Next Article in Journal
Assessment of Antioxidant and Scavenging Activities of Various Yogurts Using Different Sample Preparation Procedures
Previous Article in Journal
Kinematic Determinants of the Swimming Push Start in Competitive Swimmers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microwave Assisted Preparation of Barium Doped Titania (Ba/TiO2) as Photoanode in Dye Sensitized Solar Cells

1
Departamento de Quimica Organica, Universidad de Cordoba, Edificio Marie Curie (C-3), Ctra Nnal IV-A, Km 396, E14014 Cordoba, Spain
2
Department of Chemistry, Quaid-i-Azam University, Islamabad 43520, Pakistan
3
School of Applied Sciences and Humanity (NUSASH), National University of Technology, Islamabad 44000, Pakistan
4
Promising Centre for Sensors and Electronic Devices (PCSED), Advanced Materials and Nano-Research Centre, Najran University, Najran 11001, Saudi Arabia
5
Department of Electrical Engineering, Faculty of Engineering, Najran University, Najran 11001, Saudi Arabia
6
Empty Quarter Research Unit, Department of Chemistry, College of Science and Art in Sharurah, Najran University, Sharurah, Najran 11001, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2022, 12(18), 9280; https://doi.org/10.3390/app12189280
Submission received: 25 November 2021 / Revised: 7 January 2022 / Accepted: 21 January 2022 / Published: 16 September 2022

Abstract

:
Pure TiO2 and barium (0.5 wt%) doped TiO2 (Ba/TiO2) nanostructures have been synthesized via facile microwave irradiation method. The pure anatase phase of synthesized photoactive material was confirmed by X-ray diffraction. Ba doping in the TiO2 host structure influenced the optical band gap as confirmed by UV-visible spectroscopy. The optical band gap increased from 3.21 eV for the TiO2 to 3.26 eV for Ba/TiO2. Morphological analysis of synthesized TiO2 and Ba/TiO2 was conducted using scanning electron microscopy. Energy dispersive X-ray spectroscopy confirmed the formation of Ba/TiO2 and no impurities were observed. Electrochemical impedance spectroscopy showed that the charge transfer resistance increased for Ba/TiO2, which reduced dark current creation in a dye-sensitized solar cell. The highest power conversion efficiency (3.24%) was achieved for Ba/TiO2 photoanode compared to 2.1% for a pure TiO2 photoanode-based device.

1. Introduction

Society needs green and cost-effective technologies for the energy sector owing to the global greenhouse effect and depletion of fossil fuels in recent years. Among all existing green energy technologies such as wind turbines, fuel cells, hydropower, tidal power, solar thermal and biomass, solar cells are considered among the most promising and prominent energy sources [1,2,3,4,5]. Dye-sensitized solar cells (DSSCs) constitute an advanced generation of solar cells that mimic natural photosynthesis. With fabrication techniques including solution processable methods, the use of cheap materials and decent power conversion efficiency (PCE), DSSCs are promising candidates for the conversion of solar light into electricity [6,7,8]. The highest PCE of 13% has been achieved in DSSCs by employing ruthenium as a light-absorbing material [9]. Techniques employed to improve the performance of DSSCs aim to minimize recombination of the photoinduced electron hole pair while improving the charge transfer rate of photo-generated electrons in the transparent conducting substrate [10]. Reducing the charge recombination and enhancing charge transfer can be achieved by optimizing the band gap of the semiconducting metal oxide (MOS). The MOS in a DSSC serves as a substrate for dye uplifting as well as provides a path for photoinduced charge generation [11,12].
TiO2 nanostructures are one of the significant photoanode materials owing to their superior properties, including wide band gap, large surface area, non-toxicity, low cost and high stability [13,14]. The morphology of MOS plays a significant role in DSSCs; therefore, a fine-tuned morphology is required to deliver improved PCE in the device [15]. The photocatalytic efficiency of TiO2 is affected by its ability to respond to radiation (UV along with visible light), and thus could be improved by shifting its optical absorption from UV to the visible region. It has been observed that the photocatalytic efficiency of TiO2 is enhanced by modifying its band gap with different dopants [16]. A variety of TiO2 nano-compositions involving nanoparticles, carbon spheres, TiO2 nanoparticle composites, TiO2 micro tablets, hollow spheres, nanotube composites, and nano-fiber nanoparticle composites have been successfully synthesized and used as DSSC photoanode materials [17,18,19,20]. Doping with transition metals [21,22], rare-earth elements [23,24] and noble metals [25] into TiO2 has been widely investigated to reduce TiO2 particle size and improve its (photo)(electro)catalytic performance. Co-doping along with different doping agents including metal–nonmetal, nonmetal–nonmetal, and metal–-metal synergistically improves the optical absorption towards the visible region and reduces the recombination process [26]. Cu-N, B-N, Fe-C, Fe-N, Ni-N, and V-N have been studied as co-dopants with TiO2 [27,28,29,30,31,32,33]. Co-doping of TiO2 with N-C has been reported to modify its band gap, enhancing its photocatalytic activity towards the visible spectrum [34]. Co-doping of titania nanowires with tungsten carbon improves their electrochemical activity towards the oxygen-evolution reaction (OER) [35]. Ag/V-TiO2 was also studied to photo-catalyze the degradation of rhodamine B dye in aqueous solutions. These co-doped materials exhibited the highest photocatalytic activity in the UV-visible region [36], Similarly, it has also been previously reported that barium doping in TiO2 nanoparticles using microwave irradiation increases the optical band gap and reduces the particle size of TiO2 nanoparticles for microbiological activity [37]. Several studies have reported co-doping of TiO2 to enhance its optoelectrical properties.
The present work deals with the preparation of TiO2 and Ba/TiO2 composites using a simple microwave irradiation method. Microwave-assisted swift heating has achieved substantial significance as an emerging and favorable method for one-pot preparation of different metal nanostructures. Structural, optical, morphological, elemental, and electrochemical behaviors of synthesized photoanodes materials are investigated by X-ray diffraction (XRD), UV-visible spectroscopy, scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), and electrochemical-impedance spectroscopy (EIS). Moreover, the photovoltaic performance of synthesized TiO2 and Ba/TiO2 photoanodes has been studied. Increasing the optical band gap and surface area of Ba/TiO2 photoanodes led to a reduction of dark current and increased dye loading capacity [38]. DSSCs prepared using Ba/TiO2 composite photoanode delivered a PCE of 2.34%.

