Next Article in Journal
Dynamic Analysis of a Wiper Blade in Consideration of Attack Angle and Clarification of the Jumping Phenomenon
Next Article in Special Issue
Biomass Fly Ash Self-Hardened Adsorbent Monoliths for Methylene Blue Removal from Aqueous Solutions
Previous Article in Journal
Dose-Area Product Determination and Beam Monitor Calibration for the Fixed Beam of the Shanghai Advanced Proton Therapy Facility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sorption Potential of Different Forms of TiO2 for the Removal of Two Anticancer Drugs from Water

1
Faculty of Chemical Engineering and Technology, University of Zagreb, Marulićev trg 19, 10000 Zagreb, Croatia
2
Faculty of Metallurgy, University of Zagreb, Aleja Narodnih Heroja 3, 44000 Sisak, Croatia
3
Istrian University of Applied Sciences, Riva 6, 52100 Pula, Croatia
*
Author to whom correspondence should be addressed.
Submission received: 23 March 2022 / Revised: 13 April 2022 / Accepted: 15 April 2022 / Published: 19 April 2022

Abstract

:
Anticancer drugs pose a potential risk to the environment due to their significant consumption and biological effect even at low concentrations. They can leach into soils and sediments, wastewater, and eventually into drinking water supplies. Many conventional technologies with more effective advanced oxidation processes such as photocatalysis are being extensively studied to find an economical and environmentally friendly solution for the removal of impurities from wastewater as the main source of these pharmaceuticals. Since it is impossible to treat water by photocatalysis if there is no sorption of a contaminant on the photocatalyst, this work investigated the amount of imatinib and crizotinib sorbed from an aqueous medium to different forms of photocatalyst. In addition, based on the sorption affinity studied, the applicability of sorption as a simpler and less costly process was tested in general as a potential route to remove imatinib and crizotinib from water. Their sorption possibility was investigated determining the maximum of sorption, influence of pH, ionic strength, temperature, and sorbent dosage in form of the suspension and immobilized on the fiberglass mesh with only TiO2 and in combination with TiO2/carbon nanotubes. The sorption isotherm data fitted well the linear, Freundlich, and Langmuir model for both pharmaceuticals. An increasing trend of sorption coefficients Kd was observed in the pH range of 5–9 with CRZ, showing higher sorption affinity to all TiO2 forms, which was supported by KF values higher than 116 (μg/g)(mL/μg)1/n. The results also show a positive correlation between Kd and temperature as well as sorbent dosage for both pharmaceuticals, while CRZ sorbed less at higher salt concentration. The kinetic data were best described with a pseudo-second-order model (R2 > 0.995).

1. Introduction

Environmental pollution is a ubiquitous problem that humanity has been dealing with for many years. When considering the impact of various factors on the environment, more and more attention is paid to pharmaceuticals. These are biologically active compounds used to treat and prevent diseases, alleviate of symptoms, etc., in human and veterinary medicine [1,2]. The modernization of technology and science has led to the promotion of quality of life through the increased production of medicines whose consumption is increasing every day [3]. Since information about their fate is still limited and there are no environmental laws, they can reach different sources in the environment and undergo processes, such as sorption, photolysis, hydrolysis or even biodegradation. These processes can often produce transformation or degradation products that are more toxic than the parent compound. The fact that they are still present in the environment at low concentrations is not reassuring due to their unknown long-term effects on the aquatic and terrestrial world [1,4].
In addition to commonly studied classes of pharmaceuticals such as antibiotics [5,6,7,8,9], anticancer drugs are also a frequent research topic due to the increasing incidence of this disease and the potential carcinogenic, mutagenic, and teratogenic effects on the environment. They usually enter the environment as partially metabolized compounds via urine and feces discharged into municipal wastewater and hospitals [10,11]. Examples of such anticancer drugs include imatinib and crizotinib, tyrosine kinase inhibitors used to treat leukemia and lung cancer, respectively. Imatinib has been found in wastewater treatment plant (WWTP) effluent at concentrations ranging from 54 to 180 ng/L and is characterized as a highly persistent compound in water that does not degrade [12,13]. A similar fate is expected for crizotinib given the BIOWIN values calculated using the EPI Suite program (BIOWIN 3 = 1.0010, BIOWIN 5 = −0.4293, BIOWIN 7 = −0.3096). Moreover, these calculated modules predict the resistance of CRZ to biodegradation, aerobic and anaerobic, in terms of BIOWIN values of less than 0.5 [13]. Therefore, these two pharmaceuticals, along with other anticancer drugs, are not expected to be removed by conventional water treatment methods [14,15]. The development of advanced oxidation processes, as more efficient methods for the water treatment, is often the focus of many studies when pharmaceuticals are compounds of interest [11,16]. One such method for removing pharmaceuticals and other organic compounds from water is photocatalysis, which uses a catalyst that is excited by some type of radiation to form highly oxidative radicals (e.g., OH) that react with the pollutant to degrade ultimately into CO2 and water [17]. However, for a photocatalytic reaction to occur at all, sorption of the drug on active sites on the surface/pores of the photocatalyst is a necessary step [18]. The most commonly used photocatalyst is TiO2 due to its easy availability, low cost, non-toxicity, stability, and photochemical efficiency. It can be in the suspended form in water or immobilized on a solid support, which is also a more effective method because the immobilized catalyst can be much more easily removed from water after purification and its reuse is possible, which reduces processing costs [19,20]. Due to all advantages provided by pure TiO2, there are UV light limitations which are trying to be solved for many years by producing different composites with carbonaceous materials. One of these interesting support materials is carbon nanotubes (CNTs), which have been introduced as adsorbent and whose molecular structure allows greater surface area and more π–π interactions with aromatic organic pollutants. CNTs showed good suitability for sorption removal of some antibiotics, such as lincomycin, sulfamethoxazole, and sulfapyridine [6,21]. The creation of new materials consisting of TiO2 and CNTs improves the photocatalytic performance of TiO2, which can increase the photocatalytic activity of TiO2 due to its usually higher specific surface area, good sorbent properties, reduced h+/e recombination, and the possibility of light absorption at higher wavelengths, ultimately leading to more successful environmental applications [22,23]. Such a combination of composites has a wide range of application, from photoelectronics to water purification by photocatalysis, but high cost of CNTs still limits their wide application [6,24,25].
Considering the above, the aim of this work was to determine the potential of TiO2 as a sorbent in suspended and immobilized form for the removal of imatinib and crizotinib from water for the first time. In addition to using pure TiO2 immobilized on glass fibers, the efficiency of immobilized TiO2 in combination with carbon nanotubes, a material with a higher specific surface area, was also investigated. The sorption affinity of each compound to a given sorbent using different operating parameters (pH, ionic strength, temperature, and sorbent dosage) is described by the linear, Freundlich, and Langmuir sorption isotherms. The kinetic study of pharmaceutical sorption was determined by Lagergren’s pseudo-first-kinetic-order, pseudo-second-kinetic-order, and intraparticle diffusion model.

