Cu doped InVO4 (xCu-InVO4 (x = 0.06–0.15 wt %) was synthesized by a facile one-pot hydrothermal method for the removal of methylene blue (MB) under LED light irradiation. The X-ray photoelectron spectroscopy (XPS) analysis indicated the coexistence of V5+ and V4+ species due to the O-deficient nature of the xCu-InVO4. The synthesized photocatalysts displayed a morphology of spherical and square shaped particles (20–40 nm) and micro-sized rectangle rods with a length range of 100–200 μm. The xCu-InVO4 exhibited superior adsorption and photodegradation efficiency compared to pristine InVO4 and TiO2 due to the presence of O2 vacancies, V4+/V5+ species, and Cu dopant. The optimum reaction conditions were found to be 5 mg L−1 (MB concentration), pH 6, and 100 mg of photocatalyst mass with a removal efficiency and mineralization degree of 100% and 96.67%, respectively. The main active species responsible for the degradation of MB were OH radicals and h+. Reusability studies indicated that the 0.13Cu-InVO4 was deactivated after a single cycle of photocatalytic reaction due to significant leaching of V4+ and Cu2+ species.

  • xCu-InVO4 (x = 0.06 − 0.15 wt%) were synthesized via the hydrothermal method.

  • 0.13Cu-InVO4 active under the LED light irradiation for MB removal.

  • Optimized catalytic conditions displayed complete MB removal with >96% mineralization.

  • 0.13Cu-InVO4 deactivated after first use due to V4+ and Cu2+ species leaching.

Graphical Abstract

Graphical Abstract
Graphical Abstract

Since discovering the ability of TiO2 electrodes in water splitting by Fujishima & Honda 1972, heterogeneous semiconductor photocatalysis has been broadly utilized in various fields, including water purification, selective organic transformations, and CO2 reduction (Wen et al. 2017). However, due to the large bandgap of TiO2 (3.2 eV) and low sensitivity towards visible light (Liang et al. 2021), the potential of other semiconductors such as ZnO, ZnS, BiVO4, g-C3N4, CdS, and CdSe was explored. Zou et al. (2000) reported the suitability of InNbO4 and InTaO4 with bandgap energy of 2.5 eV and 2.6 eV as photocatalysts. This discovery has led Ye et al. (2002) to investigate whether Nb and Ta can be replaced with vanadium(V) for the synthesis of InVO4. The group successfully synthesized the InVO4 with a much narrower bandgap (2.0 eV) than InNbO4 and InTaO4. The InVO4 catalyst was able to produce H2 from water under visible light irradiation.

Due to its various potential properties such as narrow bandgap energy of ∼2.0 eV, chemical stability, non-toxicity, and photo-corrosion resistivity, InVO4 is considered as a promising visible-light active photocatalyst with great potential in various photocatalytic applications (Guo et al. 2016; Lin et al. 2016; Yuan et al. 2019). However, ineffective charge-carrier separation in its pure form severely restricts its application in photocatalysis (Chaison et al. 2017). Since the discovery of its photocatalytic activity, InVO4 has been modified by forming integrated photocatalytic systems such as BiVO4/InVO4 (Guo et al. 2015), TiO2/InVO4/RGO (Lin et al. 2015a, 2015b), In2S3/InVO4 (Yuan et al. 2019), InVO4/β-AgVO3 (Yang et al. 2019), InVO4-g-C3N4/rGO (Hafeez et al. 2020), InVO4/ZnFe2O4 (Wang et al. 2020a, 2020b), Ag-SnS2@InVO4 (He et al. 2020), and AgBr/Ag2 MoO4@InVO4 (Zhang et al. 2020a, 2020b, 2020c). Although heterojunction formation is efficient in lowering the recombination rate of photo-excited species, it still suffers from two major drawbacks. The first is that the addition of a guest photocatalyst will lead to active sites blockage and lowering the interactions between the catalyst surface and the substrate, which would finally decrease the catalytic activity. Secondly, the guest material is also believed to lower light penetration due to the shielding effect (Zhu & Wang 2017).

On the other hand, metal ion doping offers a better alternative in suppressing the rapid recombination of electron/hole (e/h+) pairs. By adding charge trapping sites (dopants), the recombination rate will be quenched, and the lifespan of the photo-excited species will be prolonged to enhance the photocatalytic activity of the semiconductor (Imam et al. 2018). In the past, InVO4 based photocatalysts were successfully modified with metals including Pt (Yan et al. 2012), Ag (Lin et al. 2015a, 2015b), Cu (Wetchakun et al. 2017), Bi (Wang et al. 2020a, 2020b), Yb and Tm (Zhang et al. 2020a, 2020b, 2020c).

In this study, different amounts of Cu were incorporated into the lattice of InVO4 via a one-pot hydrothermal method under optimized reaction conditions. The detailed physicochemical properties of the photocatalysts were systematically characterized and studied before and after the photocatalytic reaction. An in-depth investigation on the visible-light-driven photodegradation of methylene blue (MB) under a low-cost homemade LED fixed photoreactor was also carried out. The Cu-InVO4 prepared by Wetchakun et al. (2017) required higher temperature and longer heating time. The photocatalytic activities of the Cu-InVO4 were investigated using a 50 W halogen lamp with a glass filter to block the UV components. In this research, lower heating temperature (180 °C) and shorter time (12 h) were used.

Chemicals

Indium(III) nitrate hydrate (In(NO3)3.H2O, 99.9%) and sodium metavanadate (NaVO3, ≥98.0%) were purchased from Sigma-Aldrich. Copper(II) nitrate trihydrate (Cu(NO3)2.3H2O, 99.5%) and ascorbic acid (99.5%) were purchased from QREC. All chemicals were analytical grade and used without any further purification.

Preparation of InVO4

In a typical synthesis procedure, 2 mmol of In(NO3)3 was dissolved in 35 mL of deionized water under magnetic stirring followed by the addition of an equimolar amount of NaVO3. The mixture was stirred vigorously for 2 h before being transferred into a 50 mL Teflon-lined stainless-steel autoclave and was placed horizontally in an oven at 180 °C for 12 hours. The yellowish InVO4 powder was then separated from the mother liquid by centrifugation and washed alternatively with ethanol and water three times. The InVO4 powder was dried at 60 °C for 24 h.

Synthesis of Cu doped InVO4

The doping process was carried out according to the synthesis of pristine InVO4 with some modifications. The Cu(NO3)2 (2.5–10 mole%) was added to the mixture of In(NO3)3–NaVO3 solution. The photocatalysts were labelled as xCu-InVO4 (x = 0.06, 0.10, 0.13 and 0.15) based on the wt% of Cu determined by Inductively Coupled Plasma-Optical Emission Spectrometry (ICP-OES) analysis.

Characterization of the photocatalyst

The functional groups of the photocatalysts were determined using the Fourier Transform Infrared Spectroscopy (FT-IR) analysis (FT-IR 2000 Perkin Elmer). The crystal structure was determined by X-Ray diffraction (XRD) analysis (Bruker D8 Advance, current = 40 mA and voltage = 40 kV) under Cu Kα radiation (1.5406 Å). The mean crystal size of the prepared photocatalysts was evaluated using the Scherrer's Equation (Gerawork 2021), as shown in Equation (1):
(1)
where Dp is the average crystal diameter (nm), k is the Scherrer's constant (0.94), λ is the wavelength of the Cu Kα radiation (0.15406 nm), β is the full width of half maximum (FWHM) of the highest peak, which is located at 2θ ≈ 33°, and θ is the exact location of the peak in the 2θ range (Yao et al. 2009; Dianat 2018). The scanning electron microscopy (SEM) (FEI Quanta 650 FEG SEM) was used to obtain the morphology, whereas the high-resolution transmittance electron microscopy (HRTEM) with field emission (TECNAI G2 20S-TWIN, FEI HRTEM 200 kV) was utilized to study the morphology. The surface elemental composition was obtained by X-ray photoelectron spectroscopy (XPS) (Ulvac- PHI Quantera II) equipped with monochromatic Al-Kα (hv = 1,486.6 eV) X-ray power source. The adventitious carbon peak (C1s) was used as a reference and calibrated to 284.8 eV (Oswald et al. 2018). The photocatalyst's textural properties were investigated by nitrogen adsorption-desorption (NAD) analysis using Micromeritics ASAP 2020 Surface Adsorption Porosimeter. Data collected from the diffuse reflectance UV-Vis spectroscopy (UV-Vis DRS) (Perkin Elmer UV-Vis spectrophotometer, Lambda 35) was used to calculate the bandgap energy using Kubelka-Munk model (Wang et al. 2016a, 2016b). The point of zero charge (PZC) was estimated by pH drift method with some modifications (Jiao et al. 2017).

Photocatalytic study

The photodegradation of methylene blue (MB) was conducted using a homemade reactor equipped with a commercially available 36 Watts LED (Flying Butterfly Lamp company) lamp as the irradiation source. The lamp was placed directly above the reaction mixture with 8.5 cm between the lamp and the reaction mixture (Figure S1 in Supplementary Information). The MB solution (200 mL of 5 mg L−1) was stirred with the photocatalyst under the dark condition for 60 minutes to achieve adsorption-desorption equilibrium. Subsequently, the solution was irradiated with LED light. An aliquot (5 mL) was collected in 15-minute intervals and then filtered using 0.2 μm syringe filters. The reaction mixture was then analysed using a UV-Vis spectrometer (Shimadzu 2600 UV-Vis). The dye removal efficiency was calculated using Equation (2); where C0 and Ct are the initial MB concentration and MB concentration after a given time from light irradiation, respectively.
(2)
The Langmuir-Hinshelwood model was used to study the photodegradation kinetics and to determine the rate of reactions. Pseudo-first-order and pseudo-second-order kinetics were expressed using Equations (3) and (4), respectively.
(3)
(4)
where C0, Ct, t, k and k′ are the MB concentration at zero time, the MB concentration at a specific time of light irradiation, time (minutes), the pseudo-first-order rate constant (min−1), and the pseudo-second-order rate constant (L mol−1 min−1), respectively. Mineralization studies were performed using ion chromatography analysis (Metrohm 792 Basic IC) by detecting SO42− ions.