2. Materials and Methods

2.1. Materials

Titanium (IV) isopropoxide TTIP (95% purity), ruthenium dye (N3 dye), polyethylene glycol (PEG), barium chloride and Triton X- 100 were purchased from Alfa Aesar Company India. Ethanol was purchased from Merck Company India. Chloroplatinic acid (99.9%) was purchased from Himedia Company India. All of these chemicals were utilized as received, without any further treatments. All of the solutions were prepared using deionized water (DI).

2.2. Preparation of TiO2 and Ba/TiO2 Photoanode

A facile microwave irradiation method was employed to prepare Ba (0.5)TiO2 photoanode materials. An initial solution was prepared using 2 g of TTIP added with 60 mL of deionized water and stirred for 1 h at 100 °C. A separate solution was prepared using barium chloride dissolved in DI water. After that, both solutions were mixed and kept under stirring for 1 h at room temperature. The solution was kept under microwave irradiation for 5 min at 450 W. The advantage of using a microwave oven relates to a reduced synthesis time and energy consumption as compared to conventional heating. Microwaves provide homogeneous and rapid heating [39]. Afterwards, the obtained photoanode materials were calcined at 500 °C for 1 h to enhance the crystalline nature of the photoanode material. The same procedure was adopted for the preparation of TiO2 photoanode.

2.3. Photoanode Preparation

Fluorine-doped tin oxide (FTO) glass substrates were cleaned in a soap solution followed by distilled water and ethanol. Hereafter, the synthesized photoanode material was prepared using binder solutions of PEG and Triton X100. As prepared, Ba/TiO2 was deposited onto FTO using the doctor blade technique. The coated substrate was sintered at 300°C for 1 h to evaporate the binder as well as to generate porosity in the materials. The sintered substrate was immersed into 0.5 mM of ruthenium dye solution for 24 h to absorb the dye molecule. A thin layer of platinum was deposited onto the FTO glass using 5 mM chloroplatinic acid via doctor blade technique. The photo-electrode was clamped (using binder clips) with the counter-electrode to produce a sandwich-type formation. The liquid redox electrolyte composed of 0.5M potassium iodide (KI) and 0.02 M iodine (I2) dissolved in acetonitrile was filled between the photoelectrode and counter-electrode. The active surface area of the as-prepared DSSC was about 0.25 cm2 and the cross-sectional area of TiO2 and Ba/TiO2, approx. 10 μm.

2.4. Characterization

X-Ray diffraction (PANalytical-X’pert Pro powder diffractometer λ = 1.54 Å using Cu Kα radiation) analysis was conducted to determine the crystal structure of the MOS with 2θ range values from 10° to 80°. The measurement was carried out by step size of 0.0170° up to four decimal accuracies. SEM (ZEIS EVO 18 instrument with 20 kV acceleration voltage and 1.30 K magnification) was used for the morphological studies of the of photoanode materials. UV-visible spectrophotometry (Perkin-Elmer-Lambda-35 spectrophotometer) was used to investigate the optical properties. EIS was conducted by VSP300 biological instrument with a 10 mV voltage amplitude in the frequency range of 1 MHz–1 Hz. The current voltage (I-V) measurement was observed by means of an electrochemical computer-controlled workstation as well as a SAN-EI solar-stimulator. A 150-watt xenon lamp was used as a light source with an AM 1.5 G filter and an intensity of 100 mW cm−2.

