Next Article in Journal
Kinetic Modeling of Catalytic Olefin Cracking and Methanol-to-Olefins (MTO) over Zeolites: A Review
Next Article in Special Issue
Enhanced Photocatalytic Activity of Titania by Co-Doping with Mo and W
Previous Article in Journal
Biotechnological Methods of Sulfoxidation: Yesterday, Today, Tomorrow
Previous Article in Special Issue
Compositing Two-Dimensional Materials with TiO2 for Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Degradation of Estriol Using Iron-Doped TiO2 under High and Low UV Irradiation

by
Irwing M. Ramírez-Sánchez
1 and
Erick R. Bandala
2,3,*
1
Department of Civil, Architectural and Environmental Engineering, The University of Texas at Austin, Austin, TX 78712, USA
2
Desert Research Institute (DRI), 755 E. Flamingo Road, Las Vegas, NV 89119-7363, USA
3
Graduate Program Hydrologic Sciences, University of Nevada, Reno, NV 89557, USA
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(12), 625; https://doi.org/10.3390/catal8120625
Submission received: 29 September 2018 / Revised: 21 November 2018 / Accepted: 28 November 2018 / Published: 5 December 2018
(This article belongs to the Special Issue Emerging Trends in TiO2 Photocatalysis and Applications)

Abstract

:
Iron-doped TiO2 nanoparticles (Fe-TiO2) were synthesized and photocatalitically investigated under high and low fluence values of UV radiation. The Fe-TiO2 physical characterization was performed using X-ray Powder Diffraction (XRD), Brunauer–Emmett–Teller (BET) surface area analysis, Transmission Electron Microscopy (TEM), Scanning Electron Microscopy (SEM), Diffuse Reflectance Spectroscopy (DRS), and X-ray Photoelectron Spectroscopy (XPS). The XPS evidenced that the ferric ion (Fe3+) was in the TiO2 lattice and unintentionally added co-dopants were also present because of the precursors of the synthetic method. The Fe3+ concentration played a key role in the photocatalytic generation of hydroxyl radicals (OH) and estriol (E3) degradation. Fe-TiO2 accomplished E3 degradation, and it was found that the catalyst with 0.3 at.% content of Fe (0.3 Fe-TiO2) enhanced the photocatalytic activity under low UV irradiation compared with TiO2 without intentionally added Fe (zero-iron TiO2) and Aeroxide® TiO2 P25. Furthermore, the enhanced photocatalytic activity of 0.3 Fe-TiO2 under low UV irradiation may have applications when radiation intensity must be controlled, as in medical applications, or when strong UV absorbing species are present in water.

1. Introduction

In recent years, society and the scientific community have concerned of Emerging Contaminants (ECs, also called Contaminants of Emerging Concern), which are chemicals that threaten the environment, human health, and water safety and are not currently covered by existing local or international water quality regulations [1]. ECs include chemical species such as algae toxins, illegal drugs, industrial compounds, flame retardants, food additives, nanoparticles, pharmaceuticals (human and veterinary), personal care products, pesticides, biocides, steroids, synthetic and natural hormones, and surfactants [2].
Natural hormones (e.g., estrone (E1), 17β-estradiol (E2), and estriol (E3)) as ECs are susceptible of persisting and bioaccumulating in the environment, and could induce endocrine disruption in humans and wildlife (vertebrates [3,4,5] and invertebrates [6,7]). Natural attenuation, drinking water purification, and conventional municipal wastewater treatment processes are either incapable or only partially capable of removing estrogens from water [8]. As result, water treatment techniques are being developed to manage, reduce, degrade, and mineralize low-concentrated ECs (including natural estrogen) in drinking and wastewater [9]. Advanced Oxidation Processes (AOPs) are promising techniques to treat ECs in aqueous phase, which include well-known processes such as Fenton and Fenton-like processes, UV/H2O2, ozonation, and photocatalysis using semiconductors, peroxone processes (H2O2/O3), and cavitation [10,11]. Although there are many known AOPs, since Coleman’s work [12], photocatalysis using titanium dioxide (TiO2) has been identified as one of the most effective methods to degrade estrogens in water [13]. Several reports recognized that TiO2 can degrade estrogens, which prevents increases in estrogenic activity in water [14,15] and partially or completely mineralizing estrogens [14,16].
Titanium dioxide is the most commonly used photocatalyst because of its reasonable optical and electronic properties, good photocatalytic activity, insolubility in water, chemical and photochemical stability, nontoxicity, low cost, and high efficiency in pollutant mineralization [17,18,19,20]. However, the band gap energy (Eg) of TiO2, frequently reported as 3.2 eV [21], restrains the photocatalytic activation to energy sources with a portion of spectrum emission below 387.5 nm [22].
In general the photocatalytic mechanism is as shown in Figure 1. According to Density Functional Theory (DFT) computations, the valence band (VB) and conduction band (CB) of pure TiO2 are mainly composed of O2p orbitals and Ti3d orbitals, respectively. Hence, the Fermi level (EF) is located in the middle of the band gap (BG), indicating that VB is full filled while CB is empty [23]. When using photons with energy higher than 3.2 eV, photoexcitation of the semiconductor promotes electrons from VB to CB creating a charge vacancy or hole (h+) in the VB. The h+ in the VB can react with hydroxide ion to form hydroxyl radical (OH) or can also be filled by donor absorbed organic molecule (OMads). Photogenerated electrons in the CB can be transferred to acceptor of electrons and bring about OH.
Consequently, reducing the photon energy needed for TiO2 photoactivation has been the focus of the scientific community until now. Doping is one of the techniques that has been tested to control or modify the surface properties or internal structure of TiO2. Doping introduces a foreign element into TiO2 to cause an impurity state in the band gap. The most frequently used doping materials are transition-metal cations (e.g., Cr, V, Fe, and Ni) at Ti sites, and anions (e.g., N, S, and C) at O sites [24]. Among anion- and cation-dopants, the ferric ion (Fe3+) is one of the most often used because the ionic radius of Fe3+ (0.69 A) is similar to Ti4+ (0.745 A) [25]. Therefore, Fe3+ can be easily incorporated into the TiO2 crystal lattice.
The main reported effects of iron-doped TiO2 is a rapid increase in photocatalytic activity that increases with increased Fe doping, which then reaches a maximum value, and finally decreases with further increased Fe content [23,26,27,28,29,30,31,32,33,34,35,36,37]. However, detrimental effects have been also reported because of high Fe content [38,39] or agglomerated Fe-TiO2 nanoparticles [40,41].
Although several theoretical and experimental Fe-TiO2 studies have been developed, the trade-off between doping ratio and radiation intensity is scarcely mentioned. Furthermore, Fe-TiO2 photocatalyst has rarely been considered to be a useful technique for the degradation of E3 [42].
In this work, Fe-TiO2 nanoparticles were synthesized to increase the understanding of the relationship between doping ratio and radiation intensity for hydroxyl radical (OH) generation and E3 degradation. Therefore, we investigated the photocatalytic degradation of E3 using Fe-TiO2 under high and low UV irradiation. We highlight the term low UV irradiation to avoid confusion with the term “photocatalytic processes under visible light” because we did not intentionally use UV cutoff filters for the experiments.

