Next Article in Journal
High Resistance of SO2 and H2O over Monolithic Mn-Fe-Ce-Al-O Catalyst for Low Temperature NH3-SCR
Previous Article in Journal
Development of FTIR Spectroscopy Methodology for Characterization of Boron Species in FCC Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phthalocyanine-Grafted Titania Nanoparticles for Photodegradation of Ibuprofen

by
Rafal Krakowiak
1,†,
Joanna Musial
2,†,
Robert Frankowski
3,
Marcin Spychala
4,
Jadwiga Mielcarek
5,
Bernadeta Dobosz
6,
Ryszard Krzyminiewski
6,
Marek Sikorski
7,
Wioletta Bendzinska-Berus
1,
Ewa Tykarska
1,
Ryszard Blazejewski
4,
Agnieszka Zgoła-Grześkowiak
3,
Beata J. Stanisz
2,
Dariusz T. Mlynarczyk
1,* and
Tomasz Goslinski
1
1
Chair and Department of Chemical Technology of Drugs, Poznan University of Medical Sciences, Grunwaldzka 6, 60-780 Poznań, Poland
2
Chair and Department of Pharmaceutical Chemistry, Poznan University of Medical Sciences, Grunwaldzka 6, 60-780 Poznań, Poland
3
Institute of Chemistry and Technical Electrochemistry, Poznan University of Technology, Berdychowo 4, 60-965 Poznań, Poland
4
Department of Hydraulic and Sanitary Engineering, Poznan University of Life Sciences, Piątkowska 94A, 60-649 Poznań, Poland
5
Chair and Department of Inorganic and Analytical Chemistry, Poznan University of Medical Sciences, Grunwaldzka 6, 60-780 Poznan, Poland
6
Medical Physics and Radiospectroscopy Division, Faculty of Physics, Adam Mickiewicz University, Uniwersytetu Poznańskiego 2, 61-614 Poznań, Poland
7
Faculty of Chemistry, Adam Mickiewicz University, Uniwersytetu Poznańskiego 8, 61-614 Poznań, Poland
*
Author to whom correspondence should be addressed.
authors declare equal contribution.
Catalysts 2020, 10(11), 1328; https://doi.org/10.3390/catal10111328
Submission received: 19 October 2020 / Revised: 1 November 2020 / Accepted: 11 November 2020 / Published: 15 November 2020
(This article belongs to the Section Environmental Catalysis)

Abstract

:
The natural environment is constantly under threat from man-made pollution. More and more pharmaceuticals are recognized as emerging pollutants due to their growing concentration in the environment. One such chemical is ibuprofen which has been detected in processed sewage. The ineffectiveness of water methods treatment currently used raises the need for new remediation techniques, one of such is photodegradation of pollutants. In the present study, zinc(II) and copper(II) phthalocyanines were grafted onto pure anatase TiO2 nanoparticles (5 and 15 nm) to form photocatalysts for photodecomposition of ibuprofen in water. The nanoparticles were subjected to physicochemical characterization, including: thermogravimetric analysis, X-ray powder diffraction, X-ray photoelectron spectroscopy, Brunauer–Emmett–Teller surface area analysis and particle size measurements. In addition, they were assessed by means of electron spin resonance spectroscopy to evaluate the free radical generation. The materials were also tested for their photocatalytic activity under either UV (365 nm) or visible light (665 nm) irradiation. After 6 h of irradiation, almost complete removal of ibuprofen under UV light was observed, as assessed by liquid chromatography coupled to mass spectrometry. The reaction kinetics calculations revealed that the copper(II) phthalocyanine-containing nanoparticles were acting at a faster rate than those with zinc(II) derivative. The solutions after the photoremediation experiments were subjected to Microtox® acute toxicity analysis.

Graphical Abstract

1. Introduction

The natural environment all around the globe is constantly deteriorating due to human activity. The wholesale loss of animal species is associated with environmental damage as a result of decreasing natural habitats, an increase in air and seas temperatures, as well as change in ecosystems caused by alarmingly increasing concentrations of chemicals [1,2]. With the constant development of our civilization and growing industrialization, we are facing new problems within our environment. They concern an improper utilization of chemical contaminants, including industrial wastes, pharmaceuticals and cosmetics [3]. Many substances end up in the natural environment due to thoughtless disposal of wastes or abuse of drugs. Now labeled as emerging contaminants, they often pose a threat to the environment due to their inherent toxicity to plants and animals. When deposited in the ground, these wastes can travel and end up in water reservoirs spreading toxicity and endangering human health. Furthermore, this is an issue of grave importance given that technology employed in contemporary water treatment facilities cannot effectively remove all of the substances belonging to this broad group [4].
As previously stated, drugs and drug-derived molecules constitute a group of compounds that are considered an emerging threat to the environment [5]. Traces of many pharmaceutically and biologically active compounds are often reported in the environment as the result of their manufacturing, use and improper disposal (Figure 1) [6,7]. Some pharmaceuticals undergo degradation in wastewaters and therefore do not pose any human health or environmental threat but some compounds can remain unchanged for a significant time [8]. Pharmaceuticals and their metabolites escape into wastewater mostly through the domestic, hospital or veterinary use, the pharmaceutical industry or animal farming [8,9,10,11,12,13]. Since many sewage treatment plants do not possess tools to effectively remove these compounds, they remain in the effluents and, eventually, their traces can be found in drinking water [14]. One of such compounds is a non-steroid anti-inflammatory drug–ibuprofen, dispensed over the counter in the pharmacy, which is not removed entirely from the waste-water in water-treatment plants [4]. Considerable adverse effects on human health resulting from exposure to very low levels of pharmaceuticals in drinking water are very unlikely, according to the World Health Organization (WHO) report from 2011 [15]. However, the report highlighted a few issues, which remain unsolved until this day, one of which being knowledge gaps associated with long-term, low-level exposure to pharmaceuticals. Little is also known about the combined effects of their mixtures, especially for sensitive subpopulations. It is noteworthy that global spending on medicines has been growing in recent years and is predicted to exceed $1.5 trillion by 2023 [16]. This is one of the main reasons why new technologies of remediation are being pursued.
Although the WHO report provided recommendations on managing such concerns [15], technological solutions for wastewater, which would guarantee complete removal of all residues, are not yet in place [13]. Conventional wastewater treatments are often not effective enough, for example due to the low biodegradability of many pharmaceutical compounds. Therefore, attention more recently has been turned to alternative approaches, which include bioremediation, advanced oxidation processes (AOPs), ultrasonic treatment or adsorption techniques [5,17]. Among other AOPs, which employ the intermediacy of highly reactive species in order to decay the pollutant [7], photoremediation seems to be a promising method, especially when titanium(IV) oxide (titanium dioxide, TiO2) is used as the photocatalyst [18]. Various studies have already assessed this way of decomposing ibuprofen (IBU) in water solutions [4,19,20,21]. There is also evidence that TiO2 samples impregnated with photosensitizers, for example, porphyrins and phthalocyanine derivatives, enhance the degradation efficiency of organic pollutants [22].
Photoremediation is a relatively cheap and efficient method for removal of chemical contaminants. The process uses photocatalytic degradation of chemicals by exposing the photosensitizer to UV-Vis light. The photosensitizer absorbs energy from the light, which can then be transferred to the molecules present in the reaction environment. The released energy can interact with oxygen producing reactive oxygen species such as superoxides, singlet oxygen or hydroxyl radicals [23]. Metal oxides, such as zinc(II) oxide and titanium(IV) oxide, belong to chemical compounds that have so far been most commonly applied as photosensitizers in various photoremediation studies [24]. TiO2 has been used for the degradation of pharmaceuticals belonging to various groups, including antibiotics, analgesics, anticonvulsants, β-blockers, lipid regulators, non-steroid anti-inflammatory drugs, psychiatric drugs [25]. The results of broad studies are promising, with near 100% removal efficiency of pollutants in many cases being achieved. However, it is worth noting that in most experiments, the UV light, which is not an optimal source of light in water solutions, was applied [26]. This is due to the TiO2 band gap that only allows absorption of light below 387 nm. Ideally, for water remediation purposes one would require a material that is active under exposure to visible light, which is mainly dictated by its cost-effectiveness. Functionalization of TiO2 surface can significantly improve its physicochemical properties. One way of such functionalization is grafting the surface with macrocyclic porphyrinoid compounds like porphyrazines and phthalocyanines (Pc). These compounds reveal a broad excitation band that includes a visible light region. Additionally, the periphery of porphyrinoids can be modified, which even further enhances the desired characteristics [27]. By depositing porphyrinoid compounds on TiO2 surface it was possible to obtain hybrid materials that can be excited by the visible light [28].
In this work, we present the results of our studies where nanometric size TiO2 particles were functionalized with copper(II) and zinc(II) phthalocyanines (CuPc and ZnPc, respectively). The obtained materials were characterized in terms of their physicochemical properties, using particle size distribution, X-ray powder diffraction (XRPD), X-ray photoelectron spectroscopy (XPS), thermogravimetric analysis (TGA), Brunauer–Emmett–Teller (BET) surface area analysis and electron spin resonance (ESR) spectroscopy, as well as prospective utility for water remediation. Therefore, the TiO2-Pc composites were assessed for their photocatalytic activity towards photodegradation of ibuprofen, which is a common pharmaceutical water pollutant. The decomposition was monitored by liquid chromatography coupled to mass spectrometry (LC-MS/MS) and the acute toxicity of the resulting solutions was evaluated using Microtox®.