3. Results and Discussions

3.1. Structural Analysis

XRD patterns of single TiO2 as well as Ba/TiO2 nanomaterials are displayed in Figure 1. The diffraction peaks of TiO2 and Ba/TiO2 nanomaterials at 2θ values were 25.09°, 37.65°, 47.89°, 53.89°, 55.07°, 62.40°, 70.04° and 75.00°, matching well with JCPDS data (21-1272) having (101), (004), (200), (105), (211), (204) crystal planes, respectively [26]. No other characteristic peaks were reflected in the XRD pattern. Hence, both samples exhibited tetragonal structure, and corresponding lattice constant values are shown in Table 1. In the case of Ba/TiO2 nanomaterials, an increase in intensity was evidenced due to the incorporation of Ba2+ (1.42 nm) into the TiO2 host lattice because the ionic radius of Ba2+ is greater than that of Ti4+ (0.74 nm), which resulted in an increase in crystallinity. This substitution caused lattice distortion, inflation of the TiO2 nanocrystal, and internal stress due to Ba doping [40], which reduced nucleation growth (particle size controller) [41]. Another study showed that doping of substrates with Ba caused shrinkage of the crystallite size of modified nanomaterials [42].
The structural constraints of synthesized TiO2 and Ba/TiO2 nanomaterials were calculated using the following formulas [43]:
1 d 2 = h 2 + k 2 a 2 + l 2 c 2
t = 0.9 λ β cos θ
where d is the d-spacing between two consecutive lattices observed from XRD, (h,k,l) indicate the miller indices, t is the crystallite size, β is the full with half maximum, a, b, and c display as lattice constant for a = b in tetragonal structure, λ indicates the wavelength of the X-ray radiation used, and θ refers to the diffraction angle;
ρ = n M N V
S a = 6 D . ρ
where, ρ is the density of the prepared sample, n is 4 for anatase phase TiO2 and 2 for rutile phase TiO2, M is the molecular weight of the prepared nanomaterials, N corresponds to Avogadro’s constant value, V is the volume of the prepared nanomaterials, and Sa states the specific surface of the prepared sample. Table 1 shows the calculated structural values derived from the XRD patterns of TiO2 and Ba/TiO2 nanomaterials. The average crystallite sizes estimated were 25 nm for single TiO2 and 16 nm for Ba/TiO2 nanomaterials, respectively. When Ba ions were doped into the TiO2 lattice, a decrease in nucleation (and subsequent growth rate) was observed for TiO2 nanomaterials, leading to a reduction in particle size of Ba/TiO2. The reduction in particle size increases the surface area of TiO2. The specific surface area of TiO2 (61 m2/g) was smaller than that of Ba/TiO2 (95 m2/g). This favored dye adsorption in the photo electrode, thereby improving photovoltaic performance in the DSSC.

3.2. Optical Properties

Figure 2 represents the UV-visible spectra of TiO2 and Ba/TiO2 photoanode materials. Ba/TiO2 exhibited a significant visible absorption above 400 nm, which was an indication of changes occurring in the optical absorption of TiO2 photoanode material due to metal doping. Peak edges of Ba/TiO2 photoanode showed some regular and trivial blue shifts comparative to TiO2 photoanode. The blue shift was attributable to the Burstein–Moss effect, leading to the movement of Fermi level towards the conduction band due to an increase in electron concentration because of Ba doping [44,45]. It is very useful to increase the photocurrent density and also the charge transfer resistance of synthesized photoanodes. The absorption of TiO2 and Ba/TiO2 were at 415 nm and 400 nm, respectively.
Optical band gap values were determined by using Tauc plots as presented in Figure 3a. Ba/TiO2 photoanode (3.26 eV) ha a higher band gap value compared to TiO2 photoanode (3.21 eV) because of quantum confinement effects observed due to the smaller size of as-synthesized samples. According to the theory of quantum efficiency, the holes of valence band and conduction band electrons are confined by the potential barrier of the surface. As a result, the band gap energy is increased with decreasing size of synthesized particles [46,47]. The increase in band gap of doped TiO2 can be interpreted in terms of Mott’s transition effect and Burstein–Moss (BM) effects. According to Mott’ criteria, the band gap of the donor combines with the conduction band of the host material [48], while BM effects present a relationship between the band gap and concentration of carriers [49,50], as given below:
E g B M = h 2 2 λ m e 3 π 2 N e  
where λ, h, and me are the effective reduced mass of TiO2, Planck’s constant and the electron free mass, respectively. According to the BM effect, the carrier density of semiconductors increases with Ba doping, which causes the Fermi level to shift into the conduction band [51]. Due to this doping, the low energy transition is clogged, and broadening in the band gap takes place [52]. This novel effect of Ba doping on optical properties can pave the way to its extended use in optoelectronic and photovoltaic applications.
The value of n coefficient is determined by the Tauc relationship (Equation (6)). The n coefficient is related to the possible transition of electrons. For indirect allowed electronic transition, its value is 2 and 3 for indirect forbidden transitions.
αhθ = A (hθEg)n
n = l n   α h θ ln h θ     E g  
where α is the absorption coefficient, h is Plank’s constant, and A is absorption constant [53]. From the slope of ln F® vs. ln (−ν − Eg), the value of n is determined. The result listed in Table 2 represents the approximated value of n = 2 for TiO2 and Ba/TiO2, which is in agreement with values in the literature for TiO2, which is an indirect semiconductor [54]
Figure 4 presents the UV-visible absorption spectrum of dye molecule-adsorbed TiO2 and Ba/TiO2 photoanodes. The Ba/TiO2 photoanode had high absorption compared to the TiO2 photoanode. The increased absorption was due to the increasing surface area of the Ba/TiO2 photoanode, which was also confirmed by XRD results. This behavior enhances photocurrent density in DSSCs. Moreover, the shift in wavelength in the visible region further supported our findings as productive in DSSCs, because the photocatalytic efficiency of TiO2 is affected by its ability to respond to UV radiation along with visible light, which could be improved by shifting its optical absorption from UV to the visible region. Ba doping increased the absorption in the visible region.

