Next Article in Journal
Aero-TiO2 Prepared on the Basis of Networks of ZnO Tetrapods
Next Article in Special Issue
Structure Determination and Analysis of the Ceramic Material La0.987Ti1.627Nb3.307O13 by Synchrotron and Neutron Powder Diffraction and DFT Calculations
Previous Article in Journal
A Novel Nonlinear Optical Limiter Based on Stimulated Brillouin Scattering in Highly-Nonlinear Fiber
Previous Article in Special Issue
High-Energy Heavy Ion Tracks in Nanocrystalline Silicon Nitride
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanosynthesis of Mesoporous Bi-Doped TiO2: The Effect of Bismuth Doping and Ball Milling on the Crystal Structure, Optical Properties, and Photocatalytic Activity

by
Sofía Estrada-Flores
1,
Catalina M. Pérez-Berumen
1,
Tirso E. Flores-Guia
1,
Luis A. García-Cerda
2,
Joelis Rodríguez-Hernández
2,
Tzipatly A. Esquivel-Castro
1 and
Antonia Martínez-Luévanos
1,*
1
Facultad de Ciencias Químicas, Universidad Autónoma de Coahuila, Blvd. V. Carranza s/n, Saltillo 25280, Coahuila, Mexico
2
Centro de Investigación en Química Aplicada, Blvd. Enrique Reyna, Saltillo 25294, Coahuila, Mexico
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(12), 1750; https://doi.org/10.3390/cryst12121750
Submission received: 7 October 2022 / Revised: 6 November 2022 / Accepted: 9 November 2022 / Published: 3 December 2022
(This article belongs to the Special Issue Structural Investigation of Ceramic Materials)

Abstract

:
In this work, we reported obtaining mesoporous Bi-doped TiO2 by mechanosynthesis and bismuth loading of 0%, 1%, 3%, 5%, and 10% (milled TiO2, TiO2 Bi 1%, TiO2 Bi 3% TiO2 Bi 5%, and TiO2 Bi 10%, respectively). The effect of bismuth doping and ball milling on the crystal structure, optical properties, and photocatalytic performance of Bi-doped TiO2 mesoporous samples under UV, visible, and sun irradiation was investigated. According to the results of the Rietveld refinement, the estimated chemical formulas for the TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi10% samples were Ti0.99Bi0.01O2, Ti0.97Bi0.03O2, Ti0.96Bi0.04O2, and Ti0.91Bi0.09O2 respectively. The incorporation of Bi into the TiO2 lattice causes the crystallite size to decrease and, consequently, the absorption spectrum of TiO2 to extend into the visible region of the electromagnetic spectrum, resulting in a lower band gap (Eg) value. Bi-doped TiO2 mesoporous samples had Eg values of 2.90 eV, 2.83 eV, 2.77 eV, and 2.70 eV for the TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi 10% samples, respectively. Photocatalytic removal of methylene blue (MB) data fit well for second-order kinetics. Photocatalytic activity increase followed the order of TiO2 Bi 5% > TiO2 Bi 10% > TiO2 Bi 3% > TiO2 Bi 1% > pristine TiO2. The TiO2 Bi 5% sample exhibited excellent photocatalytic performance for MB photodegradation under natural sunlight (89.2%).

1. Introduction

Titanium oxide (TiO2) has now become a major material due to its applications within the environmental arena, especially in sustainable energy. Amongst these applications, its use as a photo-anode in solar cells [1], a photocatalyst in the degradation of organic pollutants [2], in the production of H2 [3], and in the manufacture of lithium batteries [4] are the most prominent. The last is mainly due to its chemical and physical properties and relatively low cost [5].
TiO2 has been widely used as a photocatalyzer to degrade a large number of organic molecules, such as dyes [6,7,8,9,10,11], drugs [12,13,14,15], and pesticides [16,17,18], amongst other molecules. However, its use is limited to processes in which the light source is in the UV range due to its forbidden energy gap or band gap value (Eg), which ranges from 3.0 to 3.2 eV, depending on its crystalline phase. The anatase crystalline phase is preferred in photocatalysis because it is obtained at lower temperatures and has a higher degree of hydroxylation on its surface with respect to rutile and brookite, which contributes to the better performance in the adsorption and photodegradation of organic molecules [5].
Numerous efforts have been made to extend the energy absorption to the visible region of the electromagnetic spectrum and decrease the value of the band gap of the TiO2 anatase phase, with the aim of achieving the activation of this semiconductor with visible and even solar light. The doping of TiO2 anatase phase with metallic and nonmetallic elements serves to achieve the modification of this optical property; some of the most-used elements are C, N, and Fe [19,20,21,22]. The degree of modification of the optical and photocatalytic properties of doped TiO2 depends on its characteristics, such as porosity, morphology, crystallite size, and the presence of other phases such as rutile and brookite; in turn, these characteristics depend a lot on the method and synthesis variables.
There are some interesting research papers in which TiO2 doping has been reported with Bi using the solvothermal and sol–gel methods, the latter being the most researched. The literature reports that the absorption spectrum of Bi-doped TiO2 has been shifted towards the visible region of the electromagnetic spectrum, significantly reducing the value of Eg and improving its photocatalytic activity; however, improving both properties strongly depends on the level of bismuth doping, morphology, and crystalline phase, and, at the same time, on the synthesis method. Wang et al. obtained photocatalysts of Bi-doped anatase TiO2 hollow thin sheets by the hydrothermal method with Eg values of 3.16 eV and 2.97, the former corresponding to a sample with 0.6 ato. % Bi, and the latter to a sample containing 0.8 at. % Bi into TiO2 and with the presence of Bi2O3 [23]. Xu et al. obtained Bi-doped titania-coated carbon spheres by the sol–gel method and using carbon spheres; they investigated the effects of Bi content on the physical structure and photocatalytic activity of hollow titanium spheres, and reported that the crystallite size of the titania becomes smaller when the doping amount of Bi ions increases, and that Bi species in the sample with Bi at. percent of 4% were present in the form of Bi2O3 [24].
There are less popular synthesis methods, but these are more simplified and more environmentally sustainable, such as mechanosynthesis, which allows for the incorporation of atoms as dopants into the crystalline structure of metal oxides, due to the high energy generated with elastic shocks between reagents and grinding media, in addition to causing defects in the crystal lattice [25]. Moreover, this method is considered a type of green synthesis, as the use of solvents is not required. It is to be expected that the incorporation of a dopant such as Bi3+ into the TiO2 anatase phase will improve the absorption of energy into the visible region of the electromagnetic spectrum, regardless of the synthesis method; however, the magnitude of such an improvement depends on the synthesis method and the degree of doping achieved. For these reasons, it is important to investigate how doping and grinding affect the crystal structure of TiO2, crystallite size, and morphology, and how these, in turn, are related to the modification of the Eg values of the samples.
This research reported the obtaining of Bi-doped TiO2 mesoporous samples by mechanosynthesis. The effect of bismuth doping and grinding on structural, morphological, and porosity characteristics, as well as their effect on optical properties and photocatalytic activity, was investigated.

2. Materials and Methods

2.1. Chemicals

TiO2 anatase phase (pristine TiO2 sample), previously obtained following the methodology of our working group [26], involved Bi(NO3)3·5H2O (98%, Aldrich, Saltillo, Mexico) and C16H18ClN3S (MB, 99%, Aldrich, Saltillo, Mexico).