2. Results and Discussion

2.1. Preparation and Characterization of the Nanoparticles

The prepared materials were based on commercially available anatase phase titanium(IV) oxide nanoparticles. The choice of this crystalline phase was based firstly on the highest photoactivity of the anatase and secondly on the best stability of anatase nanoparticles in such a small diameter [28,29,30]. The nanoparticles (NPs) were prepared using the Pc : TiO2 ratio of 1:100, based on the optimal activity observed for similar materials [31]. The researched phthalocyanines were chosen according to their high photoactivity—ZnPc [32,33] and earlier reported effectiveness in photoremediation—CuPc [22,31,34]. Prepared nanomaterials, based on 5 nm and 15 nm TiO2 grafted with zinc(II) or copper(II) phthalocyanine, are summarized in Table 1.
All the prepared materials, along with neat TiO2 NPs, were first subjected to physicochemical characterization to confirm later their potential usefulness for photoremediation. The applied techniques included: TGA, XRPD, particle size and BET surface area measurements, as well as ESR.

2.2. Thermoanalysis

The determination of the phase composition of Pc-grafted TiO2 photocatalysts was based on thermogravimetry results by analyzing the loss of mass (%), as well as the rate of mass loss during heating. In the case of unmodified commercial TiO2 (anatase), two distinct stages were observed. The first (up to 120 °C) was associated with water desorption (2.1%) [35]. In the second stage (temperature range of 120 °C-500 °C), which was accompanied by approx. 3.71% weight loss, there was a release of chemically bound water molecules observed. Water molecules in this step were formed as a result of the condensation of the surface hydroxyl groups. The obtained results are consistent with the literature data. Many authors emphasize that the temperature range of 120-500 °C is suitable for separating physically adsorbed and chemically bound water molecules [35]. It is assumed that as a result of further heating to 900 °C, the TiO2 surface becomes free of OH surface groups.
In the thermograms of the two analyzed phthalocyanines–ZnPc and CuPc–a similar effect can be observed–heating to 120 °C resulted in approx. 2.6% loss of adsorbed water. However, further heating to 670 °C, resulted in rapid decompositions of the Pcs, which were accompanied by about 62% weight loss. After exceeding this temperature, reductions in the decomposition rates (to 900 °C) were observed.
Titanium(IV) oxide composites with ZnPc and CuPc deposited by surface impregnation were also subjected to TGA. Samples containing commercial TiO2 with a particle size of 5 nm and 15 nm were analyzed. Quantitative evaluation of the obtained thermograms showed that heating in the temperature range of 20-120 °C resulted in approx. 3% weight loss caused by the desorption of adsorbed water. However, as a result of further heating, approx. 5% weight loss was observed, which could be associated with the condensation of surface hydroxyl groups (120-500 °C) accompanied by water release and the decomposition of phthalocyanine (300-500 °C). It was found that the smaller 5 nm particles expressed a higher loss of mass during the analysis than their 15 nm counterparts, regardless of their functionalization. This fact could be attributed to the higher surface area of 5 nm particles and would further translate to the number of hydroxyl groups present at the surface that could be thermally condensed with water release. However, no regularities were found when comparing bare titania nanoparticles to ZnPc or CuPc grafted materials, regardless of their sizes.
Exemplary thermograms of TiO2 nanocomposites with Pcs are presented in Figure 2.

2.3. X-ray Powder Diffraction Analysis

All the prepared materials were subjected to XRPD experiments. As presented in Figure 3, there is no difference between the diffraction patterns of neat TiO2 nanoparticles and NPs grafted with either ZnPc or CuPc. It is consistent with the literature reports since it is assumed that in such composites, macrocyclic compounds occur as amorphous substances dispersed on TiO2 surface [32,35,36]. The reflections observed in the diffractograms of 1%Cu/1%ZnPc@TiO2 correspond to the anatase phase, suggesting that the Pc deposition method used does not change the polymorph of titanium dioxide. In diffraction patterns of 5 nm TiO2 materials a decrease in peak intensity and broadening of peaks in comparison to 15 nm TiO2 materials is observed. This finding can be related to the smaller size of crystallites [37]. Calculation of the crystallite sizes using Scherrer equation showed no differences between the initial bare titania nanoparticles and the final materials prepared from the respective size particles.

2.4. XPS

All the prepared materials were subjected to X-ray photoelectron spectroscopy experiments to detect the binding energy on the bare TiO2 and TiO2-phthalocyanine catalysts. The obtained spectra are collected in Figure 4. In the spectra of all the materials, signals from Ti, O and C can be observed, as labeled in Figure 4a. The drop in intensity of the peaks is visible in the 15 nm Pc@TiO2 materials (see Figure 4d–f) in comparison to 5 nm Pc@TiO2 (see Figure 4a–c), which is related to their lower surface area. Most notably, due to lack of covalent bonding, there were no noticeable shifts found when comparing the spectra of bare TiO2 and Pc@TiO2. The probable mode of binding of the phthalocyanines to titanium dioxide nanoparticles is via the oxygen atoms of TiO2, either through hydrogen bonding or coordination of the Pc by the oxygen lone electron pairs [38]. Therefore, in the spectra, the presence of O 1s signals at 530.75 eV and 530.91 eV for a bare TiO2 (5 nm) and both Pc-grafted TiO2 (5 nm), respectively, were noted. The presence of the low-intensity C 1s carbon signal for bare TiO2 samples was attributed to the adsorption of airborne CO2 on their surface [39]. The significant increase of C 1s intensity is observed in the spectra of Pc-grafted TiO2 materials. The lack of nitrogen, as well as Zn and Cu signals in the spectra, is caused by the very low content of these elements in the studied 1%Pc@TiO2 materials [40].

2.5. Particle Size

The size of the particles was acquired using nanoparticle tracking analysis (NTA). The results are summarized in Table 2. The obtained particle sizes reveal that the materials were prone to agglomeration, which is strongly marked. The example size distribution of the NPs is shown in Figure 5. As a result of the agglomeration, none of the particles can be regarded as monodisperse. The polydispersity index of all the materials exceeds 0.2 [41].

2.6. Electron Spin Resonance Spectroscopy

To check the effect of exposure on free radical generation in samples, the materials were also studied by ESR method. The exemplary ESR spectra are shown in Figure 6.
For both types of samples, containing either zinc(II) or copper(II) phthalocyanine, a single line is visible. Typical spectroscopic parameters were determined: for zinc(II) phthalocyanine (1%ZnPc@TiO2) g=2.0032, ΔH=0.75 mT and for copper(II) phthalocyanine (1%CuPc@TiO2) g=2.0506, ΔH=4.67 mT. After the exposure to 665 nm light irradiation, the signal intensity changed, which means that the free radicals were generated. It seems that these changes depend on the time the material was irradiated: after 15 min the signal rapidly decreases and after another 15 min it increases but is still lower than non-irradiated sample (Figure 6a). Such changes were not observed for pure ZnPc, for which the signal intensity was stable (Figure 6b). In the case of differently sized TiO2 nanoparticles it was observed that the non-irradiated sample with the particle size of 5 nm gave a very weak, trace signal. In contrast, the measurements of the sample of the larger particle size resulted in a complex signal, which decreased significantly after exposure. The observed changes in signal intensity are probably associated with the surface defects that are likely to promote the rapid recombination of generated free radicals. A thorough understanding of these changes and mechanisms that take place requires further study.

