The effects of Fe2+ and Fe3+ as TiO2 cocatalysts were studied, and the experimental results showed that Fe3+ was more efficient than Fe2+, which needed an intermediate reaction to produce hydroxyl radicals. TiO2 was modified with the aim of improving its structural, optical, and adsorption properties, thus improving its photocatalytic performance. The light range of the catalyst activation process was expanded, which increased the catalyst's ability to absorb visible light. Consequently, this study exploits solar energy in photocatalysis by Fe ion doping using different methods, including impregnation, photodeposition, solvothermal doping, and hydrothermal doping, and evaluates the influence of each doping method on TiO2 optical properties and photocatalytic activity. Enhancing the catalyst adsorption capacity by morphologically modifying TiO2 nanoparticles into nanotubes using the hydrothermal method increases the catalyst surface area from 55 to 294 m2/g, as shown in the SEM and BET results. The effect of combining morphological changes and Fe3+ doping on TiO2 activity was evaluated. We observed a reduction in the TiO2 band gap from 3.29 to 3.01 eV, absorption edge widening, and an increase in the specific surface area up to 279 m2/g; thus, the synthesized catalyst eliminated Cefixime in 120 min.

  • The iron ion photocatalytic degradation mechanism is studied.

  • TiO2 nanotubes with high photocatalytic activity are explained.

  • Hydrothermal doping improves TiO2 structural, optical and physical properties.

  • A comparative study of four different Fe3+/TiO2 doping methods is performed.

  • Fe3+ use as a cocatalyst may, in some cases, be more efficient than its use as a dopant.

Graphical Abstract

Graphical Abstract
Graphical Abstract
TNT

TiO2 nanotubes

IMNP-Fex+/TiO2

iron ions loaded on TiO2 nanoparticles by impregnation

IMNT-Fe3+/TNT

Fe3+ loaded on TiO2 nanotubes by impregnation method

PDNP-Fe3+/TiO2

Fe3+ loaded on TiO2 by photodeposition method

Hydr-Fe3+/TiO2

Fe3+ loaded on TiO2 by hydrothermal method

Solv-Fe3+/TiO2

Fe3+ loaded on TiO2 by solvothermal method

TiNT

TNT drying time

TiNP

TiO2 drying time

The pharmaceutical industry generates significant pollution in aqueous environments with highly contaminated discharge, and the treatment of this water remains a major challenge (Shooshtari & Ghazi 2017). Pharmaceutical contaminants are not easily biodegradable under aerobic conditions due to the complexity of their chemical structure; therefore, aqueous effluents require specific treatment (Zavareh & Eghbalazar 2017).

Over the past few years, advanced oxidation processes (AOPs) have been widely used to eliminate residues from effluents; they are represented by hydrogen peroxide in the presence of UV radiation or by the photo-Fenton process using iron species; however, water treatment with titanium dioxide as a catalyst represents a large part of the water treatment field (Nezar & Laoufi 2018).

Unfortunately, TiO2 has two well known drawbacks. (i) Its activation is limited by the ultraviolet light range, which represents 5% of solar energy, and this drawback limits the utilization of this free and renewable energy source. Photon generation for the activation of TiO2 consumes energy, which makes photocatalytic treatment economically and environmentally costly. (ii) A crucial but overlooked TiO2 feature is the adsorption capacity, which promotes pollutant-catalyst contact. This drawback is due to the limited TiO2 adsorption capacity and the difficulty of improving this property by conventional methods. The pollutant adsorption phase is a crucial step in the photocatalytic reaction; this step induces the transfer of pollutant molecules to the catalyst surface, which is characterized by a high density of hydroxyl radicals, the species responsible for pollutant elimination.

To increase the catalyst performance, structural and morphological modifications were made. Structural modifications were made by doping TiO2 with Fe ions to reduce the band gap and expand photon absorption to the visible light range, while morphological modifications were accomplished by changing the TiO2 nanoparticle shapes into nanotubes by hydrothermal treatments to increase the specific surface area of the catalyst and therefore the adsorption capacity.

Catalyst synthesis

Synthesis of Fe2+/Fe3+-doped TiO2

Fex+/TiO2 catalysts in nanoparticle form were prepared using two doping methods, impregnation and photodeposition, while using the impregnation method for TiO2 nanotube doping (Fe3+/TNT). To prepare impregnated Fex+/TiO2 nanoparticles P25/nanotubes, pure TNP/TNT was dispersed in an aqueous solution containing an appropriate amount of Fe2+ or Fe3+ and stirred in darkness for 6 h. The product was obtained after heating at 80 °C for 12 h, and it was then ground and heat treated at 450 °C for 4 h. The products were named IMNP-Fex+/TiO2 and IMNT-Fe3+/TNT. Photodeposited Fe3+/TiO2 nanoparticles were prepared by similar steps; the unique difference in the preparation of photodeposited Fe3+/TiO2 nanoparticles is that the mixtures were stirred under solar radiation instead of in darkness, and the product was named PDNP-Fe3+/TiO2.