2.2. Synthesis of Bi-Doped TiO2

TiO2 pristine sample and a certain previously determined amount of Bi (NO3)3·5H2O were manually mixed in an agate mortar; the mixture was placed in a zirconia container with zirconia grinding media, and mechanically ground at 300 rpm using a High Energy Planetary ball mill Fritsch, Pulverisette 6 model. The load-to-balls ratio was 1:20 and the grinding time was two periods of 30 min. At the end of the grinding time, the product was heat-treated at a temperature of 450 °C for 4 h and subsequently characterized. The experiment was repeated several times by varying the amount of the bismuth precursor to obtain different degrees of doping (Ti1−xBixO2, x = 0–0.1). Samples synthesized with values of x = 0, x = 0.01, x = 0.03, x = 0.05, and x = 0.1 were identified as milled TiO2, TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi 10%, respectively.

2.3. Photocatalytic Activity Evaluation

The photocatalytic performance of the samples was evaluated using MB as a model molecule. The standard solution used for photocatalytic experiments (100 mg/L) was prepared by dissolving 100 mg of MB in 1 L of deionized water. An aqueous dilution of this dye (100 mL, 20 mg/L) was put in contact with each of the catalysts (1.0 g/L), and a solution of 0.1 M NaOH was used to adjust the pH value to 8. The mixture was stirred for 2 h in the dark to reach the adsorption equilibrium. Next, the solution with the catalyst was irradiated with visible light using a Pen-Ray® Xenon Lamp (λ = 467 nm, 1.8 Watts).
Aliquots of 2 mL were collected at the following time intervals: 5, 10, 15, 30, 45, and 60 min. Subsequently, each aliquot was centrifuged and its absorbance at 664 nm was read with a UV–Vis spectrophotometer (Jenway-6705); the MB concentration was calculated using a calibration curve that was obtained previously with standard solutions at different concentrations. A double-jacketed glass reactor and a water-cooling bath were used to keep the temperature at 25 ± 1 °C. The photolysis of MB solution was performed with visible light for 2 h. A magnetic stirrer was used to ensure the homogeneity of the solution.
The effect of the amount of catalyst (0.5 g/L, 1.0 g/L, and 2.0 g/L) and the type of irradiation (UV light, visible light, and natural sunlight) on the photodegradation efficiency of MB was investigated; an irradiation time of 1 h was used. For photodegradation experiments under UV light irradiation, a Pen-Ray® Mercury Lamp (λ = 254 nm, irradiance = 4400 µW/cm2, 300 V) was used, and for photodegradation experiments under natural sunlight, an irradiance of 320.68 W/m2 was registered. A catalyst concentration of 0.5 g/L and an irradiation time of 60 min were used. Photocatalysis and photolysis experiments were performed in triplicate.

2.4. Characterization

The infrared spectra were collected in the range of 4000 to 400 cm−1 with 4 cm−1 resolution using a Thermo Scientific Nicolet iS10 spectrometer (Waltham, MA, USA) with attenuated total reflectance (ATR) modality. X-ray photoelectron spectra (XPS) measurements were performed with a Thermo Scientific K-Alpha+ Model spectrometer, using the AlKα X-ray beam (1486.6 eV) with the C 1s (peak at 284.8 eV) as a reference. N2 adsorption–desorption measurements were performed in a Quantachrome Instruments Autosorb 1 model (Boynton Beach, FL, USA). The samples were degassed at 100 °C for 12 h; the specific surface areas were calculated using the Brunauer–Emmett–Teller (BET) theory, and the pore size distributions and pore volumes were calculated with the desorption data from adsorption-desorption isotherms (based on the Barrett–Joyner–Halenda (BJH) theory). The morphologies of the samples were investigated with a JEOL JSM-7800F electron microscope (Salem, MA, USA). The diffractograms of the samples were obtained using a Rigaku model Ultima IV equipment with D/teX detector (The Woodlands, TX, USA), Bragg–Brentano geometry, with Cu kα X-ray source at a voltage of 40 kV and current of 40 mA, and scanning in the 10–80° 2θ range at a speed of 2° min−1. Rietveld refinement of the observed XRD patterns was performed with the FULLPROF suite software, using the conditions described above. The spatial group I41/amd was used for the anatase and lattice parameters previously obtained by the Le Bail setting of the diffractograms with the FullProf software. The UV–Vis diffuse reflectance spectra and the UV–Vis absorbance spectra were obtained using a Perkin-Elmer Lambda 35 UV–Visible spectrophotometer (Houston, TX, USA).