Photocatalyst characterization

FT-IR analysis

The FT-IR spectra of the pristine InVO4 and xCu-InVO4 (x= 0.06–0.15) are shown in Figure 1(a). The IR peak located at 452 cm−1 is assigned to the V — O — V vibrations (Liu et al. 2016), whereas the peaks at 767 cm−1 and 902 cm−1 are assigned to the V — O — In vibration (Lai et al. 2018). The peak associated with the contraction and expansion vibrations of V — O bond can be seen at 950 cm−1 (Yao et al. 2009). The two bands located at 1,630 cm−1 and 3,434 cm−1 are attributed to the surface hydroxyl groups' stretching vibrations (—OH) and adsorbed water molecules, respectively (Zhang et al. 2015; Kumar et al. 2018). The IR peaks related to the CuO or Cu2O were not observed in the spectra of Cu doped InVO4 photocatalysts. The absence of these peaks suggests that the Cu was possibly incorporated within the framework of InVO4 instead of forming CuO or Cu2O species on the catalyst's surface. The IR peak at 1,386 cm−1 is attributed to the presence of Na2V6O16. This complex is usually formed when vanadium ions with ratios equal to or exceeding 3 of V/In precursors are used (Yao et al. 2009). Since the ratio used in this study was 1:1, the existence of the peak is concluded not to be caused by the presence of excess vanadium ion.

Figure 1

(a) The FT-IR spectra, (b) XRD patterns at wide range (2θ = 30° – 36°) and (c) at small range (2θ = 15° – 75°) of the prepared photocatalysts and (d) the effect of Cu doping on the orthorhombic lattice distortion and crystal size.

Figure 1

(a) The FT-IR spectra, (b) XRD patterns at wide range (2θ = 30° – 36°) and (c) at small range (2θ = 15° – 75°) of the prepared photocatalysts and (d) the effect of Cu doping on the orthorhombic lattice distortion and crystal size.

Close modal

During the photocatalyst synthesis, when NaVO3 solution was added into the In(NO3)3 solution, the colour of the solution quickly changed from colourless to dark red. The formation of the red-coloured solution indicates the presence of vanadic acid (HVO3). During the hydrothermal process, HVO3 species will be converted into [VO2(H2O)4]+ and [H2V10O28]4− ions. These two ions will undergo protonation and deprotonation during the hydrothermal synthesis, forming neutral [VO(OH)3(H2O)2]. Polycondensation of [VO(OH)3(H2O)2] species in the presence of Na+ ions will lead to the formation of Na2V6O16 (Chithaiah et al. 2012).

XRD analysis

The XRD diffractograms of the synthesized photocatalysts are shown in Figure 1(b). The diffractogram patterns are in agreement with the standard data of InVO4 in the orthorhombic phase (JPDS No.48–0898). The absence of diffraction peaks related to CuO and Cu2O further confirms the observation in the FT-IR analysis. Diffraction peaks ascribed to Na2V6O16 were not observed, possibly due to its low amount. A closer inspection at 2θ = 30–36° (Figure 1(c)) showed that the three diffraction peaks experienced shifting compared to the pristine InVO4, indicating changes in the framework of InVO4. The doped metal can either fill up the voids between the host material's atoms or replace the host material's atoms. The first is known as interstitial doping, whereas the second is known as substitutional doping. The metal doping position can be assumed based on the Hume-Rothery principle, which states that atoms’ substitution will only occur if the difference between the substituted and substituting metals' electronegativities is less than 20% (Mathew et al. 2018). Since the difference between copper electronegativity (1.95) and indium electronegativity (1.78) is 8.7%, and between copper and vanadium electronegativities (1.63) is 16.41%, it is assumed that Cu doping into InVO4 is in substitutional positions.

For 0.06Cu-InVO4, the XRD peaks shifting to the lowest angle is ascribed as an effect of Cu2+ substituting V5+ (He et al. 2018). As for 0.10Cu-InVO4, the XRD peaks shifted slightly to the right compared to 0.06Cu-InVO4, possibly due to the substitution of In3+ by Cu2+. As the doping amount increased, shifting to a lower angle was observed but to a lesser extent compared to 0.06Cu-InVO4. This observation could be due to the substitution of both In3+ and V5+ by Cu2+ but with more tendency towards replacing V5+.

The main dominant peak of InVO4 of the (112) plane located at approximately 2θ = 33.1° was used to calculate and study the effect of Cu doping on the crystal size using Scherrer's equation. Additionally, based on previous efforts, the effect of doping on the lattice distortion degree of orthorhombic structured semiconductors can be measured using the a/b lattice constants ratio (Chen et al. 2018). The lattice constants were calculated using the formula:
(5)
where dhkl is the d-spacing of a single plane. Also, h, k, and l are the Miller index of that same plane. Moreover, a, b, and c are the lattice constants.

Figure 1(d) illustrates the effect of Cu doping on the orthorhombic lattice distortion and the crystal size values obtained from the XRD analysis. It can be clearly seen that increasing the amount of doped Cu leads to an increase in the orthorhombic distortion value with no change of the crystal size until 0.13 wt.% of Cu was doped into the catalyst. However, when the doping amount further increased to 0.15 wt.%, both the orthorhombic distortion and crystal size dropped significantly. The orthorhombic distortion was minimum for 0.06Cu-InVO4 due to the substitution of solely V5+ by Cu2+ ions. The distortion becomes significant for 0.10Cu-InVO4 due to the dual replacement of In3+ and V5+ by Cu2+. In addition, the sudden drop in both orthorhombic distortion and crystal size for 0.15Cu-InVO4 can be ascribed to the excessive crystal distortion and defects, which might have caused the crystals to break down into smaller crystals.

HRTEM analysis

The TEM and HRTEM images of the photocatalysts are shown in Figure 2. The lattice fringes and interplanar spacing were determined using Fast Fourier Transformation (FFT) software. Based on the TEM images, pristine InVO4 and xCu-InVO4 (x= 0.06–0.15) photocatalysts consisted of irregularly shaped oval and spherical nanoparticles.

Figure 2

The TEM and HRTEM images of (a,b) InVO4, (c,d) 0.06Cu-InVO4, (e,f) 0.10Cu-InVO4, (g,h) 0.13Cu-InVO4, and (i,j) 0.15Cu-InVO4, respectively.

Figure 2

The TEM and HRTEM images of (a,b) InVO4, (c,d) 0.06Cu-InVO4, (e,f) 0.10Cu-InVO4, (g,h) 0.13Cu-InVO4, and (i,j) 0.15Cu-InVO4, respectively.

Close modal

The orthorhombic InVO4 XRD diffractogram contains 33 peaks and consequently 33 (hkl) planes. However, only four different d-spacing values were detected for each photocatalyst due to the different scanning areas. The pristine InVO4 was found to have d-spacing values that correlate to the hkl planes of (024), (111), (310), and (040). For 0.06Cu-InVO4, the hkl planes detected were (202), (220), (130), and (112), whereas the hkl planes of (042), (310), (222), and (200) were found in 0.10Cu-InVO4. The hkl planes present in 0.13Cu-InVO4 were (020), (040), (222), and (310). For 0.15Cu-InVO4, (004), (042), (150), and (114) hkl planes were detected.

The histograms were fitted using Gaussian distribution fitting to determine the photocatalyst's average particle size as illustrated in Figure S2. The estimated particle size obtained was found to be 44.21 ± 0.66 nm, 36.05 ± 0.32 nm, 38.46 ± 0.82 nm, 38.61 ± 1.59 nm, and 37.54 ± 1.09 nm for InVO4, 0.06Cu-InVO4, 0.10Cu-InVO4, 0.13Cu-InVO4, and 0.15Cu-InVO4, respectively. The particles obtained are bigger than the particle size reported by Wetchakun et al. (2017). Generally, higher hydrothermal reaction temperature often leads to bigger particle size. However, in this research, smaller particle size was obtained due to different ratios of the reactants used compared to Wetchakun et al. (2017).

N2 adsorption-desorption studies

Table S1 illustrates the N2 adsorption-desorption isotherms of all the synthesized catalysts. At higher relative pressure region (P/P0 = 0.8–1), pristine InVO4 and xCu-InVO4 (x= 0.06–0.15) catalysts displayed type IV isotherms with H3 hysteresis loops, suggesting the presence of mesopores in the samples (Sing et al. 1985; Hoan et al. 2020; Zhang et al. 2020a, 2020b, 2020c). To further confirm the presence of such pores, Barrett-Joyner-Halenda (BJH) pore size distribution studies were carried out.

As illustrated in Figure S3, the photocatalysts have a bimodal pore system centred on ∼3.7 nm and ∼40 nm. The relatively smaller pores (inset) are ascribed to the mesopores within the InVO4 cracks, whereas the larger pores are believed to be due to the formation of larger mesopores or macropores between the InVO4 particles (Zhang et al. 2020a, 2020b, 2020c).

As shown in Table S1, the BET surface area of the doped InVO4 was slightly lower compared to the pristine InVO4. A similar trend was observed in the total pore volume of the photocatalysts. This observation can be attributed to the low content of doped Cu2+. In addition, the small differences between the ionic radii of In3+ (0.08 nm), Cu2+ (0.073 nm), and V5+ (0.059 nm) do not significantly contribute to the changes in the surface area or pore volume. Nonetheless, Cu doping was found to enhance the pore width until 0.13wt.%, whereas a further increase in Cu amount leads to a drop in the pore width.