Hydrothermal synthesis of TiO2 nanotubes

For TiO2 nanotube preparation, 1 g of commercial TiO2 P25 was dispersed in 60 mL of concentrated sodium hydroxide solution (10 M), followed by mechanical stirring for 2 h. The mixture was transferred into a Teflon-lined stainless-steel autoclave and heated at 150 °C in an oven for 24 h The product was filtered and washed with 0.1 M HCl solution until neutral (pH = 6–7) and was then rinsed with distilled water several times (Su et al. 2018). The product was finally dried in an oven at 80 °C for 36 h, and the synthesis was completed by a calcination process carried out at 450 °C for 4 h.

Hydrothermal/solvothermal Fe3+-TiO2 doping methods

The hydrothermal and solvothermal methods of doping Fe3+-TiO2 followed the steps mentioned above, except that in the hydrothermal doping process, TiO2 and the Fe3+ precursor were dispersed in 60 mL of distilled water before adding NaOH, while in the solvothermal process, they were dispersed in ethanol (Wang et al. 2017). The resulting products were named Hydr-Fe3+/TiO2 and Solv-Fe3+/TiO2.

Experimental conditions

The experiments were carried out in a glass parallelepiped solar photoreactor (Figure 1) with dimensions of (L/W/H) = (50/40/5) cm. The reactor was connected to a reservoir with a capacity of 1 L, the flow was ensured by a peristaltic pump, and the tubing and reservoir were covered by aluminum foil to prevent light penetration. To prepare the treated solution, 10 mg of Cefixime was dissolved in 1 L of distilled water, the catalyst was added (0.2 g/L), and agitation was maintained for 10 min before starting the experiments. The variations in Cefixime concentrations during photocatalytic degradation were monitored by HPLC YL9100 using a reverse-phase column (C18).

Cefixime degradation by TiO2 was evaluated by total organic carbon measurement. The results presented in Figure 2(b) shows that 69% of TOC removal was achieved after 120 min which indicated that Cefixime is efficiently mineralized in presence of 0.2 g/L of TiO2 photocatalyst. All experiments were performed under the same operating conditions.

Figure 1

A laboratory-scale photocatalytic solar reactor.

Figure 1

A laboratory-scale photocatalytic solar reactor.

Close modal
Figure 2

Cefixime UV-visible spectra and TOC removal.

Figure 2

Cefixime UV-visible spectra and TOC removal.

Close modal

Uncertainty analysis

The experimental results are approximate values that differ from the actual values. The error study allowed us to determine the difference between these two values, which in this study did not exceed 5% within a 95% confidence interval. The experiments were repeated three times under the same conditions to increase the accuracy of the results. An uncertainty study was performed to present results with an accuracy of 95% using the following formulas (Kaloyerou 2018):
(1)
(2)
(3)
(4)
(5)
(6)
  • n: number of repetitions

  • : confidence coefficient depends on the number of results n and the desired probability, extracted from student distribution, for (Jackson 2015).

Characterization of synthesized catalysts

The nanoparticle catalysts were identified by Raman spectroscopy using HORIBA LabRAM HR Evolution. The identification of the crystallographic phases was carried out by a Panalytical X'Pert Powder X-ray diffraction system using CuKα radiation (λ = 1.5406 Å). The morphology of the prepared samples was observed using a Quanta 250 FEG scanning electron microscope (SEM), JEOL 7,600 and a JEOL JSM-7610F Plus field emission scanning electron microscope (FE-SEM) for high-resolution images. The surface area of the samples was determined by the Brunauer–Emmett–Teller (BET) equation using a micromeritics ASAP 2020 analyzer, and the pore size distribution was calculated using a method base on the Kalvin equation (Lowell et al. 2012) from the adsorption–desorption isotherms. The optical properties (UV–vis diffuse reflectance spectroscopy, DRS) were evaluated using a Jasco (V650) UV-visible spectrophotometer, the band gap was determined using the formula (7) (Boudiaf et al. 2020), the exponent n depends on the nature of the band gap transition, for direct transition n = 2 and for indirect transition n = 1/2. TiO2 P25 present direct and indirect band gap transition as it is composed of anatase an rutile phases (Zhang et al. 2014):
(7)

RAMAN characterization of (Fe2+, Fe3+)-doped TiO2 nanoparticles

The Raman spectra of TiO2 P25 and synthesized Fex+-TiO2 are shown in Figure 3. All the undoped and Fex+-doped TiO2 Raman spectra show six scattering peaks located at Δω = 142, 195, 397, 516, 517 and 638 cm−1, which belong to the Raman active vibrational modes of the anatase phase, and a peak located at 446 cm−1, which matches well with the characteristic rutile-type TiO2 peak. However, no significant difference was observed among all four Raman spectra of Fex+-TiO2 and undoped TiO2, and no Raman peaks belonging to iron oxide were detected in any of the doped catalyst spectra. This confirms that Fex+ cations are well integrated into the TiO2 substructure. It should be noted that a slight shift towards a higher frequency for the dominant band in Fex+-TiO2 Raman spectra is observed from 142.39 to 142.94 cm−1, and this small displacement is due to the minor change in the crystallinity size of the synthesized catalysts (Ramakrishnan et al. 2018).

Figure 3

Raman spectra of undoped TiO2, Fe2+- and Fe3+-doped TiO2 nanoparticles.

Figure 3

Raman spectra of undoped TiO2, Fe2+- and Fe3+-doped TiO2 nanoparticles.