3.3. Morphological Studies

The surface morphologies of the TiO2 and Ba/TiO2 nanostructures were subsequently examined by SEM as shown in Figure 5. Both TiO2 (5a) and Ba/TiO2 (5b) exhibited similar agglomerated spherical shape particles. In terms of chemical composition, TiO2 nanostructures contained Ti and O, only, which confirmed that the material was free from impurities, as observed from Figure 5c. The presence of Ba was confirmed in the Ba/TiO2 nanostructures from Figure 5d.

3.4. Electrochemical Behavior of TiO2 and Ba/TiO2

The reaction kinetics linked with DSSCs were investigated using EIS results. EIS spectra of TiO2 and Ba/TiO2 photoanodes under dark conditions with 0.6 V open circuit voltage are depicted in Figure 6. The EIS of DSSC was explained by three semicircles in the high, middle, and low frequency range, corresponding to the electrochemical output in platinized counter-electrode and electrolyte-interface (R1), WE/electrolyte (R2), and diffusion resistance of electrolyte (Rd), respectively. The displacement of the graph from origin represents the series resistance due to the contact resistance of the cell (Rs) [55]. Associated charge transfer in the electrolyte influenced the low frequency region, which was evident from the third semicircle. From the fitted data, the Ba/TiO2 photoanode charge transfer resistance value was higher than the TiO2 photoanode charge transfer resistance value. Barium doping increased the band gap value (which was also confirmed by optical analysis), and this reduces electron recombination in the Ba/TiO2/electrolyte interface. This behavior is attributed to the charge transfer value increment. The increment of charge transfer resistance was helpful to reduce charge recombination, which was evident for the photocurrent increment.

3.5. Photovoltaic Studies

The photovoltaic performance of the fabricated DSSC is shown in Figure 7. Photocurrent density of Ba/TiO2 was higher than that of the TiO2 photoanode. This increment was due to the increased photon absorption of dye molecule as compared to the TiO2 counterpart. The Ba/TiO2 photoanode had higher surface area than the TiO2 photoanode, which led to enhanced dye-loading capacity. These results were also confirmed by the XRD and UV analysis. Another important factor was that the increment in photocurrent density was supported by the recombination rate or dark current reduction. The Ba/TiO2 photoanode had the high recombination resistance compared to the TiO2 photoanode, which reduce dark current production between the MOS and electrolyte.
The Ba/TiO2 photoanode yielded higher PCE (3.4%) compared to the TiO2 photoanode (2.1%). The calculated photovoltaic parameters are summarized in Table 3. The higher PCE of the Ba/TiO2 photoanode (3.4%) corresponds to a maximum output power density (Pmax) of approximately 34.4 μW/cm2 as compared to the TiO2 photoanode’s Pmax of 26.5 μW/cm2. These findings suggest that DSSCs can be produced with the appropriate power rating to operate electronic devices such as tablets and Internet-of-Things gadgets that require low power Pmax < 100 μW/cm2 [56].

4. Conclusions

Incorporation of barium (0.5 wt%) into TiO2 was successfully accomplished using a microwave irradiation method. With respect to Ba content, its incorporation modulated the band gap of TiO2 by decreasing particle size, leads to increased specific surface area and consequently, superior dye uptake and high photocurrent, resulting in higher DSSC power conversion efficiency. The higher band gap value of Ba/TiO2 (3.26 eV) as compared to TiO2 photoanode (3.21 eV) was attributable to the quantum confinement effect, which was observed due to the reduced size of the as-synthesized samples. EDS results confirmed the presence of barium in Ba/TiO2 without any impurities; this was further confirmed by XRD results. A higher efficiency of Ba/TiO2 photoanode was observed as compared to pure TiO2 (3.4 vs 2.1%). Barium doped titania (Ba/TiO2) could be employed in DSSCs owing to its relevant photoactive properties, with further prospects for other applications including photocatalysis as well as wastewater treatment.