3. Results

3.1. Synthesis and Characterization

Figure 1 shows the ATR-FTIR spectra of Bi-doped TiO2 samples obtained by mechanosynthesis; the spectra of all samples showed the typical broad band of the Ti–O bond stretch from 800 cm−1 to 400 cm−1, as well as absorption bands of the bending and stretching of the O–H bond of the OH group in 1640 cm−1 and between 3600 and 3000 cm−1, respectively. The small absorption bands around 3725 and 3693 cm−1 correspond to the O–H bond stretch of the groups OH free on the surface of the material, which are associated with the planes with higher electronic density; in the case of anatase, the plane with the highest electronic density was (101). In addition, it has been reported that such superficial OH groups occur when oxygen vacancy exists on the surface of the material [27]. The presence of an OH group on the surface of the material is favorable because they react with the holes (H+) generated in photocatalysis when the material is activated, forming hydroxyl radicals (HO*) and making the photocatalytic process more effective [23].
The XPS technique was used to analyze the chemical structures and electronic states of pristine TiO2 and Bi-doped TiO2 samples. Figure 2A shows the XPS survey spectra of TiO2 Bi 1% and TiO2 Bi 5% samples, which contain the peaks of Ti, Bi, O, and C elements; for purposes of comparison, the XPS survey spectrum of pristine TiO2 is also shown in this figure, which contains the peaks of Ti, O, and C. The observed C 1s peak was used as an energy reference for determining the peak positions of core-level spectra.
Figure 2B–D show the high-resolution XPS spectra of the Ti 2p, Bi 4f, and O 1s regions, respectively, on the surface of TiO2 Bi 1% and TiO2 Bi 5% samples, as well as pure anatase (pristine TiO2). The Ti2p spectra are shown in Figure 2B, where it is possible to observe two peaks at 458.88 eV and 464.64 eV assigned to Ti2p3/2 and Ti2p1/2, which are characteristic of Ti4+. The spectra of doped samples show a shift of these peaks towards smaller binding energies due to the formation of Ti–O–Bi bonds and the creation of surface defects [28]; this same effect was observed by Li et al., who synthesized titanium oxide doped with Bi3+ using a modified sol–gel method [29]. The Bi4f region (Figure 2C) is formed by three peaks; the first peak located at 156.9 eV is attributed to Bi0. According to Wu et al., the reason for Bi0 species formation in samples is uncertain; the remaining carbonaceous species might have reduced partial Bi3+ to Bi0 during the calcination [30] and during the ball milling, this last considering that high temperature and energy values can be reached during grinding. Other authors report that such species also cause modifications to the X-ray diffraction pattern of samples, causing displacement at a larger 2θ degree [31]. Bi0 could be decorating the surface of the TiO2 Bi 1% and TiO2 Bi 5% samples.
The peaks at 164.6 eV and 159.3 eV for both samples can be ascribed to Bi 4f5/2 and Bi 4f7/2, respectively, corresponding to Bi3+, whereas the peaks centered at 163.4 and 158 eV for the TiO2 Bi 1% sample and the peaks at 163.1 and 158 for the TiO2 Bi 5% sample are attributed to Bi 4f5/2 and Bi 4f7/2 of Bi 4+. The peaks at 159.3 and 158 eV ascribed to Bi 4f7/2 are the signals corresponding to Bi3+ and Bi4+, respectively, indicating a partial oxidation of the centers from Bi3+ to Bi4+; Wang et al., (2016) and Hao et al., (2019) mention that the peaks at 163.7 eV and 164.5 eV, corresponding to Bi 4f5/2, and the peaks at 159.2 eV and 158.4 eV, corresponding to Bi 4f7/2, are attributed to the oxidized bismuth layers [23,32].
The O1s peak of the pristine TiO2 sample consists of two signals at 530.08 and 531.97 eV corresponding to the O–Ti bond and the O–H bond of hydroxyl groups on the surface of the material, respectively. Bi-doped samples show three signals: one around 529.78 eV corresponding to the O–Ti bond, which is displaced to lower bond energy values due to the incorporation of bismuth into the crystal lattice [33], another at 530.58 eV indicating the presence of the O–Bi bond and increasing with respect to the amount of bismuth present in the sample, and, finally, one at 531.38 eV attributed to the O–H bond, indicating the presence of hydroxyl groups on the surface of the material (Ti–OH) [29,34]. The intensity of the signal of the O–H bond decreases as the degree of doping increases, which may be an indication that hydroxyl groups interact with Bi, forming Bi–O–Ti (Figure 2D).
Figure 3A–F shows the micrographs of pristine TiO2, milled TiO2, and Bi-doped TiO2 samples. It is observed that the pristine TiO2 sample (Figure 3A) had a porous morphology, consisting of agglomerates of small particles. On the other hand, it is observed in the micrographs of the undoped sample (milled TiO2, Figure 3B) and in those of TiO2 samples doped with bismuth (Figure 3C–F) that the morphology changed markedly and had less porosity than the pristine TiO2 sample. It is well known that mechanical grinding produces changes in morphology, particle size distribution, and porosity. Porosity in photocatalyst materials is a desirable characteristic, because the number of active sites for the adsorption of molecules or compounds to be photodegraded increases with porosity. The elemental mapping of Ti, Bi, and O on the TiO2 Bi3% sample is presented in Figure 4; a uniform dispersion of the three elements in this sample is observed.
Figure 5 shows N2 adsorption–desorption curves of the TiO2 powder prepared by mechanosynthesis at different Bi-doping concentrations and calcined at 450 °C. It can be observed that the pristine TiO2, milled TiO2, and TiO2 1% samples have an isotherm of type IV, which is characteristic of mesoporous powders. On the other hand, the sample with 5% bismuth (TiO2 Bi 5% sample) presents a mixture of isotherm types II and IV. At a high relative pressure from 0.40 to 0.95, all isotherms of milled samples exhibited a hysteresis loop of type H3, indicating that they contain mesopores (2–50 nm). Table 1 presents the values of the specific surface area (SBET) determined by the BET technique, in which it can be observed that the pristine TiO2 sample had a SBET value of 138.7 m2/g and that the milled sample without doping (milled TiO2) and the milled and doped samples with bismuth (TiO2 Bi 1% and TiO2 5%) had smaller specific surface area values.
Figure 5 also shows the Barrett–Joyner–Halenda (BJH) pore size distribution of the samples, determined from the desorption branches. It can be seen that the pristine TiO2 sample had a high quantity of pores with a size less than 10 nm; on the other hand, samples subjected to mechanical grinding and bismuth doping had a wider pore distribution. However, the size of their pores is in the range of mesopores, as can be seen from the mean pore diameter values (Dp) listed in Table 1, ranging from 3.40 to 3.50 nm. The pristine TiO2 sample had an average pore size value of 5.86 nm and a relatively high mean pore volume value (Vp) of 0.300 cm3/g, compared with the ground sample and ground and doped samples. The milled TiO2 sample had two maximum peaks, and its pore size distribution was not monodisperse. The decrease in SBET value is related to the decrease in Dp and Vp; such a decrease is due to both grinding and bismuth doping, as can be seen from the SBET, Dp, and Vp values listed in Table 1. Mechanical grinding produces changes in the morphology, particle size, and porosity (Figure 3C–F); the action of the grinding media on the TiO2 pristine sample (Vp = 0.300 cm3/g) causes the partial destruction of the pores, leading to a reduction in porosity up to 33% for the TiO2 5% sample (Vp = 0.201 cm3/g). Henych et al., (2014) mentioned that the pores of a sample must be large enough that the light affecting the photocatalysis process can enter them and activate the surface [35]. The Bi-doped TiO2 samples obtained in this work by mechanosynthesis remained mesoporous, a characteristic that will allow them to have high photocatalytic activity.
Figure 6A shows the diffraction patterns of the synthesized samples; diffractograms of all samples show that the main phase obtained corresponds to anatase (PDF #21-1272). The diffractogram of the sample that was milled without the addition of bismuth salt (milled TiO2) mainly presents the diffraction peaks of the anatase phase, and a small diffraction peak at 2θ equal to 30.8°, which is assigned to the plane (121) of the TiO2 brookite phase (PDF #29-1360). It is well known that the energy produced by mechanical grinding promotes the formation of structural defects [36], which at the same time leads to the partial transformation of the anatase phase to the brookite phase. The incorporation of the Bi3+ dopant into the crystal lattice of anatase also contributes to the formation of brookite, as can be seen in the diffractograms of TiO2 samples Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi 10%. It is known that phase transformations can occur when oxygen vacations are formed, especially on the surface of the material, where the breaking of bonds is favored, thus resulting in a rear coupling of atoms [37].