Close modal

Morphological characterization of synthesis catalysts

FE-SEM and SEM studies were carried out on TiO2 nanoparticles and hydrothermally synthesized specimens to illustrate the nanometric nature of the products, and the morphologies are shown in Figure 4. The product of the hydrothermal process without acid treatment, represented in Figure 4(b), shows a structure without a clearly defined shape compared to the initial structure of TiO2 nanoparticles (Figure 4(a)). This structure is composed of nanosheets as a result of a rupture in the O–Ti–O bonds (Eslami et al. 2016), caused by the aggressive environment in which the hydrothermal reaction takes place at high pressure and temperature in an alkaline solution. Figure 4(c1,c2) reveals the formation of the TiO2 nanotubular morphology after acid treatment. The sample is composed of a three-dimensional TiO2 nanotube structure with a uniform diameter of 24 nm and a length of a few nanometres; this structure agglomerates into fibrous clusters by forming an entanglement of fine structures. This result indicates that the hydrothermal conditions (150 °C, 10 M NaOH) are adequate and that acid treatment is crucial for the formation of TiO2 nanotubes.

Figure 4

SEM/FESEM images of (a) TiO2 nanoparticles and products of the hydrothermal treatments (b) sample treated with water (c1, c2) sample treated with water and acid (nanotubes). (d) IMNT-Fe3+/TNT (e) Hydr-Fe3+/TiO2.

Figure 4

SEM/FESEM images of (a) TiO2 nanoparticles and products of the hydrothermal treatments (b) sample treated with water (c1, c2) sample treated with water and acid (nanotubes). (d) IMNT-Fe3+/TNT (e) Hydr-Fe3+/TiO2.

Close modal

Figure 4(d) and 4(e) show FE-SEM images of Fe3+-doped TiO2 nanotubes formed by hydrothermal doping methods as well as an image of Fe3+-doped nanotubes formed by impregnation. A nanotubular structure appears when using the hydrothermal doping method (Figure 4(e)), though Fe3+ cannot be spotted due to the low amount and high dispersion of the dopant. EDS spectrum of Fe3+-doped TiO2 nanotubes (Figure 5(a)) proved the existence of low amount of Fe after the doping process. As shown in Figure 4(d), the Fe3+-impregnated nanotube image is similar to the image of TNT, and the sample contains nanotubes with consistent diameters. However, a noticeable decrease in specific surface area was observed (Table 1); this decrease is probably caused by the doping process, which may have led to breakage of the weaker nanotubes because nanotubes were subjected to 6 h of continuous stirring, which may have disrupted the tubular morphology of the most fragile nanotubes. BET results show that the hydrothermal treatment caused the transformation of TiO2 nanoparticles into nanotubes, which led to an increase in the surface area with a significant decrease in the average pore size of the catalysts, in the same case nanotube doping increased average pore size and decrease the nanotubes surface area.

Table 1

Photocatalyst specific surface area

CatalystSBET (m2/g)Pore size, Dp (Å)Pore volume, Vp (cm3/g)
TiO2 55 115 0.38 
TNT 294 45 0.33 
IMNP-Fe3+/TiO2 51 120 0.38 
Solv-Fe3+/TiO2 47 113 0.36 
IMNT-Fe3+/TNT 254 65 0.35 
Hydr-Fe3+/TiO2 279 61 0.35 
CatalystSBET (m2/g)Pore size, Dp (Å)Pore volume, Vp (cm3/g)
TiO2 55 115 0.38 
TNT 294 45 0.33 
IMNP-Fe3+/TiO2 51 120 0.38 
Solv-Fe3+/TiO2 47 113 0.36 
IMNT-Fe3+/TNT 254 65 0.35 
Hydr-Fe3+/TiO2 279 61 0.35 
Figure 5

(a) EDS spectrum of Fe3+-doped TiO2 nanotubes (b) N2 adsorption–desorption isotherms (c) pore size distribution for pure and Fe3+-doped TiO2 nanoparticles and nanotubes.

Figure 5

(a) EDS spectrum of Fe3+-doped TiO2 nanotubes (b) N2 adsorption–desorption isotherms (c) pore size distribution for pure and Fe3+-doped TiO2 nanoparticles and nanotubes.

Close modal

X-ray diffraction characterization

The XRD patterns of TiO2 nanoparticles and nanotubes synthesized by the hydrothermal method before and after heat treatment are shown in Figure 6(a). The XRD pattern of the sample dried at 200 °C shows that a new intermediate phase is formed (2θ = 19°, 74°, 90°, 166°), which does not correspond to the two crystalline phases of TiO2 P25 (anatase and rutile). After the heat treatment of the sample at 450 °C, a reappearance of the anatase structure was observed, and all the sharp peaks seen in the XRD pattern of TiO2 nanotubes were attributed to the TiO2 anatase crystalline phase. Therefore, anatase becomes the major crystalline phase.

Figure 6

X-ray diffraction patterns of pure and Fe3+-doped TiO2 nanoparticles and nanotubes.

Figure 6

X-ray diffraction patterns of pure and Fe3+-doped TiO2 nanoparticles and nanotubes.