2.7. BET Surface Area Analysis

BET surface area analyses were performed for all the studied materials to assess whether the modification of titania with the phthalocyanines revealed any effect on the materials’ surface. The results are summarized in Table 3.
The difference found in the surface areas between 5 nm and 15 nm materials is evident. The 5 nm nanoparticles exhibit a surface area of around 250 m2/g, while the 15 nm ones present values close to 100 m2/g. By comparing the data, the methodology used for the modification of both TiO2 materials with 1%CuPc and 1%ZnPc caused small decreases in the materials’ surfaces, resulting from two factors. Firstly, the presence of phthalocyanines in titania nanoparticles pores. Secondly, the fact that the combination of nanoparticles with macrocycles can lead to an agglomeration of the received materials.

2.8. Photochemical Studies

2.8.1. Photodegradation of Ibuprofen

Ibuprofen is one of the most commonly used drugs in the world, for many years included in the WHO Model List of Essential Medicines [43]. Similar to other nonsteroidal anti-inflammatory drugs, it is used to treat pain, inflammation and to reduce fever both in adult and pediatric patients [44]. As an over-the-counter (OTC) drug with many indications, it is often taken without medical supervision [45]. Recent advances in analytical technology facilitate the detection of pharmaceutical compounds. The presence of ibuprofen in both wastewater influents and effluents has already been reported by numerous researchers [6,46,47,48]. The removal of ibuprofen or its metabolites in the course of the drinking water treatment process is not always efficient [49]. This raises concerns over potential risks to human health and aquatic ecosystems from exposure to ibuprofen traces in drinking water. The above considerations demonstrate that a modern, efficient technology ought to be applied for the removal of pharmaceuticals from water.
The photocatalytic systems based on Pc-TiO2 or TiO2 nanoparticle suspension as photocatalysts were used for the assessment of ibuprofen removal from water via photodegradation process. To assess the photodegradation efficiency of both nanomaterials, IBU removal using 1%CuPc@TiO2 and 1%ZnPc@TiO2 with TiO2 alone was compared. It has been previously observed that using pure TiO2 it is possible to almost completely remove IBU in 2–3 h [50,51,52,53]. Furthermore, an approach was previously undertaken, where commercially available P25 TiO2 nanoparticles (~21 nm) were used in combination with UV light to successfully degrade IBU [54]. What is more, the authors managed to identify a series of resulting IBU decomposition products. However, the authors did not use TiO2 based visible-light photocatalysts, only a neat P25 TiO2 and UV irradiation. It is worth noting, that the use of a sensitizing moiety was presented by Patterson et al., who used natural plant extracts to broaden the absorption spectrum of P25 TiO2 [55]. The composites were used to form thin films and natural sunlight was tested as the light source, which allowed to double the efficacy of the photodegradation of IBU with neat P25 TiO2. In the herein presented study the solution of IBU in water was irradiated with the photocatalyst and accompanied by careful monitoring of the IBU concentration. The results are summarized in Figure 7 and Figure 8.
In contrast to the literature findings [53], after 2 h of our experiment, the levels of IBU were around 60% when referred to the IBU level at t0. In the case of UV-light driven photocatalytic degradation with 1%CuPc@TiO2 and pure TiO2, almost complete removal of IBU was observed after six hours of our experiment. The use of the 665 nm red light did not cause any change in the IBU content in the tested experiments. Although one would expect such outcome in the case of neat TiO2 nanoparticles, ZnPc and CuPc grafted materials also did not yield any effect in these conditions despite the fact that both these phthalocyanines absorb the light at this wavelength. This is in contrast to the earlier literature reports, where ZnPc deposited on P25 TiO2 in an analogical manner was found photocatalytically active and enabled photodegradation of erythromycin [32]. Nevertheless, there are certain differences between the herein presented and the literature procedures. In our study the calcination step, following the Pc deposition on TiO2 nanoparticles, was omitted. It is believed that this step irrevocably leads to the decomposition of the phthalocyanine, as proven by the TGA measurements. Presumably, the material obtained before [32] was composed of P25 TiO2 doped with C, N or Zn on its surface with some residual phthalocyanine. Herein, in the presented study, probably due to the non-stabilized nature of the tested nanoparticles, only a portion of grafted macrocycles was present on the surface and could come in close proximity of IBU for the photocatalytic reaction to occur.
Although the factors mentioned above may have had the greatest impact on the result of the experiment, other differences will also be briefly discussed below. Factors such as the size and load of TiO2 particles, the initial IBU concentration, temperature were found to influence the photodegradation of IBU in aqueous solutions and have been explored in several studies [51]. Throughout our experiment the temperature did not exceed 35 °C, which seems to be consistent with other photocatalytic degradation research studies. As mentioned before, prior studies have also indicated the influence of the initial drug concentration on the process. The observed efficiency was usually increasing with the decrease of the initial concentration of IBU [51,56]. Additionally, the effect of the catalyst load has been reported in the literature. Several reports have shown that the optimal load of the suspended TiO2 is around 1000 mg/L, however, these values typically varied between 50-3000 mg/L [50,51,54,56,57]. Consistent with the literature, the initial IBU concentration in our study was 10 mg/L, whereas the TiO2 load was 100 mg/0.1 L. Hence, these parameters were expected to reveal similar degradation rates at similar sampling times. Another major factor, which could not be omitted, is the average particle size. In our experiment nano-grade TiO2 was used. Nano-scale particles have a higher surface-to-volume ratio than microparticles and thanks to the larger surface available for photocatalytic reactions, they are expected to be more efficient in the photodegradation studies. However, in the present study, the NPs tended to form agglomerates. Therefore, the access of IBU to the surface of the photocatalyst might have been limited, as it was suggested in other studies [51,58].
In the case of the UV irradiation experiments, another interesting fact was observed. Namely, the neat TiO2 was photocatalytically active to a similar extent as CuPc grafted titania, regardless of the nanoparticle size and the surface area presented by the nanomaterial. On the opposite, deposition of ZnPc on titanium(IV) oxide resulted in the decrease of IBU degradation rate in the case of both, 5 nm and 15 nm particles. This effect could be explained by the absorption of light by ZnPc deposited on the surface of TiO2 with the formation of Pc-derived reactive oxygen species, in particular very short-living singlet oxygen molecules. ZnPc is known to produce singlet oxygen, which would probably be dissipated before reaching IBU molecules [59]. In the case of CuPc, because of lower singlet oxygen generation quantum yields, the energy absorbed by the phthalocyanine had to be transferred to TiO2. In addition, the same results obtained for photocatalysts, regardless of the 5 or 15 nm particle size used, can be associated with the agglomeration of the non-stabilized nanoparticles [60]. As can be observed in Table 2., the measured mean sizes of the materials are much higher than those obtained for neat TiO2 nanoparticles and the wide-spread of the results is revealed by the high values of standard deviations. All these facts suggest that stabilization of the nanoparticles is necessary for the potential photocatalytic material.

2.8.2. Calculation of Photodegradation Constants for Various Conditions Applied

The kinetic parameters (degradation rate constants k of IBU decay) were calculated for experiments with UV light irradiation and for the following series of photocatalysts: 1%ZnPc@TiO2(5 nm), 1%ZnPc@TiO2(15 nm), 1%CuPc@TiO2(5 nm) 1%CuPc@TiO2(15 nm), TiO2(5 nm) and TiO2(15 nm).
The results are depicted in Table 4. and in Figure 9 and Figure 10. The studied reaction was defined as the first-order kinetics and the degradation rate constant was calculated using the formula (1):
ln Pt = ln P0–k · t,
where: Pt surface area of the sample in time t [h] in the isothermal test, P0 is the surface area of the sample in time 0 and k [s−1] is the reaction rate constant. According to the theory of first-order kinetics, the semi-logarithmic plot Pt = f(t) is linear and its slope corresponds to the magnitude of the degradation rate constant k (Figure 9 and Figure 10). Thus, the least-squares method was used to calculate the regression parameters: y = ax + b, a ± Δa, b ± Δb, standard errors Sa, Sb and the correlation coefficient r. The ±Δa, ±Δb were estimated for f = n − 2 degrees of freedom and α = 0.05. Table 4. summarizes the results for each experiment.

2.9. Acute Toxicity Assessment

The solutions of ibuprofen were subjected to ecotoxicity testing using the Microtox® test. The test is based on Aliivibrio fischeri bacterial cell suspension. The changes in bioluminescence after a sample is added can be directly correlated with the metabolism of the bacteria. So that the luminescence decreases linearly with the sample toxicity increase [61,62]. As can be seen in Figure 11a, the toxicity of the ibuprofen solutions photoremediated with UV light slightly increases in comparison with IBU solutions before the irradiation. Since ibuprofen is oxidized to other products which might express higher toxicity, these results can be associated with degradation of IBU to its photoproducts [63].
As can be seen in Figure 11b, the solutions photocatalyzed with red light by neat TiO2 nanoparticles exhibit unchanged toxicity to A. fischeri bacteria. In case of phthalocyanine-grafted titania nanoparticles, despite the UV or red light irradiation, the toxicities of both solutions are comparable. A slight ca. 10% toxicity increase of the ibuprofen irradiated solutions after 6 h of the UV light irradiation was noted. In addition, this effect was weaker for Pc@TiO2(15 nm) than for Pc@TiO2(5 nm). It indicates that thephotocatalytic activity of anatase is potentiated with the use of proper modification of titania nanoparticles.