2. Results and Discussion

2.1. Characterization of Iron-Doped TiO2

Figure 2 shows X-ray Photoelectron Spectroscopy (XPS) general spectra of TiO2 without added Fe (zero-iron TiO2) and Fe-TiO2 materials (b, c, and d). For the experimental condition used, Fe did not affect the bonding structure between titanium and oxygen because the main peaks for all samples were Ti2p and O1s with the proportion 1:2.2, which is in agreement with the atomic formula of TiO2.
XPS detected unintentionally added elements such as carbon, sulfur, and nitrogen (Table 1) as co-dopants of zero-iron TiO2 and Fe-TiO2, which were introduced into TiO2 via precursors of the synthesis. Carbon and sulfur could come from sodium dodecyl sulfate (SDS), and nitrogen could come from iron (III) nitrate (Fe(NO3)3·9H2O) and HNO3, all of them used in the synthesis process.
High-resolution XPS spectra for the iron region (Figure 3) was studied only for 1.0 Fe-TiO2 because no Fe2p signals were detected for zero-iron TiO2, 0.3 Fe-TiO2, or 0.6 Fe-TiO2. The deconvolution of high-resolution XPS spectra (Figure 3) was developed for previously reported peaks of Fe2+ and Fe3+ [43]. Shirley baseline was subtracted before peak fitting. The Gaussian–Lorentzian mix function was used with a 40% factor. Charge compensation was set by the O1s peak charge with −0.58 eV. As a result, the correlation between the experimental signal and the theoretic model (Σχ2) was 8.43 × 10−2.
According to the theoretical model (sum of fitting peaks), both Fe3+ and Fe2+ were present in the lattice of 1.0 Fe-TiO2. We suggest that Fe3+ was incorporated into the lattice of TiO2 to form Ti–O–Fe bonds, because the ionic radius of Fe3+ (0.69 A) is similar to the ionic radius of Ti4+ (0.745 A) [25]. The XPS technique detected Fe2+ because Fe3+ underwent reduction to Fe2+ during XPS measurement in vacuum [44].
The band gap energy (Eg) obtained with the Kubelka–Monk method (Figure 4) for Aeroxide® TiO2 P25 was 3.2 eV, which is consistent with the value reported previously [45]. For Aeroxide® TiO2 P25 Eg, red-shifts were detected as 0.22, 0.24, 0.25, and 0.3 eV for zero-iron TiO2, 0.3 Fe-TiO2, 0.6 Fe-TiO2, and 1.0 Fe-TiO2, respectively, which is consistent with values reported by Shi et al. of 0.25 eV [46] and with density functional theory calculations that suggested the hybridized band of Ti3d and Fe3d reduces Eg approximately 0.3–0.5 eV [44], or 0.2–0.34 eV [47].
For zero-iron TiO2, Eg for Fe-TiO2 materials (Table 2) decreased as long as the Fe content increased, so the Fe content generated red-shift. For Aeroxide® TiO2 P25 Eg, the red-shift of Fe-TiO2 agreed with previously reported values, but it agreed less for zero-iron TiO2. Therefore, red-shift was not only related to Fe content, but also to the synthesis method and unintentionally co-doped TiO2.
XRD patterns in Figure 5 revealed zero-iron TiO2 and Fe-TiO2 materials had both anatase and rutile phases. No XRD Fe2O3 peaks (2θ equal to 33.0°, 35.4°, 40.7°, 43.4°, and 49.2°) were observed, concluding that Fe3+ replaced Ti4+ in the TiO2 crystal framework [48,49]. The synthesis method allowed uniform distribution of Fe within TiO2. The anatase:rutile phase ratio calculated by Spurr and Myers’ method showed that zero-iron TiO2 and Fe-TiO2 materials were a mixture of anatase and rutile phases (Table 2). The amount of anatase was less in Fe-TiO2 materials than in Aeroxide® TiO2 P25. The smaller proportion of anatase could lead to a reduction of photocatalytic activity because the anatase phase has higher photocatalytic activity than rutile TiO2 [50,51]. However, it is accepted that the optimal photocatalytic activity of TiO2 is reached with an optimal mixture of anatase and rutile phases [52]. Moreover, the increased anatase proportion in 0.3 Fe-TiO2 and 0.6 Fe-TiO2 compared with zero-iron TiO2 could improve photocatalytic activity. The increased anatase proportion was attributable to Fe doping disturbing the arrangements of TiO2 phases [53]. This trend has also been observed when Fe-doped TiO2 was synthesized using sol-gel [54] or co-precipitation methods [32].
The average particle size of Fe-TiO2 materials obtained by Scherrer’s formula was 6.9 nm, which is less than the particle size of Aeroxide® TiO2 P25 (Table 2). Fe-TiO2 materials should increase photocatalytic activity because of their higher surface area and the short migration distance of the photogenerated charge carriers (electron/hole (e/h+)) from the bulk material to the surface.
Further BET analysis (Figure 6) confirmed that average surface area of Fe-TiO2 materials was 77.9 m2 g−1, higher than zero-iron TiO2 and Aeroxide® TiO2 P25. BET isotherms followed a type IV shape according to the Langmuir classification, which is associated with the characteristics of mesoporous material [55]. The observed hysteresis is probably due to gas cooperative adsorption or condensation inside the pores of material [56]. BET analysis showed pore sizes (Table 2) were in the mesoporous range (2–50 nm, according to IUPAC classification) for zero-iron TiO2 and 1.0 Fe-TiO2, and the microporous range (0.2–2 nm, according to IUPAC classification) for 0.3 Fe-TiO2 and 0.6 Fe-TiO2. Mesoporous pore size should facilitate the mass transfer of reactants and products in the reaction system, so photocatalytic improvement based on this property could improve zero-iron TiO2 and Fe-TiO2 materials with respect to Aeroxide® TiO2 P25 [31].
Patra et al. [49] developed a similar nanoparticle synthesis procedure, which generated surface area values ranging from 126 to 385 m2 g−1 and mesoporous size distribution values ranging from 3.1 to 3.4 nm. Particles obtained in our work were different, probably because of the application of a mild thermal treatment and the use of SDS at critical micelle concentration as a template.
Figure 7 shows SEM images of agglomerated and assembled nanoparticles of zero-iron TiO2. The different amounts of Fe in the TiO2 lattice changed neither the particle size nor the morphology of the zero-iron TiO2. Although the average pore size allowed an increase of the superficial area, agglomeration could lead to lower photocatalytic activity.
Transmission electron microscopy (TEM) images confirmed nanoparticle clusters and particle sizes of zero-iron TiO2 (Figure 8b) and 0.3 Fe-TiO2 (Figure 8a) between 5 and 10 nm (between 1.2 and 9.4 nm according to Scherrer’s formula). The lattice fringe spacing was 0.35 nm, as shown in Figure 8b, which was consistent with the d-spacing (101) of anatase [25]. The lattice fingers of the nanoparticles showed that Fe-TiO2 materials were highly crystallized.

2.2. Characterization of Irradiation Source

Figure 9 shows the emission spectra of irradiation sources used in this study. Using the main peaks reported for a fluorescent lamp (Figure 9a), the calibration of the spectrometer generated an R2 value equal to 0.999. The emission spectrum of the GE F15T8 BLB lamp (Figure 9b) was in the 356–410 nm range. However, the emission spectrum of the GE F15T8 D lamp (Figure 9c) was continuous broadband between 380 and 750 nm. The light intensity of the GE F15T8 lamp was reported to be between 3440 µW cm−2 [57] and 4000 µW cm−2 [58], from which 6% was UV radiation [59]. The intensity of the GE F15T8 lamp was 1500 µW cm−2. This lamp has an internal coating that absorbs 78% of visible light (as specified by the manufacturer) in the spectrum below 400 nm, as shown in Figure 9b. Therefore, the GE F15T8 BLB and GE F15T8 D lamps were designated as high and low UV irradiation sources, respectively.
Because Eg of Aeroxide® TiO2 P25 is 3.2 eV (387.5 nm), see Figure 9, both the GE F15T8 BLB and GE F15T8 D lamps emitted photons that could photoactivate Aeroxide® TiO2 P25. However, the proportion of the emission spectrum that Aeroxide® TiO2 P25 could use for photocatalytic activity was different. An approximation of the amount of radiative intensity used for photocatalytic activity was obtained with the area under the curve-spectrum below the Eg value. Consequently, Aeroxide® TiO2 P25 could take advantage of 36.4% of the emission spectrum of the GE F15T8 BLB lamp and 0.8% of the emission spectrum of the GE F15T8 D lamp. Table 2 lists amount of radiative spectrum used by zero-iron TiO2 and Fe-TiO2 materials according to each Eg.
Based on morphological and crystalline structure analysis, the favorable characteristics to enhance photocatalytic activity of Fe-TiO2 material are effective insertion of the Fe3+ ion into the TiO2 lattice, red-shift (2.90–2.96 eV), nanoparticle size (6.9–7.1 nm), specific surface area (73.0–83.1 nm), pore size (1.2–9.4 nm), and radiation absorbance below the equivalent Eg wavelength (8.21–10.63% of daylight lamp spectrum). Its main disadvantageous characteristics are expected to be high particle agglomeration and lower anatase phase compared with zero-iron TiO2. Further, photocatalytic activity is very sensitive to crystalline array and particle size and shape; differences in the density of hydroxyl groups on the particle surface and the number of water molecules hydrating the surface; the surface area and surface charge; differences in the number and nature of trap sites; the dopant concentration, localization, and chemical state of the dopant ions; radiation intensity; particle aggregation and superficial charge; and scavenger species in media [39,60]. Consequently, material characterization alone could not predict photocatalytic activity [28]. Therefore, in this research, we used the N,N-dimethyl-p-nitrosoaniline (pNDA) probe and E3 to evaluate the photocatalytic activity by following OH production, which is one of the most significant reactive oxygen species (ROS), and E3, which is an EC.

2.3. Hydroxyl Radical Generation under High and Low UV Irradiation

The generation of OH was measured using pNDA, which is a well-characterized OH scavenger as mentioned in Section 3.5. In brief, pNDA undergoes bleaching when reacting with OH according to Muff et al. mechanism of the oxidation of pNDA by OH [61].
In this work, pNDA bleaching followed a pseudo-first-order equation, so the apparent rate constant was calculated by ln(C/C0) = k1t, where C0 is the initial concentration, C is the reaction concentration at a given time, and k1 is the pseudo-first-order reaction rate constant. The slope of the plot after applying a linear fit represents the rate constant, k1.
Because the relationship between pNDA bleaching and OH production follows a 1:1 stoichiometry [61], the steady-state of OH generation ([OH]ss) can be considered equal to the initial velocity (r0) according to Equation (1) and reported in Table 3:
[ pNDA ] dt | t = 0 = r 0 = [ O H ] ss
Fe-TiO2 materials showed a similar anatase:rutile phase ratio, particle size, and specific surface area, and therefore the variation in r0 values was due to the difference of Fe content inside TiO2. The generation of OH radicals (r0) was feasible using zero-iron TiO2, Fe-TiO2 materials, and Aeroxide® TiO2 P25 under both high (Figure 10a) and low UV irradiation (Figure 10b).
When high UV irradiation was used, the maximum r0 was 0.58 μM•OH min−1 for 0.3 Fe-TiO2. The enhancement in photocatalytic activity of 0.3 at.% Fe-TiO2, compared with zero-iron TiO2 was by the extended lifetime values of the photogenerated charge carriers (e and h+) produced by Fe3+ ions, which played a role as charge carriers trapped at or near the particle surface. The trapping mechanisms are shown in Equations (2)–(5) [62].
Fe3+ + ecd → Fe2+      electron trap
Fe2+ + Ti4+ → Fe3+ + Ti3+   migration
Fe3+ +hvb+ → Fe4+     hole trap
Fe4+ + OH → Fe3+ + OH  migration
The mechanism suggested for OH generation is shown in Figure 11. When TiO2 contains a Fe3+ ion, the Fe3d orbitals split into two bands, one is a hybrid band (A2g) and one is midgap band (T2g), which induce a new localized BG state [23]. Therefore, when TiO2 absorbs photons with energy less than 3.2 eV, photoexcitation of the semiconductor promotes an electron from the VB to the midgap band (T2g), also called a shallow trap, creating an electron-hole pair. The hole in the valence band (VB) can react with hydroxide ions to form OH, absorbed organic molecules, or trap Fe3+ following Equations (4) and (5). Additionally, photogenerated electrons in the midgap band (T2g) can be transferred to Fe3+ following a dark redox reaction at the interface, as suggested by Neubert et al. [63] and consequently bring about OH.
Increasing the Fe3+ doping content of Fe-TiO2 to 0.6 and 1.0 at.%, Fe-TiO2 was unfavorable to the photocatalytic activity because the additional Fe3+ doping in the TiO2 sample inhibited the extended lifetime of charge carriers, acted as recombination sites and consequently decreased the photocatalytic efficiency [29], as proposed in Equations (6)–(9) [39].
Fe2+ + hvb+ → Fe3+      recombination
Fe4+ + ecd → Fe3+      recombination
Fe4+ + Fe2+ → 2Fe3+     recombination
Fe4+ + Ti3+ → Fe3+ + Ti4+  recombination
When low UV irradiation conditions were used, the r0 values for zero-iron TiO2 and Fe-TiO2 materials were lower than the value estimated for Aeroxide® TiO2 P25. Compared with the effects of high UV irradiation, the reduction in r0 value observed was related both to pNDA adsorption of UV-visible radiation (lowered the number of photons available to activate the photocatalyst), and the augmented Fe content, which increased the recombination rate.