Close modal
Indeed, there is a complex mechanism that allows the formation of titania nanotubes. The alkaline treatment of TiO2 nanoparticles in a highly concentrated NaOH solution induces alkaline titanate nanosheet formation (Figure 6(a)), which is an important intermediate compound for the formation of titania nanotubes (Bavykin et al. 2010):
(8)
Washing sodium titanate nanosheets with water and treating them with HCl activates the ion exchange of Na+ by H+, which induces nanosheet transformation into nanotubes, and the alkaline titanate nanosheets transform into hydrogen titanate nanotubes (Figure 6(a)) (Nakahira et al. 2010):
(9)
After heat treating the hydrogen titanate nanotubes at 450 °C, a reappearance of the TiO2 anatase phase pattern was observed (Figure 6(a)) because hydrogen titanate nanotubes were transformed into titania nanotubes by releasing protons in the form of H2O (López Zavala et al. 2017). During our experimental study, it was observed that nanotube drying took a much longer time than nanoparticle drying (tNT = 36 h, tNP = 12 h at T = 80 °C), which suggests that water was trapped in the nanotube structure:
(10)

The XRD patterns of TiO2 nanoparticles, nanotubes, Fe3+-doped nanotubes formed by the impregnation method (IMNT-Fe3+/TNT), Fe3+-hydrothermal and solvothermal-doped TiO2 nanotubes (Hydr-Fe3+/TiO2) and (Solv-Fe3+/TiO2) are shown in Figure 6(b). Diffraction peaks correspond to anatase crystalline phases of TiO2 P25 present in all samples, no impurities of iron oxides are detected, and the structure has not changed from those of the synthesized catalysts, indicating that Fe3+ is well integrated in the TiO2 substructure. It should be noted that the Solv-Fe3+/TiO2 XRD pattern presents peaks identical to those of pure TiO2 P25 (anatase and rutile) and with the same intensity, which suggests that solvothermal doping does not cause any changes in the TiO2 structure, unlike the hydrothermal doping that transforms the rutile phase into anatase, leading to rutile peak disappearance.

Optical characterization

Measurements of absorbance UV-visible spectroscopy were carried out on the catalysts to assist the optical properties of the doped TiO2. The results are shown in Figure 7(a). The spectra show that pure TiO2 has high absorption in the ultraviolet light range (wavelength lower than 400 nm). A typical TiO2 absorption edge is observed after 315 nm, which represents the threshold at which the catalyst can no longer absorb visible light. After doping TiO2, the absorption is redshifted towards a higher wavelength to the visible light range. This is a result of the reduced band gap from 3.29 eV for pure TiO2 to 3.26, 3.24, and 3.22 eV for IMNP-Fe2+/TiO2, IMNP-Fe3+/TiO2, and PDNP-Fe3+/TiO2, respectively, which are caused by Fe2+ and Fe3+ electrons 2d and 3d, which excite the TiO2 band gap and shrink it (Asiltürk et al. 2009). Due to the high density of electrons (as a result of ion doping), a new conduction band is formed from these electrons at a position below the conduction band of the catalyst. Therefore, the resulting bandgap is smaller than the original catalyst band gap (Jongprateep et al. 2018). We suggest that the availability of light in the photodeposition method allowed more electrons to be produced by Fe ions, which resulted in a larger density of electrons and thereby greater reduction in the TiO2 band gap. It can be observed that all the catalysts showed a red shift in absorption towards the visible light region, which is a result of Fe3+ doping and of the thermal treatment, as shown in the case of undoped TiO2 nanotubes. In a comparison of absorption capacity of the catalysts formed by the different doping methods, the solvothermal-doped TiO2 catalyst has a better visible light absorption capacity than impregnated nanotubes, hydrothermally doped TiO2, and nanotube catalysts, which means that solvothermal doping has a more positive effect on the optical properties (Table 2) of the catalysts compared to the other doping methods. The hydrothermally doped catalyst exhibits a slightly greater red shift and a higher visible light absorption intensity compared to impregnated nanotubes, while it appears that the impregnation steps adversely affected the optical properties of the final product.

Table 2

Photocatalyst band gaps and the maximum abatement rates at 95% confidence interval

CatalystDirect band gap (eV)Indirect band gap (eV)Eg ErrorX (%)X-Error
TiO2 3.287 3.025 0.18% 85.7 4.11% 
TNT 3.267 3.195 0.49% 94.9 3.86% 
IMNP-Fe2+/TiO2 3.265 2.968 0.37% 87.7 5.00% 
PDNP-Fe3+/TiO2 3.213 2.888 0.87% 87.2 3.94% 
IMNP-Fe3+/TiO2 3.242 2.919 0.34% 89.9 1.5% 
Solv-Fe3+/TiO2 3.036 2.527 1.22% 93.2 5.00% 
IMNT-Fe3+/TNT 3.210 2.764 0.83% 97.5 1.95% 
Hydr-Fe3+/TiO2 3.130 2.663 4.19% 100 3.65% 
CatalystDirect band gap (eV)Indirect band gap (eV)Eg ErrorX (%)X-Error
TiO2 3.287 3.025 0.18% 85.7 4.11% 
TNT 3.267 3.195 0.49% 94.9 3.86% 
IMNP-Fe2+/TiO2 3.265 2.968 0.37% 87.7 5.00% 
PDNP-Fe3+/TiO2 3.213 2.888 0.87% 87.2 3.94% 
IMNP-Fe3+/TiO2 3.242 2.919 0.34% 89.9 1.5% 
Solv-Fe3+/TiO2 3.036 2.527 1.22% 93.2 5.00% 
IMNT-Fe3+/TNT 3.210 2.764 0.83% 97.5 1.95% 
Hydr-Fe3+/TiO2 3.130 2.663 4.19% 100 3.65% 
Figure 7

UV–visible spectra and ‘tauc plot’ for direct and indirect transition of pure and iron ions doped TiO2 nanoparticles and nanotubes.

