Sujay Shekar
G. C.
a,
Khaled
Alkanad
a,
Gubran
Alnaggar
b,
Nabil
Al-Zaqri
c,
Mohammed Abdullah
Bajiri
d,
Thejaswini
B.
e,
M. D.
Dhileepan
f,
Bernaurdshaw
Neppolian
f and
Lokanath N.
K.
*a
aDepartment of Studies in Physics, University of Mysore, Manasagangotri, Mysuru 570006, India. E-mail: lokanath@physics.uni-mysore.ac.in
bDepartment of Studies in Chemistry, University of Mysore, Manasagangotri, Mysuru 570006, India
cDepartment of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
dDepartment of Studies and Research in Industrial Chemistry, School of Chemical Sciences, Kuvempu University, Shankaraghatta 577451, India
ePG Department of Physics, St. Philomena's College, University of Mysore, Bannimantap, Mysuru 570015, India
fEnergy and Environmental Remediation Lab, SRM-Research Institute of Science and Technology, Chennai 603203, India
First published on 4th January 2022
Semiconductor photocatalysts with surface defects display incredible light absorption bandwidth and these defects function as highly active sites for oxidation processes by interacting with the surface band structure. Accordingly, engineering the photocatalyst with surface oxygen vacancies will enhance the semiconductor nanostructure's photocatalytic efficiency. Herein, a CeO2−x nanostructure is designed under the influence of low-frequency ultrasonic waves to create surface oxygen vacancies. This approach enhances the photocatalytic efficiency compared to many heterostructures while keeping the intrinsic crystal structure intact. Ultrasonic waves induce the acoustic cavitation effect leading to the dissemination of active elements on the surface, which results in vacancy formation in conjunction with larger surface area and smaller particle size. The structural analysis of CeO2−x revealed higher crystallinity, as well as morphological optimization and the presence of oxygen vacancies is verified through Raman, X-ray photoelectron spectroscopy, temperature-programmed reduction, photoluminescence, and electron spin resonance analyses. Oxygen vacancies accelerate the redox cycle between Ce4+ and Ce3+ by prolonging photogenerated charge recombination. The ultrasound-treated pristine CeO2 sample achieved excellent hydrogen production showing a quantum efficiency of 1.125% and efficient organic degradation. Our promising findings demonstrated that ultrasonic treatment causes the formation of surface oxygen vacancies and improves photocatalytic hydrogen evolution and pollution degradation.
Semiconductor nanostructures have demonstrated promising progress in the development of photocatalysis. Whilst pristine semiconductors have a broad light-absorption range, they are frequently plagued by severe electron–hole recombination and poor adsorption on their surfaces.5 Consequently, the efficiency of charge kinetics operations is greatly reduced, and the photocatalytic activity is significantly decreased. Several techniques have been studied to enhance the photocatalytic efficiency of semiconductor nanostructures like defect engineering, doping of metals or non-metals, construction of heterojunctions, metal loading, etc.,6–8
Engineering defects in semiconductor photocatalysts is an efficient method to control and change the regional coordination of active sites in semiconductor photocatalysts.9 For altering the bandgap energies in nanostructures and improving photoexcited electron transport, defect engineering offers great potential.10 Using bulk defects in semiconductors, light absorption bandwidth may be broadened by modulating the band structure, while surface defects can act as highly active sites for oxidation processes by interacting with the surface band structure.11 The development of zero-dimensional point defects such as vacancy generation with various atomic structures in pristine nanostructures can regulate the non-stoichiometric composition of photocatalysts without the need for the introduction of co-catalysts.12 Thus, it is crucial to design photocatalysts with surface defects to increase the efficiency of pristine semiconductor nanostructures.