Author Contributions

Conceptualization, A.A. and S.K.; methodology, M.K.; software, M.K.; validation, M.K.; writing—original draft preparation, A.A.; writing—review and editing, R.L.; visualization, M.J.; supervision, R.L.; project administration, M.A.A.; funding acquisition, M.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Najran University, and the grant number is NU/IFC/INT/01/003.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors would like to acknowledge the support of the deputy for research and innovation-Ministry of Education, Kingdom of Saudi Arabia for this research through grant (NU/IFC/INT/01/003) under the institutional funding committee at Najran University, Kingdom of Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perera, F. Pollution from fossil-fuel combustion is the leading environmental threat to global pediatric health and equity: Solutions exist. Int. J. Environ. Res. Public Health 2018, 15, 16. [Google Scholar] [CrossRef] [PubMed]
  2. Khan, S.; Shah, S.S.; Anjum, M.A.R.; Khan, M.R.; Janjua, N.K. Electro-oxidation of ammonia over copper oxide impregnated γ-Al2O3 nanocatalysts. Coatings 2021, 11, 313. [Google Scholar] [CrossRef]
  3. Tiewsoh, S.L.; Jirásek, J.; Sivek, M. Electricity generation in India: Present state, future outlook and policy implications. Energies 2019, 12, 1361. [Google Scholar] [CrossRef]
  4. Khan, M.; Janjua, N.K.; Khan, S.; Qazi, I.; Ali, S.; Saad Algarni, T. Electro-oxidation of ammonia at novel Ag2O−PrO2/γ-Al2O3 catalysts. Coatings 2021, 11, 257. [Google Scholar] [CrossRef]
  5. Saleem, M.; Irfan, M.; Tabassum, S.; Albaqami, M.D.; Javed, M.S.; Hussain, S.; Pervaiz, M.; Ahmad, I.; Ahmad, A.; Zuber, M. Experimental and theoretical study of highly porous lignocellulose assisted metal oxide photoelectrodes for dye-sensitized solar cells. Arab. J. Chem. 2021, 14, 102937. [Google Scholar] [CrossRef]
  6. Ansir, R.; Shah, S.M.; Ullah, N.; Hussain, M.N. Performance of pyrocatechol violet and carminic acid sensitized ZnO/CdS nanostructured photoactive materials for dye sensitized solar cell. Solid-State Electron. 2020, 172, 107886. [Google Scholar] [CrossRef]
  7. Jacobson, M.Z.; Delucchi, M.A. Providing all global energy with wind, water, and solar power, Part I: Technologies, energy resources, quantities and areas of infrastructure, and materials. Energy Policy 2011, 39, 1154–1169. [Google Scholar] [CrossRef]
  8. Irvine, S. Solar Cells and Photovoltaics. In Springer Handbook of Electronic and Photonic Materials; Springer: Cham, The Netherlands, 2017; p. 1. [Google Scholar]
  9. Gul, M.; Kotak, Y.; Muneer, T. Review on recent trend of solar photovoltaic technology. Energy Explor. Exploit. 2016, 34, 485–526. [Google Scholar] [CrossRef]
  10. Sharma, K.; Sharma, V.; Sharma, S. Dye-sensitized solar cells: Fundamentals and current status. Nanoscale Res. Lett. 2018, 13, 381. [Google Scholar] [CrossRef]
  11. Bose, S.; Soni, V.; Genwa, K. Recent advances and future prospects for dye sensitized solar cells: A review. Int. J. Sci. Res. Publ. 2015, 5, 1–8. [Google Scholar]
  12. Mátravölgyi, B.; Hergert, T.; Thurner, A.; Varga, B.; Sangiorgi, N.; Bendoni, R.; Zani, L.; Reginato, G.; Calamante, M.; Sinicropi, A.; et al. Synthesis and Investigation of Solar-Cell Photosensitizers Having a Fluorazone Backbone. Eur. J. Org. Chem. 2017, 2017, 1843–1854. [Google Scholar] [CrossRef]
  13. Yu, Q.; Pang, Y.; Jiang, Q. NiS submicron cubes with efficient electrocatalytic activity as the counter electrode of dye-sensitized solar cells. R. Soc. Open Sci. 2018, 5, 180186. [Google Scholar] [CrossRef] [PubMed]
  14. Carella, A.; Borbone, F.; Centore, R. Research progress on photosensitizers for DSSC. Front. Chem. 2018, 6, 481. [Google Scholar] [CrossRef] [PubMed]
  15. Mishra, A.; Fischer, M.K.; Bäuerle, P. Metal-free organic dyes for dye-sensitized solar cells: From structure: Property relationships to design rules. Angew. Chem. Int. Ed. 2009, 48, 2474–2499. [Google Scholar] [CrossRef] [PubMed]
  16. Momeni, M.M. Dye-sensitized solar cell and photocatalytic performance of nanocomposite photocatalyst prepared by electrochemical anodization. Bull. Mater. Sci. 2016, 39, 1389–1395. [Google Scholar] [CrossRef]
  17. Su’ait, M.S.; Rahman, M.Y.A.; Ahmad, A. Review on polymer electrolyte in dye-sensitized solar cells (DSSCs). Sol. Energy 2015, 115, 452–470. [Google Scholar] [CrossRef]