Figure 7

UV–visible spectra and ‘tauc plot’ for direct and indirect transition of pure and iron ions doped TiO2 nanoparticles and nanotubes.

Close modal

Iron ions (Fe2+, Fe3+) effects as TiO2 cocatalysts

The effects of iron ions Fe2+ and Fe3+ as TiO2 cocatalysts in the presence and absence of H2O2 on Cefixime degradation are shown in Figure 8, and a quantity of 0.85% (M/M) of both Fe2+ and Fe3+ was examined (i.e., MFex+ = 0.85% MTiO2). The percentage of Cefixime elimination was approximately 86% in the absence of Fe2+ or Fe3+, while introducing them in solution brought the abatement to (X = 84.4% in the presence of Fe2+) and (X = 93.7% in the presence of Fe3+) after 120 min. The combinations of Fe2+/H2O2 and Fe3+/H2O2 in the presence of TiO2 were found to be more effective, and the same degradation rate was achieved for both (X = 98.9% in the presence of Fe2+) and (X = 98.9% in the presence of Fe3+). Fe2+ and Fe3+ showed different behaviors in the presence and absence of H2O2, [XTiO2 < XTiO2/Fe2+ < XTiO2/Fe3+ < (XTiO2/Fe2+/H2O2 = XTiO2/Fe3+/H2O2)], which is due to the different mechanisms that take place in the presence of each substance.

Figure 8

Cefixime solar photodegradation in presence of: TiO2, TiO2/H2O2 (3 mM), TiO2/Fex+, TiO2/Fex+/H2O2 (3 mM): 0.85% Fex+, (x = 2, 3).

Figure 8

Cefixime solar photodegradation in presence of: TiO2, TiO2/H2O2 (3 mM), TiO2/Fex+, TiO2/Fex+/H2O2 (3 mM): 0.85% Fex+, (x = 2, 3).

Close modal
In the absence of H2O2 and in the presence of FeNO3.9H2O, Fe3+ hydrolyzation (favored at (Yao et al. 2017)) begins to produce hydroxyl radicals (Zhou & Lei 2006) that attack and degrade Cefixime under solar radiation, according to mechanism:
(11)
Because Fe2+ cannot directly produce hydroxyl radicals, the reactions of Fe2+ ions are divided into two different mechanisms: (i) the stimulated amount of Fe2+ by radiation loses an electron (12) to react with oxygen, forming H2O2 (13); and (ii) the remaining Fe2+ reacts with the H2O2 formed, which gives rise to radical generation according to mechanism (14) (Wu & Deng 2000):
(12)
(13)
(14)

Introducing Fe3+ ions led to a better abatement of Cefixime compared to the introduction of Fe2+ due to the ease of radical generation from the process using Fe3+, while the production of radicals in the presence of Fe2+ ions is limited by the amount of Fe2+ remaining after the formation of H2O2.

The presence of H2O2 led to a significant increase in Cefixime degradation compared to when absent. Indeed, a new mechanism takes place that boosts hydroxyl radical generation by consuming H2O2 rapidly in the presence of superoxide produced as a result of the interaction between oxygen and electrons (Buchalska et al. 2015) according to the Haber–Weiss reaction (15), which is catalyzed by iron salts regardless of its ions (Kehrer 2000). This catalysis is the reason why no difference in Cefixime abatement was observed between Fe2+ and Fe3+ in the presence of H2O2 (Pignatello et al. 2006):
(15)

Photocatalytic performance of iron ion-doped TiO2 nanoparticles and nanotubes

The effect of Fe2+ and Fe3+ as dopants on the TiO2 nanoparticle (NP) photocatalytic performance to degrade Cefixime is represented in Figure 9(a). Photocatalytic activity enhancement is observed as a result of TiO2 doping, and this improvement is less significant in the case of IMNP-Fe2+/TiO2 than in the case of IMNP-Fe3+/TiO2, where the degradation is enhanced from 85.7% to 87.7% and 89.9% for TiO2, IMNP-Fe2+/TiO2, and IMNP-Fe3+/TiO2, respectively. This increase in the photocatalytic activity of doped TiO2 can be attributed to the shift in optical absorption to higher wavelengths towards the visible range and a decrease in the band gap energy, as shown in Figure 7. Therefore, doped TiO2 becomes more sensitive to visible light and can be better activated under solar radiation. IMNT-Fe3+/TiO2 had a higher performance than IMNT-Fe2+/TiO2 because it had a greater shift in optical absorption towards the visible range, as shown in Figure 7, and hence a greater solar radiation absorption capacity. Moreover, Fe3+ behaves as an electron and hole trap that induces electron-hole separation as a result of the higher Fe3+/Fe2+ and Fe3+/Fe4+ redox potentials (Rajagopalan et al. 2017). Fe2+ acts as a trap for electrons, and Fe4+ acts as a trap for holes, giving the electron holes sufficient time to react and generate radicals.