Point of zero charge (PZC) analysis

The PZC of the photocatalysts was determined to be in the range of pH 2.32–2.58, as illustrated in Figure S4, similar to previous reports (Lamdab et al. 2015). The protonation and deprotonation of the catalyst surface will take place depending on the pH of the solution. At pH values below the pHPZC, the catalyst surface will be positively charged, whereas at pH above the pHPZC the surface will have a total negative charge. The effect of solution pH on the surface charge of the xCu-InVO4 (x= 0.06–0.15) can be summarized in Equations (6) and (7):
(6)
(7)

UV-Vis DRS analysis

Figure 3(a) displays the absorbance spectra of the pristine InVO4 and xCu-InVO4 (x= 0.06–0.15). The absorbance edge of InVO4 was noticed to experience a slight red shift towards higher wavelength values with Cu doping, indicating increased sensitivity towards visible light. The Kubelka-Munk function was used to determine the direct bandgap of the synthesized photocatalysts, as illustrated in Figure 3(b). The bandgap values are listed in Table S2. The bandgap of pristine InVO4 was found to be 3.0 eV, which is higher than the usually reported bandgaps of InVO4 (2.0–2.4 eV) (Ye et al. 2002; Wang et al. 2016a, 2016b). The high bandgap found in this study is due to the use of different concentrations of precursors and different synthesis temperatures. These factors affect the synthesized material's optical properties (Jacob & Jerome Das 2016; Lestari et al. 2016).

Figure 3

(a) The absorbance spectra of prepared catalysts and the Kubelka-Munk function of bandgap calculations of (b) pristine and xCu-InVO4 (x= 0.06–0.15).

Figure 3

(a) The absorbance spectra of prepared catalysts and the Kubelka-Munk function of bandgap calculations of (b) pristine and xCu-InVO4 (x= 0.06–0.15).

Close modal

In addition, the redshift also indicates the formation of a new band state located between the conduction band (CB) and valence band (VB) of the InVO4. It is expected that xCu-InVO4 (x= 0.06–0.15) will have better photocatalytic performance under visible light irradiation compared to pristine InVO4 due to the smaller bandgap and enhanced Vis light-harvesting (He et al. 2018).

It is crucial to determine the CB and VB exact positions to understand the possible photocatalytic reaction mechanisms. The values of the energy bands were determined using the Equations:
(8)
(9)
where ECB and EVB are the CB and VB energy potentials, respectively. The Eg, Ee, and χ are the semiconductor's bandgap, free electrons' energy vs. hydrogen (4.5 eV), and the total electronegativity of the semiconductor (χ = 5.74 eV for InVO4), respectively. The values of ECB and EVB were calculated by substituting these values in Equations (8) and (9) and listed in Table S2.

Photoluminescence (PL) analysis

The PL analysis provides valuable information regarding electron migration and trapping efficiency. Generally, a lower PL emission intensity suggests a higher e − /h+ separation efficiency and a prolonged photo-excited charge carriers' lifespan (Priya et al. 2020). Figure 4 shows the PL spectrum of pristine and xCu-InVO4 (x= 0.06–0.15). All the photocatalysts illustrate very strong photoluminescence properties with intensity peaks mainly present in the green region of visible emission (∼550 nm). The Cu doping has significantly influenced the peak intensities.

Figure 4

The PL phenomena of pristine and xCu-InVO4 (x= 0.06–0.15).

Figure 4

The PL phenomena of pristine and xCu-InVO4 (x= 0.06–0.15).

Close modal

As proven by XRD, Cu2+ replaced V5+ sites at low doping concentrations (0.06Cu-InVO4) and replaced both In3+ and V5+ active sites at higher doping levels. Since the PL emission peak of InVO4 originates from the VO43− charge-transfer transition (Shih et al. 2015; Li & Xu 2018), the behaviour of the PL phenomena is correlated with V5+ relative concentration in the nanocomposites. Based on the obtained results, it can be concluded that substitution of V5+ ions by Cu2+ ions is highly efficient in enhancing the separation of photoinduced electron/hole pairs and provides better trapping sites compared to pristine InVO4.

The XRD analysis also indicates that 0.10Cu-InVO and 0.13Cu-InVO4 have the largest crystallite size (44 nm). It has been reported that larger crystallites have a lower number of atoms in grain boundaries and on the surface, hence increasing the light scattering efficiency compared to smaller crystals. As a result, the PL intensity of these photocatalysts appeared to be higher compared to 0.15Cu-InVO4 (35 nm). The 0.06Cu-InVO4 has lower PL intensity compared to 0.10Cu-InVO4 and 0.13Cu-InVO4, even though they have similar crystallite size. This trend is possibly due to the lower degree of orthorhombic distortion and even distribution of Cu species on the surface (Lee et al. 2015).

XPS analysis

The XPS analysis was carried out only for 0.13Cu-InVO4 as a representative for the rest of the photocatalysts. As illustrated in Figure 5(a), the survey scan of 0.13Cu-InVO4 shows the presence of indium (4d, 4p, 4s, 3d, 3p, 3s, and MNN transitions), vanadium (3p, 3s, 2p, 2s, and LMM transitions), oxygen (1s and KLL transitions), and carbon (1s transition). The MNN, LMM, and KLL are Auger peaks that provide information on the Auger transitions’ existing vacancies. Carbon originates from the instrumental environment. The existence of In3d, V2p, and O1s indicates the successful fabrication of InVO4. However, Cu transitions, which should appear at B.E. ∼932 eV and ∼953 eV for Cu 2p transitions, were not detected (Wang et al.2014; Moongraksathum et al. 2018). This is ascribed to the low amount of Cu in the sample. The XPS detection limit for most elements (including Cu) is 0.1 at.% — 1.0 at.% (Shard 2014; Lamers et al. 2018). The Cu concentration detected in 0.13Cu-InVO4 was only 0.13%.

Figure 5

The XPS spectra of 0.13Cu-InVO4. (a) Survey scan and narrow scan of (b) In3d, (c) V2p, and (d) O1s.

Figure 5

The XPS spectra of 0.13Cu-InVO4. (a) Survey scan and narrow scan of (b) In3d, (c) V2p, and (d) O1s.

Close modal

The narrow scan of In3d illustrates two peaks at 444.31 eV and 451.89 eV corresponding to In3d5/2 and In3d3/2 valence states, respectively (Figure 5(c)) (Li et al. 2009; Yang et al. 2019; Hafeez et al. 2020). Moreover, the deconvolution of V2p (Figure 5(c)) peaks shows the characteristic peaks at 517.12 eV and 524.62 eV that are assigned to V2p3/2 and V2p1/2, respectively, contributing to V5+ (Silversmit et al. 2004; Ma et al. 2015; Zhou et al. 2018). The peaks located at 516.75 eV and 523.16 eV are ascribed to the surface V4+ species (Motola et al. 2018; Han et al. 2019). Based on the Principle of Electroneutrality, the reduction of V5+ to V4+ can be ascribed to the oxygen vacancies produced during the hydrothermal synthesis (Wetchakun et al. 2017). The formation of oxygen vacancies can be attributed to two reasons, Cu2+ doping (He et al. 2018) and the high applied pressure during the hydrothermal synthesis (Abdul-Rahman et al. 2019), which would weaken the V – O bond and further facilitate the removal of the oxygen attached to V, causing a reduction in V charge to form V4+ moieties. The presence of oxygen vacancy sites and V4+ species is highly crucial in photocatalytic activities, as will be explained later. Additionally, the asymmetrical O1s (Figure 5(d)) peak can be deconvoluted into two peaks located at 529.78 eV and 530.69 eV. These peaks originated from the lattice oxygen and surface adsorbed oxygen species, respectively (Wang et al. 2013a, 2013b; He et al. 2018; Liu et al. 2020).

Table S3 illustrates the molar ratios of different species present on the surface of 0.13Cu-InVO4. The molar ratios were calculated using the area of the peaks and the relative sensitivity factors (R.S.F) of each peak. In an ideal case of InVO4, the molar ratio of V:In should be 1:1; however, in 0.13Cu-InVO4 it is slightly less. This can be ascribed to the fall in V concentration due to Cu substitutional doping. Moreover, O:In and O:V ratios should theoretically be 4:1, yet the decrease in the actual ratios further confirms the formation of extrinsic O-vacancies and establishes that the 0.13Cu-InVO4 is oxygen-deficient. Furthermore, the obtained V4+/V5+ molar ratio is higher than previously reported values for InVO4 (0.184–0.430) (Wang et al. 2013a, 2013b), indicating relatively more surface-active sites.

Photocatalytic performance

Effect of Cu doping on the photocatalytic activity of InVO4

The effect of Cu doping on the photocatalytic activity of InVO4 was tested in the photodegradation of MB under LED light irradiation. As a comparison, the reaction was also carried out using TiO2 and pristine InVO4. As shown in Figure 6(a), the removal of MB was negligible under photolysis conditions. A similar observation was noticed when TiO2 was used as the photocatalyst. The adsorption percentage of the photocatalysts were in the order of 0.13Cu-InVO4 (62.3%) > 0.10Cu-InVO4 (48.6%) > 0.15Cu-InVO4 (45.9%) > 0.06Cu-InVO4 (33.4%) > InVO4 (10.4%). The enhancement in the adsorption properties is ascribed to a relatively high V4+/V5+ ratio (He et al. 2018), oxygen vacancies and defects induced from the Cu doping (Sun et al. 2019). In addition, Cu itself (Ma et al. 2010; Tan et al. 2011) and the surface hydroxyl groups can enhance the adsorption of organic pollutants (Fu et al. 2011; He et al. 2018).

Figure 6

(a) The MB removal efficiency by TiO2, photolysis, pristine and Cu doped InVO4 with the corresponding (b) pseudo-first-order and (c) pseudo-second-order kinetics of MB photocatalytic degradation by TiO2, pristine and Cu doped InVO4 ([MB] = 5 mg L−1, catalyst dosage = 50 mg and pH = 6, LED irradiation).

Figure 6

(a) The MB removal efficiency by TiO2, photolysis, pristine and Cu doped InVO4 with the corresponding (b) pseudo-first-order and (c) pseudo-second-order kinetics of MB photocatalytic degradation by TiO2, pristine and Cu doped InVO4 ([MB] = 5 mg L−1, catalyst dosage = 50 mg and pH = 6, LED irradiation).

