Next Article in Journal
Endogenous and Exogenous Stimuli-Responsive Drug Delivery Systems for Programmed Site-Specific Release
Next Article in Special Issue
Self-Assembly of Covalently Linked Porphyrin Dimers at the Solid–Liquid Interface
Previous Article in Journal
Chemical Reactivity Theory and Empirical Bioactivity Scores as Computational Peptidology Alternative Tools for the Study of Two Anticancer Peptides of Marine Origin
Previous Article in Special Issue
Metalloporphyrin Dimers Bridged by a Peptoid Helix: Host-Guest Interaction and Chiral Recognition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Zinc Porphyrin-Functionalized Fullerenes for the Sensitization of Titania as a Visible-Light Active Photocatalyst: River Waters and Wastewaters Remediation

by
Elzbieta Regulska
1,*,
Danisha Maria Rivera-Nazario
2,
Joanna Karpinska
1,
Marta Eliza Plonska-Brzezinska
3,* and
Luis Echegoyen
2,*
1
Institute of Chemistry, University of Bialystok, Ciolkowskiego 1K, 15-245 Bialystok, Poland
2
Department of Chemistry, University of Texas at El Paso, 500 W. University Ave., El Paso, TX 79968, USA
3
Department of Organic Chemistry, Faculty of Pharmacy with the Division of Laboratory Medicine, Medical University of Bialystok, Mickiewicza 2A, 15-222 Bialystok, Poland
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(6), 1118; https://doi.org/10.3390/molecules24061118
Submission received: 3 February 2019 / Revised: 17 March 2019 / Accepted: 19 March 2019 / Published: 21 March 2019
(This article belongs to the Special Issue Spatial Organization of Multi-Porphyrins for Pre-Defined Properties)

Abstract

:
Zinc porphyrin-functionalized fullerene [C60] derivatives have been synthesized and used to prepare titania-based composites. The electrochemical properties and HOMO and LUMO levels of the photosensitizers were determined by electrochemical measurements. Raman and IR techniques were used to study chemical groups present on the titania surface. Absorption properties of the composites were measured in the solid state by diffuse reflectance UV-Vis spectra (DRS). The zeta potential and aggregate sizes were determined using dynamic light scattering (DLS) and electrophoretic light scattering (ELS) techniques. Surface areas were estimated based on Brunauer–Emmett–Teller (BET) isotherms. The photocatalytic activity of the photocatalysts was tested using two model pollutants, phenol and methylene blue. The composite with the highest photocatalytic potential (1/TiO2) was used for river and wastewater remediation. The photodegradation intermediates were identified by LC-UV/Vis-MS/MS techniques.

Graphical Abstract

1. Introduction

Since Fujishima and Honda [1] first demonstrated water splitting on titania (TiO2) electrodes, TiO2 has attracted considerable interest in the field of heterogeneous photocatalysis. The extensive usage of this semiconductor derives from its convenient properties like stability, chemical and biological inertness, no toxicity, low price, and environmental friendliness. Nevertheless, there are some significant drawbacks of titania, which preclude it from being used as an efficient photocatalyst. One limitation is the reasonably large energy band-gap of 3.2 eV, which requires ultraviolet light for its excitation, and another is the high probability for recombination of the photogenerated holes and electrons.
For that reason, the photosensitization of titania with different chromophores has been reported [2,3,4,5]. Artificial photosynthetic systems employing porphyrin-linked fullerene derivatives have been thoroughly examined in organic photovoltaic cells (OPVs). Porphyrins [6,7,8,9,10] and fullerenes [11,12,13,14,15,16,17,18,19] have also been independently employed for the preparation of modified titania. Nevertheless, to our knowledge, there are no reports involving covalently linked porphyrin–fullerene dyads as photosensitizers for titania photocatalysis.
Heterogeneous photocatalysis is considered the most effective among the Advanced Oxidation Processes (AOPs), and has high potential for polluted water treatment. The development of effective treatment technologies for surfaces as well as for wastewaters is especially important in view of the increasing growth of water pollution by chemicals arising from many industrial, agricultural, and urban human activities. Photocatalytic treatment is considered the most promising technique due to its high efficiency, and is expected to replace other AOPs currently used for that purpose.
In this article, we report the synthesis of three novel zinc porphyrin-functionalized fullerenes (see Figure 1), which were used to prepare titania-based composites and applied for the photocatalytic degradation of phenol and methylene blue as model pollutants. The photocatalyst exhibiting the highest activity was used for the remediation of selected river waters and water effluents.

2. Results and Discussion

2.1. Photoelectronic Properties of the Fullerene Derivatives

The electrochemical characterization of the synthesized fullerene derivatives was performed to examine the relation between the energies of the LUMOs of the chromophores and the conduction band of titania, and to obtain the energy band gap of each compound. The CV cycles of the compounds are presented in Figure 2.
Two waves due to the oxidation of the porphyrin macrocycle were observed in the anodic scan associated with the formation of P•+ and P•2+ [20]. Five, three, and four reduction waves were observed for the cathodic scan for 1, 2 and 3, respectively. They resulted from the reduction of porphyrin and fullerene, namely, two-electron reduction of the aromatic porphyrin ring overlapping with the reduction of the fullerene moiety. In terms of porphyrin, only the reduction of the ring is reflected on the CV curve, since the zinc(II) ion is electrochemically inactive [21]. Based on the onsets of the first oxidation and reduction peaks, the HOMO and LUMO levels of the compounds were estimated according to the following equations:
ELUMO = −(4.71 − E1/2, Fc,Fc+ + Ered, onset)
EHOMO = − (4.71 − E1/2, Fc,Fc+ + Eox, onset)
The calculated HOMO and LUMO energy values of the fullerene derivatives are presented in Figure 3. The LUMOs of the photosensitizers are all above the energy of the conduction band (CB) of TiO2, and the differences are 0.62, 0.59, and 0.51 eV for 1, 2, and 3, respectively. Therefore, efficient photosensitization of TiO2 by the fullerene derivatives should be observed.