Figure 9

Cefixime solar photodegradation in presence of pure and iron ion-doped TiO2 nanoparticles and nanotubes.

Figure 9

Cefixime solar photodegradation in presence of pure and iron ion-doped TiO2 nanoparticles and nanotubes.

Close modal

As shown in Figure 9(a), IMNP-Fe3+/TiO2 had a better photocatalytic performance in Cefixime elimination, despite having a lower shift towards the visible light range than did PDNP-Fe3+/TiO2 (Figure 6); it could be that, during photodeposition, Fe3+ ions were deposited on the TiO2 surface as Fe2+ and Fe4+, and after being reduced or oxidized by photogenerated electrons and holes, the absence of light during the impregnation process inhibited electron-hole generation, resulting in the deposition of ions as Fe3+, which is a more active iron ion.

The photocatalytic activity of TiO2 nanotubes was evaluated. Figure 9(b) shows the Cefixime photodegradation curves in which TNP and TNT are used as photocatalysts under solar radiation. In the presence of TiO2 nanotubes (TNTs), Cefixime abatement reaches 94.9% after 120 min of solar exposure, whereas it reaches 85.7% in the presence of TiO2 nanoparticles. This improved performance in Cefixime removal is mainly attributed to the morphology of titanium dioxide, which has a crucial role in catalyst performance and an important influence on photocatalytic behavior and is less attributed to the TNT structure.

Regarding the TNT structure, Figure 6 shows the disappearance of the rutile phase and the formation of a well defined anatase crystalline phase of TiO2, which is more the photoactive form of TiO2 than the rutile form. This transformation in TiO2 P25 phases is due to hydrothermal treatment.

Regarding the TNT morphology, TiO2 nanotubes offer a much larger BET surface area (294 m2/g) than nanoparticles (55 m2/g) (Table 1), which is a very important characteristic because pollutant decomposition occurs at the catalyst interface; a larger surface area means a higher number of active sites and contact surface to adsorb water for radical generation and to adsorb pollutants, which promotes pollutant-radical contact, leading to higher Cefixime oxidation. Therefore, TNT is more efficient in radical formation, which leads to the availability of a large number of hydroxyl radicals on the nanotube surface, as mentioned by Manfroi et al. (2014).

The TiO2 charge in the form of nanotubes is more exposed to pollutants and light radiation than NPs; indeed, the hollow TNT tube shape allows the transfer of pollutants into the nanotube core, which promotes pollutant diffusion. The longitudinal tube profile improves light absorption (Marien et al. 2016), which makes TNT an efficient light harvester compared to TNP. In addition, Sun et al. (2014) observed that by changing the morphology of TiO2 nanoparticles into nanotubes, the generated electrons become more efficient as their lifespan is extended and their transfer becomes faster.

Moreover, nanotubes red shift towards the visible region (Figure 7) when nanotubes are not doped, so TNT had improved solar photoactivation characteristics, which was also observed by Alfaro Cruz et al. (2019). This TNT red shift can be attributed to the tube-shaped TiO2 and the hydrothermal treatment (Camposeco et al. 2016). Considering all of the above, it is obvious that all of these features improve the photocatalytic performance of TiO2 nanotubes.

The photocatalytic activities of thermally synthesized catalysts under solar irradiation were evaluated for their efficacy in Cefixime removal, and the results are presented in Figure 8(c). All the synthesized catalysts showed higher photocatalytic activity than TiO2 P25. In 120 min., Cefixime abatement reached 85.7, 89.9, 94.9, 94.9, 97.5, and 100% for TiO2, IMNP-Fe3+/TiO2, Solv-Fe3+/TiO2, TNT, IMNT-Fe3+/TNT, and Hydr-Fe3+/TiO2, respectively. The Hydr-Fe3+/TiO2 catalyst eliminated Cefixime successfully in 120 min.

Nanoparticle catalysts doped by impregnation or by the solvothermal method achieved lower abatement than hydrothermally doped or treated catalysts due to the influence of the doping technique. Hydrothermal doping induced TiO2 P25 rutile phase conversion into anatase, a reduction in the band gap caused by the hydrothermal process itself as well as by Fe3+ doping, and nanotube formation, which made it possible to considerably increase the catalyst specific surface area. Consequently, all these properties of the hydrothermal doping process combined to provide a powerful and efficient catalyst, while these improvements were not achieved using other doping methods, such as solvothermal, impregnation, and photodeposition methods, since these methods have a limited influence on the structure (Figure 6) and physical properties (Table 2) of the catalyst, regardless of the dopant (Fe3+) influence.