Ceria, the most abundant rare-earth element, offers many desirable properties as a typical n-type semiconductor, including nontoxicity and sensitivity to photoirradiation.13 The widespread use of ceria as a catalyst for oxidation catalysis is owing to its remarkable oxygen storage capacity, which is a result of its exceptional redox characteristics, which makes it an excellent choice for photocatalytic application.14 The photocatalytic efficiency is enhanced by appropriately utilizing these properties and keeping the intrinsic crystal structure intact compared to many heterostructures.7,15 The higher Ce3+ concentration and subsequent oxygen vacancy formation in CeO2 are valuable in enhancing the photocatalytic efficiency.16 These oxygen vacancies have the potential to prolong the lifespan of charge carriers and therefore contribute to the creation of free electrons and holes, resulting in more active species serving in organic degradation. Consecutively, the OH stability through the surface oxygen vacancies supports the pathway for higher hydrogen production.17
Researchers have synthesized ceria nanostructures with surface oxygen vacancies under different methods. Bui et al. found that base treatment for CeO2 nanostructures exhibited more oxygen vacancies on the surface.5 Wang et al. reported that CeO2−x showed increased oxygen vacancy concentration with the help of Cu doping to promote photocatalytic activities.6 Hezam et al. reported a sunlight-driven synthesis method to fabricate CeO2 nanostructures with oxygen vacancies which was assisted by the use of dense convex lens.7 Choudhury et al. found a way through hydrolysis to red shift the band gap with the presence of surface defects for CeO2 nanoparticles.18 Zhang et al. showed that the self-doping of Ce3+ in CeO2 will promote the formation of oxygen vacancies.19 These synthesis methods to generate oxygen vacancies were aided with expensive instrumentations, doping of metals and various complex chemical processes leading to many drawbacks like thermal instability, dopant concentration dependency, short-term efficiency, adsorption by other coexisting species, tedious multi-step experimental procedures and average photocatalytic activity.20–22
The ultrasound-assisted approach for synthesizing photocatalysts is cost-effective and exhibits more capability than the conventional techniques which could regulate the surface defects without any dependency and displayed notable reusability. The cavitation phenomena caused by the ultrasonic waves show telekinetic effects on the surface of the photocatalyst by accelerating the dissemination of active elements.23 This fascinating cavitation effect corresponding to local high pressure and temperature, intense shear stress, and strong acoustic streaming is caused by the implosion of cavitation bubbles and turbulence resulting in breaking of the bonds allowing the reaction to occur more rapidly.24,25 Moreover, the amount of energy created by the ultrasonic reaction generates charge carriers (electron–hole pairs). A generated hole interacts with a Ce–O bond binding electron. As a consequence of the bond breaking, the single oxygen atom on the surface of CeO2 desorbs as oxygen (upon coupling with the other O atom), forming an oxygen vacancy. In the meantime, certain Ce4+ ions capture electrons and get converted to Ce3+. Besides, the ultrasound treatment enhances the specific area of nanostructures with active sites resulting in higher photocatalytic efficiency.26
In this work, CeO2 is defect-engineered to obtain surface oxygen vacancies by employing ultrasonic irradiation and used for enhanced photocatalytic hydrogen production and photocatalytic degradation of brilliant blue and phenol. The CeO2−x nanostructure is designed under the influence of low-frequency ultrasonic waves for the first time to produce surface defects. The formation of oxygen vacancies is confirmed by a variety of sophisticated analytical methods. The effect of ultrasound and the significant role of surface oxygen vacancies in higher hydrogen production is studied through the reaction pathway.
2 g of the prepared CeO2 was dispersed in 100 mL of ethanol and stirred vigorously at room temperature. The solution was kept in an ultrasonic bath operating at 20 kHz with 100 W power containing water for stabilizing the temperature. The solution is then exposed to ultrasound irradiation for 3 h. The precipitate is later purified by filtration and washed with distilled water three times. Finally, the product was dried in a hot-air oven for 90 °C for 12 h to obtain CeO2−x nanoparticles.
The photocatalytic activity of the materials was also studied by the degradation of brilliant blue dye (BB) under visible light irradiation. In a typical method, the catalyst (0.5 g L−1) and the BB (10 mg L−1) solution mixture are stirred in the dark for 30 min to accomplish adsorption–desorption equilibrium before sunlight irradiation. The absorption band of BB is detected at an equal interval of time using a UV-vis spectrophotometer to verify the productivity of the catalyst.