  18. Wang, H.; Hu, Y.H. Graphene as a counter electrode material for dye-sensitized solar cells. Energy Environ. Sci. 2012, 5, 8182–8188. [Google Scholar] [CrossRef]
  19. Labat, F.; Ciofini, I.; Hratchian, H.P.; Frisch, M.J.; Raghavachari, K.; Adamo, C. Insights into working principles of ruthenium polypyridyl dye-sensitized solar cells from first principles modeling. J. Phys. Chem. C 2011, 115, 4297–4306. [Google Scholar] [CrossRef]
  20. Corrao, R.; D’Anna, D.; Morini, M.; Pastore, L. DSSC-integrated Glassblocks for the construction of Sustainable Building Envelopes. In Advanced Materials Research; Trans Tech Publ: Zurich, Switzerland, 2014. [Google Scholar]
  21. Wu, C.-G.; Chao, C.-C.; Kuo, F.-T. Enhancement of the photo catalytic performance of TiO2 catalysts via transition metal modification. Catal. Today 2004, 97, 103–112. [Google Scholar] [CrossRef]
  22. Inturi, S.N.R.; Boningari, T.; Suidan, M.; Smirniotis, P.G. Visible-light-induced photodegradation of gas phase acetonitrile using aerosol-made transition metal (V, Cr, Fe, Co, Mn, Mo, Ni, Cu, Y, Ce, and Zr) doped TiO2. Appl. Catal. B Environ. 2014, 144, 333–342. [Google Scholar] [CrossRef]
  23. Peng, H.; Ying, J.; Zhang, J.; Zhang, X.; Peng, C.; Rao, C.; Liu, W.; Zhang, N.; Wang, X. La-doped Pt/TiO2 as an efficient catalyst for room temperature oxidation of low concentration HCHO. Chin. J. Catal. 2017, 38, 39–47. [Google Scholar] [CrossRef]
  24. Liu, H.; Yu, L.; Chen, W.; Li, Y. The Progress of TiO2 Nanocrystals Doped with Rare Earth Ions. J. Nanomater. 2012, 2012, 235879. [Google Scholar] [CrossRef]
  25. Chen, Y.; Wang, Y.; Li, W.; Yang, Q.; Hou, Q.; Wei, L.; Liu, L.; Huang, F.; Ju, M. Enhancement of photocatalytic performance with the use of noble-metal-decorated TiO2 nanocrystals as highly active catalysts for aerobic oxidation under visible-light irradiation. Appl. Catal. B Environ. 2017, 210, 352–367. [Google Scholar] [CrossRef]
  26. Momeni, M.M.; Akbarnia, M.; Ghayeb, Y. Preparation of S–W-codoped TiO2 nanotubes and effect of various hole scavengers on their photoelectrochemical activity: Alcohol series. Int. J. Hydrog. Energy 2020, 45, 33552–33562. [Google Scholar] [CrossRef]
  27. Wu, Y.; Zhang, J.; Xiao, L.; Chen, F. Properties of carbon and iron modified TiO2 photocatalyst synthesized at low temperature and photodegradation of acid orange 7 under visible light. Appl. Surf. Sci. 2010, 256, 4260–4268. [Google Scholar] [CrossRef]
  28. Sinhmar, A.; Setia, H.; Kumar, V.; Sobti, A.; Toor, A.P. Enhanced photocatalytic activity of nickel and nitrogen codoped TiO2 under sunlight. Environ. Technol. Innov. 2020, 18, 100658. [Google Scholar] [CrossRef]
  29. Lu, W.C.; Nguyen, H.D.; Wu, C.Y.; Chang, K.S.; Yoshimura, M. Modulation of physical and photocatalytic properties of (Cr, N) codoped TiO2 nanorods using soft solution processing. J. Appl. Phys. 2014, 115, 144305. [Google Scholar] [CrossRef]
  30. Liu, R.; Yang, F.; Xie, Y.; Yu, Y. Visible-light responsive boron and nitrogen codoped anatase TiO2 with exposed {0 0 1} facet: Calculation and experiment. Appl. Surf. Sci. 2019, 466, 568–577. [Google Scholar] [CrossRef]
  31. Jaiswal, R.; Bharambe, J.; Patel, N.; Dashora, A.; Kothari, D.C.; Miotello, A. Copper and Nitrogen co-doped TiO2 photocatalyst with enhanced optical absorption and catalytic activity. Appl. Catal. B Environ. 2015, 168, 333–341. [Google Scholar] [CrossRef]
  32. Zhang, S. Synergistic effects of C–Cr codoping in TiO2 and enhanced sonocatalytic activity under ultrasonic irradiation. Ultrason. Sonochem. 2012, 19, 767–771. [Google Scholar] [CrossRef]
  33. Patel, N.; Jaiswal, R.; Warang, T.; Scarduelli, G.; Dashora, A.; Ahuja, B.L.; Kothari, D.C.; Miotello, A. Efficient photocatalytic degradation of organic water pollutants using V–N-codoped TiO2 thin films. Appl. Catal. B Environ. 2014, 150, 74–81. [Google Scholar] [CrossRef]
  34. Trevisan, V.; Olivo, A.; Pinna, F.; Signoretto, M.; Vindigni, F.; Cerrato, G.; Bianchi, C.L. CN/TiO2 photocatalysts: Effect of co-doping on the catalytic performance under visible light. Appl. Catal. B Environ. 2014, 160, 152–160. [Google Scholar] [CrossRef]
  35. Cho, I.S.; Lee, C.H.; Feng, Y.; Logar, M.; Rao, P.M.; Cai, L.; Kim, D.R.; Sinclair, R.; Zheng, X. Codoping titanium dioxide nanowires with tungsten and carbon for enhanced photoelectrochemical performance. Nat. Commun. 2013, 4, 1723. [Google Scholar] [CrossRef]