According to the comparison results, Figure 9(d) shows that the photocatalytic efficiency of iron ion Fe3+ changes depending on how it was introduced into the reaction. Comparing four different Fe3+/TiO2-doping methods shows that employing Fe3+ as a TiO2 cocatalyst gives better results than impregnation, photodeposition, and solvothermal-doping processes; indeed, when Fe3+ is employed as a dopant, Cefixime is eliminated by the radicals generated through TiO2 activation, and the role of Fe3+ is to reduce the band gap and extend the electron-hole pair separation by catching the electrons. However, when Fe3+ is used as a cocatalyst, Cefixime is attacked by radicals generated not only by TiO2 activation but also by Fe3+ hydrolyzation. Thus, the photocatalytic system that used Fe3+ as a TiO2 dopant using photodeposition and impregnation-doping methods performed poorly and was shown to be ineffective compared to that in which Fe3+ is used as a cocatalyst. The doping process requires steps that consume time and energy. Therefore, when using Fe3+, it is recommended that the impregnation and photodeposition-doping methods be recognized as dispensable. Using the hydrothermal-doping method considerably promotes the photocatalytic performance of TiO2 in Cefixime elimination; this doping method combines a structural improvement represented by the transformation of the crystals from rutile to anatase form, an optical improvement due to the reduction of the band gap, and a physical improvement attributed to the formation of TiO2 nanotubes that have a large surface area and are capable of destroying the pollutant.

In this work, we were interested in defining the most effective way to exploit iron ions (Fe2+, Fe3+) to improve Cefixime removal by employing them either as titanium dioxide cocatalysts (radical generators) or as TiO2 dopants (band gap reducers).

The results suggested that there was no advantage in using Fe3+ ions as dopants by conventional methods such as impregnation, photodeposition or solvothermal methods because these methods consume time and energy; they make the photocatalytic system less efficient, so using Fe3+ ions as cocatalysts gives better results than using these doping methods.

Titanium dioxide nanotubes were prepared by a hydrothermal method, and the influence of this method on the synthesized catalyst was evaluated; it showed improved physical, structural and optical properties of TiO2. Therefore, the hydrothermal doping method was exploited to produce a highly effective catalyst with enhanced properties that allowed complete Cefixime elimination in only 90 min by taking advantage of the dopant Fe3+ and TiO2 nanotube properties.

The authors are thankful for technical help from Amar Manseri a member of the Research Center of Semi-conductor Technology for Energy, CRTSE, Algiers, Algeria and Dr Yasmina Roumila a member of the Laboratory of Electrochemistry-Corrosion, Metallurgy and Inorganic Chemistry, Faculty of Chemistry, USTHB, Algeria.

All relevant data are included in the paper or its Supplementary Information.

