Next Article in Journal
Changes in Morphology Caused by Mass Transfer Phenomenon
Previous Article in Journal
Research Progress on the Sintering Techniques of Zirconia in Prosthetic Dentistry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Binary Oxide Ceramics (TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3) for Solar Cell Applications: A Comparative and Bibliometric Analysis

1
Scientific Department, Berdyansk State Pedagogical University, 69061 Zaporizhzhia, Ukraine
2
Library, Borys Grinchenko Kyiv Metropolitan University, 04053 Kyiv, Ukraine
3
Institute of Solid State Physics, University of Latvia, 8 Kengaraga, 1063 Riga, Latvia
*
Authors to whom correspondence should be addressed.
Ceramics 2025, 8(4), 119; https://doi.org/10.3390/ceramics8040119
Submission received: 1 August 2025 / Revised: 7 September 2025 / Accepted: 22 September 2025 / Published: 23 September 2025

Abstract

Binary oxide ceramics have emerged as key materials in solar energy research due to their versatility, chemical stability, and tunable electronic properties. This study presents a comparative analysis of seven prominent oxides (TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3), focusing on their functional roles in silicon, perovskite, dye-sensitized, and thin-film solar cells. A bibliometric analysis covering over 50,000 publications highlights TiO2 and ZnO as the most widely studied materials, serving as electron transport layers, antireflective coatings, and buffer layers. Al2O3 and SiO2 demonstrate highly specialized applications in surface passivation and interface engineering, while CeO2 offers UV-blocking capability and Fe2O3 shows potential as an absorber material in photoelectrochemical systems. WO3 is noted for its multifunctionality and suitability for scalable, high-rate processing. Together, these findings suggest that binary oxide ceramics are poised to transition from supporting roles to essential components of stable, efficient, and environmentally safer next-generation solar cells.

1. Introduction

Traditionally, solar energy has been closely associated with materials such as silicon [1,2,3], cadmium telluride [4,5,6], and copper indium gallium selenide (CIGS) [7,8,9]. Silicon, in particular, has dominated the photovoltaic device market due to its abundance, relatively low cost, and well-established manufacturing processes [10,11,12,13]. Silicon-based solar cells have been the cornerstone of solar energy production, offering high efficiency and long-term stability [14,15,16]. These materials are widely recognized for their ability to effectively convert sunlight into electricity, making them the standard choice for most commercial and residential solar energy systems [17,18,19,20].
Cadmium telluride (CdTe) is another key material in the field of solar energy, particularly in thin-film solar cells [21,22,23]. CdTe has gained popularity due to its high absorption coefficient and relatively low production cost, making it a competitive alternative to silicon for specific applications [24,25,26]. However, concerns about cadmium toxicity and the limited availability of tellurium have prompted the search for safer and more sustainable alternatives [27,28].
Copper indium gallium selenide (CIGS) represents another category of thin-film solar cells that has attracted attention due to its high efficiency and flexibility [29,30,31]. CIGS cells offer higher efficiencies than other thin-film technologies and can be applied to a variety of substrates, including flexible materials, opening up new opportunities for solar energy applications [32,33,34]. Despite these advantages, the complexity of the material composition and associated manufacturing challenges have limited their widespread adoption compared to silicon-based technologies [35,36].
As the solar energy sector continues to grow and evolve, there has been a significant push to diversify the range of materials used in solar cells [37,38,39,40,41,42]. Researchers are exploring novel materials that may offer higher efficiency, lower production costs, or better performance under specific environmental conditions. This expansion has led to the investigation of organic photovoltaic materials [43,44], perovskites [45,46], and quantum dots [47,48], among others.
In particular, perovskite solar cells are attracting significant interest due to their high efficiency and ease of fabrication [47,48]. These materials have rapidly progressed from laboratory experiments to near-commercialization, offering efficiencies that rival traditional silicon solar cells. However, challenges related to long-term stability and the presence of lead in many perovskite compositions remain serious issues that must be resolved before widespread adoption is possible [49,50].
Quantum dot solar cells represent another frontier in solar energy research, offering the potential for high efficiency through multiple exciton generation [51,52]. These nanoscale semiconductor particles can be engineered to possess specific optical properties, making them easily tunable for various applications [53,54]. However, quantum dot technologies face challenges such as stability and scalability that must be overcome.
Organic photovoltaics (OPV) [55,56] have emerged as a promising field due to their potential for low-cost production and mechanical flexibility, although they currently lag behind traditional materials in terms of efficiency and stability.
This expansion of material options reflects ongoing efforts in solar energy to enhance the efficiency, affordability, and versatility of solar cells. By moving beyond traditional materials, the industry seeks to develop next-generation solar technologies capable of meeting the rising global demand for renewable energy and addressing the limitations of current systems [57].
In recent years, the solar energy field has begun to explore the potential of oxide ceramics as alternative materials for photovoltaic and related applications [58]. Known for their durability, thermal stability, and diverse electrical properties, oxide ceramics offer a promising avenue for improving the performance and longevity of solar cells [59,60,61,62]. Unlike traditional semiconductors, oxide ceramics provide a unique combination of features, including high chemical resistance and the ability to operate under extreme environmental conditions, making them ideal candidates for use in advanced solar technologies [63,64,65,66].
The growing interest in oxide ceramics for solar applications stems from their ability to play multiple roles within solar cells, such as serving as transparent conducting oxides, photoanodes, or passivation layers [67,68,69,70]. These materials can enhance light absorption, increase charge carrier mobility, and reduce recombination losses, thereby potentially improving the overall efficiency of solar cells [71,72,73,74]. Furthermore, the versatility of oxide ceramics allows them to be integrated into various types of solar cells, including dye-sensitized solar cells [75,76], perovskite solar cells [77,78], and even novel technologies such as photoelectrochemical cells for hydrogen production [79,80].
Key binary oxide ceramics (Table 1), such as titanium dioxide (TiO2) [81,82,83,84] and zinc oxide (ZnO) [85,86], have already proven to be critical components in several solar technologies [87,88,89]. TiO2, for instance, is widely used as a photoanode in dye-sensitized solar cells due to its excellent photocatalytic properties and high stability [90,91]. ZnO, with its favorable electron transport characteristics, is frequently used as a transparent electrode or photoanode in various solar cell designs [92]. These materials, along with others such as aluminum oxide (Al2O3) [93,94,95], silicon dioxide (SiO2) [96,97], and cerium dioxide (CeO2) [98,99,100], are being extensively investigated for their potential to create more efficient and durable solar cells.
Fe2O3 (hematite) is considered a promising material due to its abundance, environmental friendliness, and ability to absorb visible light [101,102,103]. The bandgap of Fe2O3 makes it particularly attractive for photoelectrochemical solar cells, where it is often used as a photoanode [104,105]. Its high chemical stability and durability enable its use in harsh environments, such as acidic or alkaline media. WO3 (tungsten trioxide) is another promising material for solar technologies due to its electrochromic and photocatalytic properties [106,107,108]. The bandgap of WO3 allows it to absorb light in the near-UV spectrum, making it useful in hybrid solar cells [109,110]. WO3 is often employed as a photoanode in photoelectrochemical cells for hydrogen production [111], as well as an active material in multilayer anti-reflective coatings for solar panels [112].
Table 1. Physico-electrical parameters of oxide ceramic materials used in photovoltaic structures *.
Table 1. Physico-electrical parameters of oxide ceramic materials used in photovoltaic structures *.
Oxide MaterialBand Gap (Eg, eV)Conductivity TypeElectron Mobility (cm2·V−1·s−1)Dielectric Constant (εr)Electron Affinity (χ, eV)Ref
TiO2 (anatase/rutile)2.9–3.4 (direct/indirect)n-type (O-vacancy, donor-doped)~0.1–1 (up to 15 in crystals)25–1000 †3.9–4.3[113,114,115,116]
ZnO (wurtzite)3.1–3.4n-type (intrinsic/doped)10–3007–12 (up to ~25 for Co/Mn-doped)4.2–4.5[117,118,119]
Al2O3 (sapphire)8.5–9.5Insulator~ (≤10−9 S cm−1)6–121.0–2.6[120,121,122]
SiO2 (quartz, glass)8.0–9.2Insulator3.7–4.30.8–1.1[123,124,125]
CeO2 (ceria)2.8–3.5n-type (Ce3+, O-vacancies)10−4–1 (small-polaron hopping)16–353.3–3.7[126,127,128,129]
Fe2O3 (hematite)1.9–2.3n-type (poor σ)10−4–0.15–1204.3–5.0[130,131,132,133]
WO3 (monoclinic)2.4–3.2n-type (O-deficient)0.1–3010–105 ‡3.2–3.6[134,135,136]
† For heavily reduced or H-implanted rutile TiO2, εr > 1000 has been reported; the listed range of 25–1000 covers both typical and “giant” values (the typical range is 25–120). WO3 near the phase transition (~16 °C) shows peak εr ≈ 105; the operational range includes the most commonly used values of 10–105. * Note: All parameters are reported as generalized ranges because their values depend on synthesis route, crystallinity, defect concentration, doping level, measurement frequency, and other experimental conditions.
The integration of oxide ceramics into solar energy systems represents a significant shift toward the development of materials that not only improve efficiency but also offer enhanced durability [106,107]. As the demand for more reliable and cost-effective solar energy solutions continues to grow, the role of oxide ceramics is expected to increase, driving further innovation in the field.
While numerous studies have been conducted on specific oxide materials or their applications in solar technologies, no comprehensive bibliometric and comparative investigation has been conducted that evaluates various oxide ceramics across a wide range of solar applications. This gap in the literature highlights the need for a holistic analysis that not only identifies current trends but also provides insight into potential future directions for research and development.
The primary objective of this study is to conduct an in-depth bibliometric and comparative analysis of oxide ceramics used in solar energy, with a focus on understanding research trends, key materials, and their impact on the field. This study aims to identify which oxide ceramics are most prominent in solar energy research, how interest in these materials has evolved, and which materials hold the most significant potential for future development in this field.
The results of this study will be valuable for researchers, industry professionals, and policymakers, as they offer a clearer understanding of the current state of the field and its future direction. By mapping the research landscape, this study will help identify the most influential work, emerging areas of interest, and potential opportunities for innovation. Furthermore, it will serve as a resource for scholars seeking to build on existing knowledge, supporting more targeted and impactful research on the use of oxide ceramics in solar energy.

2. Methodology

This study employed a combined methodology that integrates bibliometric analysis, descriptive statistics, and thematic interpretation of results. This approach enables the identification of general research trends as well as an in-depth examination of each material’s specific role in solar energy applications.
The focus of this review was restricted to seven binary oxides—TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3—based on a set of clearly defined criteria. First, each of these oxides intersects with major photovoltaic architectures (c-Si, PSC, OPV, DSSC, PEC), thereby ensuring broad technological relevance. Second, they exhibit functional complementarity across key device roles, including electron and hole transport (TiO2, ZnO, WO3), passivation and interfacial control (Al2O3, SiO2), optical filtering and stability enhancement (CeO2), and light absorption (Fe2O3, WO3). Third, they are characterized by a sufficient bibliometric corpus, providing a robust publication base for comparative statistical analysis. Other oxides such as SnO2 (a well-established TCO/ETL in PSCs) and Fe3O4 (with known roles in PV) were acknowledged as relevant but were excluded to maintain methodological consistency and comparability across the selected set. Their omission avoids introducing stoichiometric and functional heterogeneity that could complicate bibliometric interpretation. Nevertheless, we note that SnO2 and Fe3O4 represent potential candidates for future “sensitivity appendices” or extended studies, where their trajectories may be compared against the seven selected benchmarks.

2.1. Bibliometric Analysis

Bibliometric analysis is a quantitative method used to assess the structure, dynamics, and trends of scientific research through the systematic examination of publications, citations, and metadata [137,138,139]. Traditionally employed in scientometrics and library sciences, this approach is gaining relevance in materials science, where it allows researchers to uncover patterns of technological development, identify leading contributors and emerging topics, and assess the maturity and interdisciplinarity of specific material systems [140,141,142,143].
In the context of solar energy research, bibliometric tools provide valuable insights into how specific materials, such as binary metal oxides, are integrated into device architectures, studied across disciplines, and adopted by different scientific communities [144,145,146,147]. By quantitatively assessing publication trends, citation impact, and keyword co-occurrence, bibliometric analysis complements experimental and theoretical approaches, offering a macroscopic view of knowledge production and research activity [148,149,150,151].
This study applies bibliometric analysis to map the research landscape surrounding seven key oxide ceramics (TiO2, ZnO, SiO2, Al2O3, CeO2, Fe2O3, and WO3) in solar energy applications. The goal is to identify thematic concentrations, leading authors and institutions, geographical distribution, and collaboration networks, as well as to compare material-specific trends in attention and utilization.

2.1.1. Database Selection and Search Strategy

The analysis was conducted using data retrieved from the Web of Science Core Collection, which was selected due to its broad coverage of peer-reviewed literature in science and technology [152,153,154]. To ensure a systematic and comprehensive approach, a detailed search strategy was developed to identify publications related to the studied oxide materials (Figure 1).
Search queries targeted each oxide individually by combining its chemical names with keywords associated with solar energy applications. For titanium dioxide (TiO2), the query used was TS = (“TiO2” OR “titanium dioxide”) AND TS = (“solar energy” OR “solar cell*” OR “photovoltaic devices” OR “solar technology” OR “photoanode” OR “solar panel”). Similar queries were constructed for zinc oxide (ZnO), silicon dioxide (SiO2), aluminum oxide (Al2O3), cerium dioxide (CeO2), hematite or iron oxide (Fe2O3), and tungsten trioxide (WO3). These queries combined the oxide names with terms such as “solar energy,” “solar cells,” “photovoltaic devices,” “solar technologies,” “photoanode,” and “solar panel” to capture the full scope of solar-related research for each material.
Filters were applied to refine the search results and ensure consistency across datasets. The temporal scope of the study was limited to publications available up to and including the year 2023 and included only articles written in English. All document types were considered, except corrections, editorials, retractions, and reprints, in order to maintain data integrity and relevance.
To quantitatively assess the impact of research related to oxide materials, citation counts were analyzed for each material. This analysis enabled the evaluation of not only the intensity of publication activity but also the scientific weight of research within each segment.

2.1.2. Data Analysis

The retrieved datasets were processed and analyzed using VOSviewer (1.6.20 version), a bibliometric analysis software tool designed for visualizing and exploring patterns in scientific literature [155,156,157]. For each oxide material, a range of bibliometric parameters was examined to provide a detailed overview of research activity. These included temporal publication trend analysis to identify chronological patterns, distribution of document types to evaluate the nature of contributions, and distribution across research areas to understand each material’s interdisciplinary orientation.
Further analysis identified leading journals, authors, and institutions to highlight the key contributors in the field. Geographic contributions were also assessed to explore global research activity, and collaboration networks were visualized to understand international and institutional partnerships. Finally, keyword co-occurrence analysis was conducted to identify thematic trends and research priorities for each oxide material.

2.2. Statistical Analysis

To analyze publication trends related to binary oxide ceramics, descriptive statistical methods were applied to summarize and compare research activity for each material. The computed statistical indicators include the median, mean, standard deviation, coefficient of variation (CV), maximum value, first quartile (Q1), third quartile (Q3), and interquartile range (IQR) (Table 2). These indicators were selected to reflect both the central tendency and variability of the data, enabling a comprehensive assessment of publication dynamics for each material.
All statistical calculations were based on the number of publications per material for the period 1974–2023. Data for each material were sorted in ascending order, and quartiles were calculated using interpolation for noninteger positions. The resulting metrics provide detailed insights into research trends, highlighting both the intensity of scientific activity and its temporal stability.

2.3. Functional Literature Analysis

In addition to bibliometric and statistical approaches, a structured literature review was conducted to compare the functional roles of selected binary oxide ceramics across different solar cell architectures. This stage of the study aimed to link material properties to their technological relevance in device engineering.
The analysis was based on a systematic review and synthesis of peer-reviewed publications addressing
  • The physical, chemical, and optoelectronic properties of TiO2, ZnO, SiO2, Al2O3, CeO2, Fe2O3, and WO3;
  • Their specific functions in crystalline silicon (c-Si), perovskite (PSC), dye-sensitized (DSSC), thin-film chalcogenide (CIGS, CdTe, CZTS), organic (OSC), and quantum dot (QD) solar cells;
  • Comparative advantages, limitations, and integration challenges of each oxide in these technologies.
The analysis was based on a systematic review and synthesis of peer-reviewed To structure the results, the following analytical tools were used:
  • A functional role matrix mapping the oxide materials to device architectures and layer functionalities (ETL, HTL, TCO, passivation, buffer, optical interlayer);
  • A synthesis of key advantages and limitations drawn from experimental studies and review articles;
  • Cross-verification of usage trends with bibliometric co-occurrence data (e.g., TiO2 + passivation; ZnO + buffer layer).
This triangulated approach enabled a comprehensive assessment of each oxide’s contribution to modern photovoltaic engineering, highlighting both mainstream uses and cutting-edge developments.