Additionally, the photocatalytic efficiency of the prepared samples was examined by the degradation of phenol. The suspension containing 200 ml of phenol (30 mg L−1) and 0.1 g of the photocatalyst was stirred at room temperature until it reached equilibrium. After visible light irradiation, 5 ml aliquots were collected from the suspension at an equal time interval. The phenol concentration was tested using the 4-aminoantipyrine spectrophotometric method.
The Raman spectra of CeO2 and U-CeO2−x are displayed in Fig. 1b. The standard peak of CeO2 positioned at 470.6 cm−1 is ascribed to the F2g mode of a fluorite-type structure.27 Meanwhile, the presence of oxygen vacancies is demonstrated by a minor peak at 602 cm−1.28 The minor shift and the presence of Ce3+ in U-CeO2−x have confirmed the intrinsic oxygen vacancies.
The synthesized CeO2 and U-CeO2−x samples were subjected to FTIR analysis. Fig. 1c shows the FTIR spectra of the synthesized samples recorded in the range of 400–4000 cm−1. The intense peak at 630 cm−1 is assigned to Ce–O–Ce stretching vibration, confirming the formation of CeO2.29 The CO and C–O bond vibrations are observed at around 884 and 1369 cm−1, respectively. The band at 1616 cm−1 corresponds to the bending vibration of H2O absorbed from the atmosphere. The bands observed at 2007 and 2184 cm−1 are ascribed to the stretching vibration of C–H bonds. The small band at 2552 cm−1 is due to the absorption of atmospheric CO2. The weak bands at 3190 and 3903 cm−1 are attributed to the OH stretching of water.30
The morphology of the synthesized samples was investigated through SEM and TEM analysis. Fig. 1(d) shows the CeO2 nanoparticles having a porous network-like structure, and Fig. 1(e) shows U-CeO2−x having a spongy network-like structure with larger pores than its counterpart, indicating a smaller size of the particles. Fig. 1(f and g) show the TEM images of the U-CeO2−x sample demonstrating a uniform particle distribution with a smaller size. The HRTEM image shown in Fig. 1(h) shows the lattice fringes of the CeO2−x sample with 0.31 nm, 0.27 nm and 0.19 nm interlayer spacings for the (111), (200) and (220) planes, respectively, with better interplanar interface contact. This result is in agreement with the XRD studies.
EDAX with SEM was carried out to confirm the purity of the samples further. The EDAX spectrum shown in Fig. S1† confirms the Ce, O, and C elements in the photocatalyst. The weight percentages of Ce, O, and C were around 66.97, 28.94, and 4.09%, respectively.
The oxidation states and the presence of oxygen vacancies are determined by XPS analysis. The XPS spectrum of CeO2 and U-CeO2−x is shown in Fig. 2(a–d). The Ce 3d split levels are deconvoluted and fitted to analyze additional peaks and intensity variation. The peaks due to the Ce 3d3/2 level for Ce3+ ions are found at 882.7 and 896.9 eV, whereas, for the Ce4+ state, a peak at 888.1 eV is formed.31,32 At the same time, the peak owing to the Ce 3d5/2 level for Ce4+ cations has appeared at 902 and 916.9 eV along with a peak at 907.6 eV for the Ce3+ state.7,28,33 The observed peaks for the U-CeO2−x sample have the same peaks with a minor variation in their positions; nevertheless, a noteworthy change in the intensity of the peaks was observed. The ultrasound irradiation has affected CeO2 with a better proportion of Ce3+ ions [59.54%], which ultimately demonstrates as a critical factor for the formation of oxygen vacancies and improved catalytic activity.34 The signature of oxygen vacancies is ultimately traced in the XPS profile of O 1s. The conventional combustion-synthesized CeO2 showed characteristic peaks of lattice oxygen (Oa) and surface OH− (Ob) at 529.14 and 531.9 eV, respectively.35,36 On the other hand, the ultrasound-treated CeO2 along with the characteristic peaks exhibited an elevated peak corresponding to chemisorbed oxygen (Oc) at 531.2 eV.7 The chemisorbed oxygen indicates the presence of oxygen vacancies and enhances the mobility over the catalyst.28,37,38 Thus, a larger proportion of chemisorbed oxygen species [15.64%] and the significant Ce3+ ions indicate the evolution of higher oxygen vacancies. Due to the oxygen vacancies, free electrons and holes are created resulting in more active species for enhanced degradation of contaminants. Accordingly, the hydrogen evolution has been boosted due to the stability of OH caused by surface oxygen vacancies. These results are shown in Table S1.