The presence of brookite in the Bi-doped TiO2 samples has previously been observed by doping with other types of atoms such as Sn [38] and even Bi [35]. Although it is possible to form other polymorphs such as rutile, nitrate ions (which are present in bismuth nitrate salt) may facilitate the formation of brookite [39]. On the other hand, a weak diffraction peak corresponding to a bismuth oxide (Bi2O3, PDF #50-1088) in the diffractogram of the sample TiO2 Bi 10% is observed (Figure 6A), which is because a greater number of atoms of this element were used. With increasing Bi content, the peaks slightly broaden and the crystallite size is reduced, indicating a slight distortion in the crystal structure. This may be attributed to the formation of crystallographic point defects due to the substitution of Ti4+ by Bi3+ ions. Most bismuth ions might be inserted into the structures of TiO2 located at the interstices or might occupy some of the titanium sites to form a solution with bismuth–titanium solids. It has been reported that, because the ion radius of the Bi3+ (1.03 Å) is greater than that of the Ti4+ (0.61 Å) [40,41], it is first introduced in the interstitials of the crystal lattice of the anatase, causing a contraction in the lattice parameters [35]. The higher the degree of doping, the more bismuth is placed in the sites corresponding to Ti4+, which causes the expansion of the unit cell. The latter effect has previously been reported when TiO2 is doped with bismuth using the sol–gel method [42].
An amplification was performed at the highest intensity peak of the diffractograms, which corresponds to the plane (101) of the anatase (Figure 6B); from this figure, it is possible to appreciate its displacement to 2θ degrees greater than 2 with little dopant and then to smaller 2θ degrees when we increase the amount of bismuth, which is also indicative of the modification of the unit cell of the anatase. The shift to degrees greater than 2θ may also be related to the presence of Bi0 on the sample surface [31]. The TiO2 Bi 1% and TiO2 Bi 5% samples contain Bi0, based on the XPS spectra presented in Figure 2C.
To corroborate the above, an analysis by the Rietveld method of structural refinement was performed on pristine TiO2, TiO2 Bi0% (milled TiO2), and on the Bi-doped TiO2 samples. Figure 7 shows the adjustments of pure anatase, as well as samples subjected to mechanical grinding with and without doping. Table 2 shows the lattice parameters of the anatase present in each of the synthesized samples. The table shows that the TiO2 Bi 1% sample has a decrease in cell volume, while samples loaded with 3% and 5% of bismuth have an expansion in the unit cell, since, as mentioned above, the unit cell of the anatase is distorted due to the incorporation of Bi3+ atom, which is larger than that of Ti4+. According to the results of the Rietveld refinement, the quantities of the doping incorporated in the TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi10% samples were 1%, 3%, and 4%, respectively, so, the estimated chemical formula for these samples are Ti0.99Bi0.01O2, Ti0.97Bi0.03O2, Ti0.96Bi0.04O2, and Ti0.91Bi0.02O2 respectively.
The lattice parameters and cell volume of the milled TiO2 sample were smaller than those of the pristine TiO2 sample (Table 2), indicating that grinding causes distortion of the crystal lattice of the TiO2 anatase phase, causing the formation of TiO2 brookite phase by 17%, and an increase in crystallite size. Table 2 also shows that the percentage of TiO2 phase brookite was between 7% and 19.05% for doped samples with 1%, 3%, and 5% of Bi; the brookite phase was not formed in the TiO2 Bi 10% sample, but bismuth oxide (B2O3) was formed in a small amount. It has previously been reported that TiO2 phase mixtures may be beneficial for photocatalysis processes, as the recombination of electron–hollow pairs can be reduced [43].
The average crystallite sizes of TiO2 anatase in Bi-doped samples were determined from the Debye-Scherrer equation, Equation (1):
D = K λ β cos   θ
where D is the size of crystallite, K is a dimensionless shape factor (0.89), λ is the wavelength of Cu Kα radiation having a value of 1.5406 Å, β is the broadening of the peak of higher intensity (101), and θ is the angle of X-ray diffraction. The average crystallite size of all the samples was 18.28, 19.11, 18.12, 18.05, 17.78, and 17.15 nm, corresponding to the pristine TiO2, milled TiO2, TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi 10% samples, respectively. It is observed that the crystallite size of anatase increased from 18.28 nm (TiO2 pristine) to 19.11 nm (milled TiO2) by grinding action, which is because the energy that is generated causes changes in the crystal lattice of TiO2, decreasing the lattice parameter c and the cell volume; such a change in the lattice in turn causes brookite to form, as can be seen in Table 2 and Figure 6A. On the other hand, it is observed that the average crystallite size decreased with an increasing concentration of Bi. Since the ionic radius of Bi3+ is different from that of Ti4+, an increase in Bi concentration leads to reduced intensity with increased full width at the half-maximum of the Bragg peak. Furthermore, it is very challenging to substitute smaller-radius Ti4+ (0.61 Å) with larger-radius Bi3+ (1.03Å). The substitution of Ti4+ for Bi3+ might create oxygen vacancies for the charge compensation in the TiO2 lattice. Moreover, besides Bi3+ at the Ti4+ sites, a large fraction of ions may stay on the surface of particle and, hence, the doping of Bi3+ can inhibit the grain growth of the TiO2, which typically leads to a reduction in crystallite size. Ali et al., (2020) reported this same behavior in TiO2 samples doped with Mg2+ [44]. Xu et al., (2011) obtained Bi-doped TiO2 composite hollow spheres by the sol–gel method, and reported a decrease in the crystallite size of the anatase from 16.7 to 10.7 nm when increasing the content of Bi3+, which was because the bismuth species located at the edges of the crystallite prevented growth [24]. Similar crystallite sizes up to 17 nm were reported by Murcia-Lopez et al., (2011), who synthesized TiO2 doped with Bi3+ by the hydrothermal method [45]. Compared to the hydrothermal and sol–gel methods, mechanosynthesis has the advantages of being inexpensive, solvent-free, and applicable on an industrial scale.
Figure 8A presents the absorption spectra of the Bi-doped TiO2 samples; absorption spectra of pristine and milled TiO2 are also included. With the incorporation of Bi3+ into the TiO2 crystal lattice, it is possible to increase the absorption of light in the visible range of the electromagnetic spectrum; the samples with the highest absorption of light in the visible region are those containing a greater amount of the dopant. The shift of the absorption edge to the visible region indicates the modification of the Eg value; the values of this optical property were calculated using the Tauc plot (Figure 8B) considering n = 2 for allowed indirect transitions. The calculated values of Eg are listed in Table 3 and it is seen that the value of Eg decreased with an increasing concentration of Bi in the TiO2 crystal lattice. The shift in the absorption band edge in doped TiO2 samples might also be due to the shallow trap states created by oxygen vacancies associated with Ti3+ and Bi3+.
It has been reported that the presence of vacancy states just below the conduction band is related to the oxygen vacancy associated with Ti3+ and is responsible for the band narrow gap in TiO2 [46]. In this work, doping with Bi is the main reason for lowering the Eg of TiO2.
Table 3. Eg values of Bi-doped TiO2 samples obtained by mechanosynthesis compared with the reported values in the literature for Bi-doped TiO2 samples obtained with other methods.
Table 3. Eg values of Bi-doped TiO2 samples obtained by mechanosynthesis compared with the reported values in the literature for Bi-doped TiO2 samples obtained with other methods.
SampleEg (eV)Synthesis Method
TiO2 Bi 1% (This work)2.90Mechanosynthesis
TiO2 Bi 3% (This work)2.83Mechanosynthesis
TiO2 Bi 5% (This work)2.77Mechanosynthesis
TiO2 Bi 10% (This work)2.70Mechanosynthesis
Bi-doped TiO2: 0.6–1.7% [23]3.16–2.97Hydrothermal
Bi/Ti: 0.25–1 [42]3.04–2.90Sol gel/supercritical drying
Bi-doped TiO2: 0.25–5% [47]3.19–2.97Sol gel/Ultrasound assisted
Bi-doped TiO2: 0–10% [48]3.08–2.80Sol gel/assisted with propylene oxide
Table 3 also lists Eg values for Bi3+ anatase samples that were obtained by other methods. This work reported obtaining Bi-doped TiO2 samples with lower Eg values than obtained by the hydrothermal method of the sol–gel method assisted with supercritical drying and ultrasound [23,42,47,48]. The Eg values of Bi-doped TiO2 samples obtained by mechanosynthesis in this work are even lower than those obtained by this same method, using dopants such as sulfur and silver. Ji et al., (2009) prepared doped sulfide TiO2 nanoparticles by ball milling, using as commercial precursors TiO2 P25 and thiourea, and obtaining Eg values of 3.1 eV for the undoped sample and values of 2.76 and 2.83 eV for the TiO2 P25 samples doped with 5% and 11% sulfur [49]. Ellouzi et al., (2021) prepared silver-doped biphasic TiO2 composites via the glucose-assisted ball milling method; both anatase and rutile phases were identified in the obtained sample, and the best Eg value was 3.65 eV for the doped sample with 3% silver [50].
It is well known that the crystallite size of a semiconductor oxide affects its Eg value: the smaller the crystallite size, the lower the value of Eg, and vice versa. The crystallite size and the Eg value of the Bi-doped TiO2 samples decreased almost linearly with the increase in the amount of bismuth used for synthesis, as can be seen in Figure 9, which indicates that the incorporation of bismuth into the TiO2 lattice causes the crystallite size to decrease and, consequently, the absorption spectrum of TiO2 to extend into the visible region of the electromagnetic spectrum, resulting in a lower Eg value. It is also important to consider that doping promotes the creation of new electronic levels in the band structure; possibly, this also contributed to the change in Eg values.
Figure 10 presents an outline of how mechanical grinding provides the energy needed for the formation of defects, such as oxygen vacancies, into the tetragonal structure of anatase, and, together with the heating treatment, promotes the incorporation of bismuth into the crystalline lattice of TiO2, causing a distortion in the lattice parameters of the cell.