3. Materials and Methods

3.1. Materials and Instruments

All reaction mixtures were stirred using Radleys Heat-On™ heating system (Saffron Walden, United Kingdom). Solvents and all reagents were obtained from commercial suppliers (Merck, Darmstadt, Germany; Fluorochem, Hadfield, United Kingdom; Chempur, Piekary Śląskie, Poland; Avantor Performance Materials Poland S.A., Gliwice, Poland) and used without further purification. TiO2 anatase nanoparticles were purchased from US Research Nanomaterials Inc., Houston, TX, USA.
Mass spectrometry-grade methanol and ammonium acetate were purchased from Sigma–Aldrich (Saint Louis, MO, USA). Mass spectrometry-grade water was prepared by reverse osmosis in a Demiwa system from Watek (Ledec nad Sazavou, Czech Republic), followed by double distillation from a quartz apparatus.

3.2. Preparation of Nanoparticles

TiO2-Pc nanoparticles were prepared as a modification of chemical deposition method [64]. As an example, for the preparation of 1%ZnPc@TiO2 material, zinc(II) phthalocyanine (10 mg, 0.017 mmol) was dissolved in analytical grade dichloromethane in a round-bottom flask. Titanium dioxide (1000 mg) of appropriate particle size was added and the suspension was sonicated for 0.5 h and then stirred for 1 hour. After that, the solvent was removed using a rotary evaporator. The resulting blue powder was dried in air for 10 h to remove any traces of the solvent.
For the means of comparison, pure TiO2 was treated according to the same procedure but without the addition of the macrocycle.

3.3. Nanoparticle Characterization

3.3.1. Thermogravimetric Analysis

Thermogravimetric analysis was performed at the Wielkopolska Center for Advanced Technologies in Poznań, Poland, with the use of a thermogravimetric analyzer (TGA 4000, Perkin Elmer, Kraków, Poland). The analyzed materials were: TiO2 (5 nm), TiO2 (15 nm), neat ZnPc, neat CuPc and Pc@TiO2 composites. Samples of around 5 mg were heated from room temperature to 900 °C at the speed of 10 °C/min. The experiments were conducted in a nitrogen atmosphere. Temperature measurements were accurate up to 10−5 °C, while the masses with the accuracy of 10−5 mg.

3.3.2. X-ray Powder Diffraction

X-ray diffraction patterns of the samples were recorded at the Poznan University of Medical Sciences Core Facility on a Bruker AXS D2 Phaser diffractometer (Bruker AXS GmbH, Karlsruhe, Germany) with a Cu Kα anode (λ = 1.54060 Å) operating at 30 kV and 10 mA. The diffraction patterns were collected at ambient temperature over an angular measurement range of 5° to 60° 2θ with a step size of 0.02° and a counting rate of 2 s/step with the sample spinning.

3.3.3. X-ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy (XPS) experiments were performed under 10−9 mbar vacuum using SPECS Multimethod system with XPS, SPECS Surface Nano Analysis GmbH (Berlin, Germany). Al anode was used as the source of non-monochromatic X-rays. CasaXPS software was used to analyze the data.

3.3.4. Particle Size

The nanoparticle size distribution was analyzed using Malvern Panalytical (Malvern, United Kingdom) NanoSight LM10 instrument (sCMOS camera, 405 nm laser) using NTA (Nanoparticle Tracking Analysis) 3.2 Dev Build 3.2.16 software. Before the measurements, the NP dispersions were diluted with water to achieve the operating range of nanoparticle concentration. The temperature of the sample chamber was set and maintained at 25.0 ± 0.1 °C and the syringe pump infusion rate was set to 200. For each sample, three movies of 30 seconds in length were recorded.

3.3.5. Electron Spin Resonance Spectroscopy

All samples were studied using ESR spectroscopy. The measurements were carried out on an X-band Bruker EMX-10 spectrometer (Bruker, Billerica, MA, USA) with a magnetic field second modulation frequency of 100 kHz. The ESR spectra were recorded at room temperature (293 K) in two magnetic field ranges: 60 mT or 4 mT. Each sample was measured three times: before exposure, after exposure for 15 min and after subsequent exposure for another 15 min. Typical spectroscopic parameters: g-factor value, peak-to-peak line width (ΔH) and ESR signal intensity (corresponding to the concentration of free radicals in the sample) were determined for ESR spectra.

3.3.6. BET Surface Area Analysis

The Brunauer-Emmet-Teller surface area analysis was performed at the Wielkopolska Center for Advanced Technologies in Poznań, Poland, with the use of ASAP™2420 system (Micromeritics®, Norcross, GA, USA).

3.4. Photochemical Studies

3.4.1. Set-Up of the Photocatalytic Experiment

The source of light in the irradiation experiments was a set of three laser diodes with λmax = 665 nm for the red light irradiation or λmax = 365 nm for UV irradiation. Light intensity was set to 20 mW/cm2 and monitored with a RD 0.2/2 radiometer (Optel, Wrocław, Poland). Distilled water was used in all experiments.
The photodegradation experiment was carried out in a reactor, which consisted of three beakers placed on a magnetic stirrer in the middle and three UV-Vis lamps with peak emission wavelength of either 365 nm or 665 nm, positioned on a circular line (Figure 12). A layer of aluminum foil was placed behind and above the photoreactor to prevent scattering of light or heat energy and to protect the operator. To each of the beakers containing 100 mL of IBU solution at concentration 10 mg/L, 100 mg of TiO2-based material was added, sonicated for about 1 minute and stirred vigorously in the dark for 30 min. After that, LED lamps (either UV or Vis) were turned on and the mixtures were irradiated for 6 h while being constantly stirred. Samples (~2 mL) were taken at time-points: -0.5 h, 0 h, 0.5 h, 1.0 h, 2.0 h, 4.0 h, 6.0 h after the irradiation was started. The temperature of the solution during the photocatalytic reaction did not exceed 35 °C.
Collected samples were centrifuged (10000 rpm, 30 min) and the supernatant was filtered through 0.2 µm poly(tetrafluoroethylene) (PTFE) syringe filters. Before the liquid chromatography coupled to mass spectrometry (LC-MS/MS) analysis the samples were diluted 20 times in a methanol/water 1:1 (v/v) mixture.

3.4.2. LC-MS/MS Analysis of the Samples

A chromatographic system UltiMate 3000 RSLC from Dionex (Sunnyvale, CA, USA) coupled with the API 4000 QTRAP triple quadrupole mass spectrometer from AB Sciex (Foster City, CA, USA) was used and 5 µL samples were injected into a Synergi Fusion-RP column (50 mm × 2.0 mm I.D.; 2,5 µm) from Phenomenex (Torrance, CA, USA) maintained at 35 °C. The mobile phase employed in the analysis consisted of 5 × 10−3 mol L−1 ammonium acetate in water and methanol at a flow rate of 0.3 mL min−1. Gradient elution was performed by linearly increasing the percentage of organic modifier from 70 to 100% in 2.5 min and then holding it at 100% to 3 min. A pre-run time of 3 min was used before the next injection. The effluent from the LC column was introduced into the electrospray ionization source (Turbo Ion Spray) operated in negative ion mode. The following settings were used for the ion source and mass spectrometer—curtain gas 10 psi, nebulizer gas 40 psi, auxiliary gas 40 psi, temperature 400 °C, ion spray voltage −3500 V, declustering potential −35 V and collision gas set to medium. The dwell time for mass transition detected in the reaction monitoring mode was set to 100 ms. The quantitative transition was from 205 to 161 m/z at collision energy set to −10 eV.
The linearity of the method was tested in a range of 2∓1000 [µg L−1]. The instrumental limit of detection (LOD) and the instrumental limit of quantitation (LOQ) were calculated on the basis of signal to noise (S/N) ratio. The S/N = 3 was used for calculation of LOD and the S/N = 10 for calculation of LOQ. Precision and accuracy were not tested because sample preparation included only a dilution step. Therefore, only the injection precision of the instrument applies in this procedure, which was always below 1%. For ibuprofen LOD = 0.5 [µg L−1], LOQ = 1.7 [µg L−1].