2.2. Characterization of the Titania

Raman and IR spectra of the synthesized titania are shown in Figure 4. The Raman spectrum shows broad bands at the following Raman shifts: 100–260, 380–480, 500–650, and 800–900 cm−1, which were assigned to Ti–O bending, Ti–O stretching, and Ti–O–Ti stretching vibrations, respectively [22]. A band at 1020 cm−1 was assigned to the C–O group arising from the titanium precursor [22]. Typical bands arising from crystalline titania were not observed. Thus, the TiO2 prepared by the sol–gel synthesis was obtained in an amorphous phase. The IR spectrum (Figure 4) shows the presence of the hydroxyl groups on the oxide surface. The broad band observed in the range of 2500–3500 cm−1 results from O–H stretching vibrations [23]. Additionally, a band around 1710 cm−1 was assigned to the H–OH and Ti–OH bending vibrations [8,24], while the signals coming from the O–H in-plane and out-of-plane bending vibrations were observed at 410 cm−1 and in the range of 700–800 cm−1, respectively. The presence of the hydroxyl groups on the catalyst surface is crucial for its photocatalytic performance. Therefore, the prepared oxides, despite their amorphous structure, were expected to exhibit photocatalytic activity.

2.3. Characterization of the Composites

The morphologies of the composites were examined by SEM microscopy and compared with the respective components, i.e., titania and fullerene derivatives 1, 2, and 3 (see Figure 5). The pristine titania presented a porous structure (Figure 5-1), while SEM images of fullerene derivatives showed their smooth surfaces (Figure 5–2b,3b,4b), as we have previously reported for fullerenes [25,26]. The composites TiO2/1, TiO2/2, and TiO2/3 were found to form irregular bricks covered with the porous titania (Figure 5–2a,3a,4a). The introduction of the smooth fullerenes affected the specific surface areas of the composites, since the BET areas of the composites were smaller than that of the pristine titania (see Table 1).
The absorption spectra of the composites were compared with the spectrum of the pristine titania (Figure 6). The Soret and Q bands from the porphyrin macrocycles were observed for TiO2/1, TiO2/2, and TiO2/3. They appeared within the following wavelength ranges: 390–480 and 530–680 nm for Soret and Q bands, respectively. The fullerene absorption band was not observed due to overlap with the strong absorption of the titania in the range of 200–380 nm. The highest intensity of the Soret band was observed for TiO2/3, with a maximum at 420 nm. Moreover, a red-shift of the Q band was noticed for TiO2/2 and TiO2/3 when compared with the spectrum of TiO2/1. Therefore, TiO2/3 exhibits the most efficient absorption in the visible spectral range.
The specific surface areas of the titania and the corresponding composites were determined from BET isotherms. The results (Table 1) for the composites were in the range of 151–178 m2∙g−1, while the surface area of the pristine semiconductor was 269 m2∙g−1. Thus, the introduction of the fullerene derivatives in the composites resulted in decreases of the BET surface areas, as expected.
The aggregate sizes of the TiO2 and the composites were determined by dynamic light scattering (DLS) measurements (see Table 1). The sizes of the catalyst aggregates may influence the kinetics of the photocatalytic degradation. It was previously reported that upon increasing the titania diameter from 0.35 to 0.66 μm, the photocatalytic activity of the semiconductor decreased by around 4.6 times [27]. In the present study, it was found that the diameters of the composite aggregates were 3 to 7 times lower when compared to that of the pristine titania.
Additionally, electrophoretic light scattering (ELS) studies were performed to examine the dependence of the electrophoretic potential and the pH, in order to establish the range within which the catalysts form stable suspensions. The electric charge of the oxides in the solid state results from the presence of hydroxyl groups on their surface. Their amphoteric character results from their ability to react with both hydrogen and hydroxyl groups [27]. Bahnemann et al. [28] proposed the following notation for the chemisorption processes occurring on the surface of titania:
Ti O 2 + n H + Ti O 2 H n n +
Ti O 2 + n O H Ti O 2 ( O H ) n n
The high repulsion force between particles of the same charge decreases the probability of aggregation, thus increasing the stability of the sols in acidic and basic environments. The relation between the zeta potentials and the pH for TiO2 and TiO2/3, is shown in Figure 7. In both cases, a similar dependence was observed. Zero-point potentials appeared at pH levels equal to 4.9 and 5.6 for TiO2 and TiO2/3, respectively. For TiO2, in the pH range of 1.0–4.0, a 20 mV increase of the potential was observed. The decrease of the zeta potential continued until the potential reached −45 mV at pH 9.0. Stable suspensions with TiO2 are expected to be formed only between pH values of 8 and 12 pH, while within that range, absolute values of zeta potentials exceeded 30 mV. For TiO2/3, the zeta potential decreased with the increasing pH over the complete pH range. The absolute value of the zeta potential was higher than 30 mV within the pH range of 7–12. Therefore, under these conditions, the highest stability expected is for the TiO2/3 aggregate. The composite particles may tend to aggregate and not be able to form stable suspensions within the pH range of 3–7.

2.4. Photocatalytic Performance

The photocatalytic performance of the composites was tested using two model pollutants, phenol (PhOH) and methylene blue (MB). The decreases of their relative concentrations and the efficiencies (η) of their photocatalytic degradation in the presence of the catalysts are presented in Figure 8.
The highest η during the degradation of PhOH was achieved using TiO2/3 (54%). However, similar values were also obtained with TiO2/1 (50%) and TiO2/2 (53%). Introduction of 3 as a photosensitizer enhanced the photocatalytic activity of TiO2 by nearly three times (2.7). When MB was exposed to the solar-simulated light in the presence of the composites, the degradation obtained was between 44% and 46%. Again, the highest η was observed for TiO2/3, at 2.5 times compared with pristine TiO2. Slightly lower η values were observed when MB was exposed to the radiation in the presence of the photocatalysts. This was attributed to the competitive light absorption between the catalyst and the model compound. The highest photocatalytic activity was shown by TiO2/3 when both model pollutants were used. The proposed mechanism explaining the enhancement of the photocatalytic activity through the formation of the porphyrin-functionalized fullerene/titania composites versus pristine titania is presented in Figure 9. The improved performance of the synthesized catalysts derives from the utilization of the solar light for the excitation of the fullerene derivatives. We attribute the observed highest activity of TiO2/3 to both the smallest difference between the LUMO of 3 and the CB of TiO2, and to the smallest energy band-gap of 3 among the photosensitizers.

2.5. UV–Vis Absorbance Spectra of the Photodegradation Products

The photodegradation products of PhOH and MB during their exposure to solar light in the presence of TiO2/3 were examined using the LC-UV–Vis-MS technique. Figure 10 depicts a chromatogram, registered after 2 h of PhOH irradiation, that shows three peaks. Based on the UV–Vis absorption and mass spectra registered at 1.71, 2.56, and 4.05 min, the main photodegradation products were found. They were identified as 4,4’-dihydroxybiphenyl, benzoquinone, and maleic anhydride. It was also shown that under the applied conditions, complete phenol decomposition was not accomplished. Additional intermediates were identified by mass spectrometry in minor amounts (see Supporting Information).
Contrary to PhOH, MB was completely degraded under the conditions used. In Figure 11, chromatograms of MB before and after its irradiation in the presence of TiO2/3 are presented, along with the absorption spectra (Figure 11a–c) corresponding to the retention times at which pristine MB and its photodegradation intermediates were observed. The peak arising from MB did not appear in the chromatogram obtained after 2 h of photocatalytic treatment. However, two new peaks at 6.20 and 6.34 min were observed. Based on their absorption spectra (Figure 11b,c), it was found that the MB photodegradation intermediates absorb visible light. Even though MB was not completely decolorized, it is likely that optimal adjustment of irradiation times and higher light intensity can accomplish this in the future. A list of the identified MB photodegradation products is included in the Supporting Information.

2.6. Photocatalytic Performance Using Natural Samples

In order to understand the complex nature of environmental matrices, the contribution of different ions, as well as of organic compounds, needs to be considered. A series of natural matrices, including three river waters, wastewater effluent, and municipal wastewater, were used in our studies. The efficiencies of PhOH and MB photocatalytic degradation in the presence of TiO2 or TiO2/3 are presented in Figure 12. Photocatalytic degradation conducted with the natural matrix was found to be significantly slowed down compared with the one performed in the MilliQ water. This behavior is due to the complex composition of natural samples, which leads to lower light absorption efficiencies due to competition for access to light between the catalyst and the natural water ingredients. In addition, the radicals formed during the photocatalytic reactions may react with a wide variety of substances present in the matrix, rather than resulting in the direct decomposition of the model pollutant. In this regard, the following parameters are particularly important: the oxygen content, the presence of transition metal cations, the concentration of ions responsible for water hardness, dissolved organic carbon, and the color of the sample [29]. A series of parameters was determined for both the examined river waters and wastewaters, which are presented in Table 2 and Table 3, respectively. The estimated values enabled the assignment of the examined river waters to the second grade of quality [30], and for the wastewater effluents to meet the expected criteria for effluents.
Different river waters affected the decomposition of the model pollutants differently. When PhOH was subjected to photocatalytic degradation with TiO2/3, the smallest decrease of η was observed in Goldapa river water (35%) when compared to the η obtained in MilliQ water (54%). However, MB decomposition was found to decrease the least in Suprasl river water (from 46% to 37%). We found that the series of parameters, including concentration of chlorine, calcium, magnesium, and nitrate ions, along with hardness and conductivity, were found to be the highest in Suprasl river waters. However, some of the species exhibit both positive and negative effects on photodegradation.
The total hardness of the water is a factor which negatively influences the performance of the photocatalytic processes. This parameter is related to the content of carbonate and bicarbonate ions, which were estimated to be the highest for the Suprasl river. These anions react with the hydroxyl radicals to generate other less reactive radical species according to the following chemical equations:
CO32− + HO → CO3•− + HO•−
HCO32− + HO → CO3•− + H2O
It is generally believed that metal cations which are present in only one oxidation state do not affect the photocatalytic decomposition either positively or negatively. However, it was shown that the process of photocatalytic degradation can be accelerated in the presence of calcium ions. Mariquit et al. [32] showed that calcium cations support the adsorption of humic acid, thus facilitating photocatalytic oxidation. Furthermore, Yiang et al. [33], while examining typical organic pollutant 1-methylimidazole-2-thiol, indicated a faster photocatalytic degradation of the compound after the introduction of calcium ions. This parameter was the highest in the Suprasl river as well.
On the other hand, oxygen participates in the processes described below, being a source of the oxygen radicals. It can be adsorbed on the catalyst’s surface and can undergo reactions with the participation of the conduction band of TiO2 [33]:
e + O2 → O2 (ad)
O2 (ad) + H+ → HO2
The valence band of titania delivers other active forms of oxygen, like HO and O, which results from the presented equations:
h+ + H2O → HO (ad) + H+
h+ + O2 (ad) → 2 O (ad)
All of the oxygen forms, like O2, HO, HO2, and O, likely participate in the reactions leading to the degradation of the organic matter [34]. Dissolved oxygen was found to be the highest in the Sapina river, but this was not the matrix in which photodegradation was the most efficient.
The yield of the photocatalytic degradation in the presence of a river water sample may also be higher than in deionized water. Both rivers and wastewaters contain chromophores, which may also be involved in photochemical reactions leading to the formation of reactive compounds, which can initiate the decomposition of the model pollutants [35]. This was observed when MB solutions prepared in both rivers and wastewaters were irradiated in the presence of TiO2 (Figure 12b). It is believed that the presence of colored dissolved organic matter (CDOM) is responsible for the formation of singlet oxygen 1O2 and triplet excited states 3CDOM*. These forms may be even more important than the hydroxyl radical HO in the photodegradation processes. CDOM exposed to sunlight undergoes the following reactions [36]:
CDOM + hv → 1CDOM*
1CDOM*3CDOM*
3CDOM* → CDOM
3CDOM* + O2 → CDOM + O2
3CDOM* + O2 → CDOM + 1O2
3CDOM* + P → CDOM•− + P•+
CDOM•− + O2 → CDOM + O2•+
1O2 → O2
1O2 + P → Pox
where P is phenol or its derivatives.
Photocatalytic degradations in environmental matrices are very complex processes. It is really difficult, if not impossible, to predict the influence of all of the ingredients of a natural matrix on the photodegradation yields. In the presented studies, the same photodegradation products of both model compounds were identified using MilliQ water and all of the environmental matrices as reaction mediums (Supporting Information).