3. Results

3.1. Evolution of Scientific Interest in Oxide Ceramics for Solar Energy: Results of the Bibliometric Analysis

The bibliometric analysis of publications on oxide ceramics for solar energy applications provides a comprehensive overview of the research landscape. Table 3 summarizes the volume of literature identified for each of the selected binary oxide materials in the Web of Science Core Collection.
Overall, the dominance of TiO2 (22,898 publications) and ZnO (19,092) in the solar research arena is clearly evident. These are followed in popularity by SiO2 (4140), Al2O3 (3268), Fe2O3 (2633), and WO3 (2062). CeO2 received the least attention, with 491 publications.
Research on oxide ceramics in the context of solar energy has undergone significant evolution over recent decades, with notable fluctuations in interest across different materials (Figure 2). The earliest publications related to oxide ceramics appeared in the 1970s, focusing on materials such as TiO2 (1974) and WO3 (1976). However, until the late 1980s, the overall publication volume remained minimal, with only a few isolated studies. This indicates that the field had a very slow start.
The situation began to change in the early 1990s, as interest in oxide materials gradually intensified. This period saw a growth in the number of publications for TiO2, ZnO, and Fe2O3. For instance, TiO2 showed steady growth, reaching 49 publications by the end of the decade, signaling increasing recognition of this material. ZnO also garnered more attention, though at a slightly slower pace, while interest in other materials such as Al2O3 and WO3 remained low.
A true breakthrough for oxide ceramics occurred in the 2000s, when publication counts for TiO2 and ZnO rose sharply, reaching hundreds per year. TiO2 became the leading subject of research, exceeding 600 publications per year by 2010. ZnO ranked second, following a similar upward trend, though with slightly lower absolute numbers. Al2O3, Fe2O3, and WO3 also began to gain traction, albeit at more moderate rates, reaching dozens of publications annually. This reflects a broadening interest in various materials within this category.
The peak of oxide ceramics research occurred in the 2010s. TiO2 and ZnO reached their highest publication volumes, peaking in the middle of the decade (over 1600 for TiO2 and 1400 for ZnO). Other materials also hit their peaks during this period: SiO2 reached 299 publications in 2017, Al2O3 peaked at 262 in the same year, Fe2O3 at 249 in 2019, and WO3 at 182 in 2020. CeO2, despite its overall lower activity, also showed gradual growth, reaching its peak in 2024 with 63 publications. This surge of interest in different materials reflects a broad range of scientific challenges and experimental approaches associated with oxide ceramic research.
In the 2020s, the overall publication rate for most materials began to stabilize or slightly decline, likely due to saturation in certain research areas and a shift in focus toward new materials or concepts. An exception is CeO2, which continues to show growing interest, possibly due to its niche applications.
Key observations indicate that TiO2 and ZnO remain the primary subjects of research, while other materials such as Al2O3, Fe2O3, WO3, and CeO2 occupy more specialized niches.

3.2. Publication Trends and Statistical Analysis of Binary Oxide Ceramics

The statistical analysis of publications related to binary oxide ceramics revealed significant differences in research activity among the various materials, as well as substantial fluctuations in publication trends (Table 4). The results provide insight into the relative importance and developmental dynamics of each material in the research field.
TiO2 exhibits the highest research intensity, with an average of 520.41 publications and a maximum of 1665 articles in a single year.
SiO2 is characterized by lower average research indicators: a mean of 94.11 and a maximum of 299 publications. Its median (25.0) and interquartile range (IQR = 226.0) suggest that most years were marked by moderate research activity, without the surges observed for TiO2 and ZnO. The standard deviation (110.43) and coefficient of variation (117.33%) indicate relatively stable interest in the material, though without major peaks in research output.
Al2O3 showed similar trends to SiO2, with a mean of 93.37 and a maximum of 262 publications. The median (20.0) and IQR (207.0) suggest that most years had low research activity, interspersed with periodic spikes. The standard deviation (100.99) and coefficient of variation (108.16%) point to significant year-to-year variability, likely driven by developments in specific solar applications.
Fe2O3 has the lowest publication median (9.0) among the studied materials but demonstrates moderate overall activity. Its mean is 75.23, and the maximum number of publications is 271, indicating periods of increased attention. The IQR (144.5) and high coefficient of variation (131.36%) reflect uneven but occasionally intense research activity, possibly linked to specific technological innovations.
WO3 demonstrates moderate research intensity, with an average of 60.65 publications and a maximum of 192 articles. The median (16.0) and IQR (115.25) suggest a concentration of activity in the lower range, with only a few years yielding higher output. The coefficient of variation (118.36%) confirms considerable variability, highlighting that research on WO3 tends to be more focused but less consistent over time.
CeO2 has the lowest overall research activity among the materials: an average of 16.93 publications and a maximum of 63. The median (5.0) and IQR (31.0) indicate limited but gradually growing research interest in recent years. The standard deviation (19.42) and coefficient of variation (114.72%) point to a slow yet steady interest in CeO2, likely due to its niche applications.
The analysis shows that TiO2 and ZnO dominate the research landscape of binary oxide ceramics, with significantly higher publication metrics and variability compared to other materials. In contrast, SiO2, Al2O3, WO3, Fe2O3, and CeO2 exhibit more moderate research activity, often tied to specific technological breakthroughs. The high coefficients of variation across all materials emphasize the dynamic nature of research priorities in this field.

3.3. Interdisciplinary Distribution of Research on Oxide Ceramics

The Sankey diagram (Figure 3) illustrates the distribution of scientific interest in the studied materials (TiO2, ZnO, SiO2, Al2O3, CeO2, Fe2O3, and WO3) across four key research domains: Energy, Materials Science, Environmental Science, and Physics. The width of each connection corresponds to the number of publications associating a given material with the respective field. This visualization provides insight into the degree of interdisciplinarity of each material and outlines the profiles of their research involvement [158,159,160].
TiO2 and ZnO exhibit the broadest scientific activity, with strong representation in all four disciplines. Their leading role in energy and materials science stems from their wide range of functionalities—from charge transport to photocatalysis and structural stabilization in optoelectronics. In parallel, their presence in environmental and physics-related studies demonstrates their functional flexibility and adaptability to diverse technological challenges [161,162,163,164].
Fe2O3, on the other hand, displays a strong focus on energy research. Its research profile is primarily centered on photoelectrochemical water splitting and hydrogen generation [165,166,167,168], reflecting its specialization as a material for renewable energy production under harsh environmental conditions [169,170]. Its limited overlap with physics or environmental science highlights a narrow yet strategically important niche.
WO3 shows a more balanced distribution across energy, materials science, and environmental science. This structure reflects its multifunctionality: WO3 is investigated not only in the context of PEC devices [171,172], but also for applications such as electrochromic elements, sensors, and optical control coatings [173,174].
Al2O3 and SiO2 are primarily concentrated in materials science and physics. Their roles are mainly associated with passivation, dielectric separation, interface protection, and the formation of stable layers in complex solar cell architectures [174,175,176,177,178,179,180]. These applications align with their physicochemical characteristics—wide bandgap, high stability, and insulating nature [181,182,183].
CeO2, while having a smaller overall research volume, shows a relatively even contribution across all four domains. This indicates a growing interest in CeO2 as a promising material for niche photovoltaic, catalytic, and protective applications [184,185,186], particularly due to its UV absorption capabilities, redox activity, and high stability [187,188,189,190].
Thus, the diagram enables not only an assessment of the research scale for each oxide but also a contextual understanding of how scientific interest is shaped. The varying degrees of cross-sector coverage reflect differences in maturity, specialization, and transdisciplinary potential among the oxides. This has practical implications for developing research strategies and identifying priority directions for future innovations.

3.4. Comparative Analysis of the Most Cited Publications on Binary Oxide Ceramics

An analysis of citation metrics for the most influential publications on binary oxide ceramics highlights significant differences in the impact and research focus of different materials (Table 5).
TiO2 shows overwhelming dominance in citation counts, followed by ZnO and Fe2O3, while materials such as SiO2, Al2O3, WO3, and CeO2 contribute in a more specialized but still significant way. These patterns reflect both the scientific maturity of each material’s applications and their role within the broader solar energy research landscape.
TiO2 stands out as the most researched and cited material. The seminal work by O’Regan and Grätzel from 1991 [191], which introduced dye-sensitized solar cells, has garnered over 25,000 citations—far surpassing any other publication in this field. This paper laid the foundation for much of the subsequent TiO2 research and solidified its reputation as a cornerstone material in solar energy technologies. Other highly cited works on TiO2 examine its role in photoelectrochemical cells and perovskite solar cells, with citation counts ranging from 5000 to 18,000. These publications span several decades, indicating TiO2’s sustained influence across multiple generations of solar technology development.
ZnO also holds a strong position, with its most cited works receiving between 2000 and 11,000 citations. Key publications focus on its use in dye-sensitized solar cells and nanostructured applications, emphasizing its value as a versatile and cost-effective material. The overlap between ZnO and TiO2 in several studies underscores their complementary roles in similar technological areas. However, ZnO’s citation counts remain significantly lower than those of TiO2, reflecting its secondary, but still critical, role in the evolution of solar energy applications.
SiO2 shows more modest citation figures, with its most influential publications ranging from 728 to 1437 citations. These studies primarily explore SiO2’s supporting roles in solar cells, such as surface passivation and photocatalytic enhancement. While its contribution is less transformative than that of TiO2 or ZnO, SiO2 remains a fundamental component in optimizing the performance and stability of solar energy systems.
Al2O3 shows a similar trend, with top citation counts between 955 and 1886. Its impact centers on areas like charge recombination dynamics and hybrid cell stability. Though its niche applications limit its broader influence, Al2O3 has attracted significant attention in these specialized contexts.
Fe2O3 has a higher citation range (between 1400 and 2300), reflecting its role in photoelectrochemical water splitting and hydrogen generation. Publications on Fe2O3 emphasize its potential for renewable energy storage and production. WO3 shows a comparable citation profile, with its most cited works receiving between 1000 and 1600 citations. Research on WO3 highlights its stability and optical properties, particularly for applications like electrochromic films and PEC devices. Although more specialized, WO3 continues to be a valuable material in niche solar energy technologies.
CeO2 displays the lowest citation counts among the analyzed materials, with top publications ranging from 400 to 800 citations. Research on CeO2 is relatively recent and focuses on advanced topics such as defect engineering, photocatalysis, and thermochemical hydrogen production. Its growing relevance indicates emerging potential in specialized solar systems, even though its overall impact remains limited compared to other materials.
Table 6 offers insight not only into the general scientific impact of the materials but also into the evolution of research priorities over time. In the 1990s, TiO2’s dominance was undisputed: it was the material that triggered a breakthrough in solar technologies, as demonstrated by the pivotal publications of 1991 and 1993. In the 2000s, ZnO, SiO2, and Al2O3 joined the landscape, primarily as supporting or alternative components. Beginning in 2006, scientific interest expanded significantly, with key publications emerging on WO3 and Fe2O3, reflecting a growing focus on PEC technologies and photoelectrochemistry.
Simultaneously, the table illustrates the increasing research activity on CeO2 in the 2010s, marking its transition from a peripheral topic to one of active interest. Although its citation counts do not yet rival those of TiO2 or ZnO, the thematic focus of recent publications suggests strong potential in areas like hydrogen production and defect engineering. Publications from 2017 to 2018 are particularly significant indicators of which materials may form the next wave in oxide ceramic research for solar energy.

3.5. Global Trends and International Collaboration in Research on Binary Oxides for Solar Energy Applications

3.5.1. Titanium Dioxide

Bibliometric analysis (Figure 4) shows that research on TiO2 has steadily increased over the past two decades, driven by its applications in photovoltaic systems, photocatalysis, and environmental remediation. China leads the research landscape with the highest number of publications, followed by India, South Korea, and Japan. The United States is also among the leading contributors, although its node overlaps with China in the visualization, making it less visible in the map. Collaboration networks reveal strong global connections, with Germany, Italy, Spain, and France forming influential hubs in Europe. This reflects the worldwide interest in TiO2 as a key material in energy and environmental technologies.
Keyword analysis reveals major research themes such as “nanostructures,” “photoanodes,” “efficiency,” and “thin films.” The role of TiO2 in dye-sensitized solar cells (DSSC) is well established, with studies highlighting improvements in light harvesting efficiency, stability, and charge separation. Research frequently focuses on nanostructuring approaches, including nanotubes, nanorods, and mesoporous films, to increase surface area and optimize electron transport.
Additionally, TiO2 doping with elements such as nitrogen or metals has been extensively studied to extend its light absorption into the visible spectrum, making it suitable for broader solar energy applications.
Hydrogen production and water splitting represent another major research direction for TiO2, where the use of co-catalysts and heterojunction architectures enhances its integration into photoelectrochemical systems. Environmental applications, including pollutant degradation and air purification, further emphasize the versatility of TiO2, making it a sustained subject of interest across multiple disciplines.

3.5.2. Zinc Oxide

Bibliometric analysis (Figure 5) indicates a steady increase in research output on ZnO, with China, the United States, and India leading global efforts. Germany, France, and South Korea also play important roles, and the collaboration networks show strong partnerships between Asian and European countries. The growing importance of ZnO reflects its versatility in energy conversion, environmental remediation, and advanced electronics.
The keyword network is dominated by terms such as “thin films,” “nanostructures,” “dye-sensitized solar cells,” and “recombination,” highlighting ZnO’s utility in photovoltaic and catalytic systems. Studies emphasize its high electron mobility, tunable bandgap, and cost-effective synthesis. In DSSCs and hybrid perovskite solar cells, ZnO serves as a transparent conducting oxide (TCO) or electron transport layer, and advances in nanostructuring methods, such as electrospinning and hydrothermal synthesis, have improved its performance. ZnO nanowires, nanoparticles, and quantum dots are actively investigated for their enhanced surface area and light-harvesting capabilities.
ZnO’s role in photocatalysis has also drawn considerable attention. Applications include pollutant degradation, hydrogen production, and water purification. Recent studies combine ZnO with TiO2 and other materials to form heterojunctions that improve charge separation and catalytic efficiency. These innovations underscore ZnO’s critical role in sustainable technologies.

3.5.3. Silicon Dioxide

Silicon dioxide (SiO2), widely recognized for its dielectric and insulating properties, plays a crucial role in enhancing the performance and stability of solar cells. Bibliometric data (Figure 6) identify China and the United States as dominant contributors to SiO2 research, with strong collaboration networks in Germany, the Netherlands, and South Korea. Patterns of cooperation highlight the cohesion of European research efforts, anchored in partnerships between Germany and neighboring countries such as France and the Netherlands.
Keyword analysis reveals core themes such as “surface passivation,” “coating,” and “antireflective layers,” reflecting the importance of SiO2 in minimizing recombination losses and protecting solar cell components. Its applications in thermal management and optical coatings further underscore its versatility. Recent advances include nanoporous SiO2 structures for light trapping and improved heat dissipation. Sol–gel processing and plasma-enhanced chemical vapor deposition (PECVD) have emerged as prominent methods for producing uniform and durable SiO2 films.
As an integral part of tandem and thin-film solar cells, SiO2 continues to serve as a foundational material in photovoltaic research.

3.5.4. Aluminum Oxide

Aluminum oxide (Al2O3) has attracted significant attention due to its application in surface passivation, particularly in silicon-based solar cells. Bibliometric data (Figure 7) show a steady increase in research output, with China, the United States, and Germany contributing the largest share globally. The collaboration network highlights strong connections among these countries and other regions, including South Korea, India, and Australia.
Key terms such as “atomic layer deposition,” “surface passivation,” and “crystalline silicon” dominate the keyword network, illustrating Al2O3’s role in enhancing the stability and efficiency of photovoltaic systems. ALD methods are frequently used to deposit Al2O3 layers with excellent dielectric properties and conformal coverage. Recent developments include the integration of Al2O3 into perovskite solar cells and investigations of its potential as a barrier layer under harsh environmental conditions.

3.5.5. Cerium Dioxide

Cerium dioxide (CeO2) is an emerging material for solar and environmental applications due to its high oxygen storage capacity and redox properties. Bibliometric analysis (Figure 8) reveals that China, India, and the United States are the leading contributors, while collaboration across Europe and the Middle East is expanding. The cooperation network reflects a growing partnership between academic institutions and industrial stakeholders.
The analysis is dominated by keywords such as “nanoparticles,” “photocatalysis,” and “hydrogen production.” CeO2’s catalytic activity makes it ideal for solar-to-hydrogen conversion and pollutant degradation. Research focuses on the synthesis of CeO2-based nanocomposites to improve charge transport and catalytic efficiency. Hybrid systems combining CeO2 with TiO2 or ZnO have demonstrated enhanced performance in photocatalytic and thermochemical processes. CeO2’s thermal stability and optical properties further expand its potential for solar devices and energy storage systems.

3.5.6. Iron Oxide

Iron oxide (Fe2O3), particularly in its hematite form, has been widely studied for applications in photoelectrochemical water splitting and energy storage. Bibliometric data (Figure 9) identify China, the United States, and Germany as leading contributors, with broad international collaboration networks across Europe and Asia. Hematite’s abundance and stability make it an attractive material for renewable energy technologies, despite challenges such as its narrow bandgap and low conductivity.
Keywords such as “photoanodes,” “water splitting,” and “hydrogen production” emphasize Fe2O3’s central role in photoelectrochemical systems. Research focuses on overcoming material limitations through doping, nanostructuring, and the use of co-catalysts. Innovations include heterojunctions and hematite-based tandem systems that improve charge separation and photocurrent efficiency. Additionally, surface passivation strategies and hybrid designs integrating Fe2O3 with TiO2 or WO3 have demonstrated enhanced performance, solidifying its place at the center of renewable energy research.

3.5.7. Tungsten Trioxide

Research on tungsten trioxide (WO3) is driven by its applications in photoelectrochemical systems, smart windows, and environmental technologies. Bibliometric analysis (Figure 10) highlights China, the United States, and South Korea as key contributors, with strong collaboration networks in both Europe and Asia. Germany, Japan, and the United Kingdom are also notable players in advancing WO3-based technologies.
The analysis is dominated by keywords such as “photoanodes,” “thin films,” and “solar cells,” reflecting WO3’s versatility in energy and catalytic applications. Studies focus on optimizing WO3 properties through nanostructuring, doping, and integration with other materials. Recent advances include WO3-based heterojunctions for improved charge separation and the use of WO3 thin films in electrochromic devices. Its adaptability and multifunctionality position WO3 as an important material in addressing global energy and environmental challenges.

4. Comparative Analysis of Oxide Ceramics in Different Types of Solar Cells

The results of the bibliometric analysis demonstrate that binary oxide ceramic materials have diverse applications in solar energy, materials science, and photoelectrochemical technologies. The relevance of each material is defined by its physicochemical properties, its ability to enhance the efficiency of solar devices, and its potential for integration into current and future energy systems.
These oxides serve different roles, such as electron transport layers, hole-blocking layers, surface passivation coatings, or antireflective coatings, across various solar cell technologies, as outlined below. Key properties such as bandgap (electronic structure), optical transparency, chemical stability, and charge transport characteristics make them well suited for specific applications [223,224,225,226].
In the following sections, we examine how each oxide is utilized in the main types of solar cells (crystalline silicon solar cells, perovskite solar cells, dye-sensitized solar cells, thin-film chalcogenide cells, quantum dot cells, and emerging organic solar cells) (Figure 11), and explain why their intrinsic properties are advantageous in these systems.

4.1. Silicon-Based Solar Cells

Silicon-based solar cells, both monocrystalline [226,227,228] and polycrystalline [229,230], remain the predominant type of solar energy conversion devices due to their high efficiency, long-term stability, and well-established industrial manufacturing processes [231,232]. However, further performance enhancement of silicon photovoltaic devices requires minimizing losses associated with surface recombination of charge carriers and optical losses caused by light reflection [233]. Oxide materials play an important role in addressing these challenges.
One of the most effective and widely used oxide materials for silicon surface passivation is Al2O3 [234]. Its application in silicon solar cells is attributed to its excellent passivation properties and high thermal and chemical stability [235]. The deposition of thin Al2O3 layers enables the formation of a high-quality interface with silicon [236]. The primary advantage of this material lies in its ability to effectively neutralize surface defects, particularly dangling bonds on silicon atoms [237]. In addition to chemical passivation, Al2O3 exhibits a high density of fixed negative charges [238]. These charges promote field-effect passivation by repelling electrons from the interface, significantly reducing surface recombination on p-type silicon surfaces. As a result, Al2O3 layers substantially enhance the open-circuit voltage (V_OC) of silicon solar cells by reducing the defect density at the interface [236,239]. Another important characteristic of Al2O3 is its high thermal and chemical stability, which allows it to withstand high-temperature fabrication processes such as contact annealing [240].
Silicon dioxide is one of the most well-established and widely studied materials in silicon photovoltaics [241]. It is mainly used as a surface passivation layer and as a component of antireflective coatings [242]. The main advantage of SiO2 is its exceptional ability to form a very low density of surface defect states due to the formation of strong and stable bonds with silicon [243]. This is particularly important for ensuring low surface recombination velocities and, consequently, for increasing open-circuit voltage and overall solar cell efficiency. However, unlike Al2O3, SiO2 has a neutral or slightly positive fixed charge, making it less effective for field-effect passivation of p-type silicon surfaces [244,245,246]. Therefore, SiO2 is often combined with other oxide or nitride layers in passivated contact structures, such as TOPCon [247,248]. In this architecture, an ultrathin tunneling SiO2 layer is paired with a highly doped polysilicon layer, enabling efficient charge collection while maintaining excellent surface passivation [249].
In addition to its passivation function, SiO2 is widely used in optical coatings due to its excellent antireflective properties [250]. Its low refractive index allows it to effectively reduce light losses due to reflection at the surface of silicon cells [251,252]. For example, dual-layer antireflective coatings based on SiO2 combined with high-index oxides (such as TiO2) can reduce broadband reflection and significantly enhance solar cell efficiency [253].
Titanium dioxide is well known for its unique properties that make it an attractive material for photovoltaic applications, particularly in silicon solar cells as an antireflective coating or interfacial layer. The choice of TiO2 for this role is due to its wide bandgap, which ensures high transparency in the visible spectrum [254]. This makes TiO2 ideal for reducing optical losses at the surface of silicon solar cells [255]. Thanks to its high refractive index, TiO2 is often used in dual-layer antireflective coatings along with SiO2 [256]. Such combined layers significantly reduce reflection losses, positively impacting the overall efficiency of silicon photovoltaic devices.
In addition to its optical advantages, TiO2 can also passivate silicon surfaces, although less effectively than Al2O3 or SiO2 [257]. Thin TiO2 layers are used to reduce surface recombination of charge carriers, particularly on n-type silicon surfaces or on the rear side of solar cells, contributing to higher open-circuit voltage [258,259]. Due to its chemical stability and low cost, TiO2 is a favorable material in manufacturing settings [260]. However, its semiconducting nature (n-type) and moderate conductivity must be taken into account, as they may lead to unintended shunting, especially if film thickness or quality is not properly controlled [261]. It is also important to note that achieving optimal performance of TiO2 layers typically requires thermal treatment, which imposes certain limitations on their use with flexible substrates [262,263].
ZnO is one of the primary candidates for the role of a transparent conducting oxide (TCO) in silicon heterojunctions and thin-film solar cells [264]. Due to its high optical transparency, resulting from its wide bandgap, and its ability to be doped n-type (e.g., with aluminum), ZnO offers excellent properties as a transparent front electrode [265]. Aluminum-doped ZnO (AZO) layers have been successfully used as a lower-cost and more accessible alternative to the more expensive and less abundant indium tin oxide (ITO), demonstrating comparable electrical conductivity and transparency at reduced cost [266].
In addition to its role as a transparent electrode, undoped or lightly doped ZnO can also serve as a buffer layer, minimizing damage during deposition of more conductive oxide layers or facilitating energy level alignment at interfaces [267,268]. However, it is important to note that in conventional silicon solar cells with diffused junctions, the use of ZnO and TiO2 is less common, as standard technologies typically employ silicon nitride (SiNx) [269] and screen-printed contact methods [270]. Meanwhile, in advanced heterostructure-based silicon devices, especially thin-film or specialized technologies, these materials demonstrate significant potential for further development.
Despite its many advantages, ZnO has certain limitations. Its surface can react with certain materials or be sensitive to acidic environments, which affects its stability and, consequently, the durability of solar cells [271,272]. Therefore, appropriate processing conditions and additional surface passivation become critically important for the effective use of ZnO in photovoltaic devices.
CeO2 is a relatively new and experimental material in silicon photovoltaics, but it has already shown potential for improving solar cell performance. Interest in CeO2 stems from its unique combination of physical and chemical properties, particularly its wide bandgap, which provides high optical transparency across most of the solar spectrum [273]. Owing to these characteristics, CeO2 can effectively serve as a window layer in silicon heterojunctions [274], similar to materials such as amorphous silicon carbide (a-SiC) [275] or indium tin oxide (ITO) [276] in heterojunction with intrinsic Thin-layer (HIT) cell structures.
Theoretical studies and simulations confirm the promise of CeO2/n-Si heterojunctions. In such heterostructures, CeO2 acts not only as a transparent window layer but may also provide additional surface passivation by forming a high-quality interface with silicon [277]. Furthermore, a notable advantage of CeO2 is its ability to absorb ultraviolet radiation, allowing it to function as a protective layer for silicon structures, mitigating the detrimental effects of UV exposure on the stability of solar cells [278,279]. Despite these promising features, experimental data on the practical use of CeO2 in commercial devices remain limited. Further research is necessary to optimize deposition methods and improve the film quality of this oxide.
Fe2O3 and WO3 oxides are rarely discussed in the context of silicon solar cell applications, as there is currently no compelling evidence of their effectiveness for passivation or as transparent electrodes in Si-PV.
Thus, each of the oxide materials discussed has unique advantages that make it attractive for specific functions in silicon solar cells (Table 7). Al2O3 and SiO2 are traditionally the most effective materials for silicon passivation, due to their low interface defect densities and chemical inertness when in contact with silicon. Al2O3 offers the additional benefit of efficient field-effect passivation due to its negative fixed charge. For optical optimization tasks, coatings based on SiO2 and TiO2 remain leading choices due to their ability to reduce light reflection losses. ZnO, meanwhile, is a promising transparent conductive oxide because of its affordability, high electrical conductivity when doped, and suitability for use in thin-film silicon cells. At the same time, both TiO2 and ZnO may require special technological measures to overcome potential drawbacks related to photocatalytic activity or chemical instability. CeO2 shows promise as an innovative material for surface passivation and UV protection, though it still requires further investigation.

4.2. Perovskite Solar Cells (PSC)

Hybrid organic–inorganic perovskite solar cells (PSCs) have attracted significant scientific attention in recent years due to their high theoretical solar energy conversion efficiency and the potential cost-effectiveness of their production [280]. One of the crucial aspects of improving the efficiency and stability of such devices is the selection of optimal charge-selective layers, among which oxide materials play an essential role [281]. Typically, metal oxide thin films are employed in PSC architectures to facilitate selective charge transport, reduce carrier recombination, and ensure long-term device stability [282].
TiO2 is one of the most widely used and extensively studied materials for electron transport layers (ETLs) in PSCs [283], particularly in the conventional n–i–p configuration [284]. A key reason for the widespread use of TiO2 is its favorable band alignment: its conduction band matches well with that of the perovskite [285,286]. This enables efficient electron extraction from the active layer while blocking hole transport in the reverse direction, thereby reducing recombination. With its wide bandgap, TiO2 remains transparent to most of the solar spectrum, minimizing photon absorption losses [287].
PSC devices using TiO2-based mesoporous ETLs currently demonstrate power conversion efficiencies exceeding 20%, highlighting the exceptional suitability of this material [288]. TiO2 is also favored for its high chemical stability under prolonged illumination and its relatively low manufacturing cost. Despite these advantages, TiO2 also has several drawbacks. Notably, it is well known for its photocatalytic activity under UV light, which can degrade both the perovskite and adjacent organic layers in the device [289]. This creates a need for additional UV-blocking or protective layers, such as ultrathin Al2O3 or CeO2 coatings. Another significant technological limitation is the high-temperature treatment required to crystallize TiO2 films and achieve high device performance [290]. This requirement complicates its use on flexible polymer substrates, prompting efforts to develop low-temperature deposition methods, such as solution processing, chemical bath deposition, or atomic layer deposition (ALD). Despite these challenges, TiO2 remains the benchmark ETL material in PSC design due to its well-established processing protocols, stable and high performance, and ongoing improvements through fine-tuned layer engineering.
Among alternative ETL materials for PSCs, ZnO has drawn considerable attention [291]. Unlike traditional TiO2, ZnO can be processed at low temperatures, including via solution-based methods [292]. This makes ZnO particularly attractive for use on flexible and polymer substrates, where high-temperature processing (as required for TiO2) is not feasible [293]. For instance, sol–gel ZnO films can be formed at temperatures below 150 °C, greatly simplifying fabrication and reducing production costs [294].
Additionally, ZnO exhibits superior charge transport speed due to its higher electron mobility compared to TiO2. This enhances the extraction of electrons from the perovskite active layer, potentially increasing short-circuit current density (J_SC) and fill factor (FF) [295]. The energy level alignment between ZnO and perovskite materials is also favorable—ZnO has a bandgap of approximately 3.3 eV, offering high transparency in the visible spectrum [296]. Combined with its strong electron affinity, ZnO is considered a promising candidate for ETL applications [297].
Despite its many advantages, the application of ZnO in PSCs faces significant challenges, particularly due to its chemical reactivity toward perovskite materials. It has been reported that perovskite layers deposited directly onto ZnO surfaces undergo rapid degradation [293]. This is caused by chemical interactions between the ZnO surface and the organic cations in the perovskite, resulting in the deprotonation of methylammonium cations and accelerated degradation of the perovskite layer. These processes substantially reduce the stability and longevity of devices with ZnO ETLs, despite their initially high performance.
To overcome this issue, researchers are actively developing various surface modification strategies for ZnO, including the introduction of interfacial protective layers. For example, the use of bilayer structures in which ZnO is coated with a compact TiO2 layer has proven effective [298], as have coatings based on self-assembled monolayers of organic molecules or fullerene derivatives [299]. These solutions significantly enhance device stability while maintaining high initial efficiency in PSCs employing ZnO-based ETLs. Nanostructured forms of ZnO (such as nanorods or nanoparticles) further improve electron collection by increasing the effective interfacial area and providing direct pathways for charge transport [300].
Among other oxide materials that are gaining increasing prominence in perovskite photovoltaics, tin oxide (SnO2) has emerged as especially important. Although SnO2 was not part of the original list of analyzed materials, its role as an efficient ETL in PSCs warrants special attention due to its excellent performance and widespread use in modern device architectures [301]. Thanks to its combination of high stability, wide bandgap (~3.6 eV), and suitability for low-temperature processing, SnO2 has emerged as a leading ETL candidate alongside TiO2 and ZnO. Studies have shown that SnO2 provides more stable performance than TiO2 and demonstrates lower hysteresis behavior, which is crucial for the long-term operation of PSCs [302,303,304].
Aluminum oxide (Al2O3), in turn, occupies a unique niche among materials used in PSCs due to its insulating and passivating properties. In the early stages of solid-state PSC development, it was discovered that replacing the conductive mesoporous TiO2 scaffold with insulating Al2O3 did not lead to a complete drop in performance; in fact, devices with efficiencies of about 10–11% were still achieved [305]. This was possible because perovskite was able to infiltrate the mesoporous Al2O3 structure and directly transport electrons to the contact, eliminating the need for a conductive scaffold. Later research showed that mesoporous frameworks made of Al2O3 or other inert oxides could even increase the open-circuit voltage (V_OC) compared to TiO2-based devices, due to reduced surface electron recombination [306].
Al2O3 is characterized by high chemical inertness toward perovskites, preventing degradation of the active layer and significantly improving device stability. This property is effectively utilized in PSC structures with carbon electrodes, where a triple mesoporous structure is employed: a bottom TiO2 layer for efficient electron contact, a middle Al2O3 layer as an inert insulating spacer, and a top carbon layer serving as the electrode [307]. This configuration allows Al2O3 to efficiently isolate the perovskite layer from the carbon contact, reducing recombination and improving device longevity.
Another promising ETL material in PSCs under recent investigation is cerium oxide (CeO2). Its high conduction band level aligns well with the energy levels of typical perovskite materials, enabling efficient electron extraction from the active layer. In addition, CeO2 offers several unique advantages, particularly its ability to absorb ultraviolet (UV) radiation [308]. This property allows cerium oxide to serve as a protective UV-blocking layer, preventing degradation of organic–inorganic perovskites under UV exposure. As a result, the incorporation of CeO2 layers significantly enhances the long-term stability of PSCs under continuous illumination [309].
An additional advantage of CeO2 is the presence of oxygen vacancies, which not only facilitate electron transport but also enhance its chemical stability against oxygen and moisture. However, this material is still under active laboratory investigation. One of the main barriers to its widespread adoption is the difficulty of producing high-quality CeO2 thin films without high-temperature annealing. Recent studies report that optimizing deposition processes, particularly solution-based CeO2 film formation without post-deposition thermal treatment, can lead to improved device efficiency, demonstrating this material’s significant potential [310]. Nevertheless, CeO2 remains less technologically mature than conventional oxides such as TiO2 and SnO2 and requires further research before broad commercial deployment.
Another unconventional material gaining attention in perovskite photovoltaics is iron(III) oxide, commonly known as hematite (α-Fe2O3). Its appeal lies in its abundance, non-toxicity, and exceptional stability under ultraviolet and visible light exposure [311]. Hematite’s availability and low cost make it a promising alternative for use as a compact ETL in PSCs [312]. Researchers have achieved power conversion efficiencies of around 13% in devices with compact Fe2O3 layers through precise control over the fabrication process, notably by forming dense, defect-free films using solution crystallization techniques [313].
At the same time, hematite suffers from several significant drawbacks that limit its widespread use. Its main limitation is low electron mobility and a high density of trap states, which lead to severe carrier recombination [314]. Additionally, due to its relatively narrow bandgap, hematite absorbs part of the visible spectrum, reducing the overall efficiency of devices where the ETL should be transparent [315]. Currently, the use of hematite remains confined mainly to laboratory settings, where ongoing efforts are focused on improving its electronic properties through doping and nanostructuring to make it more competitive with traditional oxides such as TiO2 or SnO2.
WO3 is most commonly used as a hole transport layer (HTL), although some studies also demonstrate its applicability as an ETL in specific configurations [316]. This oxide has a wide bandgap, making it transparent across most of the solar spectrum [317]. However, the most valuable property of WO3 is its tunable work function, which varies depending on stoichiometry. In its sub-stoichiometric form (WOx), it has a high work function, making it ideal for use as an inorganic HTL, especially in inverted (p–i–n) perovskite architectures [318]. Besides serving as an HTL, WO3 is also actively investigated as an ETL material; for example, mesoporous WO3 is used in PSCs as an additional layer to enhance electron transport [319].
Al2O3 and SiO2 have limited applications in PSCs, mainly serving auxiliary functions, insulating and passivating (Al2O3) or anti-reflective (SiO2), and rarely acting as primary charge-selective layers. Their inclusion in device architecture is often aimed at improving stability and optical performance. An inert mesoporous scaffold or ultrathin Al2O3 spacer helps passivate interfacial defect states at the perovskite/ETL junction, reducing recombination rates and thereby increasing V_OC. However, as Al2O3 is non-conductive, its thickness must be strictly controlled to avoid adding series resistance. SiO2, on the other hand, is primarily used as an optical or insulating interlayer: it reduces light reflection, stabilizes the active layer morphology, and acts as a barrier to interfacial ion diffusion. Both oxides function indirectly by improving the durability and electrical performance of the cell but are not involved in selective charge transport, and therefore are excluded from the comparative table of transport oxides.
In summary, hole transport in PSCs is generally facilitated by p-type oxides with high work functions. WO3, NiO, and MoO3 (the latter two not included in this analysis) are typical representatives of such materials. The advantage of using oxide materials over organic HTLs lies in their significantly higher resistance to ultraviolet light, heat, and moisture, thereby greatly extending device lifetime. Notably, CeO2 and WO3 can also absorb UV light, protecting the perovskite layer from degradation.
However, some oxides such as TiO2 and ZnO exhibit photocatalytic activity, which, under UV exposure, can lead to degradation of the perovskite and adjacent organic layers. To address this, specific approaches are employed—either by adding protective interlayers (such as CeO2) or incorporating luminescent additives that convert UV radiation into visible light.
Therefore, the correct selection and combination of oxide materials, considering their specific properties, enables synergistic effects that enhance both efficiency and stability of perovskite solar cells. The key characteristics and application roles of these oxides in the context of PSCs are summarized in the comparative table below (Table 8).