2.4. Photocatalytic Degradation of Estriol under High and Low UV Irradiation

E3 photocatalytic degradation curves are shown in Figure 12a,b using both high and low UV irradiation, respectively. In both cases, E3 photocatalytic degradation followed a pseudo-first-order model and the rate constant, k1 (Table 4), was obtained by fitting experimental data to ln ([E3]/[E30]) = k1t. Fe content influenced k1 for both high and low UV irradiation.
Figure 13 shows the pseudo-first-order rate constant (k1) of E3 photocatalytic degradation. In general, the photocatalytic activity first increased and then decreased as the Fe concentration increased, which is similar to the behavior found with the OH probe in Section 2.3 and has been previously reported using other organic molecules [23,29,64].
Under high UV irradiation (Figure 13a), 0.6 Fe-TiO2 k1 was higher than for zero-iron TiO2, 0.3 Fe-TiO2, and 1.0 Fe-TiO2. The increase in photocatalytic performance of 0.6 Fe-TiO2 was related with the increase in the lifetime of electron-hole pairs because Fe created additional energy levels near the conduction band of TiO2, as the mechanism suggests in Figure 11.
Under low UV irradiation (Figure 13b), zero-iron TiO2, 0.3 Fe-TiO2, and 0.6 Fe-TiO2 showed more photocatalytic activity than Aeroxide® TiO2 P25 because those materials had enhanced superficial properties, such as particle size, and superficial area, as mentioned in Section 2.1. Furthermore, 0.3 Fe-TiO2 enhanced photocatalytic activities with k1 values as high as 0.005 min−1. The high photocatalytic activity of 0.3 Fe-TiO2 was due to the synergistic effect of unintentionally added co-dopants, superficial properties, and Fe content that increased the lifetime of photogenerated charge carriers and the efficiency of electron transfer.
The photocatalytic degradation rate of E3 using Aeroxide® TiO2 P25 was reported to be 0.25 min−1 [65], 0.134 min−1 [66], and 0.12 min−1 [67]. However, the experimental setups and catalyst loads were different. Besides these few studies, E3 degradation using Fe-TiO2 nanoparticles is scarcely reported. Only comparing magnitudes of k1, the first-order rates to degrade pharmaceuticals using Fe-TiO2 nanoparticles were 0.001 min−1 for ibuprofen, 0.0015 min−1 for carbamazepine, and 0.0014 min−1 for sulfamethoxazole [68], which are in the order of magnitude obtained in this work (see Table 4).
Regarding unintentionally added co-dopants, Fe-TiO2 co-doping demonstrated a synergistic effect to increase photocatalytic activity under visible light for sulfur [69], nitrogen [44], and FexTi1-xO2-yNy co-doping [70]. Surface properties of the material, such as a particle size (6.9 nm) and surface area (77.6 m2 g−1), also facilitated the mass transfer between interface, E3, and sub-products.
The relationship between the OH radical system and E3 kinetic degradation was determined via linear fit between OH initial rate generation (r0,OH) and initial E3 degradation (r0,E3). In general, the procedure to correlate r0,OH and r0,E3 was first to sort pair values (r0,OH, r0,E3), and then fit the data to linear regression, as shown Figure 14a,b.
Under high UV irradiation, the linear fit correlation was r0,E3 = 0.091 r0,OH + 0.040 with R2 = 0.197. Under low UV irradiation, the linear fit correlation was r0,E3 = 0.066 r0,OH + 0.012 with R2 = 0.975. The correlation between the pair (r0,OH, r0,E3) under high UV irradiation was too low to be considered a linear relationship. We suggest the low correlation was because not only OH caused E3 degradation, but holes (h+) or other reactive oxygen species also caused E3 degradation.
However, a linear relationship under low UV irradiation was attributable to OH being the main reactive oxygen species responsible for photocatalytic activity. Therefore, the contribution of h+ to photocatalytic activity was lower because oxidation power was lower due to reduced Eg. This suggestion supports the mechanisms proposed in Figure 11, in which adding Fe into the lattice of TiO2 reduced the Eg with a consistent reduction of redox potential, as mentioned by others [28].
The main mechanism of E3 degradation under low UV irradiation was via electron (e) transfer to give rise OH. Additionally, the enhanced photocatalytic activity of 0.3 Fe-TiO2 under low UV irradiation provides evidence that the trapping-recombination mechanism of Fe-TiO2 can be controlled by irradiation intensity. Therefore, we suggest that there is a trade-off between irradiation intensity, the trapping-recombination rate, and OH production that is worthy of further research.
The efficiency resource of the Fe-TiO2/Low UV system was obtained through dimensional analysis of the slope of the linear fit of data shown in Figure 14b. The units of slope are E3 moles degraded per OH mol generated at initial time, so 0.662 E3 molecules underwent degradation when one OH was generated for the photocatalytic system independent of Fe doping content in TiO2. A sustainable process was also achieved, for which 0.3 Fe-TiO2 since absorbed 8.21% of emission spectra of the lamp below the equivalent Eg wavelength over 0.8% or 7.64% of Aeroxide® TiO2 P25 and zero-iron TiO2, respectively.

2.5. Relationship between Fe Content and Kinetic Constant

Photonic efficiency has been suggested to increase linearly with the doping ratio due to the formation of the charge carrier trapping centers, while it concurrently decreases quadratically with the doping ratio because to the creation of recombination centers [71]. Alternatively, we suggest an empirical relationship between the E3 degradation pseudo-first-order rate constant (k1) and Fe content (at.%) in TiO2, as described in Equation (10):
k 1 ( δ ) = c [ e k e ( δ + α ) e k a ( δ + α ) ]
where k1 is the pseudo-first-order constant, ke is the electron trap constant, ka is the electron recombination constant, δ at.% is the Fe doping amount in TiO2, and c and α are system constants. To solve the model described in Equation (10), a numerical approximation by root-mean-square error minimization method was used according to Equation (11):
ε = 1 n i | [ k 1. i ] ¯ [ k 1. i ] |
where [ k 1. i ] ¯ is the theoretical k1 value, [ k 1. i ] is the experimental k1 value, n is the number of data, and ε is the root-mean-square error. The solution of Equation (10) was performed by simultaneously solving ke, ka, c, and α using Excel Solver® (Frontline Systems, NV, US). As an example, photocatalytic degradation of E3 under low UV irradiation was fitted to Equation (10), as shown in Figure 15.
The empirical model solved in Equation (12) shows that electron trap constant (ke) overcome electron recombination (ka) before optimal catalyst load. This model could lead to experimental work using iron-doped TiO2 in which the optimal content of Fe gives rise to the maximum E3 degradation.
k 1 ( δ ) = 1.99 [ e 2.81 ( δ + 0.197 ) e 2.78 ( δ + 0.197 ) ]

3. Materials and Methods

3.1. Reagents

Sigma-Aldrich (St. Louis, MO, USA) supplied estriol (E3, C18H24O3, ≥97%), titanium isopropoxide (TTIP, Ti[OCH(CH3)2]4, 97%), N,N-Dimethyl-4-nitrosoaniline (pNDA, also called RNO, C8H10N20, 97%), sodium dodecyl sulfate (SDS), and iron (III) nitrate (Fe(NO3)3·9H2O, >99.99%). Aeroxide® TiO2 P25 (formerly Degussa P25 with 50 ± 15 m2 g−1 of the specific surface area, 21 nm of average particle size, 80:20 of anatase:rutile ratio according to the manufacturer) granted by Evonik Industries (Essen, Germany) was the photocatalytic standard. Fremont (CA, USA) supplied HNO3, H2SO4, absolute ethanol, HPLC-grade methanol, and HPLC-grade water. All chemicals were used as received.

3.2. Photoreactor Setup

Figure 16 depicts the photoreactor, which was a cylindrical water-jacketed glass vessel (318 mL) with 102 mm and 63 mm of interior height and diameter, respectively. The horizontal and vertical position of the photoreactor was constant for all experiments. Lamps were set horizontally and centered above the photoreactor. Two 15 W GE F15T8 BLB lamps (also called black-light lamps, Boston, MA, USA) supplied high UV irradiation, and two 15 W GE F15T8 D lamps (also called daylight lamps) provided low UV irradiation. The overall system was in a closed box to avoid the effects of sunlight or any artificial radiation sources. Lamp emission spectra were measured using a lab-made spectrophotometer using a CMOS webcam with a diffraction grating of 1000 lines mm−1 [72,73]. Emission spectra calibration of the spectrophotometer was developed using a 9 W fluorescent lamp (Tecnolite, Jalisco, Mexico). The temperature of all experiments was set at 20 °C using a thermostatic bath with recirculation (Polystat, Cole-Palmer, Vernon Hills, IL, USA). An optical filter was not used in the experiments, so visible light condition was not simulated.