3.5. Microtox® Acute Toxicity Assessment

The samples of the remediated solutions were tested using Microtox® acute toxicity test–81.9% Screening Test which was performed using Microtox® M500 equipment, according to the protocols distributed by the producer (ModernWater plc, London, United Kingdom) [61,65]. Cell viability was calculated according to bioluminescence emitted by the Aliivibrio fischeri bacteria as measured with Microtox® M500 with Modern Water MicrotoxOmni 4.2 software.

4. Conclusions

To sum up, materials based on pure anatase titania nanoparticles and phthalocyanines were prepared by a chemical deposition method, as well as carefully analyzed by TGA, XRPD, XPS, particle size measurements and ESR. The analytical techniques revealed that the applied TiO2 modification method was suitable because of having no effect on anatase crystallographic phase. As exhibited in ESR measurements, the materials were able to generate free radicals and their formation changed upon red-light exposure. As the materials were found to agglomerate, proper stabilization of the nanoparticles will be considered in our further studies.
The experiments with ibuprofen aqueous solutions indicated that the best results can be obtained after UV light irradiation. CuPc doped TiO2 was found to be as effective as neat TiO2 when exposed to 365 nm light, whilst the ZnPc modified titania was shown to photocatalyze the degradation of ibuprofen at a slower rate. The experiments performed using the modified titania and red light at 665 nm demonstrated the lack of effectiveness of the obtained materials, as well as neat TiO2, in these conditions. The acute toxicity assessment showed only a slight increase in toxicity of the ibuprofen solutions after the photodegradation experiment in the presence of modified TiO2 nanoparticles. The results show that the use of such approach does not impede the excellent photocatalytic activity of anatase and it might be potentiated with the use of proper modification of titania nanoparticles in the form of robust molecules preventing agglomeration or the use of surfactants.

Author Contributions

Conceptualization: T.G., B.J.S., A.Z.-G., R.B., R.K. (Ryszard Krzyminiewski) and M.S. (Marek Sikorski); methodology: R.K. (Rafal Krakowiak), J.M. (Joanna Musial), R.F., J.M. (Jadwiga Mielcarek), B.D., W.B.-B. and D.T.M.; software: W.B.-B. and E.T.; validation: J.M. (Joanna Musial), R.F., A.Z.-G. and B.J.S.; formal analysis: J.M. (Joanna Musial), R.F., J.M. (Jadwiga Mielcarek), B.D., W.B.-B., A.Z.-G., B.J.S., D.T.M.; investigation: R.K. (Rafal Krakowiak), J.M. (Joanna Musial), R.F., M.S. (Marcin Spychala), J.M. (Jadwiga Mielcarek), B.D., W.B.-B., A.Z.-G., D.T.M.; resources: R.K. (Ryszard Krzyminiewski), E.T., R.B., A.Z.-G., B.J.S., T.G.; data curation: R.K. (Rafal Krakowiak), J.M. (Joanna Musial), J.M. (Jadwiga Mielcarek), B.D., W.B.-B., A.Z.-G., B.J.S., T.G., D.T.M; writing—original draft preparation: Ra.K., J.Mu., M.Sp., J.Mi., B.D., A.Z.-G., B.J.S., D.T.M.; writing—review and editing: R.K. (Rafal Krakowiak), J.M. (Joanna Musial), J.M. (Jadwiga Mielcarek), M.S. (Marek Sikorski), E.T., B.J.S., T.G., D.T.M.; visualization: R.K. (Rafal Krakowiak), J.M. (Joanna Musial), B.D., A.Z.-G., B.J.S.; supervision: T.G., R.B., E.T., R.K. (Ryszard Krzyminiewski); project administration: T.G.; funding acquisition: T.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Centre, Poland, grant number 2016/21/B/NZ9/00783.

Acknowledgments

The authors would like to acknowledge Poznan University of Medical Sciences Core Facility for providing access to D2 PHASER diffractometer. The authors thank Beata Kwiatkowska and Rita Kuba for excellent technical assistance. DTM would like to thank Agata Kaluzna-Mlynarczyk for her support. The authors acknowledge the preliminary experiments by Anna Bordzio.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript or in the decision to publish the results.