3.2. Photocatalytic Activity

The photocatalytic activity of the as-synthesized mesoporous Bi-doped TiO2 samples was investigated by degrading MB. Photodegradation of this dye with mesoporous Bi-doped samples (TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi 10%) and the pristine TiO2 sample, using visible light as a function of irradiation time, is shown in Figure 11A. MB photodegradation percentages of 38%, 54%, 60%, 74%, and 63% were obtained with pristine TiO2, TiO2 Bi 1%, TiO2 Bi 3%, TiO2 Bi 5%, and TiO2 Bi 10%, respectively, at 60 min of irradiation (Figure 11B). The first- and second-order kinetic models (Equations (2) and (3), respectively) were used to investigate the kinetics of MB photodegradation:
ln ( C t | C 0 ) = K 1   t
1 C t = K 2 t + 1 C 0
where C0 and Ct (mol/L) are the MB concentrations at times 0 and t (min), respectively; K1 (min−1) and K2 (mg−1· L· min−1) are the first- and second-order rate constants, respectively.
The variation in –ln(Co/Ct) and 1/Ct as a function of irradiation time is shown in Figure 11B,C, for first- and second-order models, respectively, by which the rate constants (K1 and K2) were calculated and listed in Table 4. Photocatalytic removal of MB data fit well for second-order reaction kinetics. The rate constant values, K2, followed the order TiO2 Bi 5% > TiO2 Bi 10% > TiO2 Bi 3% > TiO2 Bi 1% > pristine TiO2; enhanced removal of MB from solution under visible light irradiation is mainly ascribed to bismuth doping. The excitation energy was extended from the UV region to the visible region and the absorbance of the samples at a wavelength of 467 nm increased following the order TiO2 Bi 5% > TiO2 Bi 10% > TiO2 Bi 3% > TiO2 Bi 1% > pristine TiO2, which is equal to the increasing order of the photocatalytic activity. The results from photocatalytic experiments agree with the absorbance spectra (Figure 8A). The TiO2 Bi 5% sample had the highest photocatalytic activity (K2 = 2.02 × 10−3 mg−1·L·min−1) because it has a low Eg value (2.77 eV) and higher absorbance in the visible region.
Figure 12A shows that the photocatalytic activity of the TiO2 Bi 5% sample to degrade MB under visible light does not depend on the catalyst concentration in the range studied from 0.5 g/L to 2.0 g/L; it was also observed that the photocatalytic activity of this sample was higher than the P25 catalyst, reaching a photodegradation of 75.87% and 37.5%, respectively, using a catalyst concentration of 1 g/L; this is due to the fact that the P25 catalyst does not absorb photons in the visible region, it only absorbs in the UV region. The effect of the type of light used on MB photodegradation with the TiO2 Bi 5% sample is shown in Figure 12B; it is observed that photodegradation percentages of 79.5%, 74.4%, and 89.2% were reached using UV light, visible light, and natural sunlight, respectively. The highest percentage of MB photodegradation was obtained with natural sunlight irradiation, since the TiO2 Bi 5% sample had higher absorbance in the visible region of the electromagnetic spectrum, from 400 to 700 nm (Figure 8A). As already mentioned, the higher photocatalytic activity of the TiO2 Bi 5% sample is mainly associated with its higher level of bismuth doping and with its low Eg value.

4. Conclusions

In this work, mesoporous Bi-doped TiO2 catalysts were successfully prepared by mechanosynthesis. This green method allowed for the incorporation of bismuth into the crystal lattice of the TiO2 anatase phase by up to 9%, achieving interstitial doping when small amounts of bismuth are used, as well as substitute doping when the amount of this atom was increased. This was demonstrated with the results of Rietveld refinement, which indicated a contraction of the unit cell of the anatase and subsequent expansion of the cell. The presence of brookite phase in the Bi-doped TiO2 samples is due to the formation of defects caused by mechanical grinding and by the incorporation of bismuth into the TiO2 crystal lattice. The crystallite size of the TiO2 phase anatase decreased from 18.28 nm to 17.5 nm by increasing the amount of Bi used from 0% to 10%, and this in turn caused a decrease in the Eg value from 2.97 eV to 2.70 eV. The results of bond energy characterization by XPS suggest that Bi-doped TiO2 samples have Bi0 on their surface and that there are OH groups on the surface of pristine TiO2 and Bi-doped TiO2 samples. The Bi-doped TiO2 mesoporous samples obtained by mechanosynthesis in this work have good porosity and adequate Eg values that allow them to have excellent photodegradation of MB under UV, visible, and natural solar irradiation. The higher photocatalytic activity of the TiO2 Bi 5% sample under visible and natural sunlight irradiation (79.5% and 89.2%, respectively) is mainly associated with its higher level of bismuth doping (4%) and with its low Eg value (2.77 eV). The absorbance of the samples at a wavelength of 467 nm increased following the order: TiO2 Bi5% > TiO2 Bi10% > TiO2 Bi3% > TiO2 Bi1% > TiO2 pristine, which is equal to the increasing order of the photocatalytic activity.

Author Contributions

Investigation, methodology, formal analysis, writing—original draft, S.E.-F.; Supervision, conceptualization, investigation, writing—reviewing, editing, resources, and project administration, A.M.-L.; Validation and visualization, C.M.P.-B.; Resources and formal analysis, L.A.G.-C.; Resources and formal analysis, T.E.F.-G.; Software, J.R.-H.; Visualization, T.A.E.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not Applicable.