2. Materials and Methods

2.1. Chemicals and Reagents

The studied pharmaceuticals imatinib (IMT) and crizotinib (CRZ) with high purity were supplied by Pliva. Table 1 shows the structures of the active ingredients and their physicochemical properties. NaOH (p.a.) was purchased from Gram-mol, Zagreb, Croatia. HCl was purchased from VWR Chemicals, SAD, NaCl from Lach Ner, Czech Republic, acetonitrile from J. T. Baker, Netherlands, and formic acid from T.T.T. d.o.o., Sv. Nedjelja, Croatia.
The standard stock solution (1000 mg/L) of IMT was prepared by dissolving a precisely weighed amount of the powder in MilliQ water while CRZ was dissolved in acetonitrile. The working standard solutions (5; 10; 15; 20; 25 mg/L) were prepared by diluting the standard solution of pharmaceuticals in MilliQ water. HCl and NaOH were used to adjust the pH. In the experiments in which the influence of ionic strength was studied, the working standard solutions were prepared using NaCl (0.001; 0.01 and 0.1 M) instead of water.

2.2. Sorbent Preparation and Characterization

Titanium dioxide (TiO2) was supplied by Evonik (Aeroxide®, TiO2 P25 with a crystalline content of 75% anatase and 25% rutile), and the multi-walled carbon nanotubes (CNT) were obtained from Chengdu Organic Chem. Co. (outside diameter from 5 nm to 30 nm, purity > 95% and length from 10 µm to 30 µm). The suspension was prepared by weighing 0.01 g TiO2 in 10 mL of pharmaceutical aqueous solution. Before immobilization of the catalyst to the support, glass fiber (GF) meshes were cut and adapted to the dimensions of the glass vessels with an average area of 10.17 cm2. The procedure for sol-gel preparation and characterization of two types of immobilized photocatalysts (TiO2-GF and TiO2/CNT-GF) on a round-cut glass fiber mesh was the same as described in previously published papers [28,29]. The average weight of the immobilized TiO2 film was 0.0037 g/cm2 and 0.072 g/cm2 for TiO2/CNT, respectively. Adding such a mesh to a drug volume of 10 mL yielded an average concentration of 3.76 g/L. Analogous to the above calculations, the concentration of TiO2/CNT-GF mesh was 73.2 g/L, higher than that of TiO2-GF at the beginning, so it is expected that sorption will be higher with this type of catalyst.
In order to get insight in sorbent, clean GF mesh, pure TiO2, mesh with TiO2-GF and TiO2/CNT-GF were recorded by a scanning electron microscope (SEM-QUANTA FEG-250, Zaragoza, Spain) operated at 20 kV. Figure 1 shows SEM images and by comparing the surfaces of TiO2 (b) and TiO2 after binding to glass fibers (c), it can be seen that agglomeration has occurred. Furthermore, TiO2/CNT-GF (d) bonds noticeably produce even larger agglomerates. As it is well known, (I) the surface plays an important role in the sorption process, and (II) the agglomeration reduces the surface area. These are important facts in this paper.

2.3. Sorbent Studies

Batch sorption experiments were performed in glass vessels by shaking in a laboratory shaker (Innova 4080 Incubator Shaker, New Brunswick Scientific, Minneapolis, MN, USA) that allowed continuous contact (200 rpm) between a specific type of TiO2 and 10 mL of aqueous IMT/CRZ solution. After shaking, the solutions were centrifuged and filtered through 0.20-µm membrane syringe filters (Scheme 1).
To determine the time required to reach sorption equilibrium, initial experiments were performed by shaking 5, 15, and 25 mg/L IMT/CRZ solution with TiO2 at 25 °C for various time periods (10, 20, 30, 40, 50, 60, 120, 240, 360, 1080, and 1440 min). Based on these experiments, a contact time of 24 h was sufficient to establish equilibrium between IMT/CRZ and TiO2 suspension and CRZ sorbed onto TiO2/CNT, while other forms of immobilized TiO2 on fiber glass mesh were agitated for 4 h.
Five concentration levels were used (5; 10; 15; 20; 25 mg/L) for the isotherm experiments to test the influence of pH, ionic strength, temperature, and sorbent dosage. The initial pH was adjusted to pH 7 prior to each of the aforementioned experiments, except for the pH influence where pH values 3, 5, 9, and 11 were also tested. All experiments were performed in duplicate.

2.4. Instrumental Procedure

IMT and CRZ samples were analyzed with the HPLC-DAD, Agilent 1100 System (Santa Clara, CA, USA) using Kinetex C18 stationary phase (150 × 4.6 mm, particle size 3.5 µm). The mobile phase with a flow rate of 0.5 mL/min consisted of 0.1% formic acid in Mili-Q water (eluent A) and 0.1% formic acid in acetonitrile (eluent B). Gradient elution began with 80% of eluent A. Within the next 6 min, the composition of eluent A decreased to 20%. After holding 20% of A for 1 min, the initial phase composition was returned and maintained for 7 min. The injection volume was 25 µL. Detection was monitored at 258 nm for IMT and 270 nm for CRZ, respectively.

2.5. Data Anaylsis

Sorption isotherms are models that predict the nature and mechanism of interaction between sorbates and sorbent. The obtained sorption data were analyzed using the linear and two-parameter models: Freundlich and Langmuir isotherms. The isotherm constants were obtained by linearizing the model using linear regression as the simplest method for data analysis.
Through the linear model, the sorption coefficient Kd is expressed as the ratio between the amount of drug in the sorbent (qe, mg/g) and the solution (Ce, mg/L) [30]:
q e = K d × C e
The empirical Langmuir model (2) was used to describe whether the sorption process is a homogenous process that forms in a monolayer on a fixed number of active sites with no interactions between the sorbed molecules. It is given in a linear form [31]:
1 q e = 1 q m + ( 1 q m K L ) 1 C e
where qm (mg/g) and KL (L/mg) represent the maximum sorption capacity (mg/g) and the equilibrium constant referred to sorption energy and the affinity between TiO2 and pharmaceuticals. Calculating RL, a dimensionless constant, the nature of sorption can be inferred; RL values above 1 indicate an unfavorable process, whereas sorption is favored at RL between 0 and 1 [32].
The Freundlich isotherm can be applied to describe the sorption equilibrium for heterogeneous surfaces with possibility of multilayer formation, where all sorption from all sites occurs at the point with the highest binding energy [31,33].
q e = K F · C e 1 n
KF is the Freundlich constant indicating the relative sorption capacity ((μg/g)(mL/μg)1/n) and 1/n is the heterogeneity parameter which indicates the sorption intensity with values ranging from 0 to 1 [34]. Values above 1 indicate cooperative sorption, while lower values than 1 are caused by chemisorption [35].