  36. Yang, X.; Ma, F.; Li, K.; Guo, Y.; Hu, J.; Li, W.; Huo, M.; Guo, Y. Mixed phase titania nanocomposite codoped with metallic silver and vanadium oxide: New efficient photocatalyst for dye degradation. J. Hazard. Mater. 2010, 175, 429–438. [Google Scholar] [CrossRef]
  37. Devi, R.S.; Venckatesh, D.R.; Sivaraj, D.R. Synthesis of titanium dioxide nanoparticles by sol-gel technique. Int. J. Innov. Res. Sci. Eng. Technol. 2014, 3, 15206–15211. [Google Scholar] [CrossRef]
  38. Vijayalakshmi, K.; Sivaraj, D. Synergistic antibacterial activity of barium doped TiO2 nanoclusters synthesized by microwave processing. RSC Adv. 2016, 6, 9663–9671. [Google Scholar] [CrossRef]
  39. Dahiya, M.S.; Tomer, V.K.; Duhan, S. Metal–ferrite nanocomposites for targeted drug delivery. In Applications of Nanocomposite Materials in Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2018; pp. 737–760. [Google Scholar]
  40. Badawi, A.; Althobaiti, M.G.; Alharthi, S.S.; Al-Baradi, A.M. Tailoring the optical properties of CdO nanostructures via barium doping for optical windows applications. Phys. Lett. A 2021, 411, 127553. [Google Scholar] [CrossRef]
  41. Patel, S.; Gajbhiye, N.; Date, K.S. Ferromagnetism of Mn-doped TiO2 nanorods synthesized by hydrothermal method. J. Alloy. Compd. 2011, 509, S427–S430. [Google Scholar] [CrossRef]
  42. AlFaify, S.; Ganesh, V.; Haritha, L.; Shkir, M. An effect of La doping on physical properties of CdO films facilely casted by spin coater for optoelectronic applications. Phys. B: Condens. Matter 2019, 562, 135–140. [Google Scholar] [CrossRef]
  43. N’Konou, K.; Lare, Y.; Haris, M.; Baneto, M.; Amou, K.A.; Napo, K. Influence of barium doping on physical properties of zinc oxide thin films synthesized by SILAR deposition technique. Adv. Mater. 2014, 3, 63–67. [Google Scholar] [CrossRef]
  44. Cheng, G.; Xu, F.; Stadler, F.J.; Chen, R. A facile and general synthesis strategy to doped TiO2 nanoaggregates with a mesoporous structure and comparable property. RSC Adv. 2015, 5, 64293–64298. [Google Scholar] [CrossRef]
  45. Moss, T.S. The interpretation of the properties of indium antimonide. Proc. Phys. Soc. 1954, 67, 10. [Google Scholar] [CrossRef]
  46. Singh, M.; Goyal, M.; Devlal, K. Size and shape effects on the band gap of semiconductor compound nanomaterials. J. Taibah Univ. Sci. 2018, 12, 470–475. [Google Scholar] [CrossRef]
  47. Son, Y.; Park, M.; Son, Y.; Lee, J.S.; Jang, J.H.; Kim, Y.; Cho, J. Quantum confinement and its related effects on the critical size of GeO2 nanoparticles anodes for lithium batteries. Nano Lett. 2014, 14, 1005–1010. [Google Scholar] [CrossRef] [PubMed]
  48. Roth, A.; Webb, J.; Williams, D. Absorption edge shift in ZnO thin films at high carrier densities. Solid State Commun. 1981, 39, 1269–1271. [Google Scholar] [CrossRef]
  49. Velusamy, P.; Babu, R.R.; Ramamurthi, K.; Elangovan, E.; Viegas, J. Effect of La doping on the structural, optical and electrical properties of spray pyrolytically deposited CdO thin films. J. Alloy. Compd. 2017, 708, 804–812. [Google Scholar] [CrossRef]
  50. Chakraborty, P.; Datta, G.; Ghatak, K. The simple analysis of the Burstein–Moss shift in degenerate n-type semiconductors. Phys. Condens. Matter 2003, 339, 198–203. [Google Scholar] [CrossRef]
  51. Kumar, R. Structural and optical studies of Mn2+ substituted CdO nano-particles. Appl. Phys. A 2021, 127, 249. [Google Scholar]
  52. Kumar, K.C.; Rao, N.M.; Kaleemulla, S.; Rao, G.V. Structural, optical and magnetic properties of Sn doped ZnS nano powders prepared by solid state reaction. Phys. B Condens. Matter 2017, 522, 75–80. [Google Scholar] [CrossRef]
  53. Sangiorgi, N.; Aversa, L.; Tatti, R.; Verucchi, R.; Sanson, A. Spectrophotometric method for optical band gap and electronic transitions determination of semiconductor materials. Opt. Mater. 2017, 64, 18–25. [Google Scholar] [CrossRef]
  54. Reddy, K.M.; Manorama, S.V.; Reddy, A.R. Bandgap studies on anatase titanium dioxide nanoparticles. Mater. Chem. Phys. 2003, 78, 239–245. [Google Scholar] [CrossRef]
  55. Yoo, K.; Kim, J.Y.; Lee, J.A.; Kim, J.S.; Lee, D.K.; Kim, K.; Kim, J.Y.; Kim, B.; Kim, H.; Kim, W.M.; et al. Completely transparent conducting oxide-free and flexible dye-sensitized solar cells fabricated on plastic substrates. ACS Nano 2015, 9, 3760–3771. [Google Scholar] [CrossRef] [PubMed]