Acknowledgments

Sofía Estrada-Flores acknowledges CONACYT for the scholarship (446796). Antonia Martinez-Luevanos thanks Universidad Autonoma de Coahuila in Mexico for the financial support for this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Narayan, M.R. Review: Dye sensitized solar cells based on natural photosensitizers. Renew. Sustain. Energy Rev. 2012, 16, 208–215. [Google Scholar] [CrossRef]
  2. Bilgin Simsek, E. Solvothermal synthesized boron doped TiO2 catalysts: Photocatalytic degradation of endocrine disrupting compounds and pharmaceuticals under visible light irradiation. Appl. Catal. B Environ. 2017, 200, 309–322. [Google Scholar] [CrossRef]
  3. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Environmental Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  4. Yang, Z.; Choi, D.; Kerisit, S.; Rosso, K.M.; Wang, D.; Zhang, J.; Graff, G.; Liu, J. Nanostructures and lithium electrochemical reactivity of lithium titanites and titanium oxides: A review. J. Power Sources 2009, 192, 588–598. [Google Scholar] [CrossRef]
  5. Coronado, J.M.; Hernández-Alonso, M.D. The Keys of Success: TiO2 as a Benchmark Photocatalyst. In Design of Advanced Photocatalytic Materials for Energy and Environmental Applications; Coronado, J.M., Fresno, F., Hernández-alonso, M.D., Portela, R., Eds.; Springer: London, UK, 2013; pp. 85–101. ISBN 1865-3537. [Google Scholar]
  6. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M.A. Principles and mechanisms of photocatalytic dye degradation on TiO2 based photocatalysts: A comparative overview. RSC Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  7. Bian, Y.; Zheng, G.; Ding, W.; Hu, L.; Sheng, Z. Magnetic field effect on the photocatalytic degradation of methyl orange by commercial TiO2 powder. RSC Adv. 2021, 11, 6284–6291. [Google Scholar] [CrossRef] [PubMed]
  8. De Araujo Scharnberg, A.R.; Carvalho de Loreto, A.; Bender Wermuth, T.; Kopp Alves, A.; Arcaro, S.; Mantey dos Santos, P.A.; Lawisch Rodriguez, A.A. Porous ceramic supported TiO2 nanoparticles: Enhanced photocatalytic activity for Rhodamine B degradation. Bol. Soc. Esp. Ceram. Vidr. 2020, 59, 230–238. [Google Scholar] [CrossRef]
  9. Miao, J.; Zhang, R.; Zhang, L. Photocatalytic degradations of three dyes with different chemical structures using ball-milled TiO2. Mater. Res. Bull. 2018, 97, 109–114. [Google Scholar] [CrossRef]
  10. Prabhudesai, V.S.; Meshram, A.A.; Vinu, R.; Sontakke, S.M. Superior photocatalytic removal of metamitron and its mixture with Rhodamine B dye using combustion synthesized TiO2 nanomaterial. Chem. Eng. J. Adv. 2021, 5, 100084. [Google Scholar] [CrossRef]
  11. Zhang, G.; Zhang, S.; Wang, L.; Liu, R.; Zeng, Y.; Xia, X.; Liu, Y.; Luo, S. Facile synthesis of bird’s nest-like TiO2 microstructure with exposed (001) facets for photocatalytic degradation of methylene blue. Appl. Surf. Sci. 2017, 391, 228–235. [Google Scholar] [CrossRef]
  12. Bagheri, S.; Termehyousefi, A.; Do, T.O. Photocatalytic pathway toward degradation of environmental pharmaceutical pollutants: Structure, kinetics and mechanism approach. Catal. Sci. Technol. 2017, 7, 4548–4569. [Google Scholar] [CrossRef]
  13. Lin, C.-J.; Yang, W.-T. Ordered mesostructured Cu-doped TiO2 spheres as active visible-light-driven photocatalysts for degradation of paracetamol. Chem. Eng. J. 2013, 237, 131–137. [Google Scholar] [CrossRef]
  14. Meroni, D.; Jiménez-Salcedo, M.; Falletta, E.; Bresolin, B.M.; Kait, C.F.; Boffito, D.C.; Bianchi, C.L.; Pirola, C. Sonophotocatalytic degradation of sodium diclofenac using low power ultrasound and micro sized TiO2. Ultrason. Sonochem. 2020, 67, 105123. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, P.; Zhou, T.; Wang, R.; Lim, T.-T. Carbon-sensitized and nitrogen-doped TiO2 for photocatalytic degradation of sulfanilamide under visible-light irradiation. Water Res. 2011, 45, 5015–5026. [Google Scholar] [CrossRef] [PubMed]
  16. Aznar-Cervantes, S.; Aliste, M.; Garrido, I.; Yañez-Gascón, M.J.; Vela, N.; Cenis, J.L.; Navarro, S.; Fenoll, J. Electrospun silk fibroin/TiO2 mats. Preparation, characterization and efficiency for the photocatalytic solar treatment of pesticide polluted water. RSC Adv. 2020, 10, 1917–1924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Ilina, S.M.; Ollivier, P.; Slomberg, D.; Baran, N.; Pariat, A.; Devau, N.; Sani-Kast, N.; Scheringer, M.; Labille, J. Investigations into titanium dioxide nanoparticle and pesticide interactions in aqueous environments. Environ. Sci. Nano 2017, 4, 2055–2065. [Google Scholar] [CrossRef]
  18. Mirmasoomi, S.R.; Ghazi, M.M.; Galedari, M. Photocatalytic degradation of diazinon under visible light using TiO2/Fe2O3 nanocomposite synthesized by ultrasonic-assisted impregnation method. Sep. Purif. Technol. 2016, 175, 418–427. [Google Scholar] [CrossRef]
  19. Le, T.T.T.; Tran, D.T.; Danh, T.H. Remarkable enhancement of visible light driven photocatalytic performance of TiO2 by simultaneously doping with C, N, and S. Chem. Phys. 2021, 545, 111144. [Google Scholar] [CrossRef]
  20. Lin, C.J.; Liou, Y.H.; Zhang, Y.; Chen, C.L.; Dong, C.L.; Chen, S.Y.; Stucky, G.D. Mesoporous Fe-doped TiO2 sub-microspheres with enhanced photocatalytic activity under visible light illumination. Appl. Catal. B Environ. 2012, 127, 175–181. [Google Scholar] [CrossRef]
  21. Sun, Z.; Pichugin, V.F.; Evdokimov, K.E.; Konishchev, M.E.; Syrtanov, M.S.; Kudiiarov, V.N.; Li, K.; Tverdokhlebov, S.I. Effect of nitrogen-doping and post annealing on wettability and band gap energy of TiO2 thin film. Appl. Surf. Sci. 2020, 500, 144048. [Google Scholar] [CrossRef]
  22. 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] [PubMed]
  23. Wang, W.; Zhu, D.; Shen, Z.; Peng, J.; Luo, J.; Liu, X. One-Pot hydrothermal route to synthesize the Bi-doped anatase TiO2 hollow thin sheets with prior facet exposed for enhanced visible-light-driven photocatalytic activity. Ind. Eng. Chem. Res. 2016, 55, 6373–6383. [Google Scholar] [CrossRef]
  24. Xu, J.; Chen, M.; Fu, D. Study on highly visible light active Bi-doped TiO2 composite hollow sphere. Appl. Surf. Sci. 2011, 257, 7381–7386. [Google Scholar] [CrossRef]
  25. Baláž, P. Mechanochemistry in Nanoscience and Minerals Enginnering, 1st ed.; Springer: Berlin/Heidelberg, Germany, 2008; ISBN 978-3-540-74855-7. [Google Scholar]
  26. Estrada-Flores, S.; Martínez-Luévanos, A.; Perez-Berumen, C.M.; García-Cerda, L.A.; Flores-Guia, T.E. Relationship between morphology, porosity, and the photocatalytic activity of TiO2 obtained by sol–gel method assisted with ionic and nonionic surfactants. Bol. Soc. Esp. Ceram. Vidr. 2020, 59, 209–218. [Google Scholar] [CrossRef]