Close modal

The removal of MB by the synergetic effect of adsorption and photocatalysis was highest for 0.13Cu-InVO4 (78.6%) followed by 0.10Cu-InVO4 (68.2%), 0.15Cu-InVO4 (59.4%), 0.06Cu-InVO4 (45.3%), and pristine InVO4 (10.1%). The insensitivity of the pristine InVO4 towards visible light is attributed to its wide bandgap (3.0 eV), which prevented the electrons from being excited to the VB from the CB under visible light irradiation. In addition, semiconductors with a wide bandgap have a high rate of recombination of e/h+ pairs. The Cu dopant has been identified to play a few crucial roles: (1) it increases the sensitivity of the photocatalysts towards visible light absorption by reducing the bandgap, and (2) it acts as an electron reservoir, therefore prolonging the lifespan of photoinduced charge.

The kinetics of the photodegradation of MB using TiO2, pristine InVO4 and xCu-InVO4 (x= 0.06–0.15) were fitted into pseudo-first- (Figure 6(b)) and pseudo-second-order kinetics (Figure 6(c)) models. The Langmuir-Hinshelwood constants are provided in Table 1, where k, k′, and R represent pseudo-first-order reaction rate constants (min−1), the pseudo-second-order reaction rate constant (L mol−1 min−1), and correlation regression coefficient, respectively. The negligible rate constant value of pristine InVO4 and negative R2 can be explained by the desorption of some traces of the MB molecules without undergoing photodegradation when exposed to the LED light irradiation due to its relatively large bandgap (3.0 eV) and fast recombination rate of e/h+ pairs. The xCu-InVO4 (x= 0.06–0.15) photocatalysts showed good fitting using pseudo-second-order kinetics as their correlation regression coefficient (R) values were closer to 1 compared to those of pseudo-first-order. Since only a handful of researchers have reported the modification of InVO4, it is difficult to pinpoint the exact reasons for xCu-InVO4 (x= 0.06–0.15) pseudo-second-order kinetics rather than pseudo-first-order kinetics. The photocatalysts’ distinct physicochemical properties compared to InVO4 and the conditions used during the photocatalytic reaction (i.e., light source, precursor concentration, treatment temperature and time) may have favoured pseudo-second-order kinetics. All the xCu-InVO4 (x= 0.06–0.15) photocatalysts illustrated superior photo-activity, and Cu doping was found to significantly increase the rate constants compared to pristine InVO4. The maximum rate constant was obtained with 0.13Cu-InVO4 catalyst with a pseudo-second-order rate constant of 0.0197 min−1. Due to the enhanced photocatalytic activity of 0.13Cu-InVO4, it was used for the optimization of other parameters.

Table 1

The rate constants and correlation coefficient values of pseudo-first- and pseudo-second-order kinetics for the used catalysts for MB degradation

CatalystPseudo-first-order
Pseudo-second-order
k × 10−3 (min−1)R2k × 10−3
(L mol−1 min−1)
R2
TiO2 – 0.88005 – 0.37836 
InVO4 – − 0.0336 – − 0.03482 
0.06Cu-InVO4 1.82 0.98341 3.74 0.98772 
0.10Cu-InVO4 4.24 0.99552 12.29 0.99948 
0.13Cu- InVO4 4.90 0.99823 19.70 0.99956 
0.15Cu-InVO4 2.58 0.99346 6.46 0.99711 
CatalystPseudo-first-order
Pseudo-second-order
k × 10−3 (min−1)R2k × 10−3
(L mol−1 min−1)
R2
TiO2 – 0.88005 – 0.37836 
InVO4 – − 0.0336 – − 0.03482 
0.06Cu-InVO4 1.82 0.98341 3.74 0.98772 
0.10Cu-InVO4 4.24 0.99552 12.29 0.99948 
0.13Cu- InVO4 4.90 0.99823 19.70 0.99956 
0.15Cu-InVO4 2.58 0.99346 6.46 0.99711 

Effect of initial solution pH on MB removal

The effect of the solution pH on the photocatalytic degradation of MB was investigated in the range of pH 2–9. The MB removal profiles are shown in Figure 7(a). At pH values below the PZC of the 0.13Cu-InVO4 (pH 2.4), the surface of the catalyst will be protonated and becomes positively charged, whereas, at pH values above 2.4, the concentration of OH ions in the MB solution will increase and cause the deprotonation of the catalyst surface and amplify a negatively charged catalyst surface (Zawawi et al. 2017). The removal of MB was highest when the pH was increased from pH 2 to pH 6. At this range, the electrostatic interactions between the cationic MB and negatively charged 0.13Cu-InVO4 surface were strongest. As a result of the strong electrostatic interactions, more MB molecules will be adsorbed on the catalyst for photocatalytic reactions. The removal percentage started to decrease significantly when the pH was increased. The drop in the MB removal is possibly due to forming a stable hydroxyl-MB complex, reducing the electrostatic attraction between the dye molecules and the 0.13Cu-InVO4 surface (Sabar et al. 2020). The formation of the hydroxyl-MB complex was prominent at pH 9 compared to pH 7 due to the solution's high pH and a larger amount of OH ions that will compete with the catalyst over bonding with MB molecules.

Figure 7

(a) The effect of initial pH of MB solution, (b) the pseudo-second-order kinetics of effect of the light source, (c) the effect of initial MB concentration and (d) the effect of photocatalyst mass (0.13Cu-InVO4) on MB removal efficiency (right axis) and sulfate production (left axis) after 120 min.

Figure 7

(a) The effect of initial pH of MB solution, (b) the pseudo-second-order kinetics of effect of the light source, (c) the effect of initial MB concentration and (d) the effect of photocatalyst mass (0.13Cu-InVO4) on MB removal efficiency (right axis) and sulfate production (left axis) after 120 min.

Close modal

Effect of irradiation source on MB removal

The ability of 0.13Cu-InVO4 in removing MB under solar irradiation was investigated as well. The pseudo-second-order rate constant was higher when solar irradiation was used. As shown in Table 2, the intensity of visible light irradiation was five times higher in solar radiation compared to the LED light. In addition, solar light contains 29.7–34.4 W m−2 UV, whereas the UV light was not detected in the LED light. The combination of visible and UV light can produce more reactive oxygen species compared to visible light alone (Zawawi et al. 2017). Hence, higher photocatalytic activity was observed. However, the correlation coefficient (R2) for solar light is lower than that of the LED light (Figure 7(b)). The lower R2-value is ascribed to the inconsistency of solar radiation due to atmospheric and weather changes. The LED light radiation has several advantages compared to solar radiation. Besides being stable, it has no UV portion and poses fewer cancer risks to humans (Sabar et al. 2020).

Table 2

The Vis and UV intensities of different irradiation sources

Irradiation sourceVis Intensity (W m−2)UV Intensity (W m−2)
LED 68.9 0.00 
Solar 358–389 29.7–34.4 
Irradiation sourceVis Intensity (W m−2)UV Intensity (W m−2)
LED 68.9 0.00 
Solar 358–389 29.7–34.4 

In addition, the LED lights are reliable, inexpensive, and produce high light intensity under relatively low electrical power with less heat generated than fluorescence and mercury lamps. Due to these facts, the use of LED lights in photocatalytic systems has been growing rapidly during the last few years (Salehi et al. 2012; Izadifard et al. 2013; Jo & Tayade 2014; Reza et al. 2017; Alkaykh et al. 2020; Hameeda et al. 2020).

Effect of photocatalyst mass on MB removal and mineralization degree

The photocatalyst mass effect on the removal and mineralization degree of MB is shown in Figure 7(d). The MB removal increased from 76.2% to 96.0% when the photocatalyst mass was increased from 50 to 75 mg. The further increase of the mass resulted in the complete removal of the MB molecules. As the amount of photocatalyst mass is increased, more surface-active sites are available for the adsorption of the dye molecules and absorption of photons. This will eventually lead to a higher number of dye molecules reacting on the catalyst surface, and more active species will be formed (Kiwaan et al. 2020). Since sulfur is positioned at the centre of the MB triaromatic rings, the formation of SO42− ions indicates the cleavage of the resonance stabilized aromatic rings (Abdul-Rahman et al. 2019). Therefore, the mineralization degree was evaluated by monitoring the formation of SO42− traces in the reaction mixture.

The SO42− concentration was found to be 0.592, 1.303, 1.363, 0.703, and 0.05 mg L−1 when the photocatalyst mass was 50, 75, 100, 125, and 150 mg, respectively. Since the molecular weights of MB and SO42− are 319.85 g mol−1 and 90.06 g mol−1, respectively, the complete mineralization of 5 mg L−1 of MB would produce 1.41 mg L−1 of SO42− ion. Therefore, the mineralization degree of MB was found to increase significantly from 42.0% to 92.4% when the mass was increased from 50 to 75 mg and further increased to 96.7% when 100 mg of the photocatalyst was used. Although the removal of MB remained 100% when the photocatalyst mass was above 100 mg, the concentration of SO42− was significantly reduced to 0.703 mg L−1 and 0.05 mg L−1, corresponding to a mineralization degree of 49.86% and 3.55%, respectively. Excessive photocatalyst mass increases the light scattering and lowers the light penetration into the photocatalyst surface (Van de Moortel et al. 2020). Besides, particle agglomeration (particle-particle interaction), which usually takes place at a high concentration of photocatalyst particles, would lower the total surface area of the photocatalyst in the system (Tetteh et al. 2020). As a result, the mineralization process will be suppressed. At the same time, the removal remained at 100% due to higher adsorption efficiency.

Free radicals scavenging studies

Scavenging tests were performed to identify the main active species responsible for the photocatalytic degradation of MB using 0.13Cu-InVO4. Isopropyl alcohol (IPA), ethylenediaminetetraacetic acid (EDTA), and ascorbic acid (AA) were used as scavengers for OH, h+, and O2, respectively. Figure 8(a) illustrates the effect of chemical scavengers on the photocatalytic activity of 0.13Cu-InVO4. The addition of IPA and EDTA caused a slight decrease in the MB removal percentage, whereas AA did not affect the photodegradation efficiency. The results indicate that OH and h+ are the main active species in the photodegradation, whereas O2 had no contribution to the photodegradation process. In OH and h+ scavenging cases, the drop in the removal% was low, attributed to the high MB adsorption efficiency.