3.3. Synthesis of Materials

The synthesis method of iron-doped TiO2 (Fe-TiO2) materials followed the hydrothermal sol-gel synthetic approach proposed by Patra et al. with some differences in precursor and thermal treatment [49]. Our synthesis method used iron (III) nitrate instead of FeCl3 and absolute ethanol instead of isopropyl alcohol. The thermal treatment was a programmed cycle of 31 h (increasing ramp-drying-increasing ramp-calcination-decreasing ramp) instead of direct calcination for 6 h. First, solution A was prepared by dissolving 1.44 g of SDS in 10 mL of deionized water. Then, four different solutions B were prepared to dissolve iron (III) nitrate in 2 mL of absolute ethanol (≥99.8 %) and 3 mL of TTIP was added slowly. The amounts of iron (III) nitrate were 0, 0.4, 4.3, and 42.6 mg of Fe(NO3)3·9H2O identified as zero-iron TiO2, 0.3 Fe-TiO2, 0.6 Fe-TiO2, and 1.0 Fe-TiO2, respectively. Once ready, solution A was continuously stirred and solution B was slowly dropped into solution A. The pH of the resulting mixture was adjusted to 1 using concentrated HNO3 and stirred for 3 h. The mixture was kept at 3 °C for 36 h. The precipitated solid was collected by filtration using Whatman Quantitative Filter Paper Grade 42. The materials were simultaneously dried and calcinated with a programmed thermal treatment (Isotemp® Programmable Muffle Furnace, Fisher Scientific, Dubuque, IA, USA) following first the temperature increase from ambient temperature to 353 K, with a temperature ramp of 1 K min−1 that was held for 720 min. The temperature was then increased from 353 K to 773 K with a temperature ramp of 1 K min−1 that was held for 360 min. Finally, the temperature was decreased from 773 K to 353 K with a temperature ramp of −1 K min−1, and then the furnace was turned off. The materials were washed with 50:50 methanol-water and dried to 377 K overnight.

3.4. Materials Characterization

X-ray photoelectron spectroscopy (XPS) was performed using a Thermo Fisher Scientific K-Alpha X-ray photoelectron spectrometer (Waltham, MA, USA) with a monochromatized Al Kα X-ray source (1487 V). The deconvolution of high-resolution XPS spectra was developed using the software XPSpeak 4.1. (Raymund W.M. Kwok, Shatin, Hong Kong).
UV-visible reflectance spectroscopy was obtained with Video–Barrelino integrating sphere coupled to Cary 50 spectrophotometer (Varian Inc, Palo Alto, CA, USA). Diffuse reflectance spectra were transformed using the Kubelka–Munk method to obtain Eg of zero-iron TiO2 and Fe-TiO2 materials. Kubelka–Munk method plots (F(R)hv)1/2 versus hv, draws a tangent at the inflection point on the curve and estimates Eg with the hv value at the intersection with abscissa. In this case, F(R) is a reflectance function equal to (1 − R)2/2R, R is the reflectance percentage, h is the Planck’s constant, and v is frequency.
XRD patterns were recorded in a Siemens D-5000 diffractometer (Munich, Germany) using Cu Kα radiation (λ = 1.54060 Å) from 10° to 85°. The procedure for phase identification used the QualX2.0 software with database developed by Altomare et al. [74]. The cards used for identification were 00-901-5929, 00-900-1681, and 00-900-4140 for anatase, rutile, and brookite, respectively. The quantification phases followed the method proposed by Spurr and Myers according to Equation (13):
f = 1 1 + 1.26 I R I A
where f is the anatase percentage, IA is intensity at a diffraction angle 2θ of 25.36°, and IR is intensity at a diffraction angle 2θ of 27.46° [75].
The particle size was estimated by Scherrer’s formula described in Equation (14), where β is the full width at half of the maximum of the diffraction peaks (radians), k is the shape constant, λ is the wavelength of the incident Cu Kα radiation (λ = 1.54060 Å), θ is the Bragg’s angle (radians), and D is the particle size (Å).
D = k   λ β   cos θ
Brunauer–Emmett–Teller (BET) isotherms were obtained in Nova Station A equipment (Quantachrome Instruments, Boynton Beach, FL, USA). The surface morphology was observed by SEM in a JEOL ultrahigh resolution field emission electron microscope JSM-7800 F (JEOL, Tokyo, Japan) with 20 kV accelerating voltage, and 3 mm WD. Transmission electron microscopy (TEM) images were obtained in a JEM-2100 LaB6 electron microscope (JEOL, Tokyo, Japan).

3.5. Hydroxyl Radical Generation

In this study, pNDA bleaching was selected as an OH probe because pNDA was useful for measuring the photocatalytic performance of TiO2 [51,76,77] because of the following advantages: (1) it is selective of the reaction of pNDA with OH [78]; (2) its high reaction rate with OH on the order of 1010 M−1 s−1 [51,79]; (3) its easy application through observable bleaching at 440 nm following Beer’s Law, in which pNDA bleaching a yellowish solution to transparent; and (4) its 1:1 stoichiometry, meaning that one OH can bleach one pNDA molecule [51,80,81,82].
The pNDA absorption (Figure 17) measurements were obtained using a UV-visible spectrophotometer (Hatch DR/4000U, Loveland, CO, USA) at 440 nm following Beer-Lambert law. The pNDA test solution was 10 µM initial concentration and pH 6.0 ± 0.1 adjusted using NaOH or HCl when needed. No buffer solutions were used because they can compete for OH. Final pH was verified at the end of tests to discharge pH-pNDA bleaching.
The photocatalytic standard was Aeroxide® TiO2 P25, and the load was 20 mg L−1. The choice of catalyst load was based on our previous work on OH generation of Aeroxide® TiO2 P25 [16]. For zero-iron TiO2 and Fe-TiO2 materials, the catalyst load used was 320 mg L−1, which produced a OH generation rate under high UV irradiation to set a baseline. Catalyst load differences were attributable to the aggregation of lab-made TiO2, superficial properties, and optical properties of suspensions, as shown in Figure 18.
The photocatalytic experiments were conducted as follows. First, a pNDA test solution was set at 20 °C, the catalyst was added, and the suspension was mixed for 20 min without radiation. To evaluate the adsorption of pNDA on TiO2, an aliquot was withdrawn and centrifuged. Then, the system was fully illuminated, and aliquots were withdrawn after specific periods. Each sample was centrifuged at 6000 rpm for 15 min (Biofuge Primo, Sorvall, Hanau, Germany) and measured in the UV-visible spectrophotometer. Once the catalyst load was used and after the dark phase, no adsorption of pNDA was detected near the detection limit of UV-visible spectrophotometer.

3.6. Photolysis and Photocatalytic Degradation of E3

The initial E3 concentration was 10 μM because (1) this research was part of a project focused on the removal of E3 in water using sequentially coupled membrane filtration; (2) the solubility limit of E3 in water was previously reported to be 11.1 μM [83], and 45.1 μM [8,84], and (3) the sensitivity of the analytical techniques used in this work. The E3 solution was prepared to dissolve 2.88 mg of E3 in 1 L of deionized water by stirring at room conditions in the dark for six hours. Working solutions were stored in an amber flask.
Each photocatalytic experiment used 100 mL of E3 working solution. Initial pH was adjusted to obtain a similar surface charge of TiO2 [85]. Depending on the initial water conditions, the initial pH value was adjusted to 6.0 ± 0.1 using NaOH or HCl when needed. A dark period (no radiation) was allowed for 20 min. Then, similar experimental conditions were carried out as described in Section 3.5. Additionally, the aliquots withdrawn from suspension were filtered using a 0.1 μm syringe filter (MillexVV, Millipore, Billerica, MA, USA). A blank experiment without irradiation and TiO2 photocatalyst was conducted for comparison. The blank experiment showed that E3 cannot be degraded in absences of either TiO2 or UV light. Once the catalyst was loaded and after the dark phase, no adsorption of E3 was detected near the detection limit of HPLC.

3.7. Analytical Methods

The E3 concentration was monitored using an HPLC system (Waters 1515; Milford, MA, USA) equipped with a UV detector (Waters 2787) that has an injection volume of 20 μL. The analytical method was performed in isocratic analytical mode using an Inertsil® ODS-3 column (GL Science, Tokyo, Japan; 150 mm × 4.6 mm, 5 μm) thermostated at 25 °C. The wavelength was at 280 nm according to E3 maximum absorbance. The mobile phase was methanol (49%) and deionized water (51%) at a flow rate of 1 mL min−1. The retention time of E3 was 10 min, and the limit of E3 detection was 0.1 μM (0.029 mg L−1). The detection limit was obtained by developing two calibration curves: the first between 10 and 0.1 and second between 1 and 0.01. Both calibration curves followed area = 2928[E3] with R2 = 0.9899, but areas below 0.1 were not detected.

4. Conclusions

This study provided an understanding of the relationship between the Fe doping ratio and radiation intensity for OH generation and estriol (E3) degradation. The main results were that:
  • E3 degradation using 0.3 Fe-TiO2 was feasible and can be improved by controlling irradiation intensity which was found closely related with light absorption and the catalytic reaction rate;
  • the synthesis method and thermal treatment allowed nanoparticles with large superficial areas and the incorporation of iron ions into the TiO2 lattice.; and
  • changes in trapping recombination centers could be controlled with irradiation intensity to enhance the photocatalytic activity.
Therefore, our findings provide the opportunity to reconsider studies in which iron-doped TiO2 impaired photocatalytic activity and to improve an application in which irradiation should be controlled. For example, Fe-TiO2 can potentially be applied to medical uses in which low irradiation intensity should be used to avoid adverse effects in humans or wildlife, which has also been suggested by others [86]. In the field of water treatment, we propose that Fe-TiO2 is an efficient material that could harvest low-energy photons to degrade and mineralize dyes [87], biocides [88], pharmaceuticals [89], industrial chemicals [90], and estrogens—as shown in this study—to create an energetically green water treatment process.

Author Contributions

Funding acquisition, E.R.B.; Investigation, I.M.R.-S.; Project administration, E.R.B.; Supervision, E.R.B.; Writing—original draft, I.M.R.-S.; Writing—review & editing, E.R.B.

Funding

This manuscript is based on work supported in part by ConTex postdoctoral program, which is an initiative of the University of Texas System and Mexico’s National Council of Science and Technology (CONACYT). The research was partially funded by CONACYT under Project CB-2011/168285. The APC was funded by the Institutional Open Access Program (IAOP) between The University of Texas at Austin and Desert Research Institute (DRI) at Nevada.