References

  1. Ling, S.D.; Davey, A.; Reeves, S.E.; Gaylard, S.; Davies, P.L.; Stuart-Smith, R.D.; Edgar, G.J. Pollution signature for temperate reef biodiversity is short and simple. Mar. Pollut. Bull. 2018, 130, 159–169. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Q.; Yang, Z. Industrial water pollution, water environment treatment and health risks in China. Environ. Pollut. 2016, 218, 358–365. [Google Scholar] [CrossRef] [PubMed]
  3. La Farré, M.; Pérez, S.; Kantiani, L.; Barceló, D. Fate and toxicity of emerging pollutants, their metabolites and transformation products in the aquatic environment. TrAC Trends Anal. Chem. 2008, 27, 991–1007. [Google Scholar] [CrossRef]
  4. Candido, J.P.; Andrade, S.J.; Fonseca, A.L.; Silva, F.S.; Silva, M.R.A.; Kondo, M.M. Ibuprofen removal by heterogeneous photocatalysis and ecotoxicological evaluation of the treated solutions. Environ. Sci. Pollut. Res. 2016, 23, 19911–19920. [Google Scholar] [CrossRef] [PubMed]
  5. Carvalho, A.; Mestre, A.; Andrade, M.; Ania, C. Ibuprofen in the Aquatic Environment: Occurrence, Ecotoxicity and Water Remediation Technologies. In Ibuprofen: Clinical Pharmacology, Medical Uses and Adverse Effects; Nova Science Pub Inc.: Hauppauge, NY, USA, 2013; pp. 1–84. ISBN 978-1-62618-659-0. [Google Scholar]
  6. Kostich, M.S.; Batt, A.L.; Lazorchak, J.M. Concentrations of prioritized pharmaceuticals in effluents from 50 large wastewater treatment plants in the US and implications for risk estimation. Environ. Pollut. 2014, 184, 354–359. [Google Scholar] [CrossRef]
  7. Gadipelly, C.; Pérez-González, A.; Yadav, G.D.; Ortiz, I.; Ibáñez, R.; Rathod, V.K.; Marathe, K.V. Pharmaceutical Industry Wastewater: Review of the Technologies for Water Treatment and Reuse. Ind. Eng. Chem. Res. 2014, 53, 11571–11592. [Google Scholar] [CrossRef]
  8. Majumder, A.; Gupta, B.; Gupta, A.K. Pharmaceutically active compounds in aqueous environment: A status, toxicity and insights of remediation. Environ. Res. 2019, 176, 108542. [Google Scholar] [CrossRef]
  9. Godoy, A.A.; Kummrow, F.; Pamplin, P.A.Z. Occurrence, ecotoxicological effects and risk assessment of antihypertensive pharmaceutical residues in the aquatic environment - A review. Chemosphere 2015, 138, 281–291. [Google Scholar] [CrossRef]
  10. Cheng, D.; Ngo, H.H.; Guo, W.; Chang, S.W.; Nguyen, D.D.; Liu, Y.; Wei, Q.; Wei, D. A critical review on antibiotics and hormones in swine wastewater: Water pollution problems and control approaches. J. Hazard. Mater. 2020, 387, 121682. [Google Scholar] [CrossRef]
  11. Sim, W.-J.; Lee, J.-W.; Lee, E.-S.; Shin, S.-K.; Hwang, S.-R.; Oh, J.-E. Occurrence and distribution of pharmaceuticals in wastewater from households, livestock farms, hospitals and pharmaceutical manufactures. Chemosphere 2011, 82, 179–186. [Google Scholar] [CrossRef]
  12. Santos, L.H.; Gros, M.; Rodriguez-Mozaz, S.; Delerue-Matos, C.; Pena, A.; Barceló, D.; Montenegro, M.C.B.S.M. Contribution of hospital effluents to the load of pharmaceuticals in urban wastewaters: Identification of ecologically relevant pharmaceuticals. Sci. Total Environ. 2013, 461–462, 302–316. [Google Scholar] [CrossRef] [PubMed]
  13. Papageorgiou, M.; Zioris, I.; Danis, T.; Bikiaris, D.; Lambropoulou, D. Comprehensive investigation of a wide range of pharmaceuticals and personal care products in urban and hospital wastewaters in Greece. Sci. Total Environ. 2019, 694, 133565. [Google Scholar] [CrossRef] [PubMed]
  14. Behera, S.K.; Kim, H.W.; Oh, J.-E.; Park, H.-S. Occurrence and removal of antibiotics, hormones and several other pharmaceuticals in wastewater treatment plants of the largest industrial city of Korea. Sci. Total Environ. 2011, 409, 4351–4360. [Google Scholar] [CrossRef] [PubMed]
  15. World Health Organization. Pharmaceuticals in Drinking-Water; World Health Organization: Geneva, Switzerland, 2012. [Google Scholar]
  16. IQVIA Institute for Human Data Sciences. The Global Use of Medicine in 2019 and Outlook to 2023-Forecasts and Areas to Watch; IQVIA: Durham, NC, USA, 2019. [Google Scholar]
  17. Méndez-Arriaga, F.; Torres-Palma, R.A.; Pétrier, C.; Esplugas, S.; Gimenez, J.; Pulgarin, C. Ultrasonic treatment of water contaminated with ibuprofen. Water Res. 2008, 42, 4243–4248. [Google Scholar] [CrossRef]
  18. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  19. Choina, J.; Kosslick, H.; Fischer, C.; Flechsig, G.-U.; Frunza, L.; Schulz, A. Photocatalytic decomposition of pharmaceutical ibuprofen pollutions in water over titania catalyst. Appl. Catal. B Environ. 2013, 129, 589–598. [Google Scholar] [CrossRef]
  20. Lin, L.; Jiang, W.; Bechelany, M.; Nasr, M.; Jarvis, J.; Schaub, T.; Sapkota, R.R.; Miele, P.; Wang, H.; Xu, P. Adsorption and photocatalytic oxidation of ibuprofen using nanocomposites of TiO2 nanofibers combined with BN nanosheets: Degradation products and mechanisms. Chemosphere 2019, 220, 921–929. [Google Scholar] [CrossRef]
  21. Jallouli, N.; Pastrana-Martínez, L.M.; Ribeiro, A.R.; Moreira, N.F.F.; Faria, J.L.; Hentati, O.; Silva, A.M.T.; Ksibi, M. Heterogeneous photocatalytic degradation of ibuprofen in ultrapure water, municipal and pharmaceutical industry wastewaters using a TiO2/UV-LED system. Chem. Eng. J. 2018, 334, 976–984. [Google Scholar] [CrossRef]
  22. Moro, P.; Donzello, M.P.; Ercolani, C.; Monacelli, F.; Moretti, G. Tetrakis-2,3-[5,6-di-(2-pyridyl)-pyrazino]porphyrazine and its Cu(II) complex as sensitizers in the TiO2-based photo-degradation of 4-nitrophenol. J. Photochem. Photobiol. Chem. 2011, 220, 77–83. [Google Scholar] [CrossRef]
  23. Simmler, W. Air, 6. Photochemical Degradation. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2011; Volume 50, pp. 464–488. ISBN 0007-4977. [Google Scholar]
  24. Byrne, C.; Subramanian, G.; Pillai, S.C. Recent advances in photocatalysis for environmental applications. J. Environ. Chem. Eng. 2018, 6, 3531–3555. [Google Scholar] [CrossRef]
  25. Kanakaraju, D.; Glass, B.D.; Oelgemöller, M. Advanced oxidation process-mediated removal of pharmaceuticals from water: A review. J. Environ. Manag. 2018, 219, 189–207. [Google Scholar] [CrossRef] [PubMed]
  26. Mozia, S.; Morawski, A.W. The performance of a hybrid photocatalysis–MD system for the treatment of tap water contaminated with ibuprofen. Catal. Today 2012, 193, 213–220. [Google Scholar] [CrossRef]
  27. Hiroto, S.; Miyake, Y.; Shinokubo, H. Synthesis and Functionalization of Porphyrins through Organometallic Methodologies. Chem. Rev. 2017, 117, 2910–3043. [Google Scholar] [CrossRef] [PubMed]
  28. Ziental, D.; Czarczynska-Goslinska, B.; Mlynarczyk, D.T.; Glowacka-Sobotta, A.; Stanisz, B.; Goslinski, T.; Sobotta, L. Titanium Dioxide Nanoparticles: Prospects and Applications in Medicine. Nanomaterials 2020, 10, 387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Zhang, H.; Banfield, J.F. Thermodynamic analysis of phase stability of nanocrystalline titania. J. Mater. Chem. 1998, 8, 2073–2076. [Google Scholar] [CrossRef]
  30. Luttrell, T.; Halpegamage, S.; Tao, J.; Kramer, A.; Sutter, E.; Batzill, M. Why is anatase a better photocatalyst than rutile?-Model studies on epitaxial TiO2 films. Sci. Rep. 2014, 4, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Mele, G.; del Sole, R.; Vasapollo, G.; García-López, E.; Palmisano, L.; Jun, L.; Słota, R.; Dyrda, G. TiO2-based photocatalysts impregnated with metallo-porphyrins employed for degradation of 4-nitrophenol in aqueous solutions: Role of metal and macrocycle. Res. Chem. Intermed. 2007, 33, 433–448. [Google Scholar] [CrossRef]
  32. Vignesh, K.; Rajarajan, M.; Suganthi, A. Photocatalytic degradation of erythromycin under visible light by zinc phthalocyanine-modified titania nanoparticles. Mater. Sci. Semicond. Process. 2014, 23, 98–103. [Google Scholar] [CrossRef]