Alfaro Cruz
M. R.
Sanchez-Martinez
D.
Torres-Martínez
L. M.
2019
TiO2 nanorods grown by hydrothermal method and their photocatalytic activity for hydrogen production
.
Materials Letters
237
,
310
313
.
https://doi.org/10.1016/j.matlet.2018.11.040
.
Asiltürk
M.
Sayılkan
F.
Arpaç
E.
2009
Effect of Fe3+ ion doping to TiO2 on the photocatalytic degradation of Malachite Green dye under UV and vis-irradiation
.
Journal of Photochemistry and Photobiology A: Chemistry
203
(
1
),
64
71
.
Available from: https://linkinghub.elsevier.com/retrieve/pii/S1010603008005248.
Bavykin
D. V.
Carravetta
M.
Kulak
A. N.
Walsh
F. C.
2010
Application of magic-angle spinning NMR to examine the nature of protons in titanate nanotubes
.
Chemistry of Materials
22
(
8
),
2458
2465
.
Available from: https://pubs.acs.org/doi/10.1021/cm903100a.
Boudiaf
S.
Nasrallah
N.
Mellal
M.
Belabed
C.
Belhamdi
B.
Meziani
D.
Mehdi
B.
Trari
M.
2020
Synthesis and characterization of semiconductor CoAl2O4 for optical and dielectric studies: application to photodegradation of organic pollutants under visible light
.
Optik
219
,
165038
.
https://doi.org/10.1016/j.ijleo.2020.165038
.
Buchalska
M.
Kobielusz
M.
Matuszek
A.
Pacia
M.
Wojtyła
S.
Macyk
W.
2015
On oxygen activation at rutile- and anatase-TiO2
.
ACS Catalysis
5
(
12
),
7424
7431
.
Available from: https://pubs.acs.org/doi/10.1021/acscatal.5b01562.
Camposeco
R.
Castillo
S.
Navarrete
J.
Gomez
R.
2016
Synthesis, characterization and photocatalytic activity of TiO2 nanostructures: nanotubes, nanofibers, nanowires and nanoparticles
.
Catalysis Today
266
,
90
101
.
http://dx.doi.org/10.1016/j.cattod.2015.09.018
.
Eslami
H.
Moztarzadeh
F.
Kashi
T. S. J.
Khoshroo
K.
Tahriri
M.
2016
Hydrothermal synthesis and characterization of TiO2-derived nanotubes for biomedical applications
.
Synthesis and Reactivity in Inorganic, Metal-Organic, and Nano-Metal Chemistry
46
(
8
),
1149
1156
.
Available from: https://www.tandfonline.com/doi/full/10.1080/15533174.2014.988828.
Jackson
S. L.
2015
Research Methods and Statistics: A Critical Thinking Approach
.
Cengage Learning
,
Boston, MA
,
USA
.
Jongprateep
O.
Meesombad
K.
Techapiesancharoenkij
R.
Surawathanawises
K.
2018
Chemical composition, microstructure, bandgap energy and electrocatalytic activities of TiO2 and Ag-doped TiO2 powder synthesized by solution combustion technique
.
Ceramics International
44
,
S228
S232
.
https://doi.org/10.1016/j.ceramint.2018.08.108
.
Kaloyerou
P. N.
2018
Basic Concepts of Data and Error Analysis
.
Springer International Publishing
,
Cham
.
López Zavala
M. Á.
Lozano Morales
S. A.
Ávila-Santos
M.
2017
Synthesis of stable TiO2 nanotubes: effect of hydrothermal treatment, acid washing and annealing temperature
.
Heliyon
3
(
11
),
e00456
.
http://dx.doi.org/10.1016/j.heliyon.2017.e00456
.
Lowell
S.
Shields
J. E.
Thomas
M. A.
Thommes
M.
2012
Characterization of Porous Solids and Powders: Surface Area, Pore Size and Density
.
Springer Science & Business Media
, Dordrecht, The Netherlands.
Manfroi
D. C.
dos Anjos
A.
Cavalheiro
A. A.
Perazolli
L. A.
Varela
J. A.
Zaghete
M. A.
2014
Titanate nanotubes produced from microwave-assisted hydrothermal synthesis: photocatalytic and structural properties
.
Ceramics International
40
(
9
),
14483
14491
.
Available from: https://linkinghub.elsevier.com/retrieve/pii/S0272884214010402.
Marien
C. B. D.
Cottineau
T.
Robert
D.
Drogui
P.
2016
Tio2 Nanotube arrays: influence of tube length on the photocatalytic degradation of paraquat
.
Applied Catalysis B: Environmental
194
(
2
),
1
6
.
http://dx.doi.org/10.1016/j.apcatb.2016.04.040
.
Nakahira
A.
Kubo
T.
Numako
C.
2010
Formation mechanism of TiO2-derived titanate nanotubes prepared by the hydrothermal process
.
Inorganic Chemistry
49
(
13
),
5845
5852
.
Available from: https://pubs.acs.org/doi/10.1021/ic9025816.
Nezar
S.
Laoufi
N. A.
2018
Electron acceptors effect on photocatalytic degradation of metformin under sunlight irradiation
.
Solar Energy
164
(
2018
),
267
275
.
Available from: https://linkinghub.elsevier.com/retrieve/pii/S0038092X18301981.
Pignatello
J. J.
Oliveros
E.
MacKay
A.
2006
Advanced oxidation processes for organic contaminant destruction based on the Fenton reaction and related chemistry
.
Critical Reviews in Environmental Science and Technology
36
(
1
),
1
84
.
Rajagopalan
R.
Chen
B.
Zhang
Z.
Wu
X. L.
Du
Y.
Huang
Y.
Li
B.
Zong
Y.
Wang
J.
Nam
G. H.
Sindoro
M.
Dou
S. X.
Liu
H. K.
Zhang
H.
2017
Improved reversibility of Fe3+/Fe4+ redox couple in sodium super ion conductor type Na3Fe2(PO4)3 for sodium-ion batteries
.
Advanced Materials
29
(
12
).
Ramakrishnan
V. M.
Natarajan
M.
Santhanam
A.
Asokan
V.
Velauthapillai
D.
2018
Size controlled synthesis of TiO2 nanoparticles by modified solvothermal method towards effective photo catalytic and photovoltaic applications
.
Materials Research Bulletin
97
(
September 2017
),
351
360
.
Shooshtari
N. M.
Ghazi
M. M.
2017
An investigation of the photocatalytic activity of nano Α-Fe2O3/ZnO on the photodegradation of cefixime trihydrate
.
Chemical Engineering Journal
315
,
527
536
.
http://dx.doi.org/10.1016/j.cej.2017.01.058
.
Sun
K. C.
Qadir
M. B.
Jeong
S. H.
2014
Hydrothermal synthesis of TiO2 nanotubes and their application as an over-layer for dye-sensitized solar cells
.
RSC Advances
4
(
44
),
23223
.
Available from: http://xlink.rsc.org/?DOI=c4ra03266g.
Wang
Q.
Jin
R.
Zhang
M.
Gao
S.
2017
Solvothermal preparation of Fe-doped TiO2 nanotube arrays for enhancement in visible light induced photoelectrochemical performance
.
Journal of Alloys and Compounds
690
,
139
144
.
Available from: https://linkinghub.elsevier.com/retrieve/pii/S0925838816323167.
Yao
H.
Pei
J.
Wang
H.
Fu
J.
2017
Effect of Fe(II/III) on tetracycline degradation under UV/VUV irradiation
.
Chemical Engineering Journal
308
(
September
),
193
201
.
http://dx.doi.org/10.1016/j.cej.2016.09.074
.
Zavareh
S.
Eghbalazar
T.
2017
Efficient and selective removal of cefixime form aqueous solution by a modified bionanocomposite
.
Journal of Environmental Chemical Engineering
5
(
4
),
3337
3347
.
http://dx.doi.org/10.1016/j.jece.2017.06.042
.
Zhang
J.
Zhou
P.
Liu
J.
Yu
J.
2014
New understanding of the difference of photocatalytic activity among anatase, rutile and brookite TiO2
.
Physical Chemistry Chemical Physics
16
(
38
),
20382
20386
.
http://dx.doi.org/10.1039/C4CP02201G
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/)