Acknowledgments

The Aeroxide® P25 Evonik catalyst used for this work was provided by Intertrade S.A. de C.V., the supplier of Evonik Industries in Mexico. The authors thank L. Lartundo-Rojas, Raul Borja Urbi, Hugo Martinez Gutiérrez, and Joao Jairzinho Salinas Camargo for assistance in XPS spectroscopy, TEM images, SEM images, and absorption isotherms, respectively, all of whom are from Centro de Nanociencias y Micro y Nanotecnología (CNMN) of IPN, Mexico. The authors thank M.A. Quiroz Alfaro for his excellent technical help and for his permission to use materials and equipment at the UDLAP’s electrochemical lab. The authors also thank Nicole Damon (DRI) for her editorial review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dulio, V.; van Bavel, B.; Brorström-Lundén, E.; Harmsen, J.; Hollender, J.; Schlabach, M.; Slobodnik, J.; Thomas, K.; Koschorreck, J. Emerging pollutants in the EU: 10 years of NORMAN in support of environmental policies and regulations. Environ. Sci. Eur. 2018, 30, 5. [Google Scholar] [CrossRef]
  2. Mandaric, L.; Celic, M.; Marcé, R.; Petrovic, M. Introduction on Emerging Contaminants in Rivers and Their Environmental Risk. In Emerging Contaminants in River Ecosystems: Occurrence and Effects under Multiple Stress Conditions; Petrovic, M., Sabater, S., Elosegi, A., Barceló, D., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 3–25. ISBN 978-3-319-29376-9. [Google Scholar]
  3. Houtman, C.J.; Legler, J.; Thomas, K. Effect-Directed Analysis of Complex Environmental Contamination; Brack, W., Ed.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 237–265. ISBN 978-3-642-18384-3. [Google Scholar]
  4. Dimogerontas, G.; Liapi, C. Endocrine Disruptors (Xenoestrogens): An Overview. In Plastics in Dentistry and Estrogenicity: A Guide to Safe Practice; Eliades, T., Eliades, G., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 3–48. ISBN 978-3-642-29687-1. [Google Scholar]
  5. Hileman, B. Environmental Estrogens linked to Reproductive Abnormalities, Cancer. Chem. Eng. News Arch. 1994, 72, 19–23. [Google Scholar] [CrossRef]
  6. Prat, N.; Rieradevall, M.; Barata, C.; Munné, A. The combined use of metrics of biological quality and biomarkers to detect the effects of reclaimed water on macroinvertebrate assemblages in the lower part of a polluted Mediterranean river (Llobregat River, NE Spain). Ecol. Indic. 2013, 24, 167–176. [Google Scholar] [CrossRef]
  7. Souza, M.S.; Hallgren, P.; Balseiro, E.; Hansson, L.A. Low concentrations, potential ecological consequences: Synthetic estrogens alter life-history and demographic structures of aquatic invertebrates. Environ. Pollut. 2013, 178, 237–243. [Google Scholar] [CrossRef]
  8. Silva, C.P.; Otero, M.; Esteves, V. Processes for the elimination of estrogenic steroid hormones from water: A review. Environ. Pollut. 2012, 165, 38–58. [Google Scholar] [CrossRef]
  9. Rodriguez-Narvaez, O.M.; Peralta-Hernandez, J.M.; Goonetilleke, A.; Bandala, E.R. Treatment technologies for emerging contaminants in water: A review. Chem. Eng. J. 2017, 323, 361–380. [Google Scholar] [CrossRef]
  10. Gągol, M.; Przyjazny, A.; Boczkaj, G. Wastewater treatment by means of advanced oxidation processes based on cavitation—A review. Chem. Eng. J. 2018, 338, 599–627. [Google Scholar] [CrossRef]
  11. Boczkaj, G.; Fernandes, A. Wastewater treatment by means of advanced oxidation processes at basic pH conditions: A review. Chem. Eng. J. 2017, 320, 608–633. [Google Scholar] [CrossRef]
  12. Coleman; Eggins, B.; Byrne, J.A.; Palmer, F.L.; King, E. Photocatalytic degradation of 17-β-oestradiol on immobilised TiO2. Appl. Catal. B Environ. 2000, 24, L1–L5. [Google Scholar] [CrossRef]
  13. Ramirez-Sanchez, I.M.; Mendez-Rojas, M.A.; Bandala, E.R. CHAPTER 25 Photocatalytic Degradation of Natural and Synthetic Estrogens with Semiconducting Nanoparticles. In Advanced Environmental Analysis: Applications of Nanomaterials; The Royal Society of Chemistry: London, UK, 2017; Volume 2, pp. 153–177. ISBN 978-1-78262-906-1. [Google Scholar]
  14. Ohko, Y.; Iuchi, K.; Niwa, C.; Tatsuma, T.; Nakashima, T.; Iguchi, T.; Kubota, Y.; Fujishima, A. 17β-Estradiol Degradation by TiO2 Photocatalysis as a Means of Reducing Estrogenic Activity. Environ. Sci. Technol. 2002, 36, 4175–4181. [Google Scholar] [CrossRef]
  15. Coleman, H.M.; Routledge, E.J.; Sumpter, J.P.; Eggins, B.R.; Byrne, J.A. Rapid loss of estrogenicity of steroid estrogens by UVA photolysis and photocatalysis over an immobilised titanium dioxide catalyst. Water Res. 2004, 38, 3233–3240. [Google Scholar] [CrossRef]
  16. Ramírez-Sánchez, I.M.; Tuberty, S.; Hambourger, M.; Bandala, E.R. Resource efficiency analysis for photocatalytic degradation and mineralization of estriol using TiO2 nanoparticles. Chemosphere 2017, 184, 1270–1285. [Google Scholar] [CrossRef]
  17. Hashimoto, K.; Irie, H.; Fujishima, A. Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269–8285. [Google Scholar] [CrossRef]
  18. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  19. Tong, A.Y.C.; Braund, R.; Warren, D.S.; Peake, B.M. TiO2-assisted photodegradation of pharmaceuticals—A review. Cent. Eur. J. Chem. 2012, 10, 989–1027. [Google Scholar] [CrossRef]
  20. Cassaignon, S.; Colbeau-Justin, C.; Durupthy, O. Titanium dioxide in photocatalysis. In Nanomaterials: A Danger or a Promise?: A Chemical and Biological Perspective; Springer: London, UK, 2013; pp. 153–188. ISBN 9781447142133. [Google Scholar]
  21. Augugliaro, V.; Loddo, V.; Pagliaro, M.; Palmisano, G.; Palmisano, L. Clean by Light Irradiation: Practical Applications of Supported TiO2; RSC Publishing: Cambridge, UK, 2010; ISBN 1847558704. [Google Scholar]
  22. Etacheri, V.; Di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1–29. [Google Scholar] [CrossRef]
  23. Wen, L.; Liu, B.; Zhao, X.; Nakata, K.; Murakami, T.; Fujishima, A. Synthesis, Characterization, and Photocatalysis of Fe-Doped TiO2: A Combined Experimental and Theoretical Study. Int. J. Photoenergy 2012, 2012, 1–10. [Google Scholar] [CrossRef]
  24. Yu, H.; Irie, H.; Hashimoto, K. Conduction band energy level control of titanium dioxide: Toward an efficient visible-light-sensitive photocatalyst. J. Am. Chem. Soc. 2010, 132, 6898–6899. [Google Scholar] [CrossRef]
  25. Choi, W.; Termin, A.; Hoffmann, M.R. The role of metal ion dopants in quantum-sized TiO2: Correlation between photoreactivity and charge carrier recombination dynamics. J. Phys. Chem. 1994, 98, 13669–13679. [Google Scholar] [CrossRef]
  26. Kaur, T.; Sraw, A.; Wanchoo, R.K.; Toor, A.P. Visible–Light Induced Photocatalytic Degradation of Fungicide with Fe and Si Doped TiO2 Nanoparticles. Mater. Today Proc. 2016, 3, 354–361. [Google Scholar] [CrossRef]
  27. Zhao, B.; Mele, G.; Pio, I.; Li, J.; Palmisano, L.; Vasapollo, G. Degradation of 4-nitrophenol (4-NP) using Fe-TiO2 as a heterogeneous photo-Fenton catalyst. J. Hazard. Mater. 2010, 176, 569–574. [Google Scholar] [CrossRef]
  28. Yalçın, Y.; Kılıç, M.; Çınar, Z. Fe+3-doped TiO2: A combined experimental and computational approach to the evaluation of visible light activity. Appl. Catal. B Environ. 2010, 99, 469–477. [Google Scholar] [CrossRef]
  29. Cai, L.; Liao, X.; Shi, B. Using Collagen Fiber as a Template to Synthesize TiO2 and Fex/TiO2 Nanofibers and Their Catalytic Behaviors on the Visible Light-Assisted Degradation of Orange II. Ind. Eng. Chem. Res. 2010, 49, 3194–3199. [Google Scholar] [CrossRef]
  30. Li, J.; Xu, J.; Dai, W.L.; Li, H.; Fan, K. Direct hydro-alcohol thermal synthesis of special core-shell structured Fe-doped titania microspheres with extended visible light response and enhanced photoactivity. Appl. Catal. B Environ. 2009, 85, 162–170. [Google Scholar] [CrossRef]
  31. Tong, T.; Zhang, J.; Tian, B.; Chen, F.; He, D. Preparation of Fe3+-doped TiO2 catalysts by controlled hydrolysis of titanium alkoxide and study on their photocatalytic activity for methyl orange degradation. J. Hazard. Mater. 2008, 155, 572–579. [Google Scholar] [CrossRef]
  32. Ambrus, Z.; Balázs, N.; Alapi, T.; Wittmann, G.; Sipos, P.; Dombi, A.; Mogyorósi, K. Synthesis, structure and photocatalytic properties of Fe(III)-doped TiO2 prepared from TiCl3. Appl. Catal. B Environ. 2008, 81, 27–37. [Google Scholar] [CrossRef]
  33. Cong, Y.; Zhang, J.; Chen, F.; Anpo, M.; He, D. Preparation, photocatalytic activity, and mechanism of nano-TiO2 Co-doped with nitrogen and iron (III). J. Phys. Chem. C 2007, 111, 10618–10623. [Google Scholar] [CrossRef]
  34. Adán, C.; Bahamonde, A.; Fernández-García, M.; Martínez-Arias, A. Structure and activity of nanosized iron-doped anatase TiO2 catalysts for phenol photocatalytic degradation. Appl. Catal. B Environ. 2007, 72, 11–17. [Google Scholar] [CrossRef]
  35. Yamashita, H.; Harada, M.; Misaka, J.; Takeuchi, M.; Neppolian, B.; Anpo, M. Photocatalytic degradation of organic compounds diluted in water using visible light-responsive metal ion-implanted TiO2 catalysts: Fe ion-implanted TiO2. Catal. Today 2003, 84, 191–196. [Google Scholar] [CrossRef]
  36. Li, X.; Yue, P.-L.; Kutal, C. Synthesis and photocatalytic oxidation properties of iron doped titanium dioxide nanosemiconductor particles. New J. Chem. 2003, 27, 1264. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Wang, C.-C.; Zakaria, R.; Ying, J.Y. Role of Particle Size in Nanocrystalline TiO2-Based Photocatalysts. J. Phys. Chem. B 1998, 102, 10871–10878. [Google Scholar] [CrossRef]
  38. Litter, M.I.; Navío, J.A. Photocatalytic properties of iron-doped titania semiconductors. J. Photochem. Photobiol. A Chem. 1996, 98, 171–181. [Google Scholar] [CrossRef]
  39. Fàbrega, C.; Andreu, T.; Cabot, A.; Morante, J.R. Location and catalytic role of iron species in TiO2:Fe photocatalysts: An EPR study. J. Photochem. Photobiol. A Chem. 2010, 211, 170–175. [Google Scholar] [CrossRef]
  40. Seabra, M.P.; Salvado, I.M.M.; Labrincha, J.A. Pure and (zinc or iron) doped titania powders prepared by sol-gel and used as photocatalyst. Ceram. Int. 2011, 37, 3317–3322. [Google Scholar] [CrossRef]
  41. Abazović, N.D.; Mirenghi, L.; Janković, I.A.; Bibić, N.; Šojić, D.V.; Abramović, B.F.; Čomor, M.I. Synthesis and characterization of rutile TiO2 nanopowders doped with iron ions. Nanoscale Res. Lett. 2009, 4, 518–525. [Google Scholar] [CrossRef]