  33. Fadel, M.; Kassab, K.; Abdel Fadeel, D. Zinc phthalocyanine-loaded PLGA biodegradable nanoparticles for photodynamic therapy in tumor-bearing mice. Lasers Med. Sci. 2010, 25, 283–292. [Google Scholar] [CrossRef]
  34. Cabir, B.; Yurderi, M.; Caner, N.; Agirtas, M.S.; Zahmakiran, M.; Kaya, M. Methylene blue photocatalytic degradation under visible light irradiation on copper phthalocyanine-sensitized TiO2 nanopowders. Mater. Sci. Eng. B 2017, 224, 9–17. [Google Scholar] [CrossRef]
  35. Altın, İ.; Sökmen, M.; Bıyıklıoğlu, Z. Quaternized zinc(II) phthalocyanine-sensitized TiO2: Surfactant-modified sol–gel synthesis, characterization and photocatalytic applications. Desalin. Water Treat. 2016, 57, 16196–16207. [Google Scholar] [CrossRef]
  36. Flak, D.; Yate, L.; Nowaczyk, G.; Jurga, S. Hybrid ZnPc@TiO2 nanostructures for targeted photodynamic therapy, bioimaging and doxorubicin delivery. Mater. Sci. Eng. C 2017, 78, 1072–1085. [Google Scholar] [CrossRef] [PubMed]
  37. Theivasanthi, T.; Alagar, M. Titanium dioxide (TiO2) Nanoparticles XRD Analyses: An Insight. arXiv 2013, arXiv:13071091. [Google Scholar]
  38. Duan, M.; Li, J.; Li, M.; Zhang, Z.; Wang, C. Pt(II) porphyrin modified TiO2 composites as photocatalysts for efficient 4-NP degradation. Appl. Surf. Sci. 2012, 258, 5499–5504. [Google Scholar] [CrossRef]
  39. Huang, Z.; Zheng, B.; Zhu, S.; Yao, Y.; Ye, Y.; Lu, W.; Chen, W. Photocatalytic activity of phthalocyanine-sensitized TiO2–SiO2 microparticles irradiated by visible light. Mater. Sci. Semicond. Process. 2014, 25, 148–152. [Google Scholar] [CrossRef]
  40. Mali, S.S.; Kim, H.; Kim, J.H.; Patil, P.S.; Kook Hong, C. Synthesis and characterization of planar heterojunction hybrid polymer solar cells based on copper pthalocyanine (CuPc) and titanium dioxide. Ceram. Int. 2014, 40, 643–649. [Google Scholar] [CrossRef]
  41. Soema, P.C.; Willems, G.-J.; Jiskoot, W.; Amorij, J.-P.; Kersten, G.F. Predicting the influence of liposomal lipid composition on liposome size, zeta potential and liposome-induced dendritic cell maturation using a design of experiments approach. Eur. J. Pharm. Biopharm. 2015, 94, 427–435. [Google Scholar] [CrossRef] [Green Version]
  42. Kozlov, N.K.; Natashina, U.A.; Tamarov, K.P.; Gongalsky, M.B.; Solovyev, V.V.; Kudryavtsev, A.A.; Sivakov, V.; Osminkina, L.A. Recycling of silicon: From industrial waste to biocompatible nanoparticles for nanomedicine. Mater. Res. Express 2017, 4, 95026. [Google Scholar] [CrossRef] [Green Version]
  43. World Health Organization. WHO Model List of Essential Medicines, 21st ed.; World Health Organization: Geneva, Switzerland, 2019. [Google Scholar]
  44. Kanabar, D.J. A clinical and safety review of paracetamol and ibuprofen in children. Inflammopharmacology 2017, 25, 1–9. [Google Scholar] [CrossRef] [Green Version]
  45. Kaufman, D.W.; Kelly, J.P.; Battista, D.R.; Malone, M.K.; Weinstein, R.B.; Shiffman, S. Exceeding the daily dosing limit of nonsteroidal anti-inflammatory drugs among ibuprofen users. Pharmacoepidemiol. Drug Saf. 2018, 27, 322–331. [Google Scholar] [CrossRef]
  46. Tixier, C.; Singer, H.P.; Oellers, S.; Müller, S.R. Occurrence and Fate of Carbamazepine, Clofibric Acid, Diclofenac, Ibuprofen, Ketoprofen and Naproxen in Surface Waters. Environ. Sci. Technol. 2003, 37, 1061–1068. [Google Scholar] [CrossRef] [PubMed]
  47. Miège, C.; Choubert, J.M.; Ribeiro, L.; Eusèbe, M.; Coquery, M. Fate of pharmaceuticals and personal care products in wastewater treatment plants–Conception of a database and first results. Environ. Pollut. 2009, 157, 1721–1726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Pereira, A.; Silva, L.; Laranjeiro, C.; Lino, C.; Pena, A. Selected Pharmaceuticals in Different Aquatic Compartments: Part I—Source, Fate and Occurrence. Molecules 2020, 25, 1026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Simazaki, D.; Kubota, R.; Suzuki, T.; Akiba, M.; Nishimura, T.; Kunikane, S. Occurrence of selected pharmaceuticals at drinking water purification plants in Japan and implications for human health. Water Res. 2015, 76, 187–200. [Google Scholar] [CrossRef] [PubMed]
  50. Peñas-Garzón, M.; Gómez-Avilés, A.; Belver, C.; Rodriguez, J.J.; Bedia, J. Degradation pathways of emerging contaminants using TiO2-activated carbon heterostructures in aqueous solution under simulated solar light. Chem. Eng. J. 2020, 392, 124867. [Google Scholar] [CrossRef]
  51. Achilleos, A.; Hapeshi, E.; Xekoukoulotakis, N.P.; Mantzavinos, D.; Fatta-Kassinos, D. UV-A and Solar Photodegradation of Ibuprofen and Carbamazepine Catalyzed by TiO2. Sep. Sci. Technol. 2010, 45, 1564–1570. [Google Scholar] [CrossRef]
  52. Da Silva, J.C.C.; Teodoro, J.A.R.; de Cássia Franco Afonso, R.J.; Aquino, S.F.; Augusti, R. Photolysis and photocatalysis of ibuprofen in aqueous medium: Characterization of by-products via liquid chromatography coupled to high-resolution mass spectrometry and assessment of their toxicities against Artemia Salina. J. Mass Spectrom. 2014, 49, 145–153. [Google Scholar] [CrossRef]
  53. Romeiro, A.; Azenha, M.E.; Canle, M.; Rodrigues, V.H.N.; Da Silva, J.P.; Burrows, H.D. Titanium Dioxide Nanoparticle Photocatalysed Degradation of Ibuprofen and Naproxen in Water: Competing Hydroxyl Radical Attack and Oxidative Decarboxylation by Semiconductor Holes. ChemistrySelect 2018, 3, 10915–10924. [Google Scholar] [CrossRef]
  54. Jiménez-Salcedo, M.; Monge, M.; Tena, M.T. Photocatalytic degradation of ibuprofen in water using TiO2/UV and g-C3N4/visible light: Study of intermediate degradation products by liquid chromatography coupled to high-resolution mass spectrometry. Chemosphere 2019, 215, 605–618. [Google Scholar] [CrossRef]
  55. Patterson, K.; Howlett, K.; Patterson, K.; Wang, B.; Jiang, L. Photodegradation of ibuprofen and four other pharmaceutical pollutants on natural pigments sensitized TiO2 nanoparticles. Water Environ. Res. 2020. [Google Scholar] [CrossRef]
  56. Méndez-Arriaga, F.; Esplugas, S.; Giménez, J. Photocatalytic degradation of non-steroidal anti-inflammatory drugs with TiO2 and simulated solar irradiation. Water Res. 2008, 42, 585–594. [Google Scholar] [CrossRef] [PubMed]
  57. Bhatia, V.; Dhir, A. Transition metal doped TiO2 mediated photocatalytic degradation of anti-inflammatory drug under solar irradiations. J. Environ. Chem. Eng. 2016, 4, 1267–1273. [Google Scholar] [CrossRef]
  58. Agustina, T.E.; Ang, H.M.; Vareek, V.K. A review of synergistic effect of photocatalysis and ozonation on wastewater treatment. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 264–273. [Google Scholar] [CrossRef]
  59. Ogunsipe, A.; Chen, J.-Y.; Nyokong, T. Photophysical and photochemical studies of zinc(II) phthalocyanine derivatives—Effects of substituents and solvents. New J. Chem. 2004, 28, 822–827. [Google Scholar] [CrossRef]
  60. Li, G.; Lv, L.; Fan, H.; Ma, J.; Li, Y.; Wan, Y.; Zhao, X.S. Effect of the agglomeration of TiO2 nanoparticles on their photocatalytic performance in the aqueous phase. J. Colloid Interface Sci. 2010, 348, 342–347. [Google Scholar] [CrossRef]
  61. Johnson, B.T. Microtox® Acute Toxicity Test. In Small-scale Freshwater Toxicity Investigations; Blaise, C., Férard, J.-F., Eds.; Springer: Berlin/Heidelberg, Germany, 2005; pp. 69–105. ISBN 978-1-4020-3119-9. [Google Scholar]
  62. Czech, B.; Jośko, I.; Oleszczuk, P. Ecotoxicological evaluation of selected pharmaceuticals to Vibrio fischeri and Daphnia magna before and after photooxidation process. Ecotoxicol. Environ. Saf. 2014, 104, 247–253. [Google Scholar] [CrossRef]
  63. Shemer, H.; Linden, K.G. Photolysis, oxidation and subsequent toxicity of a mixture of polycyclic aromatic hydrocarbons in natural waters. J. Photochem. Photobiol. Chem. 2007, 187, 186–195. [Google Scholar] [CrossRef] [Green Version]