Figure 8

(a) The effect of active species scavengers on the degradation efficiency of MB using 0.13Cu-InVO4 ([MB] = 5 mg L−1, catalyst dosage = 50 mg, pH = 6, LED irradiation) and (b) the schematic representation of the proposed degradation mechanism of MB using 0.13Cu-InVO4 under LED light irradiation.

Figure 8

(a) The effect of active species scavengers on the degradation efficiency of MB using 0.13Cu-InVO4 ([MB] = 5 mg L−1, catalyst dosage = 50 mg, pH = 6, LED irradiation) and (b) the schematic representation of the proposed degradation mechanism of MB using 0.13Cu-InVO4 under LED light irradiation.

Close modal

Proposed photodegradation mechanism

The results obtained from scavenging studies are in good agreement with the findings from the CB and VB calculations. The CB energy (−0.23 eV) of the 0.13Cu-InVO4 was found to be not negative enough to produce O2 as compared to the reduction potential of O2 to O2, which is −0.33 eV (Xiao et al. 2013; Imam et al. 2018). However, the CB energy (−0.23 eV) is more negative than the reduction potential of O2/H2O2 (+0.695 eV — +0.682 eV). Hence, the photocatalyst can reduce O2 to H2O2, which then will react with the eCB to form OH (Jiang et al. 2012; Wen et al. 2017; Karbasi et al. 2020). The positively charged holes of VB (h+) are more positive than the reduction potential of OH/OH (+ 2.38 eV) (Hou et al. 2017; Huang et al. 2020; de Melo Santos Moura et al. 2021), which allows it to oxidize hydroxyl ions to form hydroxyl radicals. In addition, the h+ is a strong oxidizing agent that can directly degrade MB (Garcia-Segura & Brillas 2017). Based on these findings, a possible pathway for the degradation process is illustrated in Equations (10)–(16).
(10)
(11)
(12)
(13)
(14)
(15)
(16)

First, the visible light absorbed by the catalyst will lead to the migration of electrons from the VB to the CB (e), leaving positively charged holes in the VB (h+) (Equation (10)). Instead of direct recombination with h+, the photo-excited electrons will be trapped by the doped Cu2+ by reducing it to Cu+ (Cu2+/Cu+ : + 0.161 eV vs. NHE) (Equation (11)), which will extend the lifetime of the e/h+ species, allowing it to react with other species (Wetchakun et al. 2017). Then, the eCB will react with the O2 and H+ to form H2O2, which will eventually form OH (Equations (12) and (13)). On the other hand, the h+ may undergo two separate reactions; first, it will directly react with the MB molecules to oxidize and mineralize them (Equation (16)). Second, h+ will also react with the available hydroxyl anions (OH) in the solution to form OH (Equation (14)). Finally, the OH produced will react with the MB molecules and degrade them into harmless products that will eventually turn into CO2 and water (Equation (15)). The results are summarized and illustrated in Figure 8(b).

Reusability studies

Reusability studies were carried out to determine the regeneration efficiency and stability of 0.13Cu-InVO4. Prior to reusability, the 0.13Cu-InVO4 was washed with alkaline water (pH 11–12) to remove the adsorbed MB. The 0.13Cu-InVO4 was found to be deactivated after only one cycle, as shown in Figure S5(a). The adsorption efficiency of the used 0.13Cu-InVO4 was even less than that of fresh, pristine InVO4. The spent 0.13Cu-InVO4 was re-characterized using FT-IR, XRD, ICP-OES, and XPS to identify the cause of the deactivation. The complete removal of MB from the photocatalyst surface was confirmed by the FT-IR analysis (Figure S5 (b)). The XRD analysis of the spent photocatalyst did not indicate any crystal structural change compared to the fresh diffractogram as illustrated in Figure S5(c).

The ICP-OES analysis detected vanadium (3.902 mg L−1) and copper (0.326 mg L−1) in the photocatalytic reaction mixture. The significant leaching of these species is believed to be due to excessive O2 vacancies that would weaken the bonds between the atoms, causing leaching of V and Cu species. Another reason is that V4+ species are the prominent vanadium species in the 0.13Cu-InVO4 as proven by XPS, with a molar ratio of 0.704 of V4+/V5+. Since the original oxidation state of vanadium in InVO4, is V5+, the presence of high moieties of V4+ species would destabilize the structure and weaken the V–O–In bonds due to the unnatural structure of the semiconductor. In addition, crystal size and lattice distortion calculations obtained from XRD results showed that the crystal size of 0.13Cu-InVO4 increased from 43.98 nm to 73.29 nm and the orthorhombic lattice distortion (a/b ratio) slightly decreased from 0.67444 to 0.67187. The changes suggest that the crystals undergo an aggregation process to form larger crystals with less lattice distortion and better stability to overcome the significant loss of V atoms from the sample.

Figure S6(a) illustrates the comparison of the overall XPS scan between the freshly synthesized 0.13Cu-InVO4 vs. the used 0.13Cu-InVO4, and narrow scans of In3d (Figure S6(b)), V2p (Figure S6(c)), and O1s (Figure S6(d)). Table S4 provides detailed information regarding the oxidation states and elemental composition of fresh and used 0.13Cu-InVO4. As evident in Figure S6, the overall scan comparison shows that both used and fresh 0.13Cu-InVO4 comprise the same elemental composition of In, V, and O, corresponding to InVO4, and C resulting from the instrumental environment. Moreover, no peaks ascribed to sulfur, nitrogen, or chloride were observed, further confirming the complete removal of the adsorbed MB from the photocatalytic reaction. In addition, leaching of V was also confirmed as its atomic% dropped from 16.8% to 13% and V/In molar ratio significantly fell from 0.95 to 0.61 (Table 3). The decrease in O/In molar ratio can be ascribed to V leaching. When the vanadium atoms were leached, the oxygen atoms bonded to them leached as well. This also explains the decrease in the lattice oxygen portion (increase in Oads/Olatt molar ratio).

Table 3

The elemental composition of fresh and used 0.13Cu-InVO4 using XPS

In3d at.%V3p at.%O1s at.%V/InO/InOads/OlattV4+/V5+
Fresh catalyst 17.68% 16.80% 65.52% 0.95 3.71 0.51 0.70 
Used catalyst 21.33% 12.96% 65.71% 0.61 3.08 0.62 0.13 
In3d at.%V3p at.%O1s at.%V/InO/InOads/OlattV4+/V5+
Fresh catalyst 17.68% 16.80% 65.52% 0.95 3.71 0.51 0.70 
Used catalyst 21.33% 12.96% 65.71% 0.61 3.08 0.62 0.13 

The V4+/V5+ molar ratio has significantly dropped by almost 81% after the photocatalytic degradation. The drop indicates that most of the V atoms that were leached are V4+ species. As previously mentioned, V4+ surface species are highly crucial in photocatalytic reactions, as they provide active surface sites (Wang et al. 2013; He et al. 2018). Based on the obtained results, it can be concluded that the loss of adsorption and photocatalytic abilities of InVO4 are mainly attributed to the leaching of surface V4+ moieties coupled with the loss of Cu2+ dopants. These findings also confirm that V4+ species were the main contributors to the photocatalytic activity of InVO4. Additionally, considering V/In weight ratios of pristine fresh InVO4 and used 0.13Cu-InVO4 being equal to 0.6011 and 0.4734, this explains the deactivation of used 0.13Cu-InVO4.

To further investigate the chemical changes that may have occurred, the positions of deconvoluted In3d, V3p, and O1s peaks are compared and presented in Table S4. It can be noticed that In3d and O1s peaks shifted slightly to higher binding energy values, whereas V2p peaks all experienced a shift towards lower binding energies. Chemical shifts in XPS indicate changes in surface compositions and oxidation states. In general, shifting to higher binding energy occurs due to the decrease of the electron density on the respective element (Wang et al. 2013; Jia et al. 2018). Therefore, the positive shifting of O1s and In3d peaks can be explained by the following; in the bonds of In–O–V–O–In of InVO4, when V is oxidized from V4+ to V5+, the electron density around oxygen and indium atoms will drop because of the loss of an electron from the neighbouring V atoms. The O1s peak shift (+0.26 and +0.20 eV) was relatively higher than that of In3d (+0.17 and +0.15 eV) because the O atoms are directly linked to the V atom, making it more affected by the changes in the oxidation states of vanadium. As for the opposite chemical shift behaviour of V2p peaks, when V4+ moieties were lost and V5+ species became the prominent, vanadium oxidation state, therefore becoming less electron-dense, consequently, neighbouring In and O atoms became relatively more electron-dense compared to V atoms, hence causing the V2p peaks to experience a negative shift in binding energy values.

In this work, Cu doped InVO4 was successfully synthesized via a facile one-pot hydrothermal synthesis technique. The bandgap of InVO4 was found to drop with the increase in Cu content due to the formation of an intermediate energy level (Cu2+/Cu+: +0.161 eV). Due to the low Cu concentration used, the surface area, particle size, and morphology of the synthesized photocatalysts were not significantly influenced by the modifications. However, the PL phenomena and the crystal structure were found to be significantly influenced by Cu doping. The XRD analysis suggests that Cu2+ prefers to substitute the V5+ at low doping amounts, whereas at higher concentration, both In3+ and V5+ started to be substituted; however, with more tendency towards V5+ position replacement. The synergistic effect of adsorption and photodegradation due to the presence of Cu dopants, the presence of V4+ moieties, and the O-deficits on the catalyst surface significantly enhanced the photocatalytic activity of the InVO4. The photodegradation of 0.13Cu-InVO4 was best described by pseudo-second-order kinetics. The MB removal's optimum conditions were pH = 6, MB initial concentration of 5 mg L−1, and nanocomposite mass of 100 mg, which displayed a complete MB removal with >96% mineralization degree after 120 minutes of LED light radiation. The poor reusability of 0.13Cu-InVO4 is ascribed to the leaching of Cu2+ and V4+ species.