  27. Primet, M.; Pichat, P.; Mathieu, M.-V. Infrared study of the surface of titanium dioxides. J. Phys. Chem. 1971, 75, 1216–1220. [Google Scholar] [CrossRef]
  28. Li, J.-J.; Cai, S.-C.; Xu, Z.; Chen, X.; Chen, J.; Jia, H.-P.; Chen, J. Solvothermal syntheses of Bi and Zn co-doped TiO2 with enhanced electron-hole separation and efficient photodegradation of gaseous toluene under visible-light. J. Hazard. Mater. 2017, 325, 261–270. [Google Scholar] [CrossRef]
  29. Li, H.; Wang, D.; Wang, P.; Fan, H.; Xie, T. Synthesis and studies of the visible-light photocatalytic properties of near-monodisperse Bi-doped TiO2 nanospheres. Chem. A Eur. J. 2009, 15, 12521–12527. [Google Scholar] [CrossRef]
  30. Wu, Y.; Lu, G.; Li, S. The Doping Effect of Bi on TiO2 for Photocatalytic Hydrogen Generation and Photodecolorization of Rhodamine B. J. Phys. Chem. C 2009, 113, 9950–9955. [Google Scholar] [CrossRef]
  31. An’amt, M.N.; Radiman, S.; Huang, N.M.; Yarmo, M.A.; Ariyanto, N.P.; Lim, H.N.; Muhamad, M.R. Sol-gel hydrothermal synthesis of bismuth-TiO2 nanocubes for dye-sensitized solar cell. Ceram. Int. 2010, 36, 2215–2220. [Google Scholar] [CrossRef]
  32. Hao, W.; Teng, F.; Liu, Z.; Yang, Z.; Liu, Z.; Gu, W. Synergistic effect of Mott-Schottky junction with oxygen defect and dramatically improved charge transfer and separation of Bi@Bi-doped TiO2. Sol. Energy Mater. Sol. Cells 2019, 203, 110198. [Google Scholar] [CrossRef]
  33. Ma, J.; Chu, J.; Qiang, L.; Xue, J. Synthesis and structural characterization of novel visible photocatalyst Bi-TiO2/SBA-15 and its photocatalytic performance. RSC Adv. 2012, 2, 3753–3758. [Google Scholar] [CrossRef]
  34. Li, H.; Liu, J.; Qian, J.; Li, Q.; Yang, J. Preparation of Bi-doped TiO2 nanoparticles and their visible light photocatalytic performance. Chin. J. Catal. 2014, 35, 1578–1589. [Google Scholar] [CrossRef]
  35. Henych, J.; Štengl, V.; Kormunda, M.; Mattsson, A.; Österlund, L. Role of bismuth in nano-structured doped TiO2 photocatalyst prepared by environmentally benign soft synthesis. J. Mater. Sci. 2014, 49, 3560–3571. [Google Scholar] [CrossRef]
  36. Subramonian, W.; Wu, T.Y.; Chai, S. Using one-step facile and solvent-free mechanochemical process to synthesize photoactive Fe2O3 -TiO2 for treating industrial wastewater. J. Alloys Compd. 2017, 695, 496–507. [Google Scholar] [CrossRef] [Green Version]
  37. Garza-Arévalo, J.I.; García-Montes, I.; Hinojosa-Reyes, M.; Guzmán-Mar, J.L.; Rodríguez-González, V.; Hinojosa-Reyes, L. Fe doped TiO2 photocatalyst for the removal of As(III) under visible radiation and its potential application on the treatment of As-contaminated groundwater. Mater. Res. Bull. 2016, 73, 145–152. [Google Scholar] [CrossRef]
  38. Štengl, V.; Grygar, T.M.; Henych, J.; Kormunda, M. Hydrogen peroxide route to Sn-doped titania photocatalysts. Chem. Cent. J. 2012, 6, 113. [Google Scholar] [CrossRef] [Green Version]
  39. Cassaignon, S.; Koelsch, M.; Jolivet, J.P. Selective synthesis of brookite, anatase and rutile nanoparticles: Thermolysis of TiCl4 in aqueous nitric acid. J. Mater. Sci. 2007, 42, 6689–6695. [Google Scholar] [CrossRef]
  40. Delekar, S.D.; Yadav, H.M.; Achary, S.N.; Meena, S.S.; Pawar, S.H. Structural refinement and photocatalytic activity of Fe-doped anatase TiO2 nanoparticles. Appl. Surf. Sci. 2012, 263, 536–545. [Google Scholar] [CrossRef]
  41. Zhang, L.; Ma, F.; Guan, Q.; Wang, C.; Bai, C.; Sheng, L. Fluorescence enhanced ultrathin nano-plate Gd2O2SO4:Bi3+, Eu3+ transformed from layered gadolinium hydroxide. J. Alloys Compd. 2019, 802, 173–180. [Google Scholar] [CrossRef]
  42. Ma, Y.; Yang, X.; Gao, G.; Yan, Z.; Su, H.; Zhang, B.; Lei, Y.; Zhang, Y. Photocatalytic partial oxidation of methanol to methyl formate under visible light irradiation on Bi-doped TiO2: Via tuning band structure and surface hydroxyls. RSC Adv. 2020, 10, 31442–31452. [Google Scholar] [CrossRef]
  43. Henderson, M.A. A surface science perspective on TiO2 photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  44. Ali, T.; Ahmed, A.; Siddique, M.N.; Alam, U.; Muneer, M.; Tripathi, P. Influence of Mg2+ ion on the optical and magnetic properties of TiO2 nanostructures: A key role of oxygen vacancy. Optik 2020, 223, 165340. [Google Scholar] [CrossRef]
  45. Murcia-López, S.; Hidalgo, M.C.; Navío, J.A. Synthesis, characterization and photocatalytic activity of Bi-doped TiO2 photocatalysts under simulated solar irradiation. Appl. Catal. A Gen. 2011, 404, 59–67. [Google Scholar] [CrossRef]
  46. Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.; Borchardt, D.; Feng, P. Self-doped Ti3+ enhanced photocatalyst for hydrogen production under visible light. J. Am. Chem. Soc. 2010, 132, 11856–11857. [Google Scholar] [CrossRef] [PubMed]
  47. Sood, S.; Mehta, S.K.; Umar, A.; Kansal, S.K. The visible light-driven photocatalytic degradation of Alizarin red S using Bi-doped TiO2 nanoparticles. New J. Chem. 2014, 38, 3127–3136. [Google Scholar] [CrossRef]
  48. Alzamly, A.; Hamed, F.; Ramachandran, T.; Bakiro, M.; Ahmed, S.H.; Mansour, S.; Salem, S.; Abdul al, K.; Al Kaabi, N.S.; Meetani, M.; et al. Tunable band gap of Bi3+—doped anatase TiO2 for enhanced photocatalytic removal of acetaminophen under UV-visible light irradiation. J. Water Reuse Desalin. 2019, 9, 31–46. [Google Scholar] [CrossRef] [Green Version]
  49. Ji, T.; Yang, F.; Lv, Y.; Zhou, J.; Sun, J. Synthesis and visible-light photocatalytic activity of Bi-doped TiO2 nanobelts. Mater. Lett. 2009, 63, 2044–2046. [Google Scholar] [CrossRef]
  50. Ellouzi, I.; Bouddouch, A.; Bakiz, B.; Benlhachemi, A.; Abou Oualid, H. Glucose-assisted ball milling preparation of silver-doped biphasic TiO2 for efficient photodegradation of Rhodamine B: Effect of silver-dopant loading. Chem. Phys. Lett. 2021, 770, 138456. [Google Scholar] [CrossRef]
Figure 1. FTIR-ATR spectra of the pristine TiO2, milled TiO2, and Bi-doped TiO2 samples.
Figure 1. FTIR-ATR spectra of the pristine TiO2, milled TiO2, and Bi-doped TiO2 samples.
Crystals 12 01750 g001
Figure 2. XPS spectra of pristine TiO2 and Bi-doped TiO2 samples (A). Spectra of Ti 2p (B), Bi 4f (C), and O 1s (D) with deconvolution for pristine TiO2 and Bi-doped TiO2 samples.
Figure 2. XPS spectra of pristine TiO2 and Bi-doped TiO2 samples (A). Spectra of Ti 2p (B), Bi 4f (C), and O 1s (D) with deconvolution for pristine TiO2 and Bi-doped TiO2 samples.
Crystals 12 01750 g002
Figure 3. SEM images at 10,000× of the pristine TiO2 (A), milled TiO2 (B), and Bi-doped TiO2 samples (CF). A SEM image at 100,000× of the pristine TiO2 sample is inserted in (A).