† The relative portions of Ce3+ and oxygen vacancies are calculated using the formulas:
![]() | ||
Fig. 2 Ce 3d XPS spectrum of (a) CeO2 and (c) U-CeO2−x. O 1s XPS spectrum of (b) CeO2 and (d) U-CeO2−x. |
An optical reflectance study to measure the bandgap was carried out using DRS analysis. Fig. 3a shows the UV-vis DRS spectra of the CeO2 and U-CeO2−x samples. The ultrasound-treated sample showed a relatively lower reflectance rate indicating higher absorption of the corresponding wavelength. Since U-CeO2−x exhibits better absorption of UV and visible wavelength, it demonstrates an enhanced photocatalytic activity. From the spectra, the Kubelka–Munk function for measuring the bandgap can be used as,40
F(R∞)hν = A(hν − Eg)n |
The particle size distribution was analyzed by performing DLS for the samples, as shown in Fig. 3(c and d). The catalyst's aggregate size impacts the kinetics of photocatalytic activity since the smaller particle size exhibits a higher surface area leading to better photocatalytic performance.43,44 The CeO2 sample shows a size distinction of around 33.9 nm, while the U-CeO2−x sample displayed a particle diameter of about 15.6 nm. This infers that U-CeO2−x has ample surface defects exhibiting enhanced photocatalytic efficiency.
The surface area of the prepared samples was examined using a Brunauer–Emmett–Teller (BET) surface analyzer, as shown in Fig. 3e. The nitrogen absorption–desorption isotherm showed an H3-type hysteresis loop with a type III isotherm signifying a mesoporous structure.45 The specific surface area of U-CeO2−x (51.3 m2 g−1) was comparatively higher than that of the CeO2 (19.8 m2 g−1) sample. Thus, the high surface area and mesoporosity of U-CeO2−x have resulted in a higher probability of surface reaction and improved photocatalytic activity.
Catalyst | Synthesis method | Light source | H2 production (in μmol for 5 h) | Reference |
---|---|---|---|---|
U-CeO2−x | Ultrasound treatment | 300 W Xe lamp | 12![]() |
This work |
≥420 nm | ||||
rCeO2/g-C3N4 | Self-assembly | 300 W Xe lamp | 5500 | 16 |
≥420 nm | ||||
CeO2/gC3N4-5% | Self-assembly | 300 W Xe lamp | 4300 | 3 |
≥420 nm | ||||
1.5%CeO2/g-C3N4 | Facial synthesis | 300 W Xe lamp | 4150 | 47 |
≥420 nm | ||||
CeO2@N,S−C HN | Calcination | 300 W Xe lamp | 2775 | 48 |
≥420 nm | ||||
15%W18O49/CeO2 | Hydrothermal | 300 W Xe lamp | 1030 | 49 |
≥420 nm | ||||
Mn0.2Cd0.8S/CeO2 | Solvothermal | 5 W Xe lamp | 655 | 50 |
≥400 nm | ||||
1T/2H-MoS2@CeO2 | Hydrothermal | 3 W UV-LEDs | 366 | 51 |
≥420 nm | ||||
CeO2@MoS2/gC3N4 | Hydrothermal | 3 W UV-LEDs | 327 | 52 |
≥420 nm | ||||
CeO2/ZnO | Electrodeposition | 300 W Xe lamp | 13.5 | 53 |
≥420 nm |
The photocatalytic degradation test is conducted with BB dye and phenol under visible light. Fig. 4c and d show the BB and phenol degradation curves of the U-CeO2−x sample at 595 nm and 270 nm, respectively, for different irradiation times. The adsorption activity was assessed under dark conditions for both dye and phenol, where the results are displayed in Fig. 4e and f, respectively. The photocatalytic competence of the samples was calculated,
ln(C0/C) = kt |
Total organic carbon was measured to get the degree of the mineralization and it was found that phenol was mineralized better than BB dye (see Fig. S3†). TOC results were found to mineralize about 50% in 30 min and 61% in 80 min for BB dye and phenol, respectively.