The authors would like to acknowledge the financial support received from Universiti Sains Malaysia Research University Grant (RUI) (1001/PKIMIA/8011083).

The authors have no competing interest to declare.

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

Abdul-Rahman
N. R.
Muniandy
L.
Adam
F.
Iqbal
A.
Ng
E. P.
Lee
H. L.
2019
Detailed photocatalytic study of alkaline titanates and its application for the degradation of methylene blue (MB) under solar irradiation
.
Journal of Photochemistry and Photobiology A: Chemistry
375
(
February
),
219
230
.
https://doi.org/10.1016/j.jphotochem.2019.02.020.
Alkaykh
S.
Mbarek
A.
Ali-Shattle
E. E.
2020
Photocatalytic degradation of methylene blue dye in aqueous solution by MnTiO3 nanoparticles under sunlight irradiation
.
Heliyon
6
(
4
),
e03663
.
https://doi.org/10.1016/j.heliyon.2020.e03663.
Chaison
J.
Wetchakun
K.
Wetchakun
N.
2017
Investigation of the physical, optical, and photocatalytic properties of CeO2/Fe-doped InVO4 composite
.
Journal of Physics and Chemistry of Solids
111
,
95
103
.
https://doi.org/10.1016/j.jpcs.2017.07.019
Chen
Y.
Xu
J.
Xie
S.
Tan
Z.
Nie
R.
Guan
Z.
Wang
Q.
Zhu
J.
2018
Ion doping effects on the lattice distortion and interlayer mismatch of Aurivillius-type Bismuth Titanate compounds
.
Materials
11
(
5
), 821.
https://doi.org/10.3390/ma11050821
Chithaiah
P.
Chandrappa
G. T.
Livage
J.
2012
Formation of crystalline Na2V6O16·3H2O ribbons into belts and rings
.
Inorganic Chemistry
51
(
4
),
2241
2246
.
https://doi.org/10.1021/ic202260w
de Melo Santos Moura
M. M.
de Lucena
A. L. A.
Napoleão
D. C.
Duarte
M. M. M. B.
Lima
V. E.
de Melo Neto
A. A.
2021
Degradation of the mixture of the ketoprofen, meloxicam and tenoxicam drugs using TiO2/metal photocatalysers supported in polystyrene packaging waste
.
Water Science and Technology
83
(
4
),
863
876
.
https://doi.org/10.2166/wst.2021.025
Dianat
S.
2018
Visible light induced photocatalytic degradation of direct red 23 and direct brown 166 by InVO4-TiO2 nanocomposite
.
Iranian Journal of Catalysis
8
(
2
),
121
132
.
Fu
M.
Li
Y.
Wu
S.
Lu
P.
Liu
J.
Dong
F.
2011
Sol-gel preparation and enhanced photocatalytic performance of Cu-doped ZnO nanoparticles
.
Applied Surface Science
258
(
4
),
1587
1591
.
https://doi.org/10.1016/j.apsusc.2011.10.003
Fujishima
A.
Honda
K.
1972
Electrochemical photolysis of water at a semiconductor electrode
.
Nature
238
(
5358
),
37
38
.
https://doi.org/10.1038/238037a0
Garcia-Segura
S.
Brillas
E.
2017
Applied photoelectrocatalysis on the degradation of organic pollutants in wastewaters
.
Journal of Photochemistry and Photobiology C: Photochemistry Reviews
31
,
1
35
.
https://doi.org/10.1016/j.jphotochemrev.2017.01.005
Gerawork
M.
2021
Remediation of textile industry organic dye waste by photocatalysis using eggshell impregnated ZnO/CuO nanocomposite
.
Water Science and Technology
83
(
11
),
2753
2761
.
https://doi.org/10.2166/wst.2021.165
Guo
F.
Shi
W.
Lin
X.
Yan
X.
Guo
Y.
Che
G.
2015
Novel BiVO4/InVO4 heterojunctions: facile synthesis and efficient visible-light photocatalytic performance for the degradation of rhodamine B
.
Separation and Purification Technology
141
,
246
255
.
https://doi.org/10.1016/j.seppur.2014.11.026
Guo
F.
Shi
W.
Cai
Y.
Shao
S.
Zhang
T.
Guan
W.
Huang
H.
Liu
Y.
2016
Sheet-on-sphere structured Ag/AgBr@InVO4 heterojunctions and enhanced visible-light photocatalytic activity
.
RSC Advances
6
(
96
),
93887
93893
.
https://doi.org/10.1039/c6ra20657c
Hafeez
H. Y.
Lakhera
S. K.
Shankar
M. V.
Neppolian
B.
2020
Synergetic improvement in charge carrier transport and light harvesting over ternary InVO4-g-C3N4/rGO hybrid nanocomposite for hydrogen evolution reaction
.
International Journal of Hydrogen Energy
45
(
13
),
7530
7540
.
https://doi.org/10.1016/j.ijhydene.2019.05.235
Hameeda
B.
Mushtaq
A.
Saeed
M.
Munir
A.
Jabeen
U.
Waseem
A.
2020
Development of Cu-doped NiO nanoscale material as efficient photocatalyst for visible light dye degradation
.
Toxin Reviews
,
1
11
.
https://doi.org/10.1080/15569543.2020.1725578
Han
Q.
Bai
X.
Man
Z.
He
H.
Li
L.
Hu
J.
Alsaedi
A.
Hayat
T.
Yu
Z.
Zhang
W.
Wang
J.
Zhou
Y.
Zou
Z.
2019
Convincing synthesis of atomically thin, single-crystalline InVO4 sheets toward promoting highly selective and efficient solar conversion of CO2 into CO [Rapid-communication]
.
Journal of the American Chemical Society
141
(
10
),
4209
4213
.
https://doi.org/10.1021/jacs.8b13673
He
B.
Li
Z.
Zhao
D.
Liu
H.
Zhong
Y.
Ning
J.
Zhang
Z.
Wang
Y.
Hu
Y.
2018
Fabrication of porous Cu-doped BiVO4 nanotubes as efficient oxygen-evolving photocatalysts
.
ACS Applied Nano Materials
1
(
6
),
2589
2599
.
https://doi.org/10.1021/acsanm.8b00281
He
S.
Yang
Z.
Cui
X.
Zhang
X.
Niu
X.
2020
Fabrication of the novel Ag-doped SnS2@InVO4 composite with high adsorption-photocatalysis for the removal of Uranium (VI)
.
Chemosphere
260
,
127548
.
https://doi.org/10.1016/j.chemosphere.2020.127548
Hoan
N. T. V.
Minh
N. N.
Nhi
T. T. K.
Van Thang
N.
Tuan
V. A.
Nguyen
V. T.
Thanh
N. M.
Van Hung
N.
Khieu
D. Q.
2020
TiO2/Diazonium/graphene oxide composites: synthesis and visible-light-driven photocatalytic degradation of methylene blue
.
Journal of Nanomaterials
2020
.
https://doi.org/10.1155/2020/4350125
Hou
J.
Jiang
K.
Shen
M.
Wei
R.
Wu
X.
Idrees
F.
Cao
C.
2017
Micro and nano hierachical structures of BiOI/activated carbon for efficient visible-light-photocatalytic reactions
.
Scientific Reports
7
(
1
),
1
10
.
https://doi.org/10.1038/s41598-017-12266-x
Huang
Z.
Li
L.
Li
Z.
Li
H.
Wu
J.
2020
Synthesis of novel kaolin-supported g-C3N4/CeO2 composites with enhanced photocatalytic removal of ciprofloxacin
.
Materials
13
(
17
),
3811
.
https://doi.org/10.3390/ma13173811
Imam
S. S.
Adnan
R.
Kaus
N. H. M.
2018
Influence of yttrium doping on the photocatalytic activity of bismuth oxybromide for ciprofloxacin degradation using indoor fluorescent light illumination
.
Research on Chemical Intermediates
44
(
9
),
5357
5376
.
https://doi.org/10.1007/s11164-018-3427-8
Izadifard
M.
Achari
G.
Langford
C. H.
2013
Application of photocatalysts and LED light sources in drinking water treatment
.
Catalysts
3
(
3
),
726
743
.
https://doi.org/10.3390/catal3030726
Jacob
S. A.
Jerome Das
S.
2016
Effect of hydrothermal synthesis on the particulate characteristics of Nanocrystalline titanium dioxide
.
Materials Today: Proceedings
3
(
6
),
1599
1603
.
https://doi.org/10.1016/j.matpr.2016.04.048
Jia
T.
Fu
F.
Yu
D.
Cao
J.
Sun
G.
2018
Facile synthesis and characterization of N-doped TiO2/C nanocomposites with enhanced visible-light photocatalytic performance
.
Applied Surface Science
430
,
438
447
.
https://doi.org/10.1016/j.apsusc.2017.07.024
Jiang
J.
Li
H.
Zhang
L.
2012
New insight into daylight photocatalysis of AgBr@Ag: synergistic effect between semiconductor photocatalysis and plasmonic photocatalysis
.
Chemistry – A European Journal
18
(
20
),
6360
6369
.
https://doi.org/10.1002/chem.201102606
Jiao
Y.
Han
D.
Lu
Y.
Rong
Y.
Fang
L.
Liu
Y.
Han
R.
2017
Characterization of pine-sawdust pyrolytic char activated by phosphoric acid through microwave irradiation and adsorption property toward CDNB in batch mode
.
Desalination and Water Treatment
77
,
247
255
.
https://doi.org/10.5004/dwt.2017.20780
Jo
W. K.
Tayade
R. J.
2014
New generation energy-efficient light source for photocatalysis: LEDs for environmental applications
.
Industrial and Engineering Chemistry Research
53
(
6
),
2073
2084
.