3. Experimental

3.1. Methods

1H-NMR spectra were recorded on a JEOL ECA 600 NMR spectrometer (JEOL Ltd., Peabody, MA, USA) at room temperature using CDCl3 or CDCl3:CS2 as solvents. Mass spectra were obtained using a Bruker microFlex MALDI-TOF LRF spectrometer (Billerica, MA, USA) on reflector positive mode, using 1,8,9-trihydroxyanthracene or trans-2-[3-(4-tert-Butylphenyl)-2-methyl-2-propenylidene]malononitrile as the matrix.
The Attenuated Total Reflection-Fourier Transform Infra-Red (ATR-FTIR) (Thermo Scientific, Waltham, MA, USA) spectra (500–3200 cm−1) were obtained using a Nicolet Model 6700 FT-IR spectrometer with a DTGS detector. The crystal-diamond spectra were obtained with 4 cm−1 resolution, and 32 scans for each sample spectrum were obtained. Diffuse reflectance UV–Vis spectra (DRS) were recorded on a Jasco V-30 UV-Vis/NIR spectrophotometer equipped with an integrating sphere 60 mm in diameter, using BaSO4 as a reference.
Raman spectra were collected using the Renishaw Raman InVia (Renishaw, New Mills, UK) Microscope equipped with a high sensitivity ultra-low noise CCD detector. The radiation from a high-power diode laser (Renishaw, New Mills, UK) (785 nm) was used as the excitation source. The laser power and accumulation of scans were dependent upon the sample. Spectra were recorded using ×50 objective with 10 s exposure time. The instrument was calibrated using an internal silicon standard (521 cm−1).
The sizes of the composite aggregates and zeta potentials were assessed using a NanoPlus-3 zeta/nano particle analyzer (Micromeritics Instrument Corporation, Norcross, GA, USA) for aqueous solutions. The sizes of the aggregates was determined based on dynamic light scattering (DLS) measurements for solutions of pH 7, while the zeta potentials were determined based on electrophoretic light scattering (ELS) measurements for solutions within pH range 1–14. The specific surface areas were measured using low-temperature nitrogen adsorption (80 K) with an ASAP 2020 (Micromeritics Instrument Corporation, Norcross, GA, USA) instrument. Prior to the measurements, samples were out-gassed under vacuum at 393 K for 12 h. The specific surface areas were determined using the Brunauer–Emmett–Teller (BET) method.
Electrochemical measurements were performed in a mixture of MeCN:Tol = 1:4 (v/v). Tetrabutylammonium hexafluorophosphate (TBAPF6) was added as the supporting electrolyte after recrystallization from ethanol. Cyclic voltammetry (CV) experiments were performed under an argon atmosphere at room temperature, using a CH Instrument potentiostat (CHI-440B) (Metrohm AG, Herisau, Switzerland) at a scan rate of 100 mV∙s−1. A standard three-electrode set up, consisting of a glassy carbon working electrode (ALS Co., Ltd., 1 mm ID), a platinum wire (1.0 mm, Aldrich) as a counter electrode, and a silver wire (1.0 mm, Aldrich) as a pseudoreference electrode, was used. The redox couple of ferrocene/ferrocenium (Fc/Fc+) was used as an internal reference to measure the potentials.
The photocatalytic degradation experiments were investigated in a 50 mL glass cell. The reaction mixture consisted of 20 mL of the model pollutant (phenol (PhOH, 10−4 mol·L−1) or methylene blue (MB, 8 × 10−5 mol·L−1), aqueous solution, and a photocatalyst (1.5 g·L−1). Before the PhOH/MB photodegradation studies, adsorption experiments were carried out. The suspension of the model pollutant and the appropriate photocatalyst was kept in the dark while stirring for 1 h to attain equilibrium. To monitor the progress of the degradation during the irradiation (2 h), samples were taken and centrifuged every 15 min, and the supernatant was analyzed using UV–Vis spectroscopy. The UV–Vis spectra were recorded with a HITACHI U-2800A UV–Vis spectrophotometer equipped with a double monochromator and a single beam optical system (190–700 nm). A SUNTEST CPS+ (ATLAS, USA) solar simulator apparatus, emitting irradiation in the range of 380–700 nm, was used to perform photocatalytic degradation experiments. All experiments were repeated three times, and the presented results reflect the averaged data. The photon flux of solar-simulated radiation was equal to 2.32 × 10−6 Einstein s−1 for 250 W m−2. Identification of the model pollutant photodegradation products was performed using a Shimadzu Nexera ultra high-performance liquid chromatograph with a photodiode-array detector (Shimadzu Corporation, Kyoto, Japan) and a triple quadrupole mass spectrometer (Shimadzu LCMS-8040) (Shimadzu Corporation, Kyoto, Japan) equipped with an electrospray ionization (ESI) interface.

3.2. Synthesis of Fullerene Derivatives

Fullerene derivatives 1, 2, and 3 were synthesized via a Bingel–Hirsch reaction. General procedure: 1 eqv. of the porphyrin malonate 8 (see Supplementary Materials) was reacted with 1.2 eqv. of the fullerene C60 derivative (9, 13, or 14; see Supporting Information) in the presence of CBr4 (1.1 eqv.) and upon addition of 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) (1.2 eqv.). The reaction was completed within 3 h. Compounds 1, 2, and 3 were isolated through the column chromatography as a mixture of isomers. For detailed procedures, see Supporting Information.

3.3. Synthesis of Titania

2 mL of titanium isopropoxide (TIPT) was dissolved in 5 mL of anhydrous ethanol with 0.5 mL of glacial acetic acid. After 30 min of stirring, the prepared solution was added dropwise to 5 mL of 5% sodium dodecyl sulfate (SDS) solution. The obtained mixture was left for 72 h under stirring. The prepared gel was washed 3 times with ethanol and water, whereupon it was evaporated under reduced pressure. The obtained powder was dried by placing in the oven (120 °C) for 24 h.