The data obtained by several researchers as reported in Table S2† show that phenol has longer degradation time than brilliant blue dye. This is mainly due to brilliant blue dye being a photosensitizer that absorbs more light so the photoreaction takes place rapidly. Moreover, the degradation pathway of BB dye (see Fig. S6†) shows the disappearance of the reactant and breaking into two different compounds within 30 min implying that it wasn't completely mineralized. According to the conventional spectrophotometric method, the lambda max of BB dye almost disappeared within 30 min.
CBM = VBM − Eg |
The degradation mechanism was simulated in the presence of scavengers under optimal conditions in order to establish the most reactive species, namely, p-benzoquinone (BQ), sodium azide (NaN3), and isopropanol (IPA) for superoxide radicals (˙O2−), singlet oxygen (1O2), and hydroxyl radicals (˙OH), respectively. The addition of BQ into the reaction system displayed less efficiency illustrating the influence of ˙O2−. The presence of NaN3 quenched the reaction to a larger extent with the least efficiency exhibiting the impact of surface oxygen vacancies.55 In comparison, the addition of the IPA scavenger showed a better photodegradation efficiency indicating the lower contribution of ˙OH radicals in the pollutant removal. Thus, the scavenger test illustrates the significant involvement of ˙O2−, 1O2, and ˙OH in the degradation process as shown in Fig. 5a. ESR analysis was performed to confirms the presence of the above ROS by employing radical trapping substances. The generated ˙OH was detected by using DMPO as a radical trapping substance in an aqueous solution.8 Fig. S4† shows the ESR signal for the CeO2−x sample in the dark and under solar simulated light. Under dark conditions, no amount of ˙OH was detected; however, the characteristic peaks of DMPO–˙OH appeared while irradiated, suggesting the generation of ˙OH. The ˙O2− was identified using DMPO in the methanol solution and no distinctive peaks were observed in the dark, while the characteristic DMPO–˙O2− peaks were observed as shown in Fig. S5.† The 1O2 generation was confirmed by employing TEMP as a trapping substance.54 Fig. S6† shows no peaks under dark conditions and a distinctive TEMPO peak when illuminated, confirming the presence of 1O2.
The charge recombination rate and the electronic structure of the synthesized samples were studied by PL measurements. The PL spectra of CeO2 and U-CeO2−x recorded in the range of 350 nm to 700 nm with a 410 nm excitation wavelength are shown in Fig. 5b. A fine UV emission band around 359.4 nm originates due to the near bandgap emission of CeO2. A visible emission peak at 419.8 nm arises in the violet region, while significant peaks in the spectra at 467.2 and 481.4 nm emerge in the blue region. Besides, a strong green emission peak at 529.9 nm and a trifling peak at 572.16 nm are observed.56 Altogether, the visible emission bands correspond to oxygen vacancies by the transfer of charge carriers.57 The decrease in the intensity of the U-CeO2−x peaks is due to having a higher content of oxygen vacancies, which allows the excited states to have an extended lifetime. Thus, the PL spectra show efficient charge separation of U-CeO2−x for higher photocatalytic activity.