4.3. Dye-Sensitized Solar Cells (DSSC)

In dye-sensitized solar cells, the mesoporous oxide layer acts as a photoanode: it supports light-sensitive dyes and transports electrons to the transparent electrode [320,321]. The most common material in this role is TiO2 [322]. Its popularity is due to its wide bandgap, chemical inertness, non-toxicity, low cost, and ability to form highly porous nanostructures that enable efficient light harvesting [323]. The main drawback of TiO2 is the slow electron transport and the risk of recombination with oxidized electrolyte species; however, this can be mitigated through surface modification, core–shell structures, and other engineering strategies [324,325,326].
An alternative is ZnO, which has a similar bandgap and favorable energy alignment, but features higher electron mobility [327]. Due to the ease of forming nanostructures such as nanorods, ZnO provides direct pathways for electrons and can reduce recombination [328,329]. Moreover, ZnO can be deposited at low temperatures, making it suitable for flexible photovoltaic devices [330,331]. However, common dyes, especially those based on ruthenium, may interact with its surface, leading to dissolution or defect formation, thus limiting efficiency [332,333]. This issue can be addressed via interface engineering and the use of alternative dyes [334,335].
Other semiconductor oxides, such as WO3, SnO2, and Fe2O3, are also under investigation as photoanodes [336,337,338]. WO3, when combined with TiO2, can enhance UV sensitivity [339]. SnO2 is notable for its high electron mobility and its ability to increase open-circuit voltage due to a deeper conduction band. However, it requires blocking layers to counteract recombination with the electrolyte caused by its high mobility and deep conduction band [340]. Hematite absorbs visible light, but suffers from an extremely short hole diffusion length and high recombination, which limits its industrial applicability in DSSCs [341].
Inert oxides such as Al2O3 and SiO2, although not conductive, play important auxiliary roles. Ultrathin Al2O3 layers deposited on TiO2 surfaces can passivate defect states and reduce recombination, thereby increasing the open-circuit voltage [342]. SiO2, meanwhile, is used as a scattering additive in the anode or as a barrier layer that prolongs the photon path and improves light absorption [343].
Table 9 summarizes oxide materials that play key roles in the functional layers of DSSCs. CeO2 currently has limited or auxiliary applications and requires further experimental verification for widespread use.

4.4. Thin-Film Chalcogenide and Inorganic Solar Cells

In thin-film solar cells based on CIGS, CdTe, CZTS, and amorphous silicon (a-Si:H), oxide ceramics play a key role as transparent conductive oxides (TCOs), buffer layers, or passivating dielectrics [344,345,346]. The most common configuration is a bilayer TCO composed of an inner ZnO layer and an outer conductive layer providing lateral conductivity [347]. Thanks to its wide band gap, good transparency, and doping ability, AZO is widely used not only in CIGS and CZTS, but also in CdTe and silicon-based cells [348,349,350].
TiO2 is being investigated as an alternative to CdS in CIGS and CdTe for cadmium-free structures [351,352]. TiO2 is also used to passivate grain boundaries in CIGS [353]. SnO2 serves as a standard transparent electrode in CdTe and some CIGS cells. Insulating oxides such as Al2O3 and SiO2 are applied for surface passivation, grain boundary recombination suppression, and the creation of dielectric interlayers [354]. High-work-function oxides like WO3 are placed between the absorber (e.g., CdTe or CIGS) and the metal contact to improve hole extraction [355]. These materials provide better energy level alignment, reduce contact losses, and can act as recombination barriers.
Table 10 summarizes the roles and technological characteristics of oxide materials in thin-film chalcogenide and inorganic solar cells. It includes only those oxides that have demonstrated practical effectiveness as transparent contacts, buffer, or passivating layers for CIGS, CdTe, CZTS, and a-Si:H. Experimentally promising but still less commonly used oxides, such as CeO2 (due to insufficient conductivity and less mature passivation control) and Fe2O3 (high optical absorption and low electron mobility limiting its use as TCO or buffer), remain primarily at the lab stage and require further validation for widespread implementation in thin-film PV technologies.

4.5. Organic and Emerging Types of Solar Cells

In emerging photovoltaic technologies, including organic solar cells (OSCs) and quantum dot (QD) solar cells, metal oxides are used as charge-selective transport layers due to their stability, suitable energy alignment, and transparency [356,357].
ZnO and TiO2 are widely used as electron transport layers (ETLs) in inverted OSC architectures [358,359]. High-work-function oxides, such as WO3, are also commonly used in OSCs [360,361]. These materials efficiently extract holes and provide favorable alignment with the valence band of donor polymers. When oxides like Fe2O3 or WO3 are used as absorbers, tandem or multilayer structures can be developed to broaden the spectral response. The main advantages of such systems include stability, low cost, and the absence of volatile or toxic components, making them promising for use in harsh environments such as space or in solar fuel production.
Table 11 summarizes the properties of oxide layers that have already found practical application in organic, quantum dot, and “all-oxide” solar cells. Notably, Al2O3, SiO2, and CeO2 mostly serve auxiliary functions, dielectric encapsulation, passivation, or UV-barrier, and have only a limited effect on charge-selective transport in these architectures.

4.6. Application Matrix of Oxide Ceramics in Solar Cells: Analytical Summary

The universality or specialization of oxide materials in solar cells is determined not only by their electronic structure and stability, but also by how they function in the devices—as active charge transport layers or as auxiliary passivating or optical components. Table 12 and Figure 12 summarize the key roles of each oxide in various photovoltaic architectures. This application matrix allows for simultaneous evaluation of the maturity, functional flexibility, and technological relevance of the materials, visually reflecting which ones are already implemented, have limited use, or are currently under active investigation.
TiO2 and ZnO have proven to be versatile solutions for most architectures (PSC, DSSC, OSC, QD), as evidenced by their leading citation counts in the literature and frequent appearance among keywords such as electron transport layer, mesoporous, dye-sensitized, and UV stability. Their flexibility is enabled by a combination of wide bandgap, high transparency, various nanostructuring options, and stable energetic alignment with active layers.
In contrast, SiO2 and Al2O3 serve clearly defined specialized roles—passivation, dielectric separation, and optical stabilization. This is reflected in a narrower range of keywords (e.g., surface passivation, antireflective coating) and in their lower, though stable, citation levels. While they are not charge carriers, they are essential for the stability and longevity of devices.
Oxides with potential (CeO2, Fe2O3, WO3) appear in the literature as emerging materials but have not yet achieved widespread implementation. Our bibliometric analysis shows that publications on CeO2 are often accompanied by tags such as UV filter, stability enhancement, and interface engineering, whereas Fe2O3 is typically associated with PEC, low mobility, and visible light absorber. Nevertheless, major barriers remain: unfavorable band alignment, low charge mobility, and complex interfacial chemistry.
Therefore, in the next generation of architectures (such as all-oxide, tandem, and inverted designs), combined structures look promising—where the drawbacks of one material are compensated by the advantages of another (e.g., TiO2@CeO2 or ZnO/WO3). This aligns with our bibliometric findings on keyword co-occurrence, where pairings such as ZnO + buffer layer, TiO2 + passivation, and WO3 + HTL are appearing with increasing frequency.
The publication peak we identified in the 2010s aligns with the strengthening of global and regional policies in the field of decarbonization and the deployment of renewable energy. At the global level, the Paris Agreement (2015) set the framework by institutionalizing long-term climate goals and the regular review of ambition [362]. At the EU level, momentum was reinforced by Directive (EU) 2018/2001 (RED II), which established binding targets and harmonized support rules for renewable energy [363]. National mechanisms also played a key role: in China, the introduction of a nationwide feed-in tariff for solar energy in 2011 sharply improved the investment attractiveness of PV projects [364], while in the United States, the ARRA program (2009) provided substantial investments in “clean” energy, grid infrastructure, and R&D [365,366]. The combined effect of these policies was reflected in empirical series: according to IEA and IEA-PVPS data, in the 2010s the installed capacity and generation of solar energy grew at explosive rates, correlating with our bibliometric peak in studies on ETL materials, passivation, and optical stacks (TiO2, ZnO, Al2O3, SiO2), as well as with a parallel rise of interest in “emerging” oxides (WO3, Fe2O3, CeO2) [367,368].
In conclusion, a comparative analysis of physical properties, device functions, and the publication landscape indicates that oxide ceramics are no longer auxiliary components but are emerging as a full-fledged platform for next-generation photovoltaic technologies. Their flexibility, chemical stability, and multifunctionality allow for material adaptation to specific architectures, opening pathways for interface engineering, integration into hybrid structures, and the development of stable, environmentally friendly solar cells.

4.7. Future Directions

The further development of oxide ceramics for solar energy primarily hinges on refined engineering of well-established materials, particularly TiO2 and ZnO. Despite their high efficiency in perovskite, organic, and quantum dot solar cells, the stability of these oxides remains limited by surface chemical reactivity: ZnO rapidly degrades in humid or acidic environments, while mesoporous TiO2 requires careful control of porosity and crystal phase. One of the most promising strategies is the deposition of ultrathin buffer layers of Al2O3, SiO2, or CeO2, as has been implemented for other architectures [369,370,371].
Beyond binary systems, the insights gained from our systematic review can also inspire the rational design of ternary and multicomponent oxide ceramics with more complex crystal structures. In particular, ZnO–Al2O3 spinels exhibit composition-dependent tunability of electronic and mechanical properties, making them attractive as transport or buffer layers [372]. Similarly, Al2O3–SiO2 mullite phases show favorable combinations of stability and electronic structure, which could be harnessed for dielectric or passivation functions in solar devices [373]. Although these systems are more challenging to model and synthesize, their structural flexibility offers greater opportunities for fine-tuning interfacial and optoelectronic properties, thereby broadening the design space for next-generation solar cell architectures.
For the next-generation group of oxides (CeO2, Fe2O3, and WO3), the main challenge is transitioning from laboratory prototypes to stable devices. CeO2 requires optimization of its crystal phase and integration with conventional ETLs; hematite can serve as an absorber or photoelectrode only if charge transport is enhanced through doping or heterostructure design; WO3 demonstrates excellent reliability, but its properties are highly sensitive to stoichiometry, requiring precise control during deposition.
A promising direction is the concept of fully oxide-based architectures, where the same class of materials serves as both the absorbing layer and selective contacts. Combinations such as ZnO/Cu2O, TiO2/NiO, or WO3/Fe2O3 have demonstrated the feasibility of environmentally friendly and thermally stable ‘all-oxide’ solar cells, which hold promise for competing with conventional technologies in certain applications [374,375]. An additional boost is expected from the rapid advancement of SnO2, which, alongside TiO2 and ZnO, is becoming a versatile electron transport layer in perovskite solar cells.
This trajectory can be accelerated by integrating bibliometric mapping with materials science analytics: systematic analysis of keywords and co-authorship networks can help identify overlooked oxides and optimally allocate research efforts across hybrid architectures, interfacial chemistry, and long-term stability. Taken together, these approaches constitute a roadmap for the development of efficient, durable, and environmentally sustainable next-generation photovoltaic technologies.
Beyond device performance, the sustainability profile of oxide ceramics will increasingly shape their research prioritization and industrial uptake. Al2O3 offers excellent passivation but is associated with energy-intensive alumina production and the generation of bauxite residue (“red mud”), calling for low-temperature deposition routes and waste-minimization strategies at the precursor stage [376]. ZnO is widely used as an ETL/TCO; however, it is classified by ECHA as Aquatic Acute/Chronic 1 (H400/H410), which underscores the need for responsible use, effluent treatment, and end-of-life (EOL) handling of nanostructured ZnO [377]. WO3 depends on tungsten, listed among the EU critical/strategic raw materials, highlighting supply risk exposure and the importance of precursor efficiency, recycling of sputtering targets, and supplier diversification. CeO2 relies on light rare-earth element (cerium); rare earths feature on EU criticality lists due to geographical concentration of deposits, making substitution, loading minimization, and circular-economy approaches pertinent [378,379]. In contrast, Fe2O3 and SiO2 generally exhibit benign toxicological profiles and broad availability; nonetheless, high-temperature synthesis for dense or textured films raises embodied energy concerns that LCA should capture [380].
Looking ahead, these sustainability considerations should not be seen merely as constraints, but rather as opportunities to reshape material strategies in photovoltaics. The move toward low-energy and low-temperature synthesis routes—from aerosol deposition to advanced atomic-layer techniques—offers not only lower environmental footprints but also novel microstructural control. Pursuing material thrift, such as ultrathin functional layers and judicious use of dopants, can reduce costs while opening avenues for flexible or tandem designs. The vision of design-for-disassembly—enabling oxide layers to be cleanly separated and recovered at end of life—points to a future where solar modules can become part of a circular economy rather than a waste burden. Addressing the supply risks of tungsten and rare earths by diversifying sources or substituting scarce elements may also stimulate innovation in alloying and doping strategies. Finally, the integration of prospective life-cycle assessments (LCA) and criticality analyses alongside traditional performance metrics promises to transform how new oxides are evaluated, ensuring that breakthroughs in efficiency go hand in hand with environmental responsibility.
Taken together, these directions outline a pathway where oxide ceramics evolve from being reliable functional layers to becoming the backbone of sustainable, recyclable, and globally resilient photovoltaic technologies.

5. Conclusions

This study presents a comprehensive bibliometric and comparative analysis of binary oxide materials used in modern solar cells. By combining quantitative analytics (based on Web of Science data), in-depth exploration of physicochemical properties, and role-based analysis of oxides across various device architectures, we not only synthesized current knowledge but also identified emerging directions for future research.
The bibliometric findings indicate the dominance of TiO2 and ZnO in the scientific discourse, which correlates with their versatility as charge transport layers in multiple solar cell types. At the same time, oxides with more specialized functions (such as Al2O3 and SiO2) show consistent citation patterns in niche areas, reaffirming their key role in enhancing device stability and surface passivation. Trends in keyword usage and co-authorship networks reveal the formation of knowledge clusters around specific oxide functions, including charge transport, stabilization, nanostructuring, and interface engineering.
Functional analysis revealed that the effectiveness of an oxide is determined not only by its bandgap or electron affinity but also by its adaptability to the specific operating requirements of the device. In particular, CeO2, Fe2O3, and WO3 hold considerable potential but still face challenges related to interfacial compatibility, charge transport, and phase stability.
The proposed classification of materials across different solar cell architectures provides a holistic understanding of the role of oxide ceramics in modern photovoltaics. The identified future directions point toward the development of stable, efficient, and environmentally safe next-generation photovoltaic technologies, especially within the scope of all-oxide concepts and hybrid device architectures.
Altogether, this comparative review demonstrates that binary oxides are not merely auxiliary materials but are evolving into a platform for next-generation photovoltaics. The analysis highlights that their future potential lies in interface engineering, stability enhancement, and environmentally safe device design. In particular, TiO2 and ZnO will continue to dominate through nanostructuring and heterojunction engineering, while Al2O3 and SiO2 remain indispensable for passivation and optical control. CeO2, Fe2O3, and WO3, though still emerging, offer promising opportunities in UV protection, photoelectrochemical hydrogen generation, and multifunctional device integration. Beyond binary systems, the bibliometric evidence and functional mapping presented here also point to ternary and multicomponent oxides (e.g., spinels, mullites) as logical extensions where compositional tunability could unlock new device concepts. By linking bibliometric trends to functional performance gaps, this study provides a roadmap that may inspire targeted innovations and guide future experimental and theoretical research in oxide-based solar energy technologies.

Author Contributions

Conceptualization, Y.S., S.N. and A.I.P.; methodology, Y.S. and S.N.; software, S.N.; validation, Y.S. and S.N.; formal analysis, Y.S. and S.N.; investigation, Y.S., S.N. and A.I.P.; resources, S.N.; data curation, Y.S., S.N., M.K. and A.I.P.; writing—original draft preparation, Y.S. and S.N.; writing—review and editing, M.K. and A.I.P.; visualization, Y.S. and S.N.; supervision, Y.S. and A.I.P.; project administration, A.I.P.; funding acquisition, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Fund of Ukraine with the support of the University of Cambridge, Great Britain, grant number 0124U000223 “Design and Research of Oxide Heterostructures for Portable Solar Cells.”.