  42. Geissen, V.; Mol, H.; Klumpp, E.; Umlauf, G.; Nadal, M.; van der Ploeg, M.; van de Zee, S.E.A.T.M.; Ritsema, C.J. Emerging pollutants in the environment: A challenge for water resource management. Int. Soil Water Conserv. Res. 2015, 3, 57–65. [Google Scholar] [CrossRef] [Green Version]
  43. Lin, T.C.; Seshadri, G.; Kelber, J.A. A consistent method for quantitative XPS peak analysis of thin oxide films on clean polycrystalline iron surfaces. Appl. Surf. Sci. 1997, 119, 83–92. [Google Scholar] [CrossRef]
  44. Xing, M.; Wu, Y.; Zhang, J.; Chen, F. Effect of synergy on the visible light activity of B, N and Fe co-doped TiO2 for the degradation of MO. Nanoscale 2010, 2, 1233. [Google Scholar] [CrossRef] [PubMed]
  45. Lopez, R.; Gomez, R. Band-gap energy estimation from diffuse reflectance measurements on sol-gel and commercial TiO2: A comparative study. J. Sol-Gel Sci. Technol. 2012, 61, 1–7. [Google Scholar] [CrossRef]
  46. Shi, J.; Chen, G.; Zeng, G.; Chen, A.; He, K.; Huang, Z.; Hu, L.; Zeng, J.; Wu, J.; Liu, W. Hydrothermal synthesis of graphene wrapped Fe-doped TiO2 nanospheres with high photocatalysis performance. Ceram. Int. 2018, 44, 7473–7480. [Google Scholar] [CrossRef]
  47. Yu, J.; Xiang, Q.; Zhou, M. Preparation, characterization and visible-light-driven photocatalytic activity of Fe-doped titania nanorods and first-principles study for electronic structures. Appl. Catal. B Environ. 2009, 90, 595–602. [Google Scholar] [CrossRef]
  48. Goswami, P.; Ganguli, J.N. Evaluating the potential of a new titania precursor for the synthesis of mesoporous Fe-doped titania with enhanced photocatalytic activity. Mater. Res. Bull. 2012, 47, 2077–2084. [Google Scholar] [CrossRef]
  49. Patra, A.K.; Dutta, A.; Bhaumik, A. Highly ordered mesoporous TiO2-Fe2O3 mixed oxide synthesized by sol-gel pathway: An efficient and reusable heterogeneous catalyst for dehalogenation reaction. ACS Appl. Mater. Interfaces 2012, 4, 5022–5028. [Google Scholar] [CrossRef] [PubMed]
  50. 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, 4043. [Google Scholar] [CrossRef] [PubMed]
  51. Zang, L.; Qu, P.; Zhao, J.; Shen, T.; Hidaka, H. Photocatalytic bleaching of p-nitrosodimethylaniline in TiO2 aqueous suspensions: A kinetic treatment involving some primary events photoinduced on the particle surface. J. Mol. Catal. A Chem. 1997, 120, 235–245. [Google Scholar] [CrossRef]
  52. Othman, S.H.; Abdul Rashid, S.; Mohd Ghazi, T.I.; Abdullah, N. Fe-Doped TiO2 Nanoparticles Produced via MOCVD: Synthesis, Characterization, and Photocatalytic Activity. J. Nanomater. 2011, 2011, 1–8. [Google Scholar] [CrossRef]
  53. Teoh, W.Y.; Amal, R.; Mädler, L.; Pratsinis, S.E. Flame sprayed visible light-active Fe-TiO2 for photomineralisation of oxalic acid. Catal. Today 2007, 120, 203–213. [Google Scholar] [CrossRef]
  54. Pongwan, P.; Inceesungvorn, B.; Wetchakun, K.; Phanichphant, S.; Wetchakun, N. Highly efficient visible-light-induced photocatalytic activity of Fe-doped TiO2 nanoparticles. Eng. J. 2012, 16, 143–151. [Google Scholar] [CrossRef]
  55. Kruk, M.; Jaroniec, M. Gas adsorption characterization of ordered organic-inorganic nanocomposite materials. Chem. Mater. 2001, 13, 3169–3183. [Google Scholar] [CrossRef]
  56. Limousin, G.; Gaudet, J.P.; Charlet, L.; Szenknect, S.; Barthès, V.; Krimissa, M. Sorption isotherms: A review on physical bases, modeling and measurement. Appl. Geochem. 2007, 22, 249–275. [Google Scholar] [CrossRef]
  57. Carvalho, T.C.; La Cruz, T.E.; Tábora, J.E. A photochemical kinetic model for solid dosage forms. Eur. J. Pharm. Biopharm. 2017, 120, 63–72. [Google Scholar] [CrossRef] [PubMed]
  58. Daugherty, J.P.; Hixon, S.C.; Yielding, K.L. Direct in vitro photoaffinity labeling of DNA with daunorubicin, adriamycin, and rubidazone. BBA Sect. Nucleic Acids Protein Synth. 1979, 565, 13–21. [Google Scholar] [CrossRef]
  59. Hartman, P.E.; Biggley, W.H. Breakthrough of ultraviolet light from various brands of fluorescent lamps: Lethal effects on DNA repair-defective bacteria. Environ. Mol. Mutagen. 1996, 27, 306–313. [Google Scholar] [CrossRef]
  60. Serpone, N. Relative photonic efficiencies and quantum yields in heterogeneous photocatalysis. J. Photochem. Photobiol. A Chem. 1997, 104, 1–12. [Google Scholar] [CrossRef]
  61. Muff, J.; Bennedsen, L.R.; Søgaard, E.G. Study of electrochemical bleaching of p-nitrosodimethylaniline and its role as hydroxyl radical probe compound. J. Appl. Electrochem. 2011, 41, 599–607. [Google Scholar] [CrossRef] [Green Version]
  62. Zhu, J.; Zheng, W.; He, B.; Zhang, J.; Anpo, M. Characterization of Fe-TiO2 photocatalysts synthesized by hydrothermal method and their photocatalytic reactivity for photodegradation of XRG dye diluted in water. J. Mol. Catal. A Chem. 2004, 216, 35–43. [Google Scholar] [CrossRef]
  63. Neubert, S.; Mitoraj, D.; Shevlin, S.A.; Pulisova, P.; Heimann, M.; Du, Y.; Goh, G.K.L.; Pacia, M.; Kruczała, K.; Turner, S.; et al. Highly efficient rutile TiO2 photocatalysts with single Cu(II) and Fe(III) surface catalytic sites. J. Mater. Chem. A 2016. [Google Scholar] [CrossRef]
  64. Zhou, M.; Yu, J.; Cheng, B. Effects of Fe-doping on the photocatalytic activity of mesoporous TiO2 powders prepared by an ultrasonic method. J. Hazard. Mater. 2006, 137, 1838–1847. [Google Scholar] [CrossRef]
  65. Coleman, H.M.; Vimonses, V.; Leslie, G.; Amal, R. Removal of contaminants of concern in water using advance oxidation techniques. Water Sci. Technol. 2007, 55, 301–306. [Google Scholar] [CrossRef]
  66. Coleman, H.M.; Abdullah, M.I.; Eggins, B.R.; Palmer, F.L. Photocatalytic degradation of 17[beta]-oestradiol, oestriol and 17[alfa]-ethynyloestradiol in water monitored using fluorescence spectroscopy. Appl. Catal. B Environ. 2005, 55, 23–30. [Google Scholar] [CrossRef]
  67. Coleman, H.M.; Chiang, K.; Amal, R. Effects of Ag and Pt on photocatalytic degradation of endocrine disrupting chemicals in water. Chem. Eng. J. 2005, 113, 65–72. [Google Scholar] [CrossRef]
  68. Lin, L.; Wang, H.; Jiang, W.; Mkaouar, A.R.; Xu, P. Comparison study on photocatalytic oxidation of pharmaceuticals by TiO2-Fe and TiO2-reduced graphene oxide nanocomposites immobilized on optical fibers. J. Hazard. Mater. 2017, 333, 162–168. [Google Scholar] [CrossRef] [PubMed]
  69. Hamadanian, M.; Reisi-Vanani, A.; Behpour, M.; Esmaeily, A.S. Synthesis and characterization of Fe,S-codoped TiO2 nanoparticles: Application in degradation of organic water pollutants. Desalination 2011, 281, 319–324. [Google Scholar] [CrossRef]
  70. Naik, B.; Parida, K.M. Solar Light Active Photodegradation of Phenol over a FexTi1-xO2-yNy Nanophotocatalyst. Ind. Eng. Chem. Res. 2010, 49, 8339–8346. [Google Scholar] [CrossRef]
  71. Bloh, J.Z.; Dillert, R.; Bahnemann, D.W. Zinc Oxide Photocatalysis: Influence of Iron and Titanium Doping and Origin of the Optimal Doping Ratio. ChemCatChem 2013, 5, 774–778. [Google Scholar] [CrossRef]
  72. Lorenz, R.D. A simple webcam spectrograph. Am. J. Phys. 2014, 82, 169–173. [Google Scholar] [CrossRef]
  73. Widiatmoko, E.; Widayani; Budiman, M.; Abdullah, M.; Khairurrijal. A simple spectrophotometer using common materials and a digital camera. Phys. Educ. 2011, 46, 332–339. [Google Scholar] [CrossRef]
  74. Altomare, A.; Corriero, N.; Cuocci, C.; Falcicchio, A.; Moliterni, A.; Rizzi, R. QUALX2.0: A qualitative phase analysis software using the freely available database POW_COD. J. Appl. Crystallogr. 2015, 48, 598–603. [Google Scholar] [CrossRef]
  75. Spurr, R.A.; Myers, H. Quantitative Analysis of Anatase-Rutile Mixtures with an X-ray Diffractometer. Anal. Chem. 1957, 29, 760–762. [Google Scholar] [CrossRef]
  76. Kim, C.; Park, H.J.; Cha, S.; Yoon, J. Facile detection of photogenerated reactive oxygen species in TiO2 nanoparticles suspension using colorimetric probe-assisted spectrometric method. Chemosphere 2013, 93, 2011–2015. [Google Scholar] [CrossRef]
  77. Simonsen, M.E.; Muff, J.; Bennedsen, L.R.; Kowalski, K.P.; Søgaard, E.G. Photocatalytic bleaching of p-nitrosodimethylaniline and a comparison to the performance of other AOP technologies. J. Photochem. Photobiol. A Chem. 2010, 216, 244–249. [Google Scholar] [CrossRef] [Green Version]
  78. Kraljic, I.; Trumbore, C.N. p-Nitrosodimethylaniline as an OH radical scavenger in radiation chemistry. J. Am. Chem. Soc. 1965, 87, 2547–2550. [Google Scholar] [CrossRef]
  79. Farhataziz, A.B.R. Selected Specific Rates of Reactions of Transients from Water in Aqueous Solutions III: Hydroxyl Radical and Perhydroxyl Radical and Their Radical Ions; U.S. Department of Commerce: Washington, DC, USA, 1977.
  80. Martínez-Huitle, C.A.; Quiroz, M.A.; Comninellis, C.; Ferro, S.; De Battisti, A. Electrochemical incineration of chloranilic acid using Ti/IrO2, Pb/PbO2 and Si/BDD electrodes. Electrochim. Acta 2004, 50, 949–956. [Google Scholar] [CrossRef]
  81. Bors, W.; Michel, C.; Saran, M. On the nature of biochemically generated hydroxyl radicals. Studies using the bleaching of p-nitrosodimethylaniline as a direct assay method. Eur. J. Biochem. 1979, 95, 621–627. [Google Scholar] [CrossRef] [PubMed]
  82. Barashkov, N.N.; Eisenberg, D.; Eisenberg, S.; Shegebaeva, G.S.; Irgibaeva, I.S.; Barashkova, I.I. Electrochemical chlorine-free AC disinfection of water contaminated with Salmonella typhimurium bacteria. Russ. J. Electrochem. 2010, 46, 306–311. [Google Scholar] [CrossRef]
  83. Hurwitz, A.R.; Liu, S.T. Determination of aqueous solubility and pKa values of estrogens. J. Pharm. Sci. 1977, 66, 624–627. [Google Scholar] [CrossRef]
  84. Ying, G.G.; Kookana, R.S.; Ru, Y.J. Occurrence and fate of hormone steroids in the environment. Environ. Int. 2002, 28, 545–551. [Google Scholar] [CrossRef]
  85. Fernández-Ibáñez, P.; De Las Nieves, F.J.; Malato, S. Titanium Dioxide/Electrolyte Solution Interface: Electron Transfer Phenomena. J. Colloid Interface Sci. 2000, 227, 510–516. [Google Scholar] [CrossRef]
  86. George, S.; Pokhrel, S.; Ji, Z.; Henderson, B.L.; Xia, T.; Li, L.; Zink, J.I.; Nel, A.E.; Mädler, L. Role of Fe doping in tuning the band gap of TiO2 for the photo-oxidation-induced cytotoxicity paradigm. J. Am. Chem. Soc. 2011, 133, 11270–11278. [Google Scholar] [CrossRef]
  87. Bhatu, M.N.; Lavand, A.B.; Malghe, Y.S. Visible light photocatalytic degradation of malachite green using modified titania. J. Mater. Res. Technol. 2018. [Google Scholar] [CrossRef]
  88. Tabasideh, S.; Maleki, A.; Shahmoradi, B.; Ghahremani, E.; McKay, G. Sonophotocatalytic degradation of diazinon in aqueous solution using iron-doped TiO2 nanoparticles. Sep. Purif. Technol. 2017, 189, 186–192. [Google Scholar] [CrossRef]
  89. Aba-Guevara, C.G.; Medina-Ramírez, I.E.; Hernández-Ramírez, A.; Jáuregui-Rincón, J.; Lozano-Álvarez, J.A.; Rodríguez-López, J.L. Comparison of two synthesis methods on the preparation of Fe, N-Co-doped TiO2 materials for degradation of pharmaceutical compounds under visible light. Ceram. Int. 2017, 43, 5068–5079. [Google Scholar] [CrossRef]
  90. Hemmati Borji, S.; Nasseri, S.; Mahvi, A.; Nabizadeh, R.; Javadi, A. Investigation of photocatalytic degradation of phenol by Fe(III)-doped TiO2 and TiO2 nanoparticles. J. Environ. Health Sci. Eng. 2014, 12, 101. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Photocatalytic mechanism of TiO2 for OH generation. Where Eg: Band gap energy; E: photon energy; OMads: adsorbed organic molecule; and OMoxi: oxidized organic molecule.
Figure 1. Photocatalytic mechanism of TiO2 for OH generation. Where Eg: Band gap energy; E: photon energy; OMads: adsorbed organic molecule; and OMoxi: oxidized organic molecule.
Catalysts 08 00625 g001
Figure 2. X-ray Photoelectron Spectroscopy (XPS) general spectra for zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Figure 2. X-ray Photoelectron Spectroscopy (XPS) general spectra for zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Catalysts 08 00625 g002
Figure 3. High-resolution XPS spectra for the iron region for 1.0 Fe-TiO2.
Figure 3. High-resolution XPS spectra for the iron region for 1.0 Fe-TiO2.
Catalysts 08 00625 g003
Figure 4. Band gap energy (Eg) by the Kubelka–Monk method. Zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Figure 4. Band gap energy (Eg) by the Kubelka–Monk method. Zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Catalysts 08 00625 g004
Figure 5. XRD patterns for zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d), where A is Anatase and R is Rutile phases.
Figure 5. XRD patterns for zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d), where A is Anatase and R is Rutile phases.
Catalysts 08 00625 g005
Figure 6. Brunauer–Emmett–Teller (BET) isotherms for zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Figure 6. Brunauer–Emmett–Teller (BET) isotherms for zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Catalysts 08 00625 g006
Figure 7. SEM image of zero-iron TiO2 after mechanical grinding and sonication.
Figure 7. SEM image of zero-iron TiO2 after mechanical grinding and sonication.
Catalysts 08 00625 g007
Figure 8. Transmission electron microscopy (TEM) image of 0.3 Fe-TiO2 (a) and zero-iron TiO2 (b).
Figure 8. Transmission electron microscopy (TEM) image of 0.3 Fe-TiO2 (a) and zero-iron TiO2 (b).
Catalysts 08 00625 g008
Figure 9. Emission spectrum and intensity graph of the irradiation source of Tecnolite fluorescent lamp (a), GE F15T8 BLB lamp (b), and GE F15T8 D lamp (c).
Figure 9. Emission spectrum and intensity graph of the irradiation source of Tecnolite fluorescent lamp (a), GE F15T8 BLB lamp (b), and GE F15T8 D lamp (c).
Catalysts 08 00625 g009
Figure 10. OH generation (initial velocity) of zero-iron TiO2 and Fe-TiO2 under high UV irradiation (a) and low UV Irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and 1.0 Fe-TiO2 Catalysts 08 00625 i004 at pH 6 ± 0.1, and 20 °C.
Figure 10. OH generation (initial velocity) of zero-iron TiO2 and Fe-TiO2 under high UV irradiation (a) and low UV Irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and 1.0 Fe-TiO2 Catalysts 08 00625 i004 at pH 6 ± 0.1, and 20 °C.
Catalysts 08 00625 g010
Figure 11. Photocatalytic mechanism of Fe-TiO2 and OH generation. Eg is band gap energy, E is photon energy, OMads is adsorbed organic molecule, OMoxi is oxidized organic molecule.
Figure 11. Photocatalytic mechanism of Fe-TiO2 and OH generation. Eg is band gap energy, E is photon energy, OMads is adsorbed organic molecule, OMoxi is oxidized organic molecule.
Catalysts 08 00625 g011
Figure 12. Photocatalytic degradation of E3 under high UV irradiation (a), and low UV irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, Catalysts 08 00625 i004 1.0 Fe-TiO2, and Catalysts 08 00625 i005 Aeroxide® TiO2 P25; at pH 6 ± 0.1, and 20 °C.
Figure 12. Photocatalytic degradation of E3 under high UV irradiation (a), and low UV irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, Catalysts 08 00625 i004 1.0 Fe-TiO2, and Catalysts 08 00625 i005 Aeroxide® TiO2 P25; at pH 6 ± 0.1, and 20 °C.
Catalysts 08 00625 g012
Figure 13. Photocatalytic reaction rate (k1) for degradation of E3 under high UV irradiation (a), and low UV irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and Catalysts 08 00625 i004 1.0 Fe-TiO2; at pH 6 ± 0.1, and 20 °C.
Figure 13. Photocatalytic reaction rate (k1) for degradation of E3 under high UV irradiation (a), and low UV irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and Catalysts 08 00625 i004 1.0 Fe-TiO2; at pH 6 ± 0.1, and 20 °C.
Catalysts 08 00625 g013
Figure 14. Correlation between OH initial rate generation (r0,OH) and initial E3 degradation (r0,E3) under high UV irradiation (a), and low UV irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and Catalysts 08 00625 i004 1.0 Fe-TiO2; at pH 6 ± 0.1, and 20 °C.
Figure 14. Correlation between OH initial rate generation (r0,OH) and initial E3 degradation (r0,E3) under high UV irradiation (a), and low UV irradiation (b); where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and Catalysts 08 00625 i004 1.0 Fe-TiO2; at pH 6 ± 0.1, and 20 °C.
Catalysts 08 00625 g014
Figure 15. Experimental relationship between pseudo first order constant and at.% content; where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and Catalysts 08 00625 i004 1.0 Fe-TiO2; at pH 6 ± 0.1; and 20 °C.
Figure 15. Experimental relationship between pseudo first order constant and at.% content; where Catalysts 08 00625 i001 zero-iron TiO2, Catalysts 08 00625 i002 0.3 Fe-TiO2, Catalysts 08 00625 i003 0.6 Fe-TiO2, and Catalysts 08 00625 i004 1.0 Fe-TiO2; at pH 6 ± 0.1; and 20 °C.
Catalysts 08 00625 g015
Figure 16. Scheme of photoreactor used for experiments: glass reactor (1), testing solution (2), temperature probe (3), spin bar (4), lamps (5), an optical filter (if needed) (6), stirring plate (7), cooling fan (8), horizontal position template (9), and lab jack lifting platform (10).
Figure 16. Scheme of photoreactor used for experiments: glass reactor (1), testing solution (2), temperature probe (3), spin bar (4), lamps (5), an optical filter (if needed) (6), stirring plate (7), cooling fan (8), horizontal position template (9), and lab jack lifting platform (10).
Catalysts 08 00625 g016
Figure 17. Structural formula and absorbance spectrum of N,N-dimethyl-p-nitrosoaniline (pNDA).
Figure 17. Structural formula and absorbance spectrum of N,N-dimethyl-p-nitrosoaniline (pNDA).
Catalysts 08 00625 g017
Figure 18. Suspension transmittance of Fe-TiO2 material and Aeroxide TiO2 P25; where zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Figure 18. Suspension transmittance of Fe-TiO2 material and Aeroxide TiO2 P25; where zero-iron TiO2 (a), 0.3 Fe-TiO2 (b), 0.6 Fe-TiO2 (c), and 1.0 Fe-TiO2 (d).
Catalysts 08 00625 g018
Table 1. Surface elemental composition determined by XPS.
Table 1. Surface elemental composition determined by XPS.
MaterialAtomic % of Elements (at.%)
Ti2pO1sC1sFe2pS2pN1s
Zero-iron TiO224.452.921.301.4-
0.3 Fe-TiO223.851.122.90.31.10.8
0.6 Fe-TiO223.953.122.50.6--
1.0 Fe-TiO223.552.520.611.50.9
Table 2. Structural and optical properties of zero-iron TiO2, and Fe-TiO2.
Table 2. Structural and optical properties of zero-iron TiO2, and Fe-TiO2.
MaterialEgAnatase: RutileParticle SizeSurface AreaPore SizeHigh UVLow UV
eVnm%nmm2 g−1nm%%
Aeroxide® TiO2 P253.2 *387.5 *80:20 *21 *50 ± 15 *17.5 *36.40.8
Zero-iron TiO22.98416.173.1:26.96.666.58.499.267.64
0.3 Fe-TiO22.96418.977.9:21.16.977.61.299.408.21
0.6 Fe-TiO22.95420.378.8:21.27.173.01.499.428.77
1.0 Fe-TiO22.90427.676.3:23.76.983.19.499.4310.63
* According to the manufacturer.
Table 3. OH generation rate of zero-iron TiO2 and Fe-TiO2.
Table 3. OH generation rate of zero-iron TiO2 and Fe-TiO2.
Catalystat.%LoadHigh UV IrradiationLow UV Irradiation
k1R2r0k1R2r0
mg L−1min−1 μM•OH min−1min−1 μM•OH min−1
TiO2 Aeroxide® P25-200.060.9880.490.0120.9890.105
Zero-iron TiO203200.0560.9930.490.0050.9730.045
0.3 Fe-TiO20.33200.0670.9980.580.0040.9900.042
0.6 Fe-TiO20.63200.0310.9980.280.0020.9990.025
1.0 Fe-TiO213200.0040.9870.040.000020.8910.0002
Table 4. Kinetic values of E3 degradation using zero-iron TiO2 and Fe-TiO2.
Table 4. Kinetic values of E3 degradation using zero-iron TiO2 and Fe-TiO2.
CatalystLoadHigh UV IrradiationLow UV Irradiation
k1R2r0,E3k1R2r0,E3
mg L−1min−1 μME3 min−1min−1 μME3 min−1
TiO2 Aeroxide® P25200.0210.9960.210.00290.9920.030
Zero-iron TiO23200.0070.9970.0690.00450.9910.040
0.3 Fe-TiO23200.0090.9940.0900.00500.9920.042
0.6 Fe-TiO23200.0110.9970.0990.00340.9990.030
1.0 Fe-TiO23200.0030.9790.0270.00160.9870.012

Share and Cite

MDPI and ACS Style

Ramírez-Sánchez, I.M.; Bandala, E.R. Photocatalytic Degradation of Estriol Using Iron-Doped TiO2 under High and Low UV Irradiation. Catalysts 2018, 8, 625. https://doi.org/10.3390/catal8120625

AMA Style

Ramírez-Sánchez IM, Bandala ER. Photocatalytic Degradation of Estriol Using Iron-Doped TiO2 under High and Low UV Irradiation. Catalysts. 2018; 8(12):625. https://doi.org/10.3390/catal8120625

Chicago/Turabian Style

Ramírez-Sánchez, Irwing M., and Erick R. Bandala. 2018. "Photocatalytic Degradation of Estriol Using Iron-Doped TiO2 under High and Low UV Irradiation" Catalysts 8, no. 12: 625. https://doi.org/10.3390/catal8120625

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