Figure 3. SEM images at 10,000× of the pristine TiO2 (A), milled TiO2 (B), and Bi-doped TiO2 samples (CF). A SEM image at 100,000× of the pristine TiO2 sample is inserted in (A).
Crystals 12 01750 g003
Figure 4. Elemental mapping of Ti, Bi, and O in the TiO2 Bi 3% sample.
Figure 4. Elemental mapping of Ti, Bi, and O in the TiO2 Bi 3% sample.
Crystals 12 01750 g004
Figure 5. N2 adsorption–desorption isotherms and pore diameter distribution of the pristine TiO2, milled TiO2, TiO2 Bi 1%, and TiO2 Bi 5% samples.
Figure 5. N2 adsorption–desorption isotherms and pore diameter distribution of the pristine TiO2, milled TiO2, TiO2 Bi 1%, and TiO2 Bi 5% samples.
Crystals 12 01750 g005
Figure 6. XRD patterns of the TiO2 and the milled samples with different bismuth content (A), and amplification of the (101) peak from the anatase patterns (B).
Figure 6. XRD patterns of the TiO2 and the milled samples with different bismuth content (A), and amplification of the (101) peak from the anatase patterns (B).
Crystals 12 01750 g006
Figure 7. Observed (red line), calculated (black line), differences profiles (blue line), and Bragg positions (green) for the Rietveld refinement for the pristine TiO2, milled TiO2, and TiO2 samples with different bismuth content.
Figure 7. Observed (red line), calculated (black line), differences profiles (blue line), and Bragg positions (green) for the Rietveld refinement for the pristine TiO2, milled TiO2, and TiO2 samples with different bismuth content.
Crystals 12 01750 g007
Figure 8. UV–vis diffuse reflectance absorption spectra (A) and Tauc plot (B) of pristine TiO2, milled TiO2, and Bi-doped TiO2 samples.
Figure 8. UV–vis diffuse reflectance absorption spectra (A) and Tauc plot (B) of pristine TiO2, milled TiO2, and Bi-doped TiO2 samples.
Crystals 12 01750 g008
Figure 9. Effect of bismuth content loading on the crystallite size and band gap (Eg) of the Bi-doped TiO2 samples obtained by mechanosynthesis.
Figure 9. Effect of bismuth content loading on the crystallite size and band gap (Eg) of the Bi-doped TiO2 samples obtained by mechanosynthesis.
Crystals 12 01750 g009
Figure 10. Scheme of the effect of grinding and bismuth doping on the crystalline lattice of TiO2 anatase phase.
Figure 10. Scheme of the effect of grinding and bismuth doping on the crystalline lattice of TiO2 anatase phase.
Crystals 12 01750 g010
Figure 11. Photocatalytic activity (A) and photodegradation percentage (B) of MB under visible light by pristine TiO2 and Bi-doped TiO2 samples as a function of irradiation time; fitting curves of the kinetics of first- (C) and second-order (D) MB photodegradation.
Figure 11. Photocatalytic activity (A) and photodegradation percentage (B) of MB under visible light by pristine TiO2 and Bi-doped TiO2 samples as a function of irradiation time; fitting curves of the kinetics of first- (C) and second-order (D) MB photodegradation.
Crystals 12 01750 g011
Figure 12. MB photodegradation as a function of TiO2 Bi 5% photocatalyst concentration under visible light (A); MB photodegradation with P25 is included. Effect of the type of irradiation on MB photodegradation using 0.5 g/L of TiO2 Bi 5% photocatalyst and an irradiation time of 60 min (B).
Figure 12. MB photodegradation as a function of TiO2 Bi 5% photocatalyst concentration under visible light (A); MB photodegradation with P25 is included. Effect of the type of irradiation on MB photodegradation using 0.5 g/L of TiO2 Bi 5% photocatalyst and an irradiation time of 60 min (B).
Crystals 12 01750 g012
Table 1. Specific surface area (SBET), mean pore diameter (Dp) and mean pore volume (Vp) of the samples pristine TiO2, milled TiO2, TiO2 Bi 1%, and TiO2 Bi 5%.
Table 1. Specific surface area (SBET), mean pore diameter (Dp) and mean pore volume (Vp) of the samples pristine TiO2, milled TiO2, TiO2 Bi 1%, and TiO2 Bi 5%.
SampleSBET (m2/g)Dp (nm)Vp (cm3/g)
pristine TiO2 138.75.860.300
milled TiO2 74.03.500.191
TiO2 Bi 1%116.73.400.196
TiO2 Bi 5%59.13.480.201
Table 2. Parameters of the anatase structure for the samples.
Table 2. Parameters of the anatase structure for the samples.
Pristine TiO2 Milled TiO2 TiO2 Bi 1%TiO2 Bi 3%TiO2 Bi 5%TiO2 Bi 10%
Space groupI41/amdI41/amdI41/amdI41/amdI41/amdI41/amd
a (Å)3.7829 (3)3.7837 (16)3.7765 (2)3.7858 (1)3.7853 (10)3.7842 (3)
c (Å)9.4930 (9)9.4834 (6)9.4715 (3)9.4910 (6)9.4868 (10)9.482 (13)
V (Å3)135.847135.768135.08136.02135.915135.783
Formula TiO2TiO2Ti0.99Bi0.01O2Ti0.97Bi0.03O2Ti0.96Bi0.04O2Ti0.91Bi0.09O2
Rp, Rwp, Rexp8.30, 9.13, 4.347.60, 8.35, 4.5611.5, 11.3, 4.909.05, 9.44, 4.703.78, 4.94, 2.6818.4, 16.5, 13.1
X24.433.405.294.033.401.59
% Brookite01711.85719.050
Table 4. Rate constants for kinetic studies of MB photodegradation using pristine TiO2 and Bi-doped TiO2 photocatalysts under visible light (λ = 467 nm, 1.8 Watts); 100 mL of MB solution of 20 mg/L; 1.0 g/L of catalyst concentration.
Table 4. Rate constants for kinetic studies of MB photodegradation using pristine TiO2 and Bi-doped TiO2 photocatalysts under visible light (λ = 467 nm, 1.8 Watts); 100 mL of MB solution of 20 mg/L; 1.0 g/L of catalyst concentration.
SampleFirst-OrderSecond-Order
K1 (min−1) R2K2 (mg−1·L·min−1) R2
pristine TiO2 7.51 × 10−30.85025.08 × 10−40.8796
TiO2 Bi 1%8.98 × 10−30.91166.88 × 10−40.9437
TiO2 Bi 3%1.1910−20.94821.09 × 10−30.9777
TiO2 Bi 5%1.84 × 10−20.97962.02 × 10−30.9968
TiO2 Bi 10%1.4210−20.96521.37 × 10−30.9877
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Estrada-Flores, S.; Pérez-Berumen, C.M.; Flores-Guia, T.E.; García-Cerda, L.A.; Rodríguez-Hernández, J.; Esquivel-Castro, T.A.; Martínez-Luévanos, A. Mechanosynthesis of Mesoporous Bi-Doped TiO2: The Effect of Bismuth Doping and Ball Milling on the Crystal Structure, Optical Properties, and Photocatalytic Activity. Crystals 2022, 12, 1750. https://doi.org/10.3390/cryst12121750

AMA Style

Estrada-Flores S, Pérez-Berumen CM, Flores-Guia TE, García-Cerda LA, Rodríguez-Hernández J, Esquivel-Castro TA, Martínez-Luévanos A. Mechanosynthesis of Mesoporous Bi-Doped TiO2: The Effect of Bismuth Doping and Ball Milling on the Crystal Structure, Optical Properties, and Photocatalytic Activity. Crystals. 2022; 12(12):1750. https://doi.org/10.3390/cryst12121750

Chicago/Turabian Style

Estrada-Flores, Sofía, Catalina M. Pérez-Berumen, Tirso E. Flores-Guia, Luis A. García-Cerda, Joelis Rodríguez-Hernández, Tzipatly A. Esquivel-Castro, and Antonia Martínez-Luévanos. 2022. "Mechanosynthesis of Mesoporous Bi-Doped TiO2: The Effect of Bismuth Doping and Ball Milling on the Crystal Structure, Optical Properties, and Photocatalytic Activity" Crystals 12, no. 12: 1750. https://doi.org/10.3390/cryst12121750

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