Data Availability Statement

All data that was used for this research may be found at Zenodo: https://doi.org/10.5281/zenodo.15829792. Bibliometric maps (Figure 4, Figure 5, Figure 6, Figure 7, Figure 8, Figure 9 and Figure 10) can be accessed via the following permanent links: TiO2 (titanium dioxide): https://tinyurl.com/245z4qkv (accessed on 7 September 2025), https://tinyurl.com/29ecu8nc (accessed on 7 September 2025); ZnO (zinc oxide): https://tinyurl.com/2cjolpkj (accessed on 7 September 2025), https://tinyurl.com/29k59moj (accessed on 7 September 2025); SiO2 (silicon dioxide): https://tinyurl.com/23ktygy6 (accessed on 7 September 2025), https://tinyurl.com/224hqtyt (accessed on 7 September 2025); Al2O3 (aluminum oxide): https://tinyurl.com/2dceubry (accessed on 7 September 2025), https://tinyurl.com/22qy7qzs (accessed on 7 September 2025); Fe2O3 (iron oxide): https://tinyurl.com/26xze96n (accessed on 7 September 2025), https://tinyurl.com/28kszl94 (accessed on 7 September 2025); WO3 (tungsten trioxide): https://tinyurl.com/22smuu6m (accessed on 7 September 2025), https://tinyurl.com/26s56kfs (accessed on 7 September 2025); CeO2 (cerium dioxide): https://tinyurl.com/25ba3qky (accessed on 7 September 2025), https://tinyurl.com/26wxuydq (accessed on 7 September 2025).

Acknowledgments

Yana Suchikova and Serhii Nazarovets would like to thank all Ukrainian defenders for the possibility to finalize and publish this work.

Conflicts of Interest

The authors declare that they have no competing financial or non-financial interests.

References

  1. Bergmann, R. Crystalline Si thin-film solar cells: A review. Appl. Phys. A 1999, 69, 187–194. [Google Scholar] [CrossRef]
  2. Andreani, L.C.; Bozzola, A.; Kowalczewski, P.; Liscidini, M.; Redorici, L. Silicon solar cells: Toward the efficiency limits. Adv. Phys. X 2018, 4, 1548305. [Google Scholar] [CrossRef]
  3. Green, M.A. Silicon solar cells: State of the art. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2013, 371, 20110413. [Google Scholar] [CrossRef]
  4. Ferekides, C.; Balasubramanian, U.; Mamazza, R.; Viswanathan, V.; Zhao, H.; Morel, D. CdTe thin film solar cells: Device and technology issues. Sol. Energy 2004, 77, 823–830. [Google Scholar] [CrossRef]
  5. Ali, M.; Khan, Q.; Din, M.F.U.; Kasi, J.K.; Kasi, A.K.; Ali, A.; Ullah, S. Simulation-based optimization of CdS/CdTe solar cells incorporating MXene-enhanced SnO2 buffer layer: Insights from experimentally validated material properties. Sol. Energy 2025, 294, 113510. [Google Scholar] [CrossRef]
  6. Barbato, M.; Artegiani, E.; Bertoncello, M.; Meneghini, M.; Trivellin, N.; Mantoan, E.; Romeo, A.; Mura, G.; Ortolani, L.; Zanoni, E.; et al. CdTe solar cells: Technology, operation and reliability. J. Phys. D Appl. Phys. 2021, 54, 333002. [Google Scholar] [CrossRef]
  7. Mufti, N.; Amrillah, T.; Taufiq, A.; Diantoro, M.; Nur, H. Review of CIGS-based solar cells manufacturing by structural engineering. Sol. Energy 2020, 207, 1146–1157. [Google Scholar] [CrossRef]
  8. Suchikova, Y.; Bohdanov, I.; Kovachov, S.; Popov, A.I. Thin CIGS Films Obtained by Spray Pyrolysis. In Nanomaterials and Nanocomposites, Nanostructures, and Their Applications. NANO 2023; Fesenko, O., Yatsenko, L., Eds.; Springer Proceedings in Physics; Springer: Cham, Switzerland, 2024; Volume 253. [Google Scholar] [CrossRef]
  9. Lin, L.; Ravindra, N.M. Temperature dependence of CIGS and perovskite solar cell performance: An overview. SN Appl. Sci. 2020, 2, 1361. [Google Scholar] [CrossRef]
  10. Lu, S.; Chen, C.; Tang, J. Possible top cells for next-generation Si-based tandem solar cells. Front. Optoelectron. 2020, 13, 246–255. [Google Scholar] [CrossRef]
  11. Panigrahi, J.; Komarala, V.K. Progress on the intrinsic a-Si:H films for interface passivation of silicon heterojunction solar cells: A review. J. Non-Cryst. Solids 2021, 574, 121166. [Google Scholar] [CrossRef]
  12. Zhou, J.; Su, X.; Huang, Q.; Zhang, B.; Yang, J.; Zhao, Y.; Hou, G. Recent advancements in poly-Si/SiOxpassivating contacts for high-efficiency silicon solar cells: Technology review and perspectives. J. Mater. Chem. A 2022, 10, 20147–20173. [Google Scholar] [CrossRef]
  13. Raza, E.; Ahmad, Z. Review on two-terminal and four-terminal crystalline-silicon/perovskite tandem solar cells; progress, challenges, and future perspectives. Energy Rep. 2022, 8, 5820–5851. [Google Scholar] [CrossRef]
  14. Jiang, L.; Cui, S.; Sun, P.; Wang, Y.; Yang, C. Comparison of Monocrystalline and Polycrystalline Solar Modules. In Proceedings of the 2020 IEEE 5th Information Technology and Mechatronics Engineering Conference (ITOEC), Chongqing, China, 12–14 June 2020; pp. 341–344. [Google Scholar] [CrossRef]
  15. Enaganti, P.K.; Dwivedi, P.K.; Srivastava, A.K.; Goel, S. Study of solar irradiance and performance analysis of submerged monocrystalline and polycrystalline solar cells. Prog. Photovolt. Res. Appl. 2020, 28, 725–735. [Google Scholar] [CrossRef]
  16. Suchikova, Y. Provision of environmental safety through the use of porous semiconductors for solar energy sector. East.-Eur. J. Enterp. Technol. 2016, 6, 26–33. [Google Scholar] [CrossRef][Green Version]
  17. Santos, I.M.D.L.; Courel, M.; Moreno-Oliva, V.I.; Dueñas-Reyes, E.; Díaz-Cruz, E.B.; Ojeda-Martínez, M.; Pérez, L.M.; Laroze, D. The study of inorganic absorber layers in perovskite solar cells: The influence of CdTe and CIGS incorporation. Sci. Rep. 2025, 15, 10353. [Google Scholar] [CrossRef] [PubMed]
  18. Iqbal, N.; Colvin, D.J.; Schneller, E.J.; Sakthivel, T.S.; Ristau, R.; Huey, B.D.; Yu, B.X.; Jaubert, J.-N.; Curran, A.J.; Wang, M.; et al. Characterization of front contact degradation in monocrystalline and multicrystalline silicon photovoltaic modules following damp heat exposure. Sol. Energy Mater. Sol. Cells 2022, 235, 111468. [Google Scholar] [CrossRef]
  19. Wijewardane, S.; Kazmerski, L.L. Inventions, innovations, and new technologies: Flexible and lightweight thin-film solar PV based on CIGS, CdTe, and a-Si:H. Sol. Compass 2023, 7, 100053. [Google Scholar] [CrossRef]
  20. Gao, B.; Shao, Y.; Liu, W.; Xiang, H.; Yu, Y.; Liu, Z. Out-door reliability and degradation of HIT, CIGS, n-type multi-busbar, PERC, and CdTe modules in Shanghai, China. Sol. Energy Mater. Sol. Cells 2022, 236, 111490. [Google Scholar] [CrossRef]
  21. Bosio, A.; Pasini, S.; Romeo, N. The History of Photovoltaics with Emphasis on CdTe Solar Cells and Modules. Coatings 2020, 10, 344. [Google Scholar] [CrossRef]
  22. Fang, Z.; Wang, X.C.; Wu, H.C.; Zhao, C.Z. Achievements and Challenges of CdS/CdTe Solar Cells. Int. J. Photoenergy 2011, 2011, 297350. [Google Scholar] [CrossRef]
  23. Khrypunov, G.; Vambol, S.; Deyneko, N.; Sychikova, Y. Increasing the efficiency of film solar cells based on cadmium telluride. East.-Eur. J. Enterp. Technol. 2016, 6, 12–18. [Google Scholar] [CrossRef]
  24. Romeo, A.; Artegiani, E.; Menossi, D. Low substrate temperature CdTe solar cells: A review. Sol. Energy 2018, 175, 9–15. [Google Scholar] [CrossRef]
  25. Balakhayeva, R.; Akilbekov, A.; Baimukhanov, Z.; Usseinov, A.; Giniyatova, S.; Zdorovets, M.; Vlasukova, L.; Popov, A.I.; Dauletbekova, A. CdTe Nanocrystal Synthesis in SiO2/Si Ion-Track Template: The Study of Electronic and Structural Properties. Phys. Status Solidi (A) 2020, 218, 2000231. [Google Scholar] [CrossRef]
  26. Zghaibeh, M.; Okonkwo, P.C.; Emori, W.; Ahmed, T.; Mohamed, A.; Aliyu, M.; Ogunleye, G.J. CdTe solar cells fabrication and examination techniques: A focused review. Int. J. Green Energy 2022, 20, 555–570. [Google Scholar] [CrossRef]
  27. Fthenakis, V.; Athias, C.; Blumenthal, A.; Kulur, A.; Magliozzo, J.; Ng, D. Sustainability evaluation of CdTe PV: An update. Renew. Sustain. Energy Rev. 2020, 123, 109776. [Google Scholar] [CrossRef]
  28. Curtin, A.; Vail, C.; Buckley, H. CdTe in thin film photovoltaic cells: Interventions to protect drinking water in production and end-of-life. Water-Energy Nexus 2020, 3, 15–28. [Google Scholar] [CrossRef]
  29. Lin, L.; Ravindra, N.M. CIGS and perovskite solar cells—An overview. Emerg. Mater. Res. 2020, 9, 812–824. [Google Scholar] [CrossRef]
  30. Ramanujam, J.; Singh, U.P. Copper indium gallium selenide based solar cells—A review. Energy Environ. Sci. 2017, 10, 1306–1319. [Google Scholar] [CrossRef]
  31. Ong, K.H.; Agileswari, R.; Maniscalco, B.; Arnou, P.; Kumar, C.C.; Bowers, J.W.; Marsadek, M.B. Review on Substrate and Molybdenum Back Contact in CIGS Thin Film Solar Cell. Int. J. Photoenergy 2018, 2018, 9106269. [Google Scholar] [CrossRef]
  32. Romanyuk, Y.E.; Hagendorfer, H.; Stücheli, P.; Fuchs, P.; Uhl, A.R.; Sutter-Fella, C.M.; Werner, M.; Haass, S.; Stückelberger, J.; Broussillou, C.; et al. All Solution-Processed Chalcogenide Solar Cells—From Single Functional Layers Towards a 13.8% Efficient CIGS Device. Adv. Funct. Mater. 2014, 25, 12–27. [Google Scholar] [CrossRef]
  33. Feurer, T.; Reinhard, P.; Avancini, E.; Bissig, B.; Löckinger, J.; Fuchs, P.; Carron, R.; Weiss, T.P.; Perrenoud, J.; Stutterheim, S.; et al. Progress in thin film CIGS photovoltaics—Research and development, manufacturing, and applications. Prog. Photovolt. Res. Appl. 2016, 25, 645–667. [Google Scholar] [CrossRef]
  34. Zhao, C.; Yu, S.; Tang, W.; Yuan, X.; Zhou, H.; Qi, T.; Zheng, X.; Ning, D.; Ma, M.; Zhu, J.; et al. Advances in CIGS thin film solar cells with emphasis on the alkali element post-deposition treatment. Mater. Rep. Energy 2023, 3, 100214. [Google Scholar] [CrossRef]
  35. Kim, S.; Van Quy, H.; Bark, C.W. Photovoltaic technologies for flexible solar cells: Beyond silicon. Mater. Today Energy 2021, 19, 100583. [Google Scholar] [CrossRef]
  36. Dhere, N.G. Toward GW/year of CIGS production within the next decade. Sol. Energy Mater. Sol. Cells 2007, 91, 1376–1382. [Google Scholar] [CrossRef]
  37. Singh, B.P.; Goyal, S.K.; Kumar, P. Solar PV cell materials and technologies: Analyzing the recent developments. Mater. Today Proc. 2021, 43, 2843–2849. [Google Scholar] [CrossRef]
  38. Wang, M.; Obene, P.; Questianx, M.; Harris, M.; Singh, R.; Choy, K.L. Electrostatic InkJet printed silver grids for non-vacuum processed CIGS solar cells. Sci. Rep. 2025, 15, 11048. [Google Scholar] [CrossRef]
  39. Tong, Y.; Xiao, Z.; Du, X.; Zuo, C.; Li, Y.; Lv, M.; Yuan, Y.; Yi, C.; Hao, F.; Hua, Y.; et al. Progress of the key materials for organic solar cells. Sci. China Chem. 2020, 63, 758–765. [Google Scholar] [CrossRef]
  40. Li, X.; Li, P.; Wu, Z.; Luo, D.; Yu, H.-Y.; Lu, Z.-H. Review and perspective of materials for flexible solar cells. Mater. Rep. Energy 2021, 1, 100001. [Google Scholar] [CrossRef]
  41. Suchikova, Y.; Bogdanov, I.; Kovachov, S.; Kamensky, D.; Myroshnychenko, V.; Panova, N. Optimal ranges determination of morphological parameters of nanopatterned semiconductors quality for solar cells. Arch. Mater. Sci. Eng. 2020, 1, 15–24. [Google Scholar] [CrossRef]
  42. Li, J.; Aierken, A.; Liu, Y.; Zhuang, Y.; Yang, X.; Mo, J.H.; Fan, R.K.; Chen, Q.Y.; Zhang, S.Y.; Huang, Y.M.; et al. A Brief Review of High Efficiency III-V Solar Cells for Space Application. Front. Phys. 2021, 8, 631925. [Google Scholar] [CrossRef]
  43. Yao, H.; Wang, J.; Xu, Y.; Zhang, S.; Hou, J. Recent Progress in Chlorinated Organic Photovoltaic Materials. Acc. Chem. Res. 2020, 53, 822–832. [Google Scholar] [CrossRef]
  44. Fan, B.; Lin, F.; Wu, X.; Zhu, Z.; Jen, A.K.-Y. Selenium-Containing Organic Photovoltaic Materials. Acc. Chem. Res. 2021, 54, 3906–3916. [Google Scholar] [CrossRef] [PubMed]
  45. Kim, J.Y.; Lee, J.-W.; Jung, H.S.; Shin, H.; Park, N.-G. High-Efficiency Perovskite Solar Cells. Chem. Rev. 2020, 120, 7867–7918. [Google Scholar] [CrossRef] [PubMed]
  46. Kumar, N.S.; Naidu, K.C.B. A review on perovskite solar cells (PSCs), materials and applications. J. Mater. 2021, 7, 940–956. [Google Scholar] [CrossRef]
  47. Liu, L.; Najar, A.; Wang, K.; Du, M.; Liu, S. Perovskite Quantum Dots in Solar Cells. Adv. Sci. 2022, 9, 2104577. [Google Scholar] [CrossRef]
  48. Albaladejo-Siguan, M.; Baird, E.C.; Becker-Koch, D.; Li, Y.; Rogach, A.L.; Vaynzof, Y. Stability of Quantum Dot Solar Cells: A Matter of (Life)Time. Adv. Energy Mater. 2021, 11, 2003457. [Google Scholar] [CrossRef]
  49. Duan, L.; Hu, L.; Guan, X.; Lin, C.; Chu, D.; Huang, S.; Liu, X.; Yuan, J.; Wu, T. Quantum Dots for Photovoltaics: A Tale of Two Materials. Adv. Energy Mater. 2021, 11, 2100354. [Google Scholar] [CrossRef]
  50. Schileo, G.; Grancini, G. Lead or no lead? Availability, toxicity, sustainability and environmental impact of lead-free perovskite solar cells. J. Mater. Chem. C 2021, 9, 67–76. [Google Scholar] [CrossRef]
  51. Huang, Z.; Beard, M.C. Dye-Sensitized Multiple Exciton Generation in Lead Sulfide Quantum Dots. J. Am. Chem. Soc. 2022, 144, 15855–15861. [Google Scholar] [CrossRef]
  52. Sahu, A.; Garg, A.; Dixit, A. A review on quantum dot sensitized solar cells: Past, present and future towards carrier multiplication with a possibility for higher efficiency. Sol. Energy 2020, 203, 210–239. [Google Scholar] [CrossRef]
  53. Rakshit, S.; Piatkowski, P.; Mora-Seró, I.; Douhal, A. Combining Perovskites and Quantum Dots: Synthesis, Characterization, and Applications in Solar Cells, LEDs, and Photodetectors. Adv. Opt. Mater. 2022, 10, 2102566. [Google Scholar] [CrossRef]
  54. Kim, T.; Lim, S.; Yun, S.; Jeong, S.; Park, T.; Choi, J. Design Strategy of Quantum Dot Thin-Film Solar Cells. Small 2020, 16, e2002460. [Google Scholar] [CrossRef]
  55. Zhang, G.; Lin, F.R.; Qi, F.; Heumüller, T.; Distler, A.; Egelhaaf, H.-J.; Li, N.; Chow, P.C.Y.; Brabec, C.J.; Jen, A.K.-Y.; et al. Renewed Prospects for Organic Photovoltaics. Chem. Rev. 2022, 122, 14180–14274. [Google Scholar] [CrossRef]
  56. Li, Y.; Huang, X.; Sheriff, H.K.M.; Forrest, S.R. Semitransparent organic photovoltaics for building-integrated photovoltaic applications. Nat. Rev. Mater. 2022, 8, 186–201. [Google Scholar] [CrossRef]
  57. Akinoglu, B.G.; Tuncel, B.; Badescu, V. Beyond 3rd generation solar cells and the full spectrum project. Recent advances and new emerging solar cells. Sustain. Energy Technol. Assess. 2021, 46, 101287. [Google Scholar] [CrossRef]
  58. Cramer, C.L.; Ionescu, E.; Graczyk-Zajac, M.; Nelson, A.T.; Katoh, Y.; Haslam, J.J.; Wondraczek, L.; Aguirre, T.G.; LeBlanc, S.; Wang, H.; et al. Additive manufacturing of ceramic materials for energy applications: Road map and opportunities. J. Eur. Ceram. Soc. 2022, 42, 3049–3088. [Google Scholar] [CrossRef]
  59. Suchikova, Y.; Lazarenko, A.; Kovachov, S.; Usseinov, A.; Karipbaev, Z.; Popov, A.I. Formation of porous Ga2O3/GaAs layers for electronic devices. In Proceedings of the 2022 IEEE 16th International Conference on Advanced Trends in Radioelectronics, Telecommunications and Computer Engineering (TCSET), Lviv-Slavske, Ukraine, 22–26 February 2022. [Google Scholar] [CrossRef]
  60. Kadyrzhanov, K.K.; Kozlovskiy, A.A.; Konuhova, M.; Popov, A.I.; Shlimas, D.D.; Borgekov, D.B. Determination of gamma radiation shielding efficiency by radiation-resistant composite ZrO2–Al2O3—TiO2–WO3—Nb2O5 ceramics. Opt. Mater. 2024, 154, 115752. [Google Scholar] [CrossRef]
  61. Usseinov, A.; Koishybayeva, Z.; Platonenko, A.; Akilbekov, A.; Purans, J.; Pankratov, V.; Suchikova, Y.; Popov, A.I. Ab-Initio Calculations of Oxygen Vacancy in Ga2O3 Crystals. Latv. J. Phys. Tech. Sci. 2021, 58, 3–10. [Google Scholar] [CrossRef]
  62. Suchikova, Y.; Nazarovets, S.; Popov, A.I. Ga2O3 solar-blind photodetectors: From civilian applications to missile detection and research agenda. Opt. Mater. 2024, 157, 116397. [Google Scholar] [CrossRef]
  63. Zhou, S.; Pu, Y.; Zhang, X.; Shi, Y.; Gao, Z.; Feng, Y.; Shen, G.; Wang, X.; Wang, D. High energy density, temperature stable lead-free ceramics by introducing high entropy perovskite oxide. Chem. Eng. J. 2022, 427, 131684. [Google Scholar] [CrossRef]
  64. Usseinov, A.; Koishybayeva, Z.; Platonenko, A.; Pankratov, V.; Suchikova, Y.; Akilbekov, A.; Zdorovets, M.; Purans, J.; Popov, A.I. Vacancy Defects in Ga2O3: First-Principles Calculations of Electronic Structure. Materials 2021, 14, 7384. [Google Scholar] [CrossRef]
  65. Jiao, Y.; Dai, J.; Fan, Z.; Cheng, J.; Zheng, G.; Grema, L.; Zhong, J.; Li, H.-F.; Wang, D. Overview of high-entropy oxide ceramics. Mater. Today 2024, 77, 92–117. [Google Scholar] [CrossRef]
  66. Wang, Z.; Lin, R.; Huo, Y.; Li, H.; Wang, L. Formation, Detection, and Function of Oxygen Vacancy in Metal Oxides for Solar Energy Conversion. Adv. Funct. Mater. 2021, 32, 2109503. [Google Scholar] [CrossRef]
  67. Ming, W.; Jiang, Z.; Luo, G.; Xu, Y.; He, W.; Xie, Z.; Shen, D.; Li, L. Progress in Transparent Nano-Ceramics and Their Potential Applications. Nanomaterials 2022, 12, 1491. [Google Scholar] [CrossRef] [PubMed]
  68. Bondarchuk, A.N.; Corrales-Mendoza, I.; Aguilar-Martínez, J.A.; Tomás, S.A.; Gómez-Caiceros, D.A.; Hernández-Méndez, A.; Marken, F. A BiVO4 photoanode grown on porous and conductive SnO2 ceramics for water splitting driven by solar energy. Ceram. Int. 2020, 46, 9040–9049. [Google Scholar] [CrossRef]
  69. Freitas, A.L.M.; Muche, D.N.F.; Leite, E.R.; Souza, F.L. Interface engineering of nanoceramic hematite photoelectrode for solar energy conversion. J. Am. Ceram. Soc. 2020, 103, 6833–6846. [Google Scholar] [CrossRef]
  70. Ahmad, W.; Tahir, S.; Ali, A.; Mahmood, K. A novel approach to reduce both front and rear side power losses in PERC solar cells using different combinations of transparent metal oxides. Ceram. Int. 2022, 49, 2821–2828. [Google Scholar] [CrossRef]
  71. Möllmann, A.; Bialuschewski, D.; Fischer, T.; Tachibana, Y.; Mathur, S. Functional metal oxide ceramics as electron transport medium in photovoltaics and photo-electrocatalysis. In Advanced Ceramics for Energy Conversion and Storage; Elsevier: Amsterdam, The Netherlands, 2020; pp. 207–273. [Google Scholar] [CrossRef]
  72. Almakayeel, N.; Kaliyannan, G.V.; Gunasekaran, R. Enhancing power conversion efficiency of polycrystalline silicon photovoltaic cells using yttrium oxide anti-reflective coating via electro-spraying method. Ceram. Int. 2024, 50, 42392–42403. [Google Scholar] [CrossRef]
  73. Maqsood, S.; Ali, Z.; Ali, K.; Ishaq, M.; Sajid, M.; Farhan, A.; Rahdar, A.; Pandey, S. Assessment of different optimized anti-reflection coatings for ZnO/Si heterojunction solar cells. Ceram. Int. 2023, 49, 37118–37126. [Google Scholar] [CrossRef]
  74. Sadiq, I.; Ali, S.A.; Ahmad, T. Advanced Hybrid Ceramics for Nuclear and Hydrogen Energy Applications. ChemistrySelect 2023, 8, e202300837. [Google Scholar] [CrossRef]
  75. Alizadeh, A.; Roudgar-Amoli, M.; Bonyad-Shekalgourabi, S.-M.; Shariatinia, Z.; Mahmoudi, M.; Saadat, F. Dye sensitized solar cells go beyond using perovskite and spinel inorganic materials: A review. Renew. Sustain. Energy Rev. 2022, 157, 112047. [Google Scholar] [CrossRef]
  76. Kusuma, J.; Balakrishna, R.G. Ceramic grains: Highly promising hole transport material for solid state QDSSC. Sol. Energy Mater. Sol. Cells 2020, 209, 110445. [Google Scholar] [CrossRef]
  77. Mozafari, H.; Hamidinezhad, H. Effect of electrospun bilayer titanium oxide–tin oxide nanofibers on the performance of dye-sensitized solar cells. Opt. Mater. 2023, 144, 114297. [Google Scholar] [CrossRef]
  78. Szindler, M.; Szindler, M.; Drygała, A.; Lukaszkowicz, K.; Kaim, P.; Pietruszka, R. Dye-Sensitized Solar Cell for Building-Integrated Photovoltaic (BIPV) Applications. Materials 2021, 14, 3743. [Google Scholar] [CrossRef]
  79. Exeler, J.; Jüstel, T. Advances in Functional Ceramics for Water Splitting: A Comprehensive Review. Photochem 2024, 4, 271–284. [Google Scholar] [CrossRef]
  80. Hassan, N.S.; Jalil, A.A.; Khusnun, N.F.; Ahmad, A.; Abdullah, T.A.T.; Kasmani, R.M.; Norazahar, N.; Kamaroddin, M.F.A.; Vo, D.V.N. Photoelectrochemical water splitting using post-transition metal oxides for hydrogen production: A review. Environ. Chem. Lett. 2021, 20, 311–333. [Google Scholar] [CrossRef]
  81. Halfadji, A.; Bennabi, L.; Giannakis, S.; Marrani, A.G.; Bellucci, S. Sono-synthesis and characterization of next-generation antimicrobial ZnO/TiO2 and Fe3O4/TiO2 bi-nanocomposites, for antibacterial and antifungal applications. Ceram. Int. 2024, 50, 39097–39108. [Google Scholar] [CrossRef]
  82. Kaler, V.; Pandel, U.; Duchaniya, R. Development of TiO2/PVA nanocomposites for application in solar cells. Mater. Today Proc. 2018, 5, 6279–6287. [Google Scholar] [CrossRef]
  83. Serga, V.; Burve, R.; Krumina, A.; Pankratova, V.; Popov, A.I.; Pankratov, V. Study of phase composition, photocatalytic activity, and photoluminescence of TiO2 with Eu additive produced by the extraction-pyrolytic method. J. Mater. Res. Technol. 2021, 13, 2350–2360. [Google Scholar] [CrossRef]
  84. Rakhman, K.A.; Budhyantoro, A.; Aprilita, N.H.; Kartini, I. One-pot synthesis of hollow sphere TiO2/Ag nanoparticles co-sensitized with peonidin: Pelargonidin for dye-sensitized solar cells applications. J. Mater. Sci. Mater. Electron. 2024, 35, 1926. [Google Scholar] [CrossRef]
  85. Khan, M.; Ali, S.; Alwadai, N.; Haq, I.U.; Irfan, M.; Albalawi, H.; Almuqrin, A.H.; Almoneef, M.M.; Iqbal, M. Structural, electrical and optical properties of hetrostructured MoS2/ZnO thin films for potential perovskite solar cells application. J. Mater. Res. Technol. 2022, 20, 1616–1623. [Google Scholar] [CrossRef]
  86. Wibowo, A.; Marsudi, M.A.; Amal, M.I.; Ananda, M.B.; Stephanie, R.; Ardy, H.; Diguna, L.J. ZnO nanostructured materials for emerging solar cell applications. RSC Adv. 2020, 10, 42838–42859. [Google Scholar] [CrossRef]
  87. Hajji, M.; Ajili, M.; Jebbari, N.; Loreiro, A.G.; Kamoun, N.T. Photocatalytic performance and solar cell applications of coupled semiconductor CuO– ZnO sprayed thin films: Coupling effect between oxides. Opt. Mater. 2023, 140, 113798. [Google Scholar] [CrossRef]
  88. Alsmadi, A.K.M.; Salameh, B.; Alshammari, O.; Bumajdad, A.; Madkour, M. Synthesis, photocatalytic, and photoelectric performance of mesoporous Au/TiO2 and Au/TiO2/MWCNT nanocomposites. J. Phys. Chem. Solids 2025, 207, 112874. [Google Scholar] [CrossRef]
  89. Tyagi, J.; Gupta, H.; Purohit, L. Mesoporous ZnO/TiO2 photoanodes for quantum dot sensitized solar cell. Opt. Mater. 2021, 115, 111014. [Google Scholar] [CrossRef]
  90. Kaur, N.; Singh, D.P.; Mahajan, A. Plasmonic Engineering of TiO2 Photoanodes for Dye-Sensitized Solar Cells: A Review. J. Electron. Mater. 2022, 51, 4188–4206. [Google Scholar] [CrossRef]
  91. Lana, G.M.; Bello, I.T.; Adedokun, O.M.; Adenigba, V.O.; Jubu, P.R.; Adedokun, O.; Sanusi, Y.K.; Dhlamini, M.S.; Awodugba, A.O. One-Dimensional TiO2 Nanocomposite-based Photoanode for Dye-Sensitized solar Cells: A review. Sol. Energy 2024, 279, 112850. [Google Scholar] [CrossRef]
  92. Sha, S.; Lu, H.; Yang, S.; Li, T.; Wu, J.; Ma, J.; Wang, K.; Hou, C.; Sheng, Z.; Li, Y. One-step electrodeposition of ZnO/graphene composite film as photoanode for dye-sensitised solar cells. Colloids Surf. A Physicochem. Eng. Asp. 2021, 630, 127491. [Google Scholar] [CrossRef]
  93. Almakayeel, N.; Kaliyannan, G.V.; Gunasekaran, R.; Rathanasamy, R. Enhanced photovoltaic efficiency through 3D-Printed COC/Al2O3 anti-reflective coversheets. J. Mater. Res. Technol. 2024, 33, 3093–3105. [Google Scholar] [CrossRef]
  94. Shablonin, E.; Popov, A.; Prieditis, G.; Vasil’CHenko, E.; Lushchik, A. Thermal annealing and transformation of dimer F centers in neutron-irradiated Al2O3 single crystals. J. Nucl. Mater. 2021, 543, 152600. [Google Scholar] [CrossRef]
  95. Yuliantini, L.; Nursam, N.M.; Pranoto, L.M.; Hidayat, J.; Sova, R.R.; Rahayu, E.S.; Djamal, M.; Yasaka, P.; Boonin, K.; Kaewkhao, J. Photon up-conversion in Er3+ ion-doped ZnO-Al2O3-BaO-B2O3 glass for enhancing the performance of dye-sensitized solar cells. J. Alloys Compd. 2023, 954, 170163. [Google Scholar] [CrossRef]
  96. Verma, M.; Mishra, G.P. High Electric Field Sensing in Ultrathin SiO2 and Tunnel Region to Enhance GaInP/Si Dual Junction Solar Cell Performance. IEEE Sens. J. 2021, 22, 1273–1279. [Google Scholar] [CrossRef]
  97. Wang, Z.; Zhang, X.; Liu, Q.; Luo, G.; Lu, J.; Xie, Y.; Zhao, X.; Tian, S. Superhydrophilic antireflection films with excellent optical and mechanical performance for perovskite solar cells. Inorg. Chem. Commun. 2024, 168, 112876. [Google Scholar] [CrossRef]
  98. Molina-Lozano, A.E.; Lanza, M.R.V.; Ortiz, P.; Cortés, M.T. Photoelectrochemical and Structural Insights of Electrodeposited CeO2 Photoanodes. Surfaces 2024, 7, 898–919. [Google Scholar] [CrossRef]
  99. Ravi, G.; Thyagarajan, K. CoFe2O4/NiFe2O4/CeO2 nanocomposites: Structural, FTIR, XPS, BET, and magnetic properties. Appl. Phys. A 2024, 130, 726. [Google Scholar] [CrossRef]
  100. Kumar, P.; Ratan, J.K.; Divya, N.; Mummaneni, K.; Rawat, G. Investigation of CeO2/rGO Nanocomposites as Diesel Additives to Enhance Engine Performance and Reduce Exhaust Emissions. In Proceedings of the 2023 9th International Conference on Signal Processing and Communication (ICSC), Noida, India, 21–23 December 2023; pp. 714–719. [Google Scholar] [CrossRef]
  101. Nien, Y.-H.; Zhuang, S.-W.; Chou, J.-C.; Yang, P.-H.; Lai, C.-H.; Kuo, P.-Y.; Ho, C.-S.; Wu, Y.-T.; Syu, R.-H.; Chen, P.-F. Photovoltaic Measurement Under Different Illumination of the Dye-Sensitized Solar Cell With the Photoanode Modified by Fe2O3/g-C3N4/TiO2 Heterogeneous Nanofibers Prepared by Electrospinning With Dual Jets. IEEE Trans. Semicond. Manuf. 2023, 36, 291–297. [Google Scholar] [CrossRef]
  102. Wu, S.; Cheng, C.-H.; Hsiao, Y.-J.; Juang, R.-C.; Wen, W.-F. Fe2O3 films on stainless steel for solar absorbers. Renew. Sustain. Energy Rev. 2016, 58, 574–580. [Google Scholar] [CrossRef]
  103. Kamil, A.F.; Abdullah, H.I.; Rheima, A.M.; Mohammed, S.H. UV-Irradiation synthesized α-Fe2O3 nanoparticles based dye-sensitized solar cells. Mater. Today Proc. 2022, 61, 820–825. [Google Scholar] [CrossRef]
  104. Shinde, S.S.; Bansode, R.A.; Bhosale, C.H.; Rajpure, K.Y. Physical properties of hematite α-Fe2O3thin films: Application to photoelectrochemical solar cells. J. Semicond. 2011, 32, 013001. [Google Scholar] [CrossRef]
  105. Chen, L.; Wu, S.; Ma, D.; Shang, A.; Li, X. Optoelectronic modeling of the Si/α-Fe2O3 heterojunction photoanode. Nano Energy 2018, 43, 177–183. [Google Scholar] [CrossRef]
  106. Dadkhah, M.; Nine, J.; Purasinhala, K.; Sandhu, G.S.; Losic, D. Nanostructure-dependent colouration efficiency of electrochromic coatings using 0D, 1D, and 2D WO3 for smart windows. Nano Mater. Sci. 2024. [Google Scholar] [CrossRef]
  107. Wang, L.; Zhai, Z.; Li, L. Rapid Fabrication of Tungsten Oxide-Based Electrochromic Devices through Femtosecond Laser Processing. Micromachines 2024, 15, 785. [Google Scholar] [CrossRef] [PubMed]
  108. Aldirham, S.H.; Helal, A.; Shkir, M.; Sayed, M.A.; Ali, A.M. Enhancement Study of the Photoactivity of TiO2 Photocatalysts during the Increase of the WO3 Ratio in the Presence of Ag Metal. Catalysts 2024, 14, 633. [Google Scholar] [CrossRef]
  109. Matalkeh, M.; Nasrallah, G.K.; Shurrab, F.M.; Al-Absi, E.S.; Mohammed, W.; Elzatahry, A.; Saoud, K.M. Visible Light Photocatalytic Activity of Ag/WO3 Nanoparticles and its Antibacterial Activity Under Ambient Light and in The Dark. Results Eng. 2022, 13, 100313. [Google Scholar] [CrossRef]
  110. Ghattavi, S.; Nezamzadeh-Ejhieh, A. Nanoscale AgI-WO3 binary photocatalyst: Synthesis, brief characterization, and investigation of its photocatalytic activity. Mater. Res. Bull. 2022, 158, 112085. [Google Scholar] [CrossRef]
  111. Hedayati, M.; Fouladvand, M.; Rouhollahi, A. Fabrication of WO3/Co3O4 nanorods as a p-n heterojunction photoanode for efficient photoelectrochemical oxygen evolution. Int. J. Hydrogen Energy 2024, 84, 288–295. [Google Scholar] [CrossRef]
  112. Lee, D.; Chae, M.; Ahmad, I.; Kim, J.-R.; Kim, H.-D. Influence of WO3-Based Antireflection Coatings on Current Density in Silicon Heterojunction Solar Cells. Nanomaterials 2023, 13, 1550. [Google Scholar] [CrossRef] [PubMed]
  113. Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48, 53–229. [Google Scholar] [CrossRef]
  114. Imanishi, A.; Tsuji, E.; Nakato, Y. Dependence of the Work Function of TiO2 (Rutile) on Crystal Faces, Studied by a Scanning Auger Microprobe. J. Phys. Chem. C 2007, 111, 2128–2132. [Google Scholar] [CrossRef]
  115. Kadoshima, M.; Hiratani, M.; Shimamoto, Y.; Torii, K.; Miki, H.; Kimura, S.; Nabatame, T. Rutile-type TiO2 thin film for high-k gate insulator. Thin Solid Film. 2003, 424, 224–228. [Google Scholar] [CrossRef]
  116. Dewasi, A.; Mitra, A. Effect of variation of thickness of TiO2 on the photovoltaic performance of n-TiO2/p-Si heterostructure. J. Mater. Sci. Mater. Electron. 2017, 28, 18075–18084. [Google Scholar] [CrossRef]
  117. Falson, J.; Kozuka, Y.; Uchida, M.; Smet, J.H.; Arima, T.-H.; Tsukazaki, A.; Kawasaki, M. MgZnO/ZnO heterostructures with electron mobility exceeding 1 × 106 cm2/Vs. Sci. Rep. 2016, 6, 26598. [Google Scholar] [CrossRef]
  118. Ozgür, Ü.; Alivov, Y.I.; Liu, C.; Teke, A.; Reshchikov, M.A.; Doğan, S.; Avrutin, V.; Cho, S.-J.; Morkoç, H. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 041301. [Google Scholar] [CrossRef]
  119. Look, D. Recent advances in ZnO materials and devices. Mater. Sci. Eng. B 2001, 80, 383–387. [Google Scholar] [CrossRef]
  120. Kim, S.K.; Lee, S.W.; Hwang, C.S.; Min, Y.-S.; Won, J.Y.; Jeong, J. Low Temperature (<100 °C) Deposition of Aluminum Oxide Thin Films by ALD with O3 as Oxidant. J. Electrochem. Soc. 2006, 153, F69–F76. [Google Scholar] [CrossRef]
  121. Liu, D.; Robertson, J. Oxygen vacancy levels and interfaces of Al2O3. Microelectron. Eng. 2009, 86, 1668–1671. [Google Scholar] [CrossRef]
  122. Costina, I.; Franchy, R. Band gap of amorphous and well-ordered Al2O3 on Ni3Al(100). Appl. Phys. Lett. 2001, 78, 4139–4141. [Google Scholar] [CrossRef]
  123. Guerreiro, A.N.; Costa, I.B.; Vale, A.B.; Braga, M.H. Distinctive Electric Properties of Group 14 Oxides: SiO2, SiO, and SnO2. Int. J. Mol. Sci. 2023, 24, 15985. [Google Scholar] [CrossRef] [PubMed]
  124. Li, Z.; Zhang, Z.; Tian, J.; Wu, G.; He, Y.; Yu, B.; Zhan, F.; Wang, Y.; Sun, M.; Yang, W.; et al. Efficient and Dense Electron Emission from a SiO2 Tunneling Diode with Low Poisoning Sensitivity. Nano Lett. 2022, 22, 1270–1277. [Google Scholar] [CrossRef] [PubMed]
  125. Kim, T.; Oh, C.; Park, S.H.; Lee, J.W.; Lee, S.I.; Kim, B.S. Electrical properties of low-temperature SiO2 thin films prepared by plasma-enhanced atomic layer deposition with different plasma times. AIP Adv. 2021, 11, 115126. [Google Scholar] [CrossRef]
  126. Kusmierek, E. A CeO2 Semiconductor as a Photocatalytic and Photoelectrocatalytic Material for the Remediation of Pollutants in Industrial Wastewater: A Review. Catalysts 2020, 10, 1435. [Google Scholar] [CrossRef]
  127. Wang, B.; Zhu, B.; Yun, S.; Zhang, W.; Xia, C.; Afzal, M.; Cai, Y.; Liu, Y.; Wang, Y.; Wang, H. Fast ionic conduction in semiconductor CeO2-δ electrolyte fuel cells. NPG Asia Mater. 2019, 11, 51. [Google Scholar] [CrossRef]
  128. Zhang, N.; Zhang, G.; Chong, S.; Zhao, H.; Huang, T.; Zhu, J. Ultrasonic impregnation of MnO2/CeO2 and its application in catalytic sono-degradation of methyl orange. J. Environ. Manag. 2018, 205, 134–141. [Google Scholar] [CrossRef] [PubMed]
  129. Avellaneda, C.O.; Berton, M.A.; Bulhões, L.O. Optical and electrochemical properties of CeO2 thin film prepared by an alkoxide route. Sol. Energy Mater. Sol. Cells 2008, 92, 240–244. [Google Scholar] [CrossRef]
  130. Erlebach, A.; Kurland, H.-D.; Grabow, J.; Müller, F.A.; Sierka, M. Structure evolution of nanoparticulate Fe2O3. Nanoscale 2014, 7, 2960–2969. [Google Scholar] [CrossRef]
  131. Cong-Zhou, L.; Wei-Pin, C. High dielectric constants of hematite (α-Fe2O3) and its anomalous characteristics. Ferroelectrics 1981, 38, 926. [Google Scholar] [CrossRef]
  132. Shaheen, W.; Hong, K. Thermal characterization and physicochemical properties of Fe2O3–Mn2O3/Al2O3 system. Thermochim. Acta 2002, 381, 153–164. [Google Scholar] [CrossRef]
  133. Mahmoud, H.R.; El-Molla, S.A.; Saif, M. Improvement of physicochemical properties of Fe2O3/MgO nanomaterials by hydrothermal treatment for dye removal from industrial wastewater. Powder Technol. 2013, 249, 225–233. [Google Scholar] [CrossRef]
  134. Walter, C.W.; Hertzler, C.F.; Devynck, P.; Smith, G.P.; Peterson, J.R. Photodetachment of WO3: The electron affinity of WO3. J. Chem. Phys. 1991, 95, 824–827. [Google Scholar] [CrossRef]
  135. Vuong, N.M.; Hieu, H.N.; Kim, D. An edge-contacted pn-heterojunction of a p-SWCNT/n-WO3 thin film. J. Mater. Chem. C 2013, 1, 5153–5160. [Google Scholar] [CrossRef]
  136. Sanchez, F.; Marot, L.; Dmitriev, A.; Antunes, R.; Steiner, R.; Meyer, E. WO3 work function enhancement induced by filamentous films deposited by resistive heating evaporation technique. J. Alloys Compd. 2023, 968, 171888. [Google Scholar] [CrossRef]
  137. Passas, I. Bibliometric Analysis: The Main Steps. Encyclopedia 2024, 4, 1014–1025. [Google Scholar] [CrossRef]
  138. Donthu, N.; Kumar, S.; Mukherjee, D.; Pandey, N.; Lim, W.M. How to conduct a bibliometric analysis: An overview and guidelines. J. Bus. Res. 2021, 133, 285–296. [Google Scholar] [CrossRef]
  139. Ellegaard, O.; Wallin, J.A. The bibliometric analysis of scholarly production: How great is the impact? Scientometrics 2015, 105, 1809–1831. [Google Scholar] [CrossRef]
  140. Suchikova, Y.; Nazarovets, S.; Popov, A.I. Electrochemical Etching vs. Electrochemical Deposition: A Comparative Bibliometric Analysis. Electrochem 2025, 6, 18. [Google Scholar] [CrossRef]
  141. Menéndez-Manjón, A.; Moldenhauer, K.; Wagener, P.; Barcikowski, S. Nano-energy research trends: Bibliometrical analysis of nanotechnology research in the energy sector. J. Nanoparticle Res. 2011, 13, 3911–3922. [Google Scholar] [CrossRef]
  142. Sales, M.B.; Neto, J.G.L.; Braz, A.K.D.S.; Junior, P.G.D.S.; Melo, R.L.F.; Valério, R.B.R.; Serpa, J.d.F.; Lima, A.M.D.S.; De Lima, R.K.C.; Guimarães, A.P.; et al. Trends and Opportunities in Enzyme Biosensors Coupled to Metal-Organic Frameworks (MOFs): An Advanced Bibliometric Analysis. Electrochem 2023, 4, 181–211. [Google Scholar] [CrossRef]
  143. Zhang, F.; Xu, W.; Deng, Z.; Huang, J. A bibliometric and visualization analysis of electrochemical biosensors for early diagnosis of eye diseases. Front. Med. 2025, 11, 1487981. [Google Scholar] [CrossRef]
  144. Dong, B.; Xu, G.; Luo, X.; Cai, Y.; Gao, W. A bibliometric analysis of solar power research from 1991 to 2010. Scientometrics 2012, 93, 1101–1117. [Google Scholar] [CrossRef]
  145. Obaideen, K.; Olabi, A.G.; Al Swailmeen, Y.; Shehata, N.; Abdelkareem, M.A.; Alami, A.H.; Rodriguez, C.; Sayed, E.T. Solar Energy: Applications, Trends Analysis, Bibliometric Analysis and Research Contribution to Sustainable Development Goals (SDGs). Sustainability 2023, 15, 1418. [Google Scholar] [CrossRef]
  146. Fauzi, M.A.; Abidin, N.H.Z.; Suki, N.M.; Budiea, A.M.A. Residential rooftop solar panel adoption behavior: Bibliometric analysis of the past and future trends. Renew. Energy Focus 2023, 45, 1–9. [Google Scholar] [CrossRef]
  147. Yousef, B.A.; Obaideen, K.; AlMallahi, M.N.; Alajmi, N.; Radwan, A.; Al-Shihabi, S.; Elgendi, M. On the contribution of concentrated solar power (CSP) to the sustainable development goals (SDGs): A bibliometric analysis. Energy Strat. Rev. 2024, 52, 101356. [Google Scholar] [CrossRef]
  148. David, T.M.; Rizol, P.M.S.R.; Machado, M.A.G.; Buccieri, G.P. Future research tendencies for solar energy management using a bibliometric analysis, 2000–2019. Heliyon 2020, 6, e04452. [Google Scholar] [CrossRef] [PubMed]
  149. Du, H.; Li, N.; Brown, M.A.; Peng, Y.; Shuai, Y. A bibliographic analysis of recent solar energy literatures: The expansion and evolution of a research field. Renew. Energy 2014, 66, 696–706. [Google Scholar] [CrossRef]
  150. Zwane, N.; Tazvinga, H.; Botai, C.; Murambadoro, M.; Botai, J.; de Wit, J.; Mabasa, B.; Daniel, S.; Mabhaudhi, T. A Bibliometric Analysis of Solar Energy Forecasting Studies in Africa. Energies 2022, 15, 5520. [Google Scholar] [CrossRef]
  151. Saikia, K.; Vallès, M.; Fabregat, A.; Saez, R.; Boer, D. A bibliometric analysis of trends in solar cooling technology. Sol. Energy 2020, 199, 100–114. [Google Scholar] [CrossRef]
  152. Mongeon, P.; Paul-Hus, A. The journal coverage of Web of Science and Scopus: A comparative analysis. Scientometrics 2016, 106, 213–228. [Google Scholar] [CrossRef]
  153. Archambault, É.; Campbell, D.; Gingras, Y.; Larivière, V. Comparing bibliometric statistics obtained from the Web of Science and Scopus. J. Am. Soc. Inf. Sci. Technol. 2009, 60, 1320–1326. [Google Scholar] [CrossRef]
  154. Franceschet, M. A comparison of bibliometric indicators for computer science scholars and journals on Web of Science and Google Scholar. Scientometrics 2009, 83, 243–258. [Google Scholar] [CrossRef]
  155. Arruda, H.; Silva, E.R.; Lessa, M.; Domício Proença, J.; Bartholo, R. VOSviewer and Bibliometrix. J. Med. Libr. Assoc. 2022, 110, 392–395. [Google Scholar] [CrossRef]
  156. Wong, D. VOSviewer. Tech. Serv. Q. 2018, 35, 219–220. [Google Scholar] [CrossRef]
  157. Bukar, U.A.; Sayeed, S.; Razak, S.F.A.; Yogarayan, S.; Amodu, O.A.; Mahmood, R.A.R. A method for analyzing text using VOSviewer. MethodsX 2023, 11, 102339. [Google Scholar] [CrossRef] [PubMed]
  158. Schmidt, M. The Sankey Diagram in Energy and Material Flow Management. J. Ind. Ecol. 2008, 12, 82–94. [Google Scholar] [CrossRef]
  159. Lupton, R.; Allwood, J. Hybrid Sankey diagrams: Visual analysis of multidimensional data for understanding resource use. Resour. Conserv. Recycl. 2017, 124, 141–151. [Google Scholar] [CrossRef]
  160. Daniel, D.; West-Mitchell, K. The Sankey diagram: An exploratory application of a data visualization tool. Transfusion 2024, 64, 967–968. [Google Scholar] [CrossRef]
  161. Muruganandham, M.; Zhang, Y.; Suri, R.; Lee, G.-J.; Chen, P.-K.; Hsieh, S.-H.; Sillanpyää, M.; Wu, J.J. Environmental Applications of ZnO Materials. J. Nanosci. Nanotechnol. 2015, 15, 6900–6913. [Google Scholar] [CrossRef]
  162. Gonçalves, R.A.; Toledo, R.P.; Joshi, N.; Berengue, O.M. Green Synthesis and Applications of ZnO and TiO2 Nanostructures. Molecules 2021, 26, 2236. [Google Scholar] [CrossRef] [PubMed]
  163. Šebesta, M.; Kolenčík, M.; Sunil, B.R.; Illa, R.; Mosnáček, J.; Ingle, A.P.; Urík, M. Field Application of ZnO and TiO2 Nanoparticles on Agricultural Plants. Agronomy 2021, 11, 2281. [Google Scholar] [CrossRef]
  164. Navidpour, A.H.; Hosseinzadeh, A.; Zhou, J.L.; Huang, Z. Progress in the application of surface engineering methods in immobilizing TiO2 and ZnO coatings for environmental photocatalysis. Catal. Rev. 2021, 65, 822–873. [Google Scholar] [CrossRef]
  165. Najaf, Z.; Nguyen, D.L.T.; Chae, S.Y.; Joo, O.-S.; Shah, A.U.H.A.; Vo, D.-V.N.; Nguyen, V.-H.; Van Le, Q.; Rahman, G. Recent trends in development of hematite (α-Fe2O3) as an efficient photoanode for enhancement of photoelectrochemical hydrogen production by solar water splitting. Int. J. Hydrogen Energy 2021, 46, 23334–23357. [Google Scholar] [CrossRef]
  166. Tahir, A.A.; Wijayantha, K.G.U.; Saremi-Yarahmadi, S.; Mazhar, M.; McKee, V. Nanostructured α-Fe2O3Thin Films for Photoelectrochemical Hydrogen Generation. Chem. Mater. 2009, 21, 3763–3772. [Google Scholar] [CrossRef]
  167. Emin, S.; de Respinis, M.; Mavrič, T.; Dam, B.; Valant, M.; Smith, W. Photoelectrochemical water splitting with porous α-Fe2O3 thin films prepared from Fe/Fe-oxide nanoparticles. Appl. Catal. A Gen. 2016, 523, 130–138. [Google Scholar] [CrossRef]
  168. Kalantarian, K.; Sheibani, S. Ag and carbon quantum dot-modified Fe2O3/g-C3N4 nanocomposites for efficient photocatalytic degradation of organic pollutants and hydrogen production. Int. J. Hydrogen Energy 2025, 140, 343–361. [Google Scholar] [CrossRef]
  169. Mohamed, H.O.; Obaid, M.; Poo, K.-M.; Abdelkareem, M.A.; Talas, S.A.; Fadali, O.A.; Kim, H.Y.; Chae, K.-J. Fe/Fe2O3 nanoparticles as anode catalyst for exclusive power generation and degradation of organic compounds using microbial fuel cell. Chem. Eng. J. 2018, 349, 800–807. [Google Scholar] [CrossRef]
  170. Ashraf, M.; Khan, I.; Usman, M.; Khan, A.; Shah, S.S.; Khan, A.Z.; Saeed, K.; Yaseen, M.; Ehsan, M.F.; Tahir, M.N.; et al. Hematite and Magnetite Nanostructures for Green and Sustainable Energy Harnessing and Environmental Pollution Control: A Review. Chem. Res. Toxicol. 2019, 33, 1292–1311. [Google Scholar] [CrossRef]
  171. Suchea, M.P.; Tudose, I.V.; Romanitan, C.; Pachiu, C.; Popescu, M.; Mouratis, K.; Manica, M.; Antohe, S.; Couris, S.; Zisopol, D.G.; et al. Study of evolution for 3D structured surface with nano-balls and walls-like features with thickness variation for WO3 thin films made by spray deposition. Sci. Rep. 2025, 15, 4275. [Google Scholar] [CrossRef]
  172. Phogat, P.; Rawat, S.; Thakur, J.; Jha, R.; Singh, S. Influence of metal ion doping on the photo-electrochemical detection performance of WO3. Sens. Actuators A Phys. 2024, 382, 116150. [Google Scholar] [CrossRef]
  173. Yao, Y.; Sang, D.; Zou, L.; Wang, Q.; Liu, C. A Review on the Properties and Applications of WO3 Nanostructure−Based Optical and Electronic Devices. Nanomaterials 2021, 11, 2136. [Google Scholar] [CrossRef]
  174. Buch, V.R.; Chawla, A.K.; Rawal, S.K. Review on electrochromic property for WO3 thin films using different deposition techniques. Mater. Today Proc. 2016, 3, 1429–1437. [Google Scholar] [CrossRef]
  175. Vezzoli, G.; Chen, M.; Caslavsky, J. New high dielectric constant materials: Micro/nanocomposites of suspended Au clusters in SiO2/SiO2, Al2O3, Li2O gels and relevant theory for high capacitance applications. Nanostruct. Mater. 1994, 4, 985–1009. [Google Scholar] [CrossRef]
  176. Na Kim, R.; Yun, H.W.; Lee, J.; Kim, W.-B. Dipole Formation and Electrical Properties According to SiO2 Layer Thickness at an Al2O3/SiO2 Interface. J. Phys. Chem. C 2021, 125, 14486–14492. [Google Scholar] [CrossRef]
  177. Kim, D.; Choi, J.; Ryu, S.W.; Kim, W. Improved Interface and Electrical Properties by Inserting an Ultrathin SiO2 Buffer Layer in the Al2O3/Si Heterojunction. Adv. Funct. Mater. 2019, 29, 1807271. [Google Scholar] [CrossRef]
  178. Kim, A.S.; Serko, N.A.; Khakuashev, P.E.; Kolky, A.N.; Yurchuck, S.Y. Application of Al2O3 Film for Stabilization of the Charge Properties of the SiO2/p-Si Interface. Russ. Microelectron. 2023, 52, 835–841. [Google Scholar] [CrossRef]
  179. Sen, M.T.S.K.A.; Bronsveld, P.; Weeber, A. Thermally stable MoOx hole selective contact with Al2O3 interlayer for industrial size silicon solar cells. Sol. Energy Mater. Sol. Cells 2021, 230, 111139. [Google Scholar] [CrossRef]
  180. Bullock, J.; Yan, D.; Cuevas, A. Passivation of aluminium–n+ silicon contacts for solar cells by ultrathin Al2O3 and SiO2 dielectric layers. Phys. Status Solidi (RRL)—Rapid Res. Lett. 2013, 7, 946–949. [Google Scholar] [CrossRef]
  181. Yang, Y.; He, J.; Wu, G.; Hu, J. “Thermal Stabilization Effect” of Al2O3 nano-dopants improves the high-temperature dielectric performance of polyimide. Sci. Rep. 2015, 5, 16986. [Google Scholar] [CrossRef] [PubMed]
  182. Bartzsch, H.; Glöß, D.; Böcher, B.; Frach, P.; Goedicke, K. Properties of SiO2 and Al2O3 films for electrical insulation applications deposited by reactive pulse magnetron sputtering. Surf. Coat. Technol. 2003, 174–175, 774–778. [Google Scholar] [CrossRef]
  183. Ren, Y.; Zhang, B.; Zhong, Z.; Ye, J.; Zhang, J.; Fang, Z.; Ye, F. Reusable, high specific surface areas, and excellent thermal stability Al2O3–SiO2 aerogel composites as high-temperature thermal insulators for radome applications. J. Alloys Compd. 2024, 984, 173990. [Google Scholar] [CrossRef]
  184. Gayathri, P.; Akila, T.; Sibu, G.A.; Selvaraj, M.; Assiri, M.A.; Matheswaran, P.; Balasubramani, V. Enhancing photovoltaic applications through precipitating agents in ITO/CIS/CeO2/Al heterojunction solar cell. Inorg. Chem. Commun. 2024, 168, 112866. [Google Scholar] [CrossRef]
  185. Aklalouch, M.; Calleja, A.; Granados, X.; Ricart, S.; Boffa, V.; Ricci, F.; Puig, T.; Obradors, X. Hybrid sol–gel layers containing CeO2 nanoparticles as UV-protection of plastic lenses for concentrated photovoltaics. Sol. Energy Mater. Sol. Cells 2014, 120, 175–182. [Google Scholar] [CrossRef]
  186. Li, Z.; Tong, L.; Ma, Y.; Zhao, L. Fabrication of CeO2/CaO nanocomposite as ultraviolet screening agent and its application in sunscreen. Ceram. Int. 2024, 50, 20431–20440. [Google Scholar] [CrossRef]
  187. Zhao, Y.; Gao, Z.; Lv, C.; Zhao, H.; Li, J.; Lv, Y. Synthesis and applications of polymorphic CeO2 and AgI/CeO2 nanocomposites. J. Mater. Sci. Mater. Eng. 2025, 20, 56. [Google Scholar] [CrossRef]
  188. Saha, S.; Keerthi, C.J.; Roy, S.; Sahatiya, P.; Dan, S.S.; Pal, S. Engineering Oxygen Vacancies in CeO2 for High-Performance UV-Vis Photodetector. IEEE Photon-Technol. Lett. 2023, 35, 1207–1210. [Google Scholar] [CrossRef]
  189. Semwal, N.; Mahar, D.; Arya, M.C. Response surface methodology-driven optimization of dye-sensitized Fe-doped CeO2 photocatalysts for efficient phenol degradation. J. Mol. Struct. 2025, 1341, 142639. [Google Scholar] [CrossRef]
  190. Qiu, G.; Yang, X.; Zhang, Y.; Zhao, J.; Liu, Q. Highly dispersed MoC quantum dots assist CeO2 for photocatalytic hydrogen production under visible light. Int. J. Hydrogen Energy 2025, 106, 423–431. [Google Scholar] [CrossRef]
  191. O’Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  192. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  193. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef]
  194. Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J.E.; et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, 591. [Google Scholar] [CrossRef]
  195. Nazeeruddin, M.K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.; Mueller, E.; Liska, P.; Vlachopoulos, N.; Graetzel, M. Conversion of light to electricity by cis-X2bis(2,2′-bipyridyl-4,4′-dicarboxylate)ruthenium(II) charge-transfer sensitizers (X = Cl-, Br-, I-, CN-, and SCN-) on nanocrystalline titanium dioxide electrodes. J. Am. Chem. Soc. 1993, 115, 6382–6390. [Google Scholar] [CrossRef]
  196. Law, M.; Greene, L.E.; Johnson, J.C.; Saykally, R.; Yang, P. Nanowire dye-sensitized solar cells. Nat. Mater. 2005, 4, 455–459. [Google Scholar] [CrossRef]
  197. Grätzel, M. Dye-sensitized solar cells. J. Photochem. Photobiol. C Photochem. Rev. 2003, 4, 145–153. [Google Scholar] [CrossRef]
  198. Liu, D.; Kelly, T.L. Perovskite solar cells with a planar heterojunction structure prepared using room-temperature solution processing techniques. Nat. Photonics 2013, 8, 133–138. [Google Scholar] [CrossRef]
  199. Liu, B.; Aydil, E.S. Growth of Oriented Single-Crystalline Rutile TiO2 Nanorods on Transparent Conducting Substrates for Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2009, 131, 3985–3990. [Google Scholar] [CrossRef] [PubMed]
  200. Kay, A.; Cesar, I.; Grätzel, M. New Benchmark for Water Photooxidation by Nanostructured α-Fe2O3 Films. J. Am. Chem. Soc. 2006, 128, 15714–15721. [Google Scholar] [CrossRef]
  201. Cushing, S.K.; Li, J.; Meng, F.; Senty, T.R.; Suri, S.; Zhi, M.; Li, M.; Bristow, A.D.; Wu, N. Photocatalytic Activity Enhanced by Plasmonic Resonant Energy Transfer from Metal to Semiconductor. J. Am. Chem. Soc. 2012, 134, 15033–15041. [Google Scholar] [CrossRef] [PubMed]
  202. Palomares, E.; Clifford, J.N.; Haque, S.A.; Lutz, T.; Durrant, J.R. Control of Charge Recombination Dynamics in Dye Sensitized Solar Cells by the Use of Conformally Deposited Metal Oxide Blocking Layers. J. Am. Chem. Soc. 2002, 125, 475–482. [Google Scholar] [CrossRef]
  203. Zou, S.; Liu, Y.; Li, J.; Liu, C.; Feng, R.; Jiang, F.; Li, Y.; Song, J.; Zeng, H.; Hong, M.; et al. Stabilizing Cesium Lead Halide Perovskite Lattice through Mn(II) Substitution for Air-Stable Light-Emitting Diodes. J. Am. Chem. Soc. 2017, 139, 11443–11450. [Google Scholar] [CrossRef]
  204. Aberle, A.G. Surface passivation of crystalline silicon solar cells: A review. Prog. Photovolt. Res. Appl. 2000, 8, 473–487. [Google Scholar] [CrossRef]
  205. Mor, G.K.; Varghese, O.K.; Paulose, M.; Shankar, K.; Grimes, C.A. A review on highly ordered, vertically oriented TiO2 nanotube arrays: Fabrication, material properties, and solar energy applications. Sol. Energy Mater. Sol. Cells 2006, 90, 2011–2075. [Google Scholar] [CrossRef]
  206. Huang, Z.; Geyer, N.; Werner, P.; de Boor, J.; Gösele, U. Metal-Assisted Chemical Etching of Silicon: A Review. Adv. Mater. 2010, 23, 285–308. [Google Scholar] [CrossRef]
  207. Malinkiewicz, O.; Yella, A.; Lee, Y.H.; Espallargas, G.M.; Graetzel, M.; Nazeeruddin, M.K.; Bolink, H.J. Perovskite solar cells employing organic charge-transport layers. Nat. Photonics 2013, 8, 128–132. [Google Scholar] [CrossRef]
  208. Niu, G.; Li, W.; Meng, F.; Wang, L.; Dong, H.; Qiu, Y. Study on the stability of CH3NH3PbI3films and the effect of post-modification by aluminum oxide in all-solid-state hybrid solar cells. J. Mater. Chem. A 2013, 2, 705–710. [Google Scholar] [CrossRef]
  209. Sivula, K.; Le Formal, F.; Grätzel, M. Solar Water Splitting: Progress Using Hematite (α-Fe2O3) Photoelectrodes. ChemSusChem 2011, 4, 432–449. [Google Scholar] [CrossRef] [PubMed]
  210. Osterloh, F.E. Inorganic nanostructures for photoelectrochemical and photocatalytic water splitting. Chem. Soc. Rev. 2012, 42, 2294–2320. [Google Scholar] [CrossRef] [PubMed]
  211. Kango, S.; Kalia, S.; Celli, A.; Njuguna, J.; Habibi, Y.; Kumar, R. Surface modification of inorganic nanoparticles for development of organic–inorganic nanocomposites—A review. Prog. Polym. Sci. 2013, 38, 1232–1261. [Google Scholar] [CrossRef]
  212. Park, J.H.; Kim, S.; Bard, A.J. Novel Carbon-Doped TiO2 Nanotube Arrays with High Aspect Ratios for Efficient Solar Water Splitting. Nano Lett. 2005, 6, 24–28. [Google Scholar] [CrossRef]
  213. Wang, C.-C.; Li, J.-R.; Lv, X.-L.; Zhang, Y.-Q.; Guo, G. Photocatalytic organic pollutants degradation in metal–organic frameworks. Energy Environ. Sci. 2014, 7, 2831–2867. [Google Scholar] [CrossRef]
  214. Bak, T.; Nowotny, J.; Rekas, M.; Sorrell, C.C. Photo-electrochemical hydrogen generation from water using solar energy. Materials-related aspects. Int. J. Hydrogen Energy 2002, 27, 991–1022. [Google Scholar] [CrossRef]
  215. Granqvist, C. Electrochromic tungsten oxide films: Review of progress 1993–1998. Sol. Energy Mater. Sol. Cells 2000, 60, 201–262. [Google Scholar] [CrossRef]
  216. Meyer, J.; Hamwi, S.; Kröger, M.; Kowalsky, W.; Riedl, T.; Kahn, A. Transition Metal Oxides for Organic Electronics: Energetics, Device Physics and Applications. Adv. Mater. 2012, 24, 5408–5427. [Google Scholar] [CrossRef]
  217. Baetens, R.; Jelle, B.P.; Gustavsen, A. Properties, requirements and possibilities of smart windows for dynamic daylight and solar energy control in buildings: A state-of-the-art review. Sol. Energy Mater. Sol. Cells 2010, 94, 87–105. [Google Scholar] [CrossRef]
  218. Liu, X.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.; Zhao, S.; Li, Z.; Lin, Z. Noble metal–metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2016, 10, 402–434. [Google Scholar] [CrossRef]
  219. Corma, A.; Atienzar, P.; García, H.; Chane-Ching, J.-Y. Hierarchically mesostructured doped CeO2 with potential for solar-cell use. Nat. Mater. 2004, 3, 394–397. [Google Scholar] [CrossRef]
  220. Ou, G.; Xu, Y.; Wen, B.; Lin, R.; Ge, B.; Tang, Y.; Liang, Y.; Yang, C.; Huang, K.; Zu, D.; et al. Tuning defects in oxides at room temperature by lithium reduction. Nat. Commun. 2018, 9, 1302. [Google Scholar] [CrossRef]
  221. Abanades, S.; Flamant, G. Thermochemical hydrogen production from a two-step solar-driven water-splitting cycle based on cerium oxides. Sol. Energy 2006, 80, 1611–1623. [Google Scholar] [CrossRef]
  222. Boyjoo, Y.; Sun, H.; Liu, J.; Pareek, V.K.; Wang, S. A review on photocatalysis for air treatment: From catalyst development to reactor design. Chem. Eng. J. 2017, 310, 537–559. [Google Scholar] [CrossRef]
  223. Talebi, H.; Rad, R.R.; Emami, F. Synergistic effects of SiO2 and Au nanostructures for enhanced broadband light absorption in perovskite solar cells. Sci. Rep. 2025, 15, 11548. [Google Scholar] [CrossRef]
  224. Dkhili, M.; Khalifa, M.; Aouida, S.; Ezzaouia, H.; Brown, T.M. Unveiling interface dynamics in perovskite solar cells with CeO2/SnO2 bilayer: Insights from electrochemical impedance spectroscopy. Mater. Today Commun. 2025, 46, 112786. [Google Scholar] [CrossRef]
  225. Haque, M.M.; Mahjabin, S.; Islam, M.A.; Selvanathan, V.; Putthisigamany, Y.; Abdullah, H.B.; Arith, F.; Akhtaruzzaman, M.; Ibrahim, M.A.; Chelvanathan, P. Modulation of optoelectronic properties of WO3 thin film via Cr doping through RF co-sputtering. Inorg. Chem. Commun. 2025, 177, 114300. [Google Scholar] [CrossRef]
  226. Liu, H.; Du, Y.; Yin, X.; Bai, M.; Liu, W. Micro/Nanostructures for Light Trapping in Monocrystalline Silicon Solar Cells. J. Nanomater. 2022, 2022, 8139174. [Google Scholar] [CrossRef]
  227. Saga, T. Advances in crystalline silicon solar cell technology for industrial mass production. NPG Asia Mater. 2010, 2, 96–102. [Google Scholar] [CrossRef]
  228. Green, M.A. Crystalline and thin-film silicon solar cells: State of the art and future potential. Sol. Energy 2003, 74, 181–192. [Google Scholar] [CrossRef]
  229. Rath, J. Low temperature polycrystalline silicon: A review on deposition, physical properties and solar cell applications. Sol. Energy Mater. Sol. Cells 2003, 76, 431–487. [Google Scholar] [CrossRef]
  230. Schropp, R.E.; Carius, R.; Beaucarne, G. Amorphous Silicon, Microcrystalline Silicon, and Thin-Film Polycrystalline Silicon Solar Cells. MRS Bull. 2007, 32, 219–224. [Google Scholar] [CrossRef]
  231. Zhang, H.; Wang, X.; Chen, X.; Zhang, Y. Research progress of compound-based dopant-free asymmetric heterogeneous contact silicon solar cell. Nano Energy 2025, 136, 110715. [Google Scholar] [CrossRef]
  232. Liang, B.; Chen, X.; Wang, X.; Yuan, H.; Sun, A.; Wang, Z.; Hu, L.; Hou, G.; Zhao, Y.; Zhang, X. Progress in crystalline silicon heterojunction solar cells. J. Mater. Chem. A 2024, 13, 2441–2477. [Google Scholar] [CrossRef]
  233. Kirchartz, T.; