (b) NPSO, (c) LFM, and (d) IPM plots of CIP removal onto 10%MnTe@β and 20%MnTe@β from 100 mg L−1 solution.
Processes 13 02357 g006
Figure 7. (a,b) the concentration–temperature impacts on CIP sorption by 10%MnTe@β and 20%MnTe@β; (c,d) collective LIM and FIM fittings for CIP sorption by 10%MnTe@β and 20%MnTe@β at 293K; (e,f) the thermodynamic plot for 10%MnTe@β and 20%MnTe@β.
Figure 7. (a,b) the concentration–temperature impacts on CIP sorption by 10%MnTe@β and 20%MnTe@β; (c,d) collective LIM and FIM fittings for CIP sorption by 10%MnTe@β and 20%MnTe@β at 293K; (e,f) the thermodynamic plot for 10%MnTe@β and 20%MnTe@β.
Processes 13 02357 g007
Figure 8. (a) Application of 20%MnTe@β in the decontamination of natural water samples spiked with 5 and 10 mg L−1 CIP concentrations and (b) the reusability of 20%MnTe@β.
Figure 8. (a) Application of 20%MnTe@β in the decontamination of natural water samples spiked with 5 and 10 mg L−1 CIP concentrations and (b) the reusability of 20%MnTe@β.
Processes 13 02357 g008
Table 1. The lattice parameters of the β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials.
Table 1. The lattice parameters of the β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials.
Phase NameD (nm)A (Å)B (Å)C (Å)α (deg)β (deg)γ (deg)
β-C3N40.657.3437.3432.990.0090.00120.00
MnTeO3 of 10%MnTe@β19.013.9210.2311.7890.0092.990.00
β-C3N4 of 10%MnTe@β9.074.4594.4596.47090.0090.00120.00
MnTeO3 of 20%MnTe@β8.06.427.554.68390.0090.0090.00
β-C3N4 of 20%MnTe@β3.286.1656.1652.69490.0090.00120.00
Table 2. The FT-IR results of β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials.
Table 2. The FT-IR results of β-C3N4, 10%MnTe@β, and 20%MnTe@β nanomaterials.
The Absorbing Groupβ-C3N410%MnTe@β20%MnTe@β
Adsorbed moisture O-H-~3444 cm−1~3444 cm−1
N—H stretching3080–3280 cm−13130–3308 cm−13120–3334 cm−1
C-H stretching2933 cm−12933 cm−12933 cm−1
C≡N stretching,2368 cm−12374 cm−12364 cm−1
C—N stretching1234–1642 cm−11234–1642 cm−11234–1642 cm−1
Out-of-plane bending of the triazine ring803–812 cm−1803–812 cm−1803–812 cm−1
Te—O stretching-755 cm−1760 cm−1
Mn—O stretching-584 cm−1575 cm−1
Table 3. The surface characteristics β-C3N4, 10%MnTe@β, and 20%MnTe@β.
Table 3. The surface characteristics β-C3N4, 10%MnTe@β, and 20%MnTe@β.
SorbentSA (m2 g−1)PD (Å)PV (cm3 g−1)
β-C3N485.8612.610.29
10%MnTe@β97.4013.990.22
20%MnTe@β109.5410.420.37
Table 5. Kinetic results of CIP sorption onto 10%MnTe@β and 20%MnTe@β nanocomposites.
Table 5. Kinetic results of CIP sorption onto 10%MnTe@β and 20%MnTe@β nanocomposites.
Adsorption Rate Order
Sorbentqmax exp
(mg. g−1)
NPFONPSO
qe (mg. g−1)K1R2X2RSSqe (mg. g−1)K2R2X2RSS
10%MnTe@β48.8945.1580.65550.96828.6318260.4227347.85550.673440.99192.1965715.376
20%MnTe@β77.4174.1920.51170.972919.09689133.6782377.27661.25120.99443.963627.745
Adsorption rate mechanism
SorbentIPMLFM
KIP (mg. g−1 min1/2)C (mg. g−1)R2RSSKLF (min−1)R2RSS
10%MnTe@β2.66629.1850.84542.642330.067470.8502.2715
20%MnTe@β2.9788857.35520.83457.767730.07250.9530.74024
Table 6. Isotherms and thermodynamic outputs of CIP sorption by 10%MnTe@β and 20%MnTe@β.
Table 6. Isotherms and thermodynamic outputs of CIP sorption by 10%MnTe@β and 20%MnTe@β.
Adsorption Isotherms
Isotherm modelLangmuirFreundlich
SorbentR2KLqmR2Kf1/n
10%MnTe@β0.993970.01606136.0070.996033.168230.77782
20%MnTe@β0.97793.1375 × 10571715.20.968351.63751.10947
Thermodynamic results 10%MnTe@β
Conc. (mg L−1)ΔH°ΔS°ΔG°(293 K)ΔG°-303 KΔG°-313 KΔG°-323 KR2
10.0−7.698−0.025−0.479−0.2320.0140.0140.999
20.0−9.455−0.0320.0570.3820.7061.0310.859
30.0−8.939−0.0320.3750.6931.0111.3290.603
50.0−11.770−0.0441.0131.4491.8852.3220.995
Thermodynamic results 20%MnTe@β
Conc. (mg L−1)ΔH°ΔS°ΔG°(293 K)ΔG°-303 KΔG°-313 KΔG°-323 KR2
10.0−9.840−0.032−0.428−0.1070.2140.2140.853
20.0−9.782−0.0340.1920.5320.8731.210.872
30.0−12.206−0.0440.6321.0701.5081.9460.952
50.0−3.323−0.0161.2211.3761.5311.6860.937
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

Elamin, M.R.; Elamin, N.Y.; Ibrahim, T.G.; Salih, M.; Albadri, A.; Ramadan, R.; Abdulkhair, B.Y. Facile Synthesis of β-C3N4 and Its Novel MnTeO3 Nanohybrids for Remediating Water Contaminated by Pharmaceuticals. Processes 2025, 13, 2357. https://doi.org/10.3390/pr13082357

AMA Style

Elamin MR, Elamin NY, Ibrahim TG, Salih M, Albadri A, Ramadan R, Abdulkhair BY. Facile Synthesis of β-C3N4 and Its Novel MnTeO3 Nanohybrids for Remediating Water Contaminated by Pharmaceuticals. Processes. 2025; 13(8):2357. https://doi.org/10.3390/pr13082357

Chicago/Turabian Style

Elamin, Mohamed R., Nuha Y. Elamin, Tarig G. Ibrahim, Mutaz Salih, Abuzar Albadri, Rasha Ramadan, and Babiker Y. Abdulkhair. 2025. "Facile Synthesis of β-C3N4 and Its Novel MnTeO3 Nanohybrids for Remediating Water Contaminated by Pharmaceuticals" Processes 13, no. 8: 2357. https://doi.org/10.3390/pr13082357

APA Style

Elamin, M. R., Elamin, N. Y., Ibrahim, T. G., Salih, M., Albadri, A., Ramadan, R., & Abdulkhair, B. Y. (2025). Facile Synthesis of β-C3N4 and Its Novel MnTeO3 Nanohybrids for Remediating Water Contaminated by Pharmaceuticals. Processes, 13(8), 2357. https://doi.org/10.3390/pr13082357

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