3.4. Synthesis of Composites

0.018 g of 1, 2, or 3 were placed in a round-bottom flask with 5 mL of CHCl3 and 5 mL of 5% aqueous SDS solution. The obtained mixture was sonicated for 12 h (solution A). Independently, an ethanolic (20 mL) solution of TIPT (666 μL) and glacial acetic acid (1 mL) (solution B) was prepared. Solution B was left under stirring for 30 min, whereupon it was added dropwise to solution A. The mixture of A and B was placed in a round-bottom flask and stirred for 2 h. The obtained gel was washed 3 times with ethanol and water, whereupon it was evaporated under reduced pressure. The obtained powder was dried by placing in the oven (120 °C) for 24 h.

4. Conclusions

Three novel fullerene derivatives covalently linked to a zinc porphyrin were synthesized and applied for the sensitization of titania. This paper represents the first report of the application of porphyrin-functionalized fullerene/titania composites as photocatalysts. It was found that the LUMO levels of the modified fullerenes are appropriate for proper electron injection into the conduction band of titania. Compound 3, which exhibits the smallest energy difference with respect to the CB of TiO2, was the most active catalyst. The photocatalytic activities of the composites were tested against MB and PhOH in aqueous solutions as well as in a series of environmental matrices. The photocatalytic activities of all of the composites were comparable. Nevertheless, the highest photodegradation efficiency was observed for TiO2/3. It was shown that it presents photocatalytic activity in all of the examined river waters, along with municipal wastewater and wastewater effluents. It was observed that it can form stable suspensions in solutions over a pH range of 7–12. Therefore, it can be used for both river and wastewater remediation, since rivers and wastewaters show neutral and basic pH levels, respectively. It should also be stressed that both colorless and colored substances, like PhOH and MB, undergo degradation in the presence of the composites. Therefore, it can be assumed that they could be used not only for remediation of rivers and municipal wastewaters, but also of industrial wastewaters. In particular, the textile industry releases significant amounts of dyes, which are difficult to remediate. Accordingly, we believe that the results presented here show potential applications for the decontamination of a series of complex environmental and industrial matrices, using solar light as an irradiation source.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/24/6/1118/s1, Figure S1: Synthetic route to porphyrin (8): (i) MeOH, HCl, 65 °C, (ii) pyrrole, 4-tert-butyl benzaldehyde, TFA, DDQ, CHCl3, (iii) Zn(OA)2, CHCl3, (iv) LiALH4, THF, (v) TEA, ethyl malonyl chloride, THF, CH2Cl2; Figure S2: Synthetic route to fullerene (1): (i) (4), sarcosine, o-DCB, 180 °C, (ii) (8), DBU, CH2Cl2, Tol; Figure S3: Synthetic route to fullerene (2): (i) TEA, CH2Cl2, (ii) C60, CBr4, DBU, Tol, CH2Cl2, (iii) (8), DBU, CH2Cl2, Tol; Figure S4: Synthetic route to fullerene (3): (i) HCl, MeOH, 65 °C, (ii) TsNHNH2, MeOH, 65 °C, (iii) C60, NaOMe, pyridine, o-DCB, 180 °C, (iv) (8), DBU, CH2Cl2, Tol; Figure S5: PXRD images of the as-synthesized titania (a) and the annealed at 450 °C (b); Table S1: List of identified compounds obtained in the photocatalytic degradation of PhOH; Table S2: List of identified compounds obtained in the photocatalytic degradation of MB.

Author Contributions

Investigation, E.R. and D.M.R.-N.; Writing original draft, E.R.; Editing, M.E.P.-B., L.E. and J.K.; Supervision, M.E.P.-B. and L.E.