The impedance Nyquist plots were obtained to confirm the electron–hole separation efficiency of the ceria samples, as shown in Fig. 5c. The intended response of a simple circuit is shown by the semicircle arc in the Nyquist plot, which explains the charge transfer resistance. The contributions from the charge transfer resistance (Rct) and constant phase element from the sample's interface are ascribed to the semicircular arcs in the EIS spectra.58 A substantial shift in the Rct value is observed due to the formation of surface defects on the U-CeO2−x nanostructure.59 Accordingly, the semicircle radius of U-CeO2−x in the EIS is smaller, indicating a lower resistance and effective separation of photogenerated electron–hole pairs and faster interfacial charge transfer. Thus, EIS evidences that the surface defect-induced U-CeO2−x sample has higher charge carrier separation efficiency and better photocatalytic activity.
Fig. 5d shows the photocurrent densities of the CeO2 and U-CeO2−x samples, which were measured in the dark and under visible light irradiation to examine the photoreactivity. The improvement in the photocurrent response of U-CeO2−x is attributed to an increase in light absorption capacity caused by surface defects that generate a mid-gap state, which assists in narrowing the bandgap.60 In general, a high photocurrent indicates that the sample possesses a significant capacity to generate and transmit the photoexcited charge carrier when exposed to light radiation. The photocurrent of U-CeO2−x is greater than those of its counterpart under the same parameters, indicating that U-CeO2−x has a better capacity in the separation of electron–hole pairs. Thus, the enhanced photocatalytic activity of the U-CeO2−x sample can be attributed to the excellent light-harvesting capacity and the efficient separation of photogenerated charge carriers.
The ESR signals were studied to probe the oxidation states of the ceria samples, and Fig. 5e shows the ESR spectra of the CeO2 and U-CeO2−x samples. The ESR spectra of the samples were recorded at room temperature under a 100 kHz modular frequency. A strong signal is observed at g = 1.96 and mainly attributed to Ce3+ ions in the synthesized samples.60 The U-CeO2−x sample shows an increased intensity signal, owing to the reduction of Ce4+ to Ce3+ ions. These Ce3+ ion's trigonal sites are recognized as near-surface oxygen defects.61 Accordingly, the two samples show a single ESR signal corresponding to oxygen vacancies, and the signal intensity is very high for the U-CeO2−x sample. Thus, the U-CeO2−x sample showed the presence of Ce3+ ions and a higher concentration of oxygen vacancies.
Based on the position of the VBM and CBM of the samples, the band energy alignment and the photocatalytic mechanism are shown in Fig. 5f. When the sample is exposed to visible light irradiation, electron and hole pairs are generated. The photogenerated electrons accumulated on the CBM with a reduction potential higher than H+/H2 resulting in photocatalytic H2 production.17 The U-CeO2−x sample was comprised of Ce3+ and oxygen vacancies on the surface layer assisting the H2 production pathway. On the other hand, when Ce4+ oxidizes to Ce3+, oxygen vacancies (VO) may develop on the surface of CeO2 and O2 gets adsorbed on the surface to produce ˙O2−.28 Similar to the Haber–Weiss reaction, the self-combination of ˙O2− leads to the generation of 1O2. Besides, the generated ˙OH reacts with ˙O2− to produce 1O2 signifying its major role in degrading pollutants.62 The three reactive oxygen species (ROS), ˙O2−, 1O2, and ˙OH, reduce the organic pollutants into intermediates leading to the decomposition.
VO(2Ce3+) + O2 → [Ce4+ + Ce3+] + ˙O2− |
H2O + h+ → H+ + ˙OH |
2˙O2− + 2H2O → 1O2 + H2O2 + 2OH− |
˙O2− + ˙OH → 1O2 + OH− |
1O2 + pollutants → Intermediate products |
When the ceria solution is subjected to low ultrasonic irradiation, bubbles are formed and squashed by the solution's sonic field. This acoustic cavitation effect leads to a larger surface area, high crystallinity, relatively smaller particle size and phase purity of the sample.66 Furthermore, the ultrasound treatment energizes the chemical reactions in the liquids by producing significant collisions and shock waves, which contribute to particle aggregation and surface defects leading to oxygen vacancy formation.24 Thus, the ultrasound treatment enhances photocatalytic hydrogen evolution and pollutant degradation.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1cy01940f |
This journal is © The Royal Society of Chemistry 2022 |