https://doi.org/10.1021/ie404176 g
Karbasi
M.
Karimzadeh
F.
Raeissi
K.
Rtimi
S.
Kiwi
J.
Giannakis
S.
Pulgarin
C.
2020
Insights into the photocatalytic bacterial inactivation by flower-like Bi2WO6 under solar or visible light, through in situ monitoring and determination of reactive oxygen species (ROS)
.
Water (Switzerland)
12
(
4
),
1099
.
https://doi.org/10.3390/W12041099
Kiwaan
H. A.
Atwee
T. M.
Azab
E. A.
El-Bindary
A. A.
2020
Photocatalytic degradation of organic dyes in the presence of nanostructured titanium dioxide
.
Journal of Molecular Structure
1200
,
127115
.
https://doi.org/10.1016/j.molstruc.2019.127115
Kumar
A.
Prajapati
P. K.
Pal
U.
Jain
S. L.
2018
Ternary rGO/InVO4/Fe2O3 Z-Scheme Heterostructured photocatalyst for CO2 reduction under visible light irradiation
.
ACS Sustainable Chemistry and Engineering
6
(
7
),
8201
8211
.
https://doi.org/10.1021/acssuschemeng.7b04872
Lai
Y. S.
Yang
C. H.
Jehng
J. M.
2018
The formation of (NH4)2V6O16 phase in the synthesized InVO4 for the hydrogen evolving applications
.
Catalysis Communications
103
,
19
23
.
https://doi.org/10.1016/j.catcom.2017.09.008
Lamdab
U.
Wetchakun
K.
Phanichphant
S.
Kangwansupamonkon
W.
Wetchakun
N.
2015
Highly efficient visible light-induced photocatalytic degradation of methylene blue over InVO4/BiVO4 composite photocatalyst
.
Journal of Materials Science
50
(
17
),
5788
5798
.
https://doi.org/10.1007/s10853-015-9126-6
Lamers
M.
Fiechter
S.
Friedrich
D.
Abdi
F. F.
Van De Krol
R.
2018
Formation and suppression of defects during heat treatment of BiVO4 photoanodes for solar water splitting
.
Journal of Materials Chemistry A
6
(
38
),
18694
18700
.
https://doi.org/10.1039/c8ta06269b
Lee
S. H.
Choi
J. I.
Kim
Y. J.
Han
J. K.
Ha
J.
Novitskaya
E.
Talbot
J. B.
McKittrick
J.
2015
Comparison of luminescent properties of Y2O3:Eu3+and LaPO4:Ce3+, Tb3+ phosphors prepared by various synthetic methods
.
Materials Characterization
103
,
162
169
.
https://doi.org/10.1016/j.matchar.2015.03.027
Lestari
A.
Iwan
S.
Djuhana
D.
Imawan
C.
Harmoko
A.
Fauzia
V.
2016
Effect of precursor concentration on the structural and optical properties of ZnO nanorods prepared by hydrothermal method
.
AIP Conference Proceedings
1729
.
https://doi.org/10.1063/1.4946930
Li
Y.
Xu
S.
2018
The contribution of Eu3+ doping concentration on the modulation of morphology and luminescence properties of InVO4:Eu3+
.
RSC Advances
8
(
56
),
31905
31910
.
https://doi.org/10.1039/c8ra02716a
Li
Y.
Cao
M.
Feng
L.
2009
Hydrothermal synthesis of mesoporous InVO4 hierarchical microspheres and their photoluminescence properties
.
Langmuir
25
(
3
),
1705
1712
.
https://doi.org/10.1021/la803682d
Liang
H.
Wang
S.
Lu
Y.
Ren
P.
Li
G.
Yang
F.
Chen
Y.
2021
Highly efficient and cheap treatment of dye by graphene-doped TiO2 microspheres
.
Water Science and Technology
83
(
1
),
223
232
.
https://doi.org/10.2166/wst.2020.545
Lin
X.
Guo
X.
Shi
W.
Zhao
L.
Yan
Y.
Wang
Q.
2015a
Ternary heterostructured Ag-BiVO4/InVO4 composites: synthesis and enhanced visible-light-driven photocatalytic activity
.
Journal of Alloys and Compounds
635
,
256
264
.
https://doi.org/10.1016/j.jallcom.2015.02.063
Lin
X.
Xu
D.
Lin
Z.
Jiang
S.
Chang
L.
2015b
Construction of heterostructured TiO2/InVO4/RGO microspheres with dual-channels for photo-generated charge separation
.
RSC Advances
5
(
103
),
84372
84380
.
https://doi.org/10.1039/c5ra17676j
Lin
X.
Xu
D.
Zheng
J.
Song
M.
Che
G.
Wang
Y.
Yang
Y.
Liu
C.
Zhao
L.
Chang
L.
2016
Graphitic carbon nitride quantum dots loaded on leaf-like InVO4/BiVO4 nanoheterostructures with enhanced visible-light photocatalytic activity
.
Journal of Alloys and Compounds
688
,
891
898
.
https://doi.org/10.1016/j.jallcom.2016.07.275
Liu
Z.
Lu
Q.
Guo
E.
Liu
S.
2016
Electrospinning synthesis of InVO4/BiVO4 heterostructured nanobelts and their enhanced photocatalytic performance
.
Journal of Nanoparticle Research
18
(
8
),
1
11
.
https://doi.org/10.1007/s11051-016-3555-2
Liu
G.
Li
F.
Zhu
Y.
Li
J.
Sun
L.
2020
Cobalt doped BiVO4 with rich oxygen vacancies for efficient photoelectrochemical water oxidation
.
RSC Advances
10
(
48
),
28523
28526
.
https://doi.org/10.1039/d0ra01961e
Ma
Y.
Zhang
J.
Tian
B.
Chen
F.
Wang
L.
2010
Synthesis and characterization of thermally stable Sm,N co-doped TiO2 with highly visible light activity
.
Journal of Hazardous Materials
182
(
1–3
),
386
393
.
https://doi.org/10.1016/j.jhazmat.2010.06.045
Ma
D.
Zhang
Y.
Gao
M.
Xin
Y.
Wu
J.
Bao
N.
2015
RGO/InVO4 hollowed-out nanofibers: electrospinning synthesis and its application in photocatalysis
.
Applied Surface Science
353
,
118
126
.
https://doi.org/10.1016/j.apsusc.2015.06.067
Mathew
S.
Ganguly
P.
Rhatigan
S.
Kumaravel
V.
Byrne
C.
Hinder
S. J.
Bartlett
J.
Nolan
M.
Pillai
S. C.
2018
Cu-Doped TiO2: visible light assisted photocatalytic antimicrobial activity
.
Applied Sciences (Switzerland)
8
(
11
), 2067.
https://doi.org/10.3390/app8112067
Moongraksathum
B.
Shang
J. Y.
Chen
Y. W.
2018
Photocatalytic antibacterial effectiveness of Cu-doped TiO2 thin film prepared via the peroxo sol-gel method
.
Catalysts
8
(
9
), 352.
https://doi.org/10.3390/catal8090352
Motola
M.
Satrapinskyy
L.
Čaplovicová
M.
Roch
T.
Gregor
M.
Grančič
B.
Greguš
J.
Čaplovič
Ľ.
Plesch
G.
2018
Enhanced photocatalytic activity of hydrogenated and vanadium doped TiO2 nanotube arrays grown by anodization of sputtered Ti layers
.
Applied Surface Science
434
,
1257
1265
.
https://doi.org/10.1016/j.apsusc.2017.11.253
Oswald
S.
Thoss
F.
Zier
M.
Hoffmann
M.
Jaumann
T.
Herklotz
M.
Nikolowski
K.
Scheiba
F.
Kohl
M.
Giebeler
L.
Mikhailova
D.
Ehrenberg
H.
2018
Binding energy referencing for XPS in alkali metal-based battery materials research (II): application to complex composite electrodes
.
Batteries
4
(
3
),
36
.
https://doi.org/10.3390/batteries4030036
Priya
A.
Senthil
R. A.
Selvi
A.
Arunachalam
P.
Senthil Kumar
C. K.
Madhavan
J.
Boddula
R.
Pothu
R.
Al-Mayouf
A. M.
2020
A study of photocatalytic and photoelectrochemical activity of as-synthesized WO3/g-C3N4 composite photocatalysts for AO7 degradation
.
Materials Science for Energy Technologies
3
,
43
50
.
https://doi.org/10.1016/j.mset.2019.09.013
Reza
K. M.
Kurny
A.
Gulshan
F.
2017
Parameters affecting the photocatalytic degradation of dyes using TiO2: a review
.
Applied Water Science
7
(
4
),
1569
1578
.
https://doi.org/10.1007/s13201-015-0367-y
Sabar
S.
Aziz
H. A.
Yusof
N. H.
Subramaniam
S.
Foo
K. Y.
Wilson
L. D.
Lee
H. K.
2020
Preparation of sulfonated chitosan for enhanced adsorption of methylene blue from aqueous solution
.
Reactive and Functional Polymers
0
,
104584
.
https://doi.org/10.1016/j.reactfunctpolym.2020.104584
Salehi
M.
Hashemipour
H.
Mirzaee
M.
2012
Experimental study of influencing factors and kinetics in catalytic removal of methylene blue with TiO2 nanopowder
.
American Journal of Environmental Engineering
2
(
1
),
1
7
.
https://doi.org/10.5923/j.ajee.20120201.01
Shard
A. G.
2014
Detection limits in XPS for more than 6000 binary systems using Al and Mg Kα X-rays
.
Surface and Interface Analysis
46
(
3
),
175
185
.
https://doi.org/10.1002/sia.5406
Shih
H. R.
Liu
K. T.
Teoh
L. G.
Wei
L. K.
Chang
Y. S.
2015
Synthesis and photoluminescence properties of (La, Pr) co-doped InVO4 phosphor
.
Microelectronic Engineering
148
,
10
13
.
https://doi.org/10.1016/j.mee.2015.07.007
Silversmit
G.
Depla
D.
Poelman
H.
Marin
G. B.
De Gryse
R.
2004
Determination of the V2p XPS binding energies for different vanadium oxidation states (V5+ to V0+)
.
Journal of Electron Spectroscopy and Related Phenomena
135
(
2–3
),
167
175
.
https://doi.org/10.1016/j.elspec.2004.03.004
Sing
K. S. W.
Everett
D. H.
Haul
R. A. W.
Pierotti
R. A.
Mouscou
R.
Moscou
L.
Siemieniewska
T.
1985
Reporting Physisorption Data for Gas Solid. Systems with Special Reference to the Determination of Surface Area and Porosity
.
Pure & Applied Chemistry
57
,
603
619
. https:// doi.org/10.1351/pac198557040603.
Sun
X.-Y.
Zhang
X.
Sun
X.
Qian
N.-X.
Wang
M.
Ma
Y.-Q.
2019
Improved adsorption and degradation performance by S-doping of (001)-TiO2
.
Beilstein Journal of Nanotechnology
10
,
2116
2127
.
https://doi.org/10.3762/bjnano.10.206
Tan
Y. N.
Wong
C. L.
Mohamed
A. R.
2011
An overview on the photocatalytic activity of nano-doped- TiO2 in the degradation of organic pollutants
.
ISRN Materials Science
2011
,
1
18
.
https://doi.org/10.5402/2011/261219
Tetteh
E. K.
Rathilal
S.
Naidoo
D. B.
2020
Photocatalytic degradation of oily waste and phenol from a local South Africa oil refinery wastewater using response methodology
.
Scientific Reports
10
(
1
),
1
12
.
https://doi.org/10.1038/s41598-020-65480-5
Van de Moortel
W.
Kamali
M.
Sniegowski
K.
Braeken
L.
Degrève
J.
Luyten
J.
Dewil
R.
2020
How photocatalyst dosage and ultrasound application influence the photocatalytic degradation rate of phenol in water: elucidating the mechanisms behind
.
Water (Switzerland)
12
(
6
),
1672
.
https://doi.org/10.3390/W12061672
Wang
M.
Liu
Q.
Che
Y.
Zhang
L.
Zhang
D.
2013a
Characterization and photocatalytic properties of N-doped BiVO4 synthesized via a sol-gel method
.
Journal of Alloys and Compounds
548
,
70
76
.
https://doi.org/10.1016/j.jallcom.2012.08.140
Wang
Y.
Dai
H.
Deng
J.
Liu
Y.
Zhao
Z.
Li
X.
Arandiyan
H.
2013b
Three-dimensionally ordered macroporous InVO4: fabrication and excellent visible-light-driven photocatalytic performance for methylene blue degradation
.
Chemical Engineering Journal
226
,
87
94
.
https://doi.org/10.1016/j.cej.2013.04.032
Wang
Y.
Duan
W.
Liu
B.
Chen
X.
Yang
F.
Guo
J.
2014
The effects of doping copper and mesoporous structure on photocatalytic properties of TiO
.
Journal of Nanomaterials
2014
.
https://doi.org/10.1155/2014/178152
Wang
M.
Nie
B.
Yee
K.-K.
Bian
H.
Lee
C.
Lee
H.
Zheng
B.
Lu
J.
Luo
L.
Li
Y.
2016a
Low-temperature fabrication of brown TiO2 with enhanced photocatalytic activities under visible light
.
Chemical Communications
52
(
14
),
2988
2991
.
https://doi.org/10.1039/c5cc09176d
Wang
Y.
Long
Y.
Zhang
D.
2016b
Novel bifunctional V2O5/BiVO4 nanocomposite materials with enhanced antibacterial activity
.
Journal of the Taiwan Institute of Chemical Engineers
68
,
387
395
.
https://doi.org/10.1016/j.jtice.2016.10.001
Wang
J.
Hua
C.
Dong
X.
Wang
Y.
Zheng
N.
2020a
Synthesis of plasmonic bismuth metal deposited InVO4 nanosheets for enhancing solar light-driven photocatalytic nitrogen fixation
.
Sustainable Energy and Fuels
4
(
4
),
1855
1862
.
https://doi.org/10.1039/c9se01136f
Wang
Z.
Wang
J.
Pan
Y.
Liu
F.
Lai
Y.
Li
J.
Jiang
L.
2020b
Preparation and characterization of a novel and recyclable InVO4/ZnFe2O4 composite for methylene blue removal by adsorption and visible-light photocatalytic degradation
.
Applied Surface Science
501
,
https://doi.org/10.1016/j.apsusc.2019.144006
Wen
J.
Xie
J.
Chen
X.
Li
X.
2017
A review on g-C3N4-based photocatalysts
.
Applied Surface Science
391
,
72
123
.
https://doi.org/10.1016/j.apsusc.2016.07.030
Wetchakun
N.
Wanwaen
P.
Phanichphant
S.
Wetchakun
K.
2017
Influence of Cu doping on the visible-light-induced photocatalytic activity of InVO4
.
RSC Advances
7
(
23
),
13911
13918
.
https://doi.org/10.1039/c6ra27138c
Xiao
X.
Jiang
J.
Zhang
L.
2013
Selective oxidation of benzyl alcohol into benzaldehyde over semiconductors under visible light: the case of Bi12O17Cl2 nanobelts
.
Applied Catalysis B: Environmental
142–143
,
487
493
.
https://doi.org/10.1016/j.apcatb.2013.05.047
Yan
Y.
Liu
X.
Fan
W.
Lv
P.
Shi
W.
2012
InVO4 microspheres: preparation, characterization and visible-light-driven photocatalytic activities
.
Chemical Engineering Journal
200–202
,
310
316
.
https://doi.org/10.1016/j.cej.2012.05.102
Yang
J.
Hao
J.
Xu
S.
Wang
Q.
Dai
J.
Zhang
A.
Pang
X.
2019
InVO4/β-AgVO3 nanocomposite as a direct Z-Scheme photocatalyst toward efficient and selective visible-light-driven CO2 reduction
.
ACS Applied Materials and Interfaces
11
(
35
),
32025
32037
.
https://doi.org/10.1021/acsami.9b10758
Yao
J. M.
Lee
C. K.
Yang
S. J.
Hwang
C. S.
2009
Characterization of nano-InVO4 powders synthesized by the hydrothermal process on various In/V molar ratio and soaking conditions
.
Journal of Alloys and Compounds
481
(
1–2
),
740
745
.
https://doi.org/10.1016/j.jallcom.2009.03.093
Ye
J.
Zou
Z.
Oshikiri
M.
Matsushita
A.
Shimoda
M.
Imai
M.
Shishido
T.
2002
A novel hydrogen-evolving photocatalyst InVO4 active under visible light irradiation
.
Chemical Physics Letters
356
(
3–4
),
221
226
.
https://doi.org/10.1016/S0009-2614(02)00254-3
Yuan
X.
Jiang
L.
Liang
J.
Pan
Y.
Zhang
J.
Wang
H.
Leng
L.
Wu
Z.
Guan
R.
Zeng
G.
2019
In-situ synthesis of 3D microsphere-like In2S3/InVO4 heterojunction with efficient photocatalytic activity for tetracycline degradation under visible light irradiation
.
Chemical Engineering Journal
356
(
May 2018
),
371
381
.
https://doi.org/10.1016/j.cej.2018.09.079
Zawawi
A.
Ramli
R.
Yub Harun
N.
2017
Photodegradation of 1-Butyl-3-methylimidazolium Chloride [Bmim]Cl via synergistic effect of adsorption–Photodegradation of Fe-TiO2/AC
.
Technologies
5
(
4
),
82
.
https://doi.org/10.3390/technologies5040082
Zhang
Y.
Ma
D.
Wu
J.
Zhang
Q.
Xin
Y.
Bao
N.
2015
One-step preparation of CNTs/InVO4 hollow nanofibers by electrospinning and its photocatalytic performance under visible light
.
Applied Surface Science
353
,
1260
1268
.
https://doi.org/10.1016/j.apsusc.2015.06.143
Zhang
J.
Wang
J.
Zhu
Q.
Zhang
B.
Xu
H.
Duan
J.
Hou
B.
2020a
Fabrication of a Novel AgBr/Ag2MoO4@InVO4 composite with excellent visible light photocatalytic property for antibacterial use
.
Nanomaterials
10
(
8
),
1541
.
https://doi.org/10.3390/nano10081541
Zhang
K.
Guan
J.
Mu
P.
Yang
K.
Xie
Y.
Li
X.
Zou
L.
Huang
W.
Yu
C.
Dai
W.
2020b
Visible and near-infrared driven Yb3+/Tm3+ co-doped InVO4 nanosheets for highly efficient photocatalytic applications
.
Dalton Transactions
49
(
40
),
14030
.
https://doi.org/10.1039/d0dt02318c
Zhang
S.
Yu
T.
Wen
H.
Guo
R.
Xu
J.
Zhong
R.
Li
X.
You
J.
2020c
Enhanced photocatalytic activity of a visible-light-driven ternary WO3/Ag/Ag3PO4 heterojunction: a discussion on electron transfer mechanisms
.
RSC Advances
10
(
29
),
16892
16903
.
https://doi.org/10.1039/d0ra01731 k
Zhou
Y.
Liu
L.
Wu
T.
Yuan
G.
Li
J.
Ding
Q.
Qi
F.
Zhu
W.
Ouyang
X.
Wang
Y.
2018
Flake-like InVO4 modified TiO2 nanofibers with longer carrier lifetimes for visible-light photocatalysts
.
RSC Advances
8
(
48
),
27073
27079
.
https://doi.org/10.1039/c8ra04344b
Zhu
S.
Wang
D.
2017
Photocatalysis: basic principles, diverse forms of implementations and emerging scientific opportunities
.
Advanced Energy Materials
7
(
23
),
1
24
.
https://doi.org/10.1002/aenm.201700841
Zou
Z.
Ye
J.
Arakawa
H.
2000
Structural properties of InNbO4 and InTaO4: correlation with photocatalytic and photophysical properties
.
Chemical Physics Letters
332
(
3–4
),
271
277
.
https://doi.org/10.1016/S0009-2614(00)01265-3
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/).

Supplementary data