Funding

E.R. thanks the National Science Centre, Poland (projects: 2012/05/N/ST5/01479 and 2014/12/T/ST5/00118). L.E. thanks the Robert A. Welch Foundation for an endowed chair, grant #AH-0033, and the US NSF, grants: PREM program (DMR-1205302) and CHE-1801317. LC–MS, porosimeter, IR and Raman spectrometers, UV–Vis/NIR spectrophotometer, and zeta/nano particle analyzer were funded by EU, as part of the Operational Programme Development of Eastern Poland, projects Nr POPW.01.03.00-20-034/09-00 and POPW.01.03.00-20-004/11-00.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Fujishima, A.; Honda, K. Photolysis-decomposition of water at the surface of an irradiated semiconductor. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Savinkina, E.; Obolenskaya, L.; Kuzmicheva, G. Efficiency of sensitizing nano-titania with organic dyes and peroxo complexes. Appl. Nanosci. 2015, 5, 125–133. [Google Scholar] [CrossRef]
  3. Guarisco, C.; Palmisano, G.; Calogero, G.; Ciriminna, R.; di Marco, G.; Loddo, V.; Pagliaro, M.; Parrino, F. Visible-light driven oxidation of gaseous aliphatic alcohols to the corresponding carbonyls via TiO2 sensitized by a perylene derivative. Environ. Sci. Pollut. Res. 2014, 21, 11135–11141. [Google Scholar] [CrossRef] [PubMed]
  4. Rao, V.G.; Dhital, B.; He, Y.; Lu, H.P. Single-Molecule Interfacial Electron Transfer Dynamics of Porphyrin on TiO2 Nanoparticles: Dissecting the Complex Electronic Coupling Dependent Dynamics. J. Phys. Chem. C 2014, 118, 20209–20221. [Google Scholar] [CrossRef]
  5. Machado, A.E.H.; França, M.D.; Velani, V.; Magnino, G.A.; Velani, H.M.M.; Freitas, F.S.; Müller, P.S.; Sattler, C.; Schmücker, M. Characterization and Evaluation of the Efficiency of TiO2/Zinc Phthalocyanine Nanocomposites as Photocatalysts for Wastewater Treatment Using Solar Irradiation. Int. J. Photoenergy 2008, 2008, 482373. [Google Scholar] [CrossRef]
  6. Ardo, S.; Achey, D.; Morris, A.J.; Abrahamsson, M.; Meyer, G.J. Non-Nernstian Two-Electron Transfer Photocatalysis at Metalloporphyrin–TiO2 Interfaces. J. Am. Chem. Soc. 2011, 133, 16572–16580. [Google Scholar] [CrossRef]
  7. Ren, X.-B.; Chen, M.; Qian, D.-J. Pd(II)-Mediated Triad Multilayers with Zinc Tetrapyridylporphyrin and Pyridine-Functionalized Nano-TiO2 as Linkers: Assembly, Characterization, and Photocatalytic Properties. Langmuir 2012, 28, 7711–7719. [Google Scholar] [CrossRef]
  8. Duan, M.; Li, J.; Mele, G.; Wang, C.; Lü, X.; Vasapollo, G.; Zhang, F. Photocatalytic Activity of Novel Tin Porphyrin/TiO2 Based Composites. J. Phys. Chem. C 2010, 114, 7857–7862. [Google Scholar] [CrossRef]
  9. Li, D.; Dong, W.; Sun, S.; Shi, Z.; Feng, S. Photocatalytic Degradation of Acid Chrome Blue K with Porphyrin-Sensitized TiO2 under Visible Light. J. Phys. Chem. C 2008, 112, 14878–14882. [Google Scholar] [CrossRef]
  10. Mele, G.; del Sole, R.; Vasapollo, G.; Marcì, G.; Garcìa-Lòpez, E.; Palmisano, L.; Coronado, J.M.; Hernández-Alonso, M.D.; Malitesta, C.; Guascito, M.R. TRMC, XPS, and EPR Characterizations of Polycrystalline TiO2 Porphyrin Impregnated Powders and Their Catalytic Activity for 4-Nitrophenol Photodegradation in Aqueous Suspension. J. Phys. Chem. B 2005, 109, 12347–12352. [Google Scholar] [CrossRef]
  11. Meng, Z.-D.; Zhu, L.; Choi, J.-G.; Chen, M.-L.; Oh, W.-C. Effect of Pt treated fullerene/TiO2 on the photocatalytic degradation of MO under visible light. J. Mater. Chem. 2011, 21, 7596. [Google Scholar] [CrossRef]
  12. Krishna, V.; Noguchi, N.; Koopman, B.; Moudgil, B. Enhancement of titanium dioxide photocatalysis by water-soluble fullerenes. J. Colloid Interface Sci. 2006, 304, 166–171. [Google Scholar] [CrossRef] [PubMed]
  13. Krishna, V.; Yanes, D.; Imaram, W.; Angerhofer, A.; Koopman, B.; Moudgil, B. Mechanism of enhanced photocatalysis with polyhydroxy fullerenes. Appl. Catal. B Environ. 2008, 79, 376–381. [Google Scholar] [CrossRef]
  14. Lin, J.; Zong, R.; Zhou, M.; Zhu, Y. Photoelectric catalytic degradation of methylene blue by C60-modified TiO2 nanotube array. Appl. Catal. B Environ. 2009, 89, 425–431. [Google Scholar] [CrossRef]
  15. Apostolopoulou, V.; Vakros, J.; Kordulis, C.; Lycourghiotis, A. Preparation and characterization of [60] fullerene nanoparticles supported on titania used as a photocatalyst. Colloids Surf. Physicochem. Eng. Asp. 2009, 349, 189–194. [Google Scholar] [CrossRef]
  16. Oh, W.-C.; Jung, A.-R.; Ko, W.-B. Preparation of Fullerene/TiO2 Composite and Its Photocatalytic Effect. J. Ind. Eng. Chem. 2007, 13, 1208–1214. [Google Scholar]
  17. Katsumata, K.; Matsushita, N.; Okada, K. Preparation of TiO2-Fullerene Composites and Their Photocatalytic Activity under Visible Light. Int. J. Photoenergy 2012, 2012, 256096. [Google Scholar] [CrossRef]
  18. Oh, W.-C.; Zhang, F.-J.; Meng, Z.-D.; Zhang, K. Relative Photonic Properties of Fe/TiO2-Nanocarbon Catalysts for Degradation of MB Solution under Visible Light. Bull. Korean Chem. Soc. 2010, 31, 1128–1134. [Google Scholar] [CrossRef]
  19. Yang, M.-Q.; Zhang, N.; Xu, Y.-J. Synthesis of Fullerene–, Carbon Nanotube–, and Graphene–TiO2 Nanocomposite Photocatalysts for Selective Oxidation: A Comparative Study. ACS Appl. Mater. Interfaces 2013, 5, 1156–1164. [Google Scholar] [CrossRef]
  20. Zhao, Z.; Ozoemena, K.I.; Maree, D.M.; Nyokong, T. Synthesis and electrochemical studies of a covalently linked cobalt(II) phthalocyanine-cobalt(II) porphyrin conjugate. Dalton Trans. 2005, 7, 1241–1248. [Google Scholar] [CrossRef]
  21. Kadish, K.M.; van Caemelbecke, E. Electrochemistry of porphyrins and related macrocycles. J. Solid State Electrochem. 2003, 7, 254–258. [Google Scholar] [CrossRef]
  22. Fernández-García, M.; Wang, X.; Belver, C.; Hanson, J.C.; Rodriguez, J.A. Anatase-TiO2 Nanomaterials: Morphological/Size Dependence of the Crystallization and Phase Behavior Phenomena. J. Phys. Chem. C 2007, 111, 674–682. [Google Scholar] [CrossRef]
  23. Burgos, M.; Langlet, M. The sol-gel transformation of TIPT coatings: A FTIR study. Thin Solid Films 1999, 349, 19–23. [Google Scholar] [CrossRef]
  24. Zhao, D.; Peng, T.; Liu, M.; Lu, L.; Cai, P. Fabrication, characterization and photocatalytic activity of Gd3+-doped titania nanoparticles with mesostructured. Microporous Mesoporous Mater. 2008, 114, 166–174. [Google Scholar] [CrossRef]
  25. Regulska, E.; Karpinska, J. Investigation of Photocatalytic Activity of C60/TiO2 Nanocomposites Produced by Evaporation Drying Method. Pol. J. Environ. Stud. 2014, 23, 2175–2182. [Google Scholar]
  26. Regulska, E.; Karpinska, J. Investigation of novel material for effective photodegradation of bezafibrate in aqueous samples. Environ. Sci. Pollut. Res. 2014, 21, 5242–5248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Su, C.; Hong, B.-Y.; Tseng, C.-M. Sol–gel preparation and photocatalysis of titanium dioxide. Catal. Today 2004, 96, 119–126. [Google Scholar] [CrossRef]
  28. Bahnemann, D.; Henglein, A.; Spanhel, L. Detection of the intermediates of colloidal TiO2-catalysed photoreactions. Faraday Discuss. Chem. Soc. 1984, 78, 151–163. [Google Scholar] [CrossRef]
  29. Chung, I.; Lee, B.; He, J.; Chang, R.P.H.; Kanatzidis, M.G. All-solid-state dye-sensitized solar cells with high efficiency. Nature 2012, 485, 486–489. [Google Scholar] [CrossRef]
  30. Dz. U 2014 poz. 1482—Regulation of the Minister of the Environment on the Method of Classifying the Status of Uniform Bodies of Surface Water and Environmental Quality Standards for Priority Substances (Journal of Laws, Item 1482, 2014); 2014.
  31. Dz. U 2014 poz. 1800—Regulation of the Minister of the Environment on the Conditions to Be Met When Discharging Sewage to Waters or to the Soil and on Substances of Particular Adverse Impact on the Water Environment (Journal of Laws, Item 1800, 2014); 2014.
  32. Mariquit, E.G.; Salim, C.; Hinode, H. The Role of Calcium Ions in the Photocatalytic Oxidation of Humic Acid at Neutral pH. Ann. N. Y. Acad. Sci. 2008, 1140, 389–393. [Google Scholar] [CrossRef]
  33. Jiang, Y.; Luo, Y.; Lu, Z.; Huo, P.; Xing, W.; He, M.; Li, J.; Yan, Y. Influence of Inorganic Ions and pH on the Photodegradation of 1-Methylimidazole-2-thiol with TiO2 Photocatalyst Based on Magnetic Multi-walled Carbon Nanotubes. Bull. Korean Chem. Soc. 2014, 35, 76–82. [Google Scholar] [CrossRef]
  34. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269–8285. [Google Scholar] [CrossRef]
  35. Martínez-Zapata, M.; Aristizábal, C.; Peñuela, G. Photodegradation of the endocrine-disrupting chemicals 4n-nonylphenol and triclosan by simulated solar UV irradiation in aqueous solutions with Fe(III) and in the absence/presence of humic acids. J. Photochem. Photobiol. Chem. 2013, 251, 41–49. [Google Scholar] [CrossRef]
  36. Canonica, S. Oxidation of Aquatic Organic Contaminants Induced by Excited Triplet States. Chim. Int. J. Chem. 2007, 61, 641–644. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from E.R.
Figure 1. Structures of the fullerene derivatives: 1, 2 and 3.
Figure 1. Structures of the fullerene derivatives: 1, 2 and 3.
Molecules 24 01118 g001
Figure 2. Cyclic voltammograms of 1 (a), 2 (b), 3 (c) in TBAPF6/(MeCN:Tol = 1:4, v/v), at a scan rate of 50 mV∙s−1.
Figure 2. Cyclic voltammograms of 1 (a), 2 (b), 3 (c) in TBAPF6/(MeCN:Tol = 1:4, v/v), at a scan rate of 50 mV∙s−1.
Molecules 24 01118 g002
Figure 3. HOMO–LUMO energy levels for the synthesized fullerene derivatives (13) and TiO2.
Figure 3. HOMO–LUMO energy levels for the synthesized fullerene derivatives (13) and TiO2.
Molecules 24 01118 g003
Figure 4. Raman (black line) and Attenuated Total Reflection-Fourier Transform Infra-Red (ATR-FTIR) (grey line) spectra of the synthesized titania.
Figure 4. Raman (black line) and Attenuated Total Reflection-Fourier Transform Infra-Red (ATR-FTIR) (grey line) spectra of the synthesized titania.
Molecules 24 01118 g004
Figure 5. SEM images of titania (1), the composites TiO2/1 (2a), TiO2/2 (3a), and TiO2/3 (4a) and the fullerene derivates 1 (2b), 2 (3b), and 3 (4b).
Figure 5. SEM images of titania (1), the composites TiO2/1 (2a), TiO2/2 (3a), and TiO2/3 (4a) and the fullerene derivates 1 (2b), 2 (3b), and 3 (4b).
Molecules 24 01118 g005
Figure 6. Kubelka–Munk plots of the titania-based composites with the fullerene derivatives: 1, 2 and 3.
Figure 6. Kubelka–Munk plots of the titania-based composites with the fullerene derivatives: 1, 2 and 3.
Molecules 24 01118 g006
Figure 7. The efficiencies of phenol and methylene blue photocatalytic degradation in the presence of TiO2 and the given composites.
Figure 7. The efficiencies of phenol and methylene blue photocatalytic degradation in the presence of TiO2 and the given composites.
Molecules 24 01118 g007
Figure 8. The decrease of phenol (a) and methylene blue (b) relative concentrations, and the efficiencies (insets) of their photocatalytic degradation in the presence of TiO2 and the given composites.
Figure 8. The decrease of phenol (a) and methylene blue (b) relative concentrations, and the efficiencies (insets) of their photocatalytic degradation in the presence of TiO2 and the given composites.
Molecules 24 01118 g008
Figure 9. The proposed mechanism of the photocatalytic activity of the porphyrin-functionalized fullerenes/titania composites.
Figure 9. The proposed mechanism of the photocatalytic activity of the porphyrin-functionalized fullerenes/titania composites.
Molecules 24 01118 g009
Figure 10. The chromatograms of phenol (PhOH) initial solution (dotted line) and after 2 h of photocatalytic treatment (solid line) with TiO2/3, recorded at 239 nm. Absorption spectra corresponding to the peaks are registered in the chromatograms at 1.71 (a), 2.56 (b), and 4.05 min (c). Insets represents mass spectra registered at respective retention times.
Figure 10. The chromatograms of phenol (PhOH) initial solution (dotted line) and after 2 h of photocatalytic treatment (solid line) with TiO2/3, recorded at 239 nm. Absorption spectra corresponding to the peaks are registered in the chromatograms at 1.71 (a), 2.56 (b), and 4.05 min (c). Insets represents mass spectra registered at respective retention times.
Molecules 24 01118 g010
Figure 11. The chromatogram of the initial methylene blue (MB) solution (dotted line) and after 2 h of photocatalytic treatment (solid line) with TiO2/3 recorded at 610 nm. Absorption spectra corresponding to the peaks are registered in the chromatogram at 5.61 (a), 6.20 (b), and 6.34 min (c). Insets are the mass spectra registered at the respective retention times.
Figure 11. The chromatogram of the initial methylene blue (MB) solution (dotted line) and after 2 h of photocatalytic treatment (solid line) with TiO2/3 recorded at 610 nm. Absorption spectra corresponding to the peaks are registered in the chromatogram at 5.61 (a), 6.20 (b), and 6.34 min (c). Insets are the mass spectra registered at the respective retention times.
Molecules 24 01118 g011
Figure 12. Efficiencies (η) of PhOH (a) and MB (b) photocatalytic degradation in the presence of TiO2 or TiO2/3, and the following matrices: MilliQ water, Goldapa, Sapina, and Suprasl river waters, wastewater effluent, and municipal wastewater.
Figure 12. Efficiencies (η) of PhOH (a) and MB (b) photocatalytic degradation in the presence of TiO2 or TiO2/3, and the following matrices: MilliQ water, Goldapa, Sapina, and Suprasl river waters, wastewater effluent, and municipal wastewater.
Molecules 24 01118 g012
Table 1. Brunauer–Emmett–Teller (BET) surface areas and diameters determined for titania and titania-based composites.
Table 1. Brunauer–Emmett–Teller (BET) surface areas and diameters determined for titania and titania-based composites.
CatalystTiO2TiO2/1TiO2/2TiO2/3
BET surface area/m2∙g−1269178151158
Diameter/μm3.241.070.860.48
Table 2. River water parameters.
Table 2. River water parameters.
ParametrGoldapa RiverSapina RiverSuprasl RiverThe Range of Acceptable Values for Rivers [30]Norms
pH7.87.87.56.5-8.5PN-90/C-04540/01
Acidity/mg CaCO3 ∙ L−1111.699.280.6ndPN-90/C-04540/03
Alkalinity/mg CaCO3 ∙ L−1240.2260.2180.2≤150 (I) *;
≤250 (II) **
PN-90/C-04540/03
COD ***/mg O2 ∙ L−11.92.32.3≤6 (I);
≤12 (II)
PN-85/C-04578/02
Conductivity/S ∙ cm−1245270360≤1000 (I);
≤1500 (II)
Chlorine (Cl)/mg Cl ∙ L−1156.0113.4170.2≤200 (I);
≤300 (II)
PN-75/C-04617/02
Calcium (Ca2+)/
mg Ca2+ ∙ L−1
56.174.1104.2≤100 (I);
≤200 (II)
PN-91/C-04551/01
Magnesium (Mg2+)/
mg Mg2+ ∙ L−1
27.316.035.3≤50 (I);
≤100 (II)
PN-71/C-04554/10
Hardness/mg CaCO3 ∙ L−1207.6223.6347.3≤300 (I);
≤500 (II)
PN-71/C-04554/02
Dissolved oxygen/mg O2 ∙ L−10.92.21.0≥7 (I);
≥5 (II)
ISO 5813:1983
Nitrate (V)/mg ∙ L−13.73.469.3≤2.2 (I);
≤5.0 (II)
PN-82/C-04576/08
Phosphate/mg ∙ L−111.57.3nd≤0.20 (I);
0.31 (II)
PN-88/C-04537/04
* First grade of quality, ** Second grade of quality, *** Chemical oxygen demand.
Table 3. Municipal wastewater and wastewater effluents parameters.
Table 3. Municipal wastewater and wastewater effluents parameters.
ParameterMunicipal WastewaterWastewater EffluentsThe range of Acceptable Values for Effluents [31]Norms
pH10.810.76.5-9.0PN-90/C-04540/01
Acidity/mg CaCO3 ∙ L−12.31.1ndPN-90/C-04540/03
Alkalinity/mg CaCO3 ∙ L−1580.0285.0150-350PN-90/C-04540/03
COD/mg O2 ∙ L−189.836.7<125PN-85/C-04578/02
Conductivity/S ∙ cm−185007000nd
Chlorine (Cl)/mg Cl ∙ L−1368.7326.1<1000PN-75/C-04617/02
Calcium (Ca2+)/mg Ca2+ ∙ L−178.686.7ndPN-91/C-04551/01
Magnesium (Mg2+)/mg Mg2+ ∙ L−141.717.6>300PN-71/C-04554/10
Hardness/mg CaCO3 ∙ L−1300.2260.2ndPN-71/C-04554/02
Dissolved oxygen/mg O2 ∙ L−1nd1.4ndISO 5813:1983
Nitrate (V)/mg ∙ L−11226.345.0<30PN-82/C-04576/08
Phosphate/mg ∙ L−13418.518.1ndPN-88/C-04537/04

Share and Cite

MDPI and ACS Style

Regulska, E.; Rivera-Nazario, D.M.; Karpinska, J.; Plonska-Brzezinska, M.E.; Echegoyen, L. Zinc Porphyrin-Functionalized Fullerenes for the Sensitization of Titania as a Visible-Light Active Photocatalyst: River Waters and Wastewaters Remediation. Molecules 2019, 24, 1118. https://doi.org/10.3390/molecules24061118

AMA Style

Regulska E, Rivera-Nazario DM, Karpinska J, Plonska-Brzezinska ME, Echegoyen L. Zinc Porphyrin-Functionalized Fullerenes for the Sensitization of Titania as a Visible-Light Active Photocatalyst: River Waters and Wastewaters Remediation. Molecules. 2019; 24(6):1118. https://doi.org/10.3390/molecules24061118

Chicago/Turabian Style

Regulska, Elzbieta, Danisha Maria Rivera-Nazario, Joanna Karpinska, Marta Eliza Plonska-Brzezinska, and Luis Echegoyen. 2019. "Zinc Porphyrin-Functionalized Fullerenes for the Sensitization of Titania as a Visible-Light Active Photocatalyst: River Waters and Wastewaters Remediation" Molecules 24, no. 6: 1118. https://doi.org/10.3390/molecules24061118

Article Metrics

Back to TopTop