  64. Lü, X.; Li, J.; Wang, C.; Duan, M.; Luo, Y.; Yao, G.; Wang, J.-L. Enhanced photoactivity of CuPp-TiO2 photocatalysts under visible light irradiation. Appl. Surf. Sci. 2010, 257, 795–801. [Google Scholar] [CrossRef]
  65. Recillas, S.; García, A.; González, E.; Casals, E.; Puntes, V.; Sánchez, A.; Font, X. Use of CeO2, TiO2 and Fe3O4 nanoparticles for the removal of lead from water. Desalination 2011, 277, 213–220. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of sources of emerging contaminants.
Figure 1. Schematic representation of sources of emerging contaminants.
Catalysts 10 01328 g001
Figure 2. Thermograms of (a) 1%CuPc@TiO2 (5 nm) (b) 1%CuPc@TiO2 (15 nm).
Figure 2. Thermograms of (a) 1%CuPc@TiO2 (5 nm) (b) 1%CuPc@TiO2 (15 nm).
Catalysts 10 01328 g002
Figure 3. X-ray powder diffraction (XRPD) patterns of the prepared materials.
Figure 3. X-ray powder diffraction (XRPD) patterns of the prepared materials.
Catalysts 10 01328 g003
Figure 4. X-ray photoelectron spectra obtained for the materials: (a) TiO2 5 nm; (b) 1%CuPc@TiO2 5 nm; (c) 1%ZnPc@TiO2 5 nm; (d) TiO2 15 nm; (e) 1%CuPc@TiO2 15 nm; (f) 1%ZnPc@TiO2 15 nm.
Figure 4. X-ray photoelectron spectra obtained for the materials: (a) TiO2 5 nm; (b) 1%CuPc@TiO2 5 nm; (c) 1%ZnPc@TiO2 5 nm; (d) TiO2 15 nm; (e) 1%CuPc@TiO2 15 nm; (f) 1%ZnPc@TiO2 15 nm.
Catalysts 10 01328 g004
Figure 5. Particle size distribution patterns for (a) TiO2 (5 nm) and (b) TiO2 (15 nm).
Figure 5. Particle size distribution patterns for (a) TiO2 (5 nm) and (b) TiO2 (15 nm).
Catalysts 10 01328 g005
Figure 6. Exemplary electron spin resonance (ESR) spectra: (a) for 1%CuPc@TiO2 (15 nm) and (b) for ZnPc.
Figure 6. Exemplary electron spin resonance (ESR) spectra: (a) for 1%CuPc@TiO2 (15 nm) and (b) for ZnPc.
Catalysts 10 01328 g006
Figure 7. The changes in ibuprofen concentration after irradiation (365 nm light) of the solution containing a photocatalyst.
Figure 7. The changes in ibuprofen concentration after irradiation (365 nm light) of the solution containing a photocatalyst.
Catalysts 10 01328 g007
Figure 8. The changes in ibuprofen concentration after irradiation (665 nm light) of the solution containing a photocatalyst.
Figure 8. The changes in ibuprofen concentration after irradiation (665 nm light) of the solution containing a photocatalyst.
Catalysts 10 01328 g008
Figure 9. Comparison of ibuprofen (IBU) decomposition rates under UV irradiation with different photocatalysts.
Figure 9. Comparison of ibuprofen (IBU) decomposition rates under UV irradiation with different photocatalysts.
Catalysts 10 01328 g009
Figure 10. Comparison of degradation rate constants (k, s−1) of IBU under UV irradiation with different photocatalysts.
Figure 10. Comparison of degradation rate constants (k, s−1) of IBU under UV irradiation with different photocatalysts.
Catalysts 10 01328 g010
Figure 11. Cell viability decrease of A. fischeri cells upon addition of ibuprofen solutions before (0 h) and after (6 h) photodegradation experiment with different photocatalysts, when irradiated with (a) UV light (λmax = 365 nm), (b) red light (λmax = 665 nm).
Figure 11. Cell viability decrease of A. fischeri cells upon addition of ibuprofen solutions before (0 h) and after (6 h) photodegradation experiment with different photocatalysts, when irradiated with (a) UV light (λmax = 365 nm), (b) red light (λmax = 665 nm).
Catalysts 10 01328 g011
Figure 12. Setup of the photodegradation experiment (a) as seen from the front and (b) as seen from the above. 1. Beaker containing 100 mg of Pc@TiO2 and 100 mL of IBU solution; 2. magnetic stirrer; 3. UV or Vis lamp; 4. aluminum foil.
Figure 12. Setup of the photodegradation experiment (a) as seen from the front and (b) as seen from the above. 1. Beaker containing 100 mg of Pc@TiO2 and 100 mL of IBU solution; 2. magnetic stirrer; 3. UV or Vis lamp; 4. aluminum foil.
Catalysts 10 01328 g012
Table 1. Hybrid TiO2-Pc materials applied in the study.
Table 1. Hybrid TiO2-Pc materials applied in the study.
TiO2 Nanoparticle SizePc, Pc : TiO2 Ratio (m/m)Symbol of the Material
5 nmCuPc, 1 : 1001%CuPc@TiO2(5 nm)
5 nmZnPc, 1 : 1001%ZnPc@TiO2(5 nm)
15 nmCuPc, 1 : 1001%CuPc@TiO2(15 nm)
15 nmZnPc, 1 : 1001%ZnPc@TiO2(15 nm)
Table 2. Particle size measured for TiO2 based materials.
Table 2. Particle size measured for TiO2 based materials.
MaterialParticle SizePolydispersity Index a
TiO2 (5 nm)145.7 ± 68.4 nm0.22
1%CuPc@TiO2 (5 nm)103.5 ± 87.8 nm0.72
1%ZnPc@TiO2 (5 nm)158.7 ± 74.4 nm0.22
TiO2 (15 nm)201.6 ± 139.8 nm0.48
1%CuPc@TiO2 (15 nm)144.5 ± 129.1 nm0.80
1%ZnPc@TiO2 (15 nm)236.2 ± 128.1 nm0.29
a–calculated according to the formula PDI = (SD/mean diameter)2 [42].
Table 3. Brunauer–Emmett–Teller (BET) surface area measured for the materials prepared.
Table 3. Brunauer–Emmett–Teller (BET) surface area measured for the materials prepared.
MaterialBET Surface Area [m2/g]
TiO2 (5 nm)264.8 ± 1.2
1%CuPc@TiO2 (5 nm)249.7 ± 1.5
1%ZnPc@TiO2 (5 nm)244.3 ± 1.5
TiO2 (15 nm)104.6 ± 0.3
1%CuPc@TiO2 (15 nm)96.8 ± 0.3
1%ZnPc@TiO2 (15 nm)100.4 ± 0.3
Table 4. The regression and kinetic parameters of IBU degradation rate constants (k, s−1) under UV irradiation with 1%ZnPc@TiO2 (5 nm), 1%ZnPc@TiO2 (15 nm), 1%CuPc@TiO2 (5 nm) 1%CuPc@TiO2 (15 nm), TiO2 (5 nm) and TiO2 (15 nm) as a photocatalysts.
Table 4. The regression and kinetic parameters of IBU degradation rate constants (k, s−1) under UV irradiation with 1%ZnPc@TiO2 (5 nm), 1%ZnPc@TiO2 (15 nm), 1%CuPc@TiO2 (5 nm) 1%CuPc@TiO2 (15 nm), TiO2 (5 nm) and TiO2 (15 nm) as a photocatalysts.
SamplesRegression Parameters aKinetic Parameters [k (s−1)] b
1%ZnPc@TiO2(5 nm)a−0.1214k2.024 × 10−3s−1
Δα0.0303Δκ5.051 × 10−4s−1
b = y(0)1 504 745.2
Δβ1.098
sa0.0117
sb0.0364
sy0.0611
r−0.982
1%ZnPc@TiO2(15 nm)a−0.1548k2.580 × 10−3s−1
Δα0.0136Δκ2.279 × 10−4s−1
b = y(0)912 073.1
Δβ1.043
sa0.00531
sb0.0164
sy0.02757
r−0.998
1%CuPc@TiO2(5 nm) a−0.3832k6.387 × 10−3s−1
Δα0.0601Δκ1.002 × 10−3s−1
b = y(0)1 160 993.9
Δβ1.204
sa0.0233
sb0.0722
sy0.121
r−0.993
1%CuPc@TiO2(15 nm) a−0.3649k6.083 × 10−3s−1
Δα0.0548Δκ9.149 × 10−4s−1
b =y(0)1 242 276.5
Δβ1.184
sa0.0213
sb0.0659
sy0.11
r−0.993
TiO2(5 nm)a−0.391k6.517 × 10−3s−1
Δα0.0292Δκ4.876 × 10−4s−1
b = y(0)1 158 957.5
Δβ1.0945
sa0.011381
sb0.0351
sy0.0589
r−0.998
TiO2(15 nm)a−0.359k5.988 × 10−3s−1
Δα0.0252Δκ4.202 × 10−4s−1
b = y(0)1 111 784.1
Δβ1.08
sa0.0098
sb0.0302
sy0.0508
r−0.999
a The parameters: a, sa, b, sb and r were calculated by the least square regression and stand for slope, the standard deviation of the slope, intercept, the standard deviation of intercept and correlation coefficient, respectively. b The kinetic parameters k [s−1] is the photodegradation reaction of IBU constant rate.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Krakowiak, R.; Musial, J.; Frankowski, R.; Spychala, M.; Mielcarek, J.; Dobosz, B.; Krzyminiewski, R.; Sikorski, M.; Bendzinska-Berus, W.; Tykarska, E.; et al. Phthalocyanine-Grafted Titania Nanoparticles for Photodegradation of Ibuprofen. Catalysts 2020, 10, 1328. https://doi.org/10.3390/catal10111328

AMA Style

Krakowiak R, Musial J, Frankowski R, Spychala M, Mielcarek J, Dobosz B, Krzyminiewski R, Sikorski M, Bendzinska-Berus W, Tykarska E, et al. Phthalocyanine-Grafted Titania Nanoparticles for Photodegradation of Ibuprofen. Catalysts. 2020; 10(11):1328. https://doi.org/10.3390/catal10111328

Chicago/Turabian Style

Krakowiak, Rafal, Joanna Musial, Robert Frankowski, Marcin Spychala, Jadwiga Mielcarek, Bernadeta Dobosz, Ryszard Krzyminiewski, Marek Sikorski, Wioletta Bendzinska-Berus, Ewa Tykarska, and et al. 2020. "Phthalocyanine-Grafted Titania Nanoparticles for Photodegradation of Ibuprofen" Catalysts 10, no. 11: 1328. https://doi.org/10.3390/catal10111328

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