2.6. Sorption Kinetic Models

The Lagergren’s pseudo-first kinetic model (Equation (4)), Ho’s pseudo-second kinetic model (Equation (5)), and Weber and Morris intraparticle diffusion model (Equation (6)) were used to evaluate the mechanism and describe the phases of the sorption process.
l n ( q e q t ) = l n q e k 1
t q t = 1 k 2 q e 2 + t q e
q t = k i d t 1 / 2
qt and qe represent the amount of sorbed analyte at time t and at equilibrium state (µg/g), k1 is the constant rate of pseudo-first-order (min−1), k2 is the constant rate of pseudo-second-order kinetic model (g/µg min), while kid is the intraparticle diffusion rate constant (μg/g min1/2) [33].
The sorption process can be described by a few stages. The first stage involves the transfer of sorbate near the surface of the sorbent. The second stage is called external diffusion followed by intraparticle diffusion as the third stage. The last stage includes chemical and physical reactions on the surface of the sorbent [36].

3. Results and Discussion

3.1. Kinetic Study

The maximum time required to reach sorption equilibrium was examined by shaking pharmaceutical solutions with three initial pharmaceutical concentrations (5, 15, and 25 mg/L) for 11 different time periods. The TiO2 suspension reached equilibrium with both pharmaceuticals after 24 h. IMT and CRZ were not significantly sorbed on immobilized TiO2 after 4 h, whereas CRZ was required to remain in contact with TiO2/CNT-GF for 24 h. Review of the literature revealed similar results for multi-walled CNTs as sorbents in the case of sulfapyridine and sulfamethoxazole, which also reached equilibrium after 4 h, while ciprofloxacin moved even longer than CRZ (72 h) [21,37]. For all three types of sorbents, sorption in the initial phase is a dominant and fast process, while sorption after reaching equilibrium was a slow process over time. The initial concentrations did not affect the contact time between pharmaceutical and TiO2.
Table 2 and Table 3 show the parameters from the kinetic study in which the pseudo-first- and pseudo-second-order models were tested for IMT and CRZ, respectively. For the pseudo-first-order model, the linear correlation coefficient R2 does not show good agreement, while the R2 for pseudo-second-order kinetic model was very high (R2 > 0.99). The theoretical values qe,cal were closer to the experimental values qe,exp obtained at all concentrations used. Thus, the pseudo-second-order model is the better model for describing the sorption kinetics of the two pharmaceuticals on all three types of TiO2 photocatalyst.
The intraparticle diffusion constants obtained and values related to the boundary layer thickness indicate that the sorption process occurs mainly in three stages. After contact with an aqueous solution of pharmaceutical and the solid phase, the compound migrates into the sorbent boundary layer. The second stage involves intraparticle diffusion between the porous material and the IMT/CRZ. Sorption at the active sites of the sorbents and equilibrium state are usually reached in the third stage, which is confirmed by the smallest slope, kp3 (Table 4 and Table 5) [33,38].
This step is very fast and does not control the rate of the whole sorption process. From the obtained results, it can be concluded that the sorption of IMT and CRZ on TiO2 samples is a complex process controlled mainly by external and intraparticle diffusion. The C values increased from the first to the third stage, with values greater than 0 confirming the first step of surface sorption as the rate-limiting step [39].

3.2. Modeling of Sorption Isotherms—Influence of pH

The sorption parameters of the three types of sorbents fitted to the linear, Freundlich, and Langmuir isotherms are shown in Table 6 for IMT and Table 7 for CRZ, respectively.
The linear isotherm showed the best fit to the obtained data for all TiO2 experiments at different pH values (R2 > 0.99) suggesting the domination of partition interactions [40]. The sorption coefficient Kd showed an increasing trend with increasing pH values for IMT and CRZ on TiO2 in three forms (Figure 2). According to the literature [41,42], the zeta potential of TiO2 is between 6 and 6.5, so TiO2 is positively charged at pH 5; it is probably predominantly negatively charged at pH 7 and 9. The pKa value of IMT nitrogen atom of piperazine is about 7.8–8.07 [27,43], which means that IMT is positively charged in acidic medium and predominantly negatively charged under alkaline conditions. At pH 5, electrostatic repulsive interactions predominated causing the least sorption, while at higher pH values, electrostatic attractive forces were promoted between neutral IMT and more negatively charged TiO2. At pH 9, the sorption affinity of the pharmaceutical was highest, while at pH 7 (value close to the isoelectric point of the sorbent), agglomeration was possible on the TiO2 surface resulting in lower sorption affinity [41,44]. A similar trend in sorption affinity was obtained with CRZ which has two pKa values: 5.6 and 9.4. It can exist in positive, neutral, and negative forms under different pH conditions [26]. Therefore, the increase in sorption coefficient is accompanied by an increase in pH, which is due to the different species present in the medium causing repulsive forces at pH 5 or attractive forces at pH 9. It has been shown that the efficiency of drug removal can be strongly influenced by the presence of different ionizable species of the drug by changing the process conditions, i.e., sorption can be positively influenced by an increase in pH as in this study or negatively influenced in the case of ciprofloxacin and multi-walled CNT (above pH 7) [37], and cefdinir in the pH range of 4–10 using TiO2 as a sorbent [45].
To investigate the sorption capacity and multilayer formation on heterogeneous surfaces, the parameters n and KF were calculated from the logarithmic form of the Freundlich isotherm [46,47]. In the case of sorption of both pharmaceuticals to TiO2 in suspension, the n values are higher than 1, indicating that this type of sorbent is favorable for the removal of IMT and CRZ at low concentrations [48,49]. The sorption of IMT on TiO2-GF mesh is described by n close to 1 (at pH 5 and 9), confirming the linearity of the isotherm and the constant affinity for sorption in the applied concentration range. At the surface of TiO2/CNT-GF, sorption was a cooperative process with the same n values for each pH showing the decrease in affinity of IMT sorption due to the filled active binding sites [50]. CRZ showed a constant sorption potential with higher sorbate concentration on both immobilized catalysts with n values close to unity. The sorption capacity coefficients KF were highest for the suspension, confirming the highest affinity for sorption of IMT in this form. Other KF values are relatively low (19.3–27.2 (μg/g)(mL/μg)1/n), indicating the mobility of pharmaceutical in the sorbent [51]. CRZ generally showed higher tendency to sorb to all TiO2 forms. The Langmuir model in conjunction with the Freundlich isotherm also fitted the experimental data very well, as in the case of cefdinir and TiO2 suspension [45], indicating the complexity of sorption considered as a possible monolayer, and heterogenous multilayer process [45,48,52]. At pH 9, the highest values of maximum saturated monolayer sorption capacity qm (except for immobilized TiO2 at pH 7) of the three TiO2 modifications were obtained, indicating a higher sorption capacity at the mentioned pH, which is in agreement with the determined Kd values. In general, the TiO2 suspension showed the better capacity to sorb both pharmaceuticals, which was expected due to the higher number of active sites on the catalyst surface and easier mass transfer while immobilization of material by trapping active sites may decrease the surface activity [22]. Higher mass of TiO2 does not necessarily bring higher sorption performance, rather the performance of the sorbent or catalyst plays a major role. The results are also consistent with the results of the SEM analysis (Figure 1), which shows the agglomeration occurred during the binding of TiO2 and TiO2/CNT to GF meshes, resulting in a decrease in the active surface area of the sorbent. However, to avoid problems in sample reuse and filtration, it is better to immobilize only TiO2 and not to use it in combination with nanotubes because the different active sites of the two materials may cause heterogeneous sorption [45].

3.3. Influence of Ionic Strength

Standard solutions of pharmaceuticals at a natural pH 7 were prepared in the presence of 0.001 M, 0.01 M, and 0.1 M NaCl instead of MilliQ water to determine the sorption capacity by changing the concentration of the inorganic salt.
A slight increasing trend of IMT sorption with higher ionic strength for all three sorbent samples (Figure 3) can be attributed to possible neutralization of the TiO2 surface by positive ions, which allows non-electrostatic interactions between pharmaceutical and sorbent [53]. Increased sorption may also be the result of decreased solubility of the pharmaceutical due to addition of soluble salts in aqueous solution [48]. The higher concentration of Na+ ions affecting the decrease in CRZ sorption on TiO2 may be the result of decreased active sites due to competitive sorption between Na+ and CRZ, and weak electrostatic repulsion interactions due to equal charges of sorbent and sorbate, including ambient pH [54]. Determining the effect of ionic strength provides important information on how much the removal of organic contaminants from water is affected by the presence or addition of different ions. These results may also indicate that the sorption of IMT and CRZ is also affected by different substances in the water matrix, which needs to be confirmed by additional experiments.

3.4. Effect of Sorbent Dosage

For each type of sorbent, experiments were performed with three different sorbent dosages (Figure 4); the lower mass and the higher mass of the sorbent used in the experiments described above. The percentage of sorption was not affected too much by higher dosage of sorbent; a slight increase in Kd with higher mass was observed, which was due to an increase in active sites for both types of immobilized TiO2 [55]. When the TiO2 is in the form of a suspension, a linear increase in IMT sorption was observed. A minor influence of sorbent dosage was also observed in the sorption of two sulfonamide antibiotics on multi-walled CNTs, which showed the opposite trend, i.e., a higher mass of sorbent does not necessarily increase the removal efficiency due to the occurrence of saturation [21,55].

3.5. Sorption Thermodynamics

To describe the nature of the sorption process by determining thermodynamic parameters, experiments were performed at temperatures of 25, 30, and 35 °C (298; 303; 308 K). The pH of the solutions was adjusted to a neutral environment (pH 7). It was found that temperature has different influences on sorption to TiO2 in different forms (Table 8). The sorption of IMT on the suspension decreased with increasing temperature. For the sorption of IMT on TiO2-GF, the effect of temperature was not noticeable due to a slight increase in active sites and Kd values. A similar temperature effect was found for the sorption of CRZ on all TiO2 forms; increasing the temperature led to an increase in molecular diffusion to the surface of the photocatalyst and then promoted pharmaceutical binding [56]. Similar sorption thermodynamics depending on the type of sorbent used as for IMT was also reported for ciprofloxacin [37].
Thermodynamic parameters, including Gibbs free energy, ∆G°, enthalpy, ∆H°, and entropy, ∆S°, were calculated using the following equations [57]:
Δ G ° = R T l n K d
l n K d = Δ H ° R T + Δ S ° R
The sorption of IMT and CRZ on all sorbents is favorable and spontaneous physisorption process, which is confirmed by negative values of ∆G°, ranging from −1.8 to −14.8 kJ/mol for both pharmaceuticals [58]. The more negative the ∆G° values, the higher sorption affinity of the pharmaceuticals, as a result of increased driving molecular forces such as Van der Waals forces [59]. The positive ∆H° and ∆S° values indicate that the sorption of CRZ onto three forms of TiO2 and of IMT onto both types of immobilized TiO2, is an endothermic reaction [60], where an increase in temperature leads to a higher sorption capacity. Based on the negative values of the thermodynamic parameters ∆H° and ∆S°, it can be assumed that IMT sorption in suspension is an exothermic reaction in which randomness at the interface of TiO2 and IMT solution decreases [33,59].

3.6. Statistical Analysis

Statistical analysis was performed by calculating correlations between the parameters affecting sorption as independent variables (pH, ionic strength, temperature, sorbent dosage) and the dependent one—the sorption coefficient. Pearson’s correlation coefficient r can range from −1 to +1, while − or + indicate the direction of correlation [61]. The sorption of IMT to all types of TiO2 under different influences showed strong linear correlations with r higher than 0.84. A slightly weaker correlation was found for the influence of ionic strength on IMT sorption to TiO2-GF (Figure 5). A similar trend was also observed for CRZ, where a strong positive correlation (r = 0.87–0.98) indicated a proportional relationship between Kd and pH, temperature, and sorbent dosage. The sorption of CRZ showed a negative medium-strong correlation when the ionic strength of the solution was changed.

4. Conclusions

In this manuscript, preliminary data on the removal of IMT and CRZ from an aqueous solution by sorption processes with TiO2 suspension, TiO2-GF, and immobilized TiO2/CNT-GF were presented. The influence of contact time, solution pH, initial pharmaceutical concentration, ionic strength, sorbent dosage, and ambient temperature was studied to determine the affinity of IMT and CRZ for the photocatalyst used. Statistical analysis confirmed that the complexity of the sorption process is strongly influenced by the various parameters mentioned above. Since most of the pharmaceuticals present in the environment are expected to be in ionized form, the experiments were carried out at different pH values in the range of 5–9, to show what interactions may occur due to the physicochemical properties of sorbent and sorbate during the sorption process. The Kd values obtained from the linear isotherm, the best model to describe the sorption processes for both pharmaceuticals (R2 > 0.99), increased with increasing alkalinity of the aqueous medium. When inorganic salt was added to the solution, an opposite behavior of the drugs was observed. The sorption affinity of CRZ decreased with increasing concentration of Na-cations, while IMT was slightly more sorbed on all three catalyst types. Increasing the mass of TiO2 also increased the active sites on the catalysts, resulting in higher sorption affinity of the two drugs. Thermodynamic parameters indicated spontaneity of CRZ and IMT sorption on all catalyst forms used. Higher ambient temperature had positive effect on the sorption of pharmaceuticals on both types of immobilized TiO2 and on the suspension for CRZ, while IMT added in suspension showed an exothermic response. The pseudo-second-order model best described the kinetics of the sorption process with R2 values close to one. Despite the fact that the suspended form of TiO2 had a higher sorption capacity, when the efficiency of the removal process, sorption or photocatalysis was considered together with the operating costs, including the preparation of the sorbent, it can be concluded that the suspension is not cost-effective in real water treatment applications at higher levels, considering the time-consuming and expensive separation without the possibility of reuse.
CRZ and IMT showed a similar and relatively good tendency to sorb to the TiO2 photocatalyst, which can be considered for future research as a potential sorbent for pharmaceuticals removal from various types of water matrices. If a compound has a good affinity for the sorption catalyst, the sorption process should be considered as an economically and environmentally efficient method for the removal of contaminants, rather than a process that requires higher energy consumption.

Author Contributions

Conceptualization, D.M.P. and K.T.Č.; methodology, D.M.P. and K.T.Č.; validation, K.T.Č. and D.M.P.; formal analysis, D.M.P., K.T.Č., K.D., M.P., I.F., I.J. and V.Š.; investigation, D.M.P., K.T.Č., K.D., M.P. and I.F.; data curation, D.M.P. and K.T.Č.; writing—original draft preparation, K.T.Č.; writing—review and editing, K.T.Č., D.M.P. and I.B.; supervision, D.M.P. and K.T.Č. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Periša, M.; Babić, S. Farmaceutici u okolišu. Kem. Ind. 2015, 65, 471–482. [Google Scholar] [CrossRef]
  2. Jones, O.A.H.; Voulvoulis, N.; Lester, J.N. Human Pharmaceuticals in Wastewater Treatment Processes. Crit. Rev. Environ. Sci. Technol. 2005, 35, 401–427. [Google Scholar] [CrossRef]
  3. González Peña, O.I.; López Zavala, M.A.; Cabral Ruelas, H. Pharmaceuticals Market, Consumption Trends and Disease Incidence Are Not Driving the Pharmaceutical Research on Water and Wastewater. Int. J. Environ. Res. Public Health 2021, 18, 2532. [Google Scholar] [CrossRef] [PubMed]
  4. Nikolaou, A.; Meric, S.; Fatta, D. Occurrence patterns of pharmaceuticals in water and wastewater environments. Anal. Bioanal. Chem. 2007, 387, 1225–1234. [Google Scholar] [CrossRef] [PubMed]
  5. Coates, A.R.M.; Halls, G.; Hu, Y. Novel classes of antibiotics or more of the same? Br. J. Pharmacol. 2011, 163, 184–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W. Adsorptive removal of antibiotics from water and wastewater: Progress and challenges. Sci. Total Environ. 2015, 532, 112–126. [Google Scholar] [CrossRef] [PubMed]
  7. Hassan, M.; Zhu, G.; Lu, Y.-Z.; Al-Falahi, A.H.; Lu, Y.; Huang, S.; Wan, Z. Removal of antibiotics from wastewater and its problematic effects on microbial communities by bioelectrochemical Technology: Current knowledge and future perspectives. Environ. Eng. Res. 2021, 26, 190405. [Google Scholar] [CrossRef]
  8. 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]
  9. Dutta, J.; Mala, A.A. Removal of antibiotic from the water environment by the adsorption technologies: A review. Water Sci. Technol. 2020, 82, 401–426. [Google Scholar] [CrossRef]
  10. Booker, V.; Halsall, C.; Llewellyn, N.; Johnson, A.; Williams, R. Prioritising anticancer drugs for environmental monitoring and risk assessment purposes. Sci. Total Environ. 2014, 473–474, 159–170. [Google Scholar] [CrossRef] [Green Version]
  11. Pieczyńska, A.; Fiszka Borzyszkowska, A.; Ofiarska, A.; Siedlecka, E.M. Removal of cytostatic drugs by AOPs: A review of applied processes in the context of green technology. Crit. Rev. Environ. Sci. Technol. 2017, 47, 1282–1335. [Google Scholar] [CrossRef]
  12. Gouveia, T.I.A.; Alves, A.; Santos, M.S.F. New insights on cytostatic drug risk assessment in aquatic environments based on measured concentrations in surface waters. Environ. Int. 2019, 133, 105236. [Google Scholar] [CrossRef] [PubMed]
  13. Mihçiokur, H. Environmental risk assessment of commonly used anti-cancer drugs. Cumhur. Sci. J. 2021, 42, 310–320. [Google Scholar] [CrossRef]
  14. Zhang, J.; Chang, V.W.C.; Giannis, A.; Wang, J.-Y. Removal of cytostatic drugs from aquatic environment: A review. Sci. Total Environ. 2013, 445–446, 281–298. [Google Scholar] [CrossRef]
  15. Mansouri, F.; Chouchene, K.; Roche, N.; Ksibi, M. Removal of Pharmaceuticals from Water by Adsorption and Advanced Oxidation Processes: State of the Art and Trends. Appl. Sci. 2021, 11, 6659. [Google Scholar] [CrossRef]
  16. Sousa, M.A.; Gonçalves, C.; Vilar, V.J.P.; Boaventura, R.A.R.; Alpendurada, M.F. Suspended TiO2-assisted photocatalytic degradation of emerging contaminants in a municipal WWTP effluent using a solar pilot plant with CPCs. Chem. Eng. J. 2012, 198–199, 301–309. [Google Scholar] [CrossRef]
  17. Čizmić, M.; Ljubas, D.; Ćurković, L.; Škorić, I.; Babic, S. Kinetics and degradation pathways of photolytic and photocatalytic oxidation of the anthelmintic drug praziquantel. J. Hazard. Mater. 2017, 323, 500–512. [Google Scholar] [CrossRef]
  18. Bhatkhande, D.S.; Pangarkar, V.G.; Beenackers, A.A.C.M. Photocatalytic degradation for environmental applications-a review. J. Chem. Technol. Biotechnol. 2001, 77, 102–116. [Google Scholar] [CrossRef]
  19. Rachel, A.; Subrahmanyam, M.; Boule, P. Comparison of photocatalytic efficiencies of TiO2 in suspended and immobilized form for the photocatalytic degradation of nitrobenzene sulfonic acids. Appl. Catal. B 2002, 37, 301–308. [Google Scholar] [CrossRef]
  20. Babić, S.; Zrnčić, M.; Ljubas, D.; Ćurković, L.; Škorić, I. Photolytic and thin TiO2 film assisted photocatalytic degradation of sulfamethazine in aqueous solution. Environ. Sci. Pollut. Res. 2015, 22, 11372–11386. [Google Scholar] [CrossRef]
  21. Tian, Y.; Gao, B.; Chen, H.; Wang, Y.; Li, H. Interactions between carbon nanotubes and sulfonamide antibiotics in aqueous solutions under various physicochemical conditions. J. Environ. Sci. Health A 2013, 48, 1136–1144. [Google Scholar] [CrossRef] [PubMed]
  22. Awfa, D.; Ateia, M.; Fujii, M.; Johnson, M.S.; Yoshimura, C. Photodegradation of pharmaceuticals and personal care products in water treatment using carbonaceous-TiO2 composites: A critical review of recent literature. Water Res. 2018, 142, 26–45. [Google Scholar] [CrossRef] [PubMed]
  23. Li, Z.; Gaoa, B.; Zheng Chena, G.; Mokaya, R.; Sotiropoulos, S.; Li Puma, G. Carbon nanotube/titanium dioxide (CNT/TiO2) core–shell nanocomposites with tailored shell thickness, CNT content and photocatalytic/photoelectrocatalytic properties. Appl. Catal. B 2011, 110, 50. [Google Scholar] [CrossRef]
  24. Silva, C.G.; Faria, J.L. Photocatalytic oxidation of benzene derivatives in aqueous suspensions: Synergic effect induced by the introduction of carbon nanotubes in a TiO2 matrix. Appl. Catal. B 2010, 101, 81–89. [Google Scholar] [CrossRef]
  25. Liu, X.; Wang, M.; Zhang, S.; Pan, B. Application potential of carbon nanotubes in water treatment: A review. J. Environ. Sci. 2013, 25, 1263–1280. [Google Scholar] [CrossRef]
  26. Bandla, J.; Ganapaty, S. Stability indicating UPLC method development and validation for the determination of crizotinib in pharmaceutical dosage forms. Int. J. Sci. Prog. Res. 2018, 9, 1493–1498. [Google Scholar]
  27. Mioduszewska, K.; Dołżonek, J.; Wyrzykowski, D.; Kubik, Ł.; Wiczling, P.; Sikorska, C.; Toński, M.; Kaczyński, Z.; Stepnowski, P.; Białk-Bielińska, A. Overview of experimental and computational methods for the determination of the pKa values of 5-fluorouracil, cyclophosphamide, ifosfamide, imatinib. TrAC Trends Anal. Chem. 2017, 97, 283–296. [Google Scholar] [CrossRef]
  28. Grčić, I.; Marčec, J.; Radetić, L.; Radovan, A.M.; Melnjak, I.; Jajčinović, I.; Brnardić, I. Ammonia and methane oxidation on TiO2 supported on glass fiber mesh under artificial solar irradiation. Environ. Sci. Pollut. Res. 2020, 28, 18354–18367. [Google Scholar] [CrossRef]
  29. Malinowski, S.; Presečki, I.; Jajčinović, I.; Brnardić, I.; Mandić, V.; Grčić, I. Intensification of dihydroxybenzenes degradation over immobilized TiO2 based photocatalysts under simulated solar light. Appl. Sci. 2020, 10, 7571. [Google Scholar] [CrossRef]
  30. Kim, Y.; Lim, S.; Han, M.; Cho, J. Sorption characteristics of oxytetracycline, amoxicillin, and sulfathiazole in two different soil types. Geoderma 2012, 185–186, 97–101. [Google Scholar] [CrossRef]
  31. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  32. Webber, T.W.; Chakkravorti, R.K. Pore and solid diffusion models for fixed-bedadsorbers. AlChE J. 1974, 20, 228–238. [Google Scholar] [CrossRef]
  33. Mutavdžić Pavlović, D.; Ćurković, L.; Macan, J.; Žižek, K. Eggshell as a New Biosorbent for the Removal of Pharmaceuticals From Aqueous Solutions. Clean-Soil Air Water 2017, 45, 1700082. [Google Scholar] [CrossRef]
  34. Park, J.Y.; Huwe, B. Effect of pH and soil structure on transport of sulfonamide antibiotics in agricultural soils. Environ. Pollut. 2016, 213, 561–570. [Google Scholar] [CrossRef] [PubMed]
  35. Haghseresht, F.; Lu, G.Q. Adsorption Characteristics of Phenolic Compounds onto Coal-Reject-Derived Adsorbents. Energy Fuels 1998, 12, 1100–1107. [Google Scholar] [CrossRef]
  36. Glavaš, Z.; Štrkalj, A. Kinetics of adsorption process for the system waste mold sand/Cu (II) ions. Hrvat. Vode 2015, 23, 185. [Google Scholar]
  37. Li, H.; Zhang, D.; Han, X.; Xing, B. Adsorption of antibiotic ciprofloxacin on carbon 496 nanotubes: pH dependence and thermodynamics. Chemosphere 2014, 95, 150–155. [Google Scholar] [CrossRef] [PubMed]
  38. Ramachandran, P.; Vairamuthu, R.; Ponnusamy, S. Adsorption isotherms, kinetics, thermodynamics and desorption studies of reactive orange 16 on activated carbon derived from Ananas comosus (L.) carbon. ARPN J. Eng. Appl. Sci. 2011, 6, 15–26. [Google Scholar]
  39. Pholosi, A.; Naidoo, E.B.; Ofomaja, A.E. Intraparticle diffusion of Cr(VI) through biomass and magnetite coated biomass: A comparative kinetic and diffusion study. South Afr. J. Chem. Eng. 2020, 32, 39–55. [Google Scholar] [CrossRef]
  40. Mamitiana Razanajatovoa, R.; Dinga, J.; Zhanga, S.; Jianga, H.; Zoua, Z. Sorption and desorption of selected pharmaceuticals by polyethylene microplastics. Mar. Pollut. Bull. 2018, 136, 516–523. [Google Scholar] [CrossRef]
  41. Gora, S.L.; Andrews, S.A. Adsorption of natural organic matter and disinfection byproduct precursors from surface water onto TiO2 nanoparticles: pH effects, isotherm modelling and implications for using TiO2 for drinking water treatment. Chemosphere 2017, 174, 363–370. [Google Scholar] [CrossRef] [PubMed]
  42. Qamar, M.; Muneer, M.; Bahnemann, D. Heterogeneous photocatalysed degradation of two selected pesticide derivatives, triclopyr and daminozid in aqueous suspensions of titanium dioxide. J. Environ. Manag. 2006, 80, 99–106. [Google Scholar] [CrossRef] [PubMed]
  43. Secrétan, P.H.; Karoui, M.; Sadou Yayé, H.; Levi, Y.; Tortolano, L.; Solgadi, A.; Yagoubi, N.; Do, B. Imatinib: Major photocatalytic degradation pathways in aqueous media and the relative toxicity of its transformation products. Sci. Total Environ. 2019, 655, 547–556. [Google Scholar] [CrossRef] [PubMed]
  44. Azeez, F.; Al-Hetlani, E.; Arafa, M.; Abdelmonem, Y.; Nazeer, A.A.; Amin, M.O.; Madkour, M. The efect of surface charge on photocatalytic degradation of methylene blue dye using chargeable titania nanoparticles. Sci. Rep. 2018, 8, 20174. [Google Scholar] [CrossRef] [PubMed]
  45. Yener, H.B. Removal of Cefdinir from Aqueous Solution Using Nanostructure Adsorbents of TiO2, SiO2 and TiO2/SiO2: Equilibrium, thermodynamic and kinetic studies. Chem. Biochem. Eng. Q. 2019, 33, 235–248. [Google Scholar] [CrossRef]
  46. Theivarasu, C.; Mylsamy, S. Equilibrium and kinetic adsorption studies of rhodamine-B from aqueous solutions using cocoa (Theobroma cacao) shell as a new adsorbent. Int. J. Eng. Sci. Technol. 2010, 2, 6284–6292. [Google Scholar]
  47. Tor, A.; Cengeloglu, Y. Removal of congo red from aqueous solution by adsorption onto acid activated red mud. J. Hazard. Mater. 2006, 138, 409–415. [Google Scholar] [CrossRef]
  48. Toński, M.; Dołżonek, J.; Paszkiewicz, M.; Wojsławski, J.; Stepnowski, P.; Białk-Bielińska, A. Preliminary evaluation of the application of carbon nanotubes as potential adsorbents for the elimination of selected anticancer drugs from water matrices. Chemosphere 2018, 201, 32–40. [Google Scholar] [CrossRef]
  49. Santanu Mukherjee, S.; Weihermüller, L.; Tappe, W.; Hofmann, D.; Köppchen, S.; Laabs, V.; Vereecken, H.; Burauel, P. Sorption–desorption behaviour of bentazone, boscalid and pyrimethanilin biochar and digestate based soil mixtures for biopurification systems. Sci. Tot. Environ. 2016, 559, 63–73. [Google Scholar] [CrossRef]
  50. Tolić, K.; Mutavdžić Pavlović, D.; Stankir, N.; Runje, M. Biosorbents from Tomato, Tangerine, and Maple Leaves for the Removal of Ciprofloxacin from Aqueous Media. Water Air Soil Pollut. 2021, 232, 1–16. [Google Scholar] [CrossRef]
  51. Doretto, K.M.; Rath, S. Sorption of sulfadiazine on Brazilian soils. Chemosphere 2013, 90, 2027–2034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Yi, Z.; Yao, J.; Zhu, M.; Chen, H.; Wang, F.; Liu, X. Kinetics, equilibrium, and thermodynamics investigation on the adsorption of lead (II) by coal-based activated carbon. Springer Plus 2016, 5, 1160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Aguedach, A.; Brosillon, S.; Morvan, J.; Lhadi, E.K. Influence of ionic strength in the adsorption and during photocatalysis of reactive black 5 azo dye on TiO2 coated on non woven paper with SiO2 as a binder. J. Hazard. Mater. 2008, 150, 250–256. [Google Scholar] [CrossRef] [PubMed]
  54. Golet, E.M.; Xifra, I.; Siegrist, H.; Alder, A.C.; Giger, W. Environmental Exposure Assessment of Fluoroquinolone Antibacterial Agents from Sewage to Soil. Environ. Sci. Technol. 2003, 37, 3243–3249. [Google Scholar] [CrossRef] [PubMed]
  55. Akhtar, J.; Amin, N.A.S.; Shahzad, K. A review on removal of pharmaceuticals from water by adsorption. Desalin. Water Treat. 2016, 57, 12842–12860. [Google Scholar] [CrossRef]
  56. Sarkar, S.; Das, R.; Choi, H.; Bhattacharjee, C. Involvement of process parameters and various modes of application of TiO2 nanoparticles in heterogeneous photocatalysis of pharmaceutical wastes–a short review. RSC Adv. 2014, 4, 57250–57266. [Google Scholar] [CrossRef]
  57. Awala, H.A.; El Jamal, M.M. Equilibrium and kinetics study of adsorption of some dyes onto feldspar. J. Chem. Technol. Metall. 2011, 46, 45–52. [Google Scholar]
  58. Malakootian, M.; Nasiri, A.; Gharaghani, M.A. Photocatalytic degradation of ciprofloxacin antibiotic by TiO2 nanoparticles immobilized on a glass plate. Chem. Eng. Commun. 2019, 207, 56–72. [Google Scholar] [CrossRef]
  59. Yu, F.; Wu, Y.; Li, X.; Ma, J. Kinetic and Thermodynamic Studies of Toluene, Ethylbenzene, and m-Xylene Adsorption from Aqueous Solutions onto KOH-Activated Multiwalled Carbon Nanotubes. J. Agric. Food Chem. 2012, 60, 12245–12253. [Google Scholar] [CrossRef]
  60. Saha, P.; Chowdhury, S.; Gupta, S.; Kumar, I. Insight into adsorption equilibrium, kinetics and thermodynamics of Malachite Green onto clayey soil of Indian origin. Chem. Eng. J. 2010, 165, 874–882. [Google Scholar] [CrossRef]
  61. Udovičić, M.; Baždarić, K.; Bilić-Zulle, L.; Petrovečki, M. What we need to know when calculating the coefficient of correlation? Biochem. Med. 2007, 17, 10–15. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The SEM images: (a) clean GF mesh, (b) TiO2, (c) TiO2-GF mesh, and (d) TiO2/CNT-GF mesh.
Figure 1. The SEM images: (a) clean GF mesh, (b) TiO2, (c) TiO2-GF mesh, and (d) TiO2/CNT-GF mesh.
Applsci 12 04113 g001
Scheme 1. Schematic description of sorption experiment procedure.
Scheme 1. Schematic description of sorption experiment procedure.
Applsci 12 04113 sch001
Figure 2. pH influence on IMT and CRZ sorption (1—suspended TiO2, 2—TiO2-GF, 3—TiO2/CNT-GF).
Figure 2. pH influence on IMT and CRZ sorption (1—suspended TiO2, 2—TiO2-GF, 3—TiO2/CNT-GF).
Applsci 12 04113 g002
Figure 3. Ionic strength as influencing parameter on IMT and CRZ sorption (1—suspended TiO2, 2—TiO2-GF, 3—TiO2/CNT-GF).
Figure 3. Ionic strength as influencing parameter on IMT and CRZ sorption (1—suspended TiO2, 2—TiO2-GF, 3—TiO2/CNT-GF).
Applsci 12 04113 g003
Figure 4. Effect of sorbent dosage (1—suspended TiO2, 2—TiO2-GF, 3—TiO2/CNT-GF. 1: min—0.5 g/L, average—1.0 g/L, max—1.5 g/L; mass layer of 2: min—2.3 g/L, average—3.7 g/L, max—5.2 g/L; mass layer of 3: min—47.8 g/L, average—73.2 g/L, max—107.6 g/L).
Figure 4. Effect of sorbent dosage (1—suspended TiO2, 2—TiO2-GF, 3—TiO2/CNT-GF. 1: min—0.5 g/L, average—1.0 g/L, max—1.5 g/L; mass layer of 2: min—2.3 g/L, average—3.7 g/L, max—5.2 g/L; mass layer of 3: min—47.8 g/L, average—73.2 g/L, max—107.6 g/L).
Applsci 12 04113 g004
Figure 5. Determination of Pearson’s correlation coefficients between sorption influencing parameters.
Figure 5. Determination of Pearson’s correlation coefficients between sorption influencing parameters.
Applsci 12 04113 g005
Table 1. Physicochemical properties of pharmaceuticals [14,26,27].
Table 1. Physicochemical properties of pharmaceuticals [14,26,27].
PhAcIMTCRZ
Applsci 12 04113 i001 Applsci 12 04113 i002
FormulaC29H31N7OC21H22Cl2FN5O
CAS152459-95-5877399-52-5
Mr493.6450.1
pKa8.07; 3.73; 2.56; 1.525.6; 9.4
log Kow2.89/
Table 2. Sorption kinetic parameters for IMT.
Table 2. Sorption kinetic parameters for IMT.
IMTInitial Concentration, mg/Lqe,exp, μg/gPseudo-First-OrderPseudo-Second-Order
qe,calc, μg/gk1, min−1R2qe,calc, μg/gk2, g/μg minR2
suspended
TiO2
514.628.470.00590.948814.902.03 × 10−40.9990
1528.408.505 × 10−50.623829.674.26 × 10−40.9981
2532.4710.763 × 10−50.673134.253.13 × 10−40.9969
TiO2-GF527.374.730.00020.643028.578.36 × 10−40.9989
15118.506.520.00040.63851252.96 × 10−40.9999
25168.258.940.00020.6516174.441.67 × 10−40.9995
TiO2/CNT-GF532.354.870.00030.573436.761.72 × 10−40.9792
1568.188.230.00020.610085.474.20 × 10−40.8151
25104.6510.480.00020.6309161.291.05 × 10−50.5285
Table 3. Sorption kinetic parameters for CRZ.
Table 3. Sorption kinetic parameters for CRZ.
CRZInitial Concentration, mg/Lqe,exp, μg/gPseudo-First-OrderPseudo-Second-Order
qe,calc, μg/gk1, min−1R2qe,calc, μg/gk2, g/μg minR2
suspended
TiO2
515.888.000.00590.946015.779.87 × 10−30.9997
1546.327.914 × 10−50.623046.301.13 × 10−30.9996
2567.3910.215 × 10−50.720068.034.12 × 10−40.9989
TiO2-GF542.392.860.00020.422242.557.15 × 10−41.000
15120.575.720.00020.7360121.955.37 × 10−40.9996
25207.938.990.00040.9133212.777.43 × 10−40.9954
TiO2/CNT-GF530.293.906 × 10−50.342730.407.07 × 10−31.000
1586.527.350.00020.64888.503.39 × 10−40.9996
25121.439.960.00020.7592128.211.07 × 10−40.9954
Table 4. Intraparticle diffusion model for IMT.
Table 4. Intraparticle diffusion model for IMT.
Samplec0, mg/LIntraparticle Diffusion
First StageSecond StageThird Stage
kp1,
µg/g min1/2
R2C1kp2,
µg/g min1/2
R2C2kp3,
µg/g min1/2
R2C3
suspended
TiO2
52.4840.97315.7570.11390.892810.730.10180.980010.92
153.2570.97518.8840.59560.985312.350.23950.999719.34
253.4750.95779.5250.48540.990115.420.41810.978817.096
TiO2-GF50.74960.98287.6751.62480.99152.0120.01210.884227.19
1512.3170.99424.8014.84790.987344.820.17610.9614115.88
2512.7380.984040.468.40910.986735.570.08670.9994166.9
TiO2/CNT-GF51.7320.96234.4912.77140.91008.4610.00180.818732.33
155.3380.955718.0165.50060.985015.580.04020.971267.58
255.0180.937317.19910.2830.969451.270.11270.9920102.87
Table 5. Intraparticle diffusion model for CRZ.
Table 5. Intraparticle diffusion model for CRZ.
Samplec0, mg/LIntraparticle Diffusion
First StageSecond StageThird Stage
kp1,
µg/g min1/2
R2C1kp2,
µg/g min1/2
R2C2kp3,
µg/g min1/2
R2C3
suspended
TiO2
50.49130.953911.000.03300.904614.51---
152.3180.977617.810.64520.993531.130.15360.950840.21
252.8140.946520.051.7790.939527.350.38870.997152.47
TiO2-GF51.2610.953030.340.17870.870939.170.00450.997642.22
155.4410.952460.150.64950.981298.460.02201.000119.73
2514.9140.99180.76234.1380.916880.252.8030.9994102.09
TiO2/CNT-GF52.0090.840115.590.15060.955426.730.02100.943629.54
1517.5320.92692.1241.9480.992843.260.39570.932472.35
2510.6900.960917.213.7130.997931.721.0480.934583.87
Table 6. Sorption isotherm parameters for IMT as an influence of pH.
Table 6. Sorption isotherm parameters for IMT as an influence of pH.
Isotherm
Parameter
Suspended
TiO2
TiO2-GFTiO2/CNT-GF
579579579
Linear
Kd (mL/g)162.2213.8672.322.136.4160.575.482.3113.2
R20.99600.99800.99100.99300.99700.99100.99000.99000.9920
Langmuir
qm (mg/g)5.05.0205.01.15.00.450.620.67
KL (mL/g)0.50.40.070.0040.150.030.0510.0490.059
R20.88960.91990.99890.99900.91900.99000.94300.95400.9440
Freundlich
n3.02.61.40.951.880.910.640.680.66
KF (μg/g)(mL/μg)1/n2027.21944.51563.517.79.39.814.221.127.2
R20.97500.99900.99900.99800.96900.99100.97200.97700.9680
Table 7. Sorption isotherm parameters for CRZ as an influence of pH.
Table 7. Sorption isotherm parameters for CRZ as an influence of pH.
Isotherm
Parameter
Suspended
TiO2
TiO2-GFTiO2/CNT-GF
579579579
Linear
Kd (mL/g)103.5233.8234.4117.12179.18208.91197.74234.12329.67
R20.99650.99700.99930.99800.99300.99220.99110.99360.9940
Langmuir
qm (mg/g)10.05.020.01005.010.050166.733.3
KL (mL/g)0.0150.50.50.0010.0740.0360.0040.00130.0097
R20.99680.86650.90720.99800.99760.99900.99110.99940.9986
Freundlich
n1.22.62.70.991.31.20.990.970.98
KF (μg/g)(mL/μg)1/n174.22025.82105.7116.5399.9382.6204.2208.8329.7
R20.99820.94820.97070.99840.99740.99940.99650.99750.9967
Table 8. Thermodynamic parameters for IMT and CRZ sorption.
Table 8. Thermodynamic parameters for IMT and CRZ sorption.
Sample Kd, mL/gΔ, kJ/molΔ, kJ/molΔ, J/mol
IMT298 K303 K308 K298 K303 K308 K
suspended
TiO2
213.861.949.7−13.3−10.4−10.0−57.9−151.0
TiO2-GF36.440.347.4−8.8−9.3−9.510.464.6
TiO2/CNT-GF82.382.588.7−10.9−11.1−1.82.946.4
CRZ
suspended
TiO2
233.8268.8271.7−13.5−14.3−14.816.7101.7
TiO2-GF179.2180.9197.6−10.1−11.6−11.93.8255.8
TiO2/CNT-GF234.1325.2357.0−8.1−8.7−9.75.9765.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tolić Čop, K.; Mutavdžić Pavlović, D.; Duić, K.; Pranjić, M.; Fereža, I.; Jajčinović, I.; Brnardić, I.; Špada, V. Sorption Potential of Different Forms of TiO2 for the Removal of Two Anticancer Drugs from Water. Appl. Sci. 2022, 12, 4113. https://0-doi-org.brum.beds.ac.uk/10.3390/app12094113

AMA Style

Tolić Čop K, Mutavdžić Pavlović D, Duić K, Pranjić M, Fereža I, Jajčinović I, Brnardić I, Špada V. Sorption Potential of Different Forms of TiO2 for the Removal of Two Anticancer Drugs from Water. Applied Sciences. 2022; 12(9):4113. https://0-doi-org.brum.beds.ac.uk/10.3390/app12094113

Chicago/Turabian Style

Tolić Čop, Kristina, Dragana Mutavdžić Pavlović, Katarina Duić, Minea Pranjić, Iva Fereža, Igor Jajčinović, Ivan Brnardić, and Vedrana Špada. 2022. "Sorption Potential of Different Forms of TiO2 for the Removal of Two Anticancer Drugs from Water" Applied Sciences 12, no. 9: 4113. https://0-doi-org.brum.beds.ac.uk/10.3390/app12094113

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