  56. Saeed, M.A.; Kang, H.C.; Yoo, K.; Asiam, F.K.; Lee, J.J.; Shim, J.W. Cosensitization of metal-based dyes for high-performance dye-sensitized photovoltaics under ambient lighting conditions. Dye. Pigment. 2021, 194, 109624. [Google Scholar] [CrossRef]
  57. Amarsingh Bhabu, K.; Kalpana Devi, A.; Theerthagiri, J.; Madhavan, J.; Balu, T.; Rajasekaran, T.R. Tungsten doped titanium dioxide as a photoanode for dye sensitized solar cells. J. Mater. Sci. Mater. Electron. 2017, 28, 3428–3439. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ba/TiO2, TiO2 photoanodes along with the reference JCPDS data 21-1272.
Figure 1. XRD patterns of Ba/TiO2, TiO2 photoanodes along with the reference JCPDS data 21-1272.
Applsci 12 09280 g001
Figure 2. UV-visible spectra of synthesized TiO2 and Ba/TiO2 photoanodes.
Figure 2. UV-visible spectra of synthesized TiO2 and Ba/TiO2 photoanodes.
Applsci 12 09280 g002
Figure 3. (a) Tauc plots of synthesized TiO2 and Ba/TiO2 photoanodes (b) Graphical representation of Equation (7), inset relates to Ba/TiO2.
Figure 3. (a) Tauc plots of synthesized TiO2 and Ba/TiO2 photoanodes (b) Graphical representation of Equation (7), inset relates to Ba/TiO2.
Applsci 12 09280 g003
Figure 4. UV-visible absorption spectra of synthesized TiO2 and Ba/TiO2 photoanodes with N3 dye molecule.
Figure 4. UV-visible absorption spectra of synthesized TiO2 and Ba/TiO2 photoanodes with N3 dye molecule.
Applsci 12 09280 g004
Figure 5. SEM analysis of synthesized (a) TiO2 and (b) Ba/TiO2, and EDS analysis of synthesized (c) TiO2 and (d) Ba/TiO2 nanopowders.
Figure 5. SEM analysis of synthesized (a) TiO2 and (b) Ba/TiO2, and EDS analysis of synthesized (c) TiO2 and (d) Ba/TiO2 nanopowders.
Applsci 12 09280 g005
Figure 6. EIS analysis of synthesized (a) TiO2, (b) Ba/TiO2 photoanode and along with the equivalent circuit of corresponding EIS.
Figure 6. EIS analysis of synthesized (a) TiO2, (b) Ba/TiO2 photoanode and along with the equivalent circuit of corresponding EIS.
Applsci 12 09280 g006
Figure 7. Photovoltaic response of synthesized TiO2 and Ba/TiO2 photoanodes.
Figure 7. Photovoltaic response of synthesized TiO2 and Ba/TiO2 photoanodes.
Applsci 12 09280 g007
Table 1. Structural parameters derived from XRD patterns of the synthesized TiO2 and Ba/TiO2 photoanode.
Table 1. Structural parameters derived from XRD patterns of the synthesized TiO2 and Ba/TiO2 photoanode.
Photoactive MaterialFWHM (2θ)d-Spacing (Å)Crystallite Size(D) (nm)Lattice Constant (Å)Unit Cell Volume V (Å)3Density (ρ) (g/cm3)Specific Surface Area (m2/g)
a = bC
TiO20.568 ± 0.0263.494253.778 ± 0.0059.480 ± 0.012135.313.961
Ba/TiO20.876 ± 0.0383.504163.781 ± 0.0029.470 ± 0.004135.383.995
Table 2. n coefficient determination and band gap.
Table 2. n coefficient determination and band gap.
SampleEg (eV)n Coefficient
TiO23.211.72 ± 0.003
Ba/TiO23.261.61 ± 0.003
Table 3. Photovoltaic parameters derived from current voltage characteristic curves of synthesized TiO2 and Ba/TiO2 photoanodes.
Table 3. Photovoltaic parameters derived from current voltage characteristic curves of synthesized TiO2 and Ba/TiO2 photoanodes.
SampleShort Circuit Current Density
Jsc (mA/cm2)
Open Circuit Voltage
Voc (mV)
Fill Factor
FF (%)
Efficiency (η) (%)Ref
TiO25.96761482.1Present work
Ba/TiO28.70740533.4Present work
TiO22.53650631.03[57]
W-TiO24.77730712.47[57]
Fe-TiO21.526800.4630.47[26]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmad, A.; Khan, S.; Khan, M.; Luque, R.; Jalalah, M.; Alsaiari, M.A. Microwave Assisted Preparation of Barium Doped Titania (Ba/TiO2) as Photoanode in Dye Sensitized Solar Cells. Appl. Sci. 2022, 12, 9280. https://doi.org/10.3390/app12189280

AMA Style

Ahmad A, Khan S, Khan M, Luque R, Jalalah M, Alsaiari MA. Microwave Assisted Preparation of Barium Doped Titania (Ba/TiO2) as Photoanode in Dye Sensitized Solar Cells. Applied Sciences. 2022; 12(18):9280. https://doi.org/10.3390/app12189280

Chicago/Turabian Style

Ahmad, Awais, Safia Khan, Mariam Khan, Rafael Luque, Mohammed Jalalah, and Mabkhoot A. Alsaiari. 2022. "Microwave Assisted Preparation of Barium Doped Titania (Ba/TiO2) as Photoanode in Dye Sensitized Solar Cells" Applied Sciences 12, no. 18: 9280. https://doi.org/10.3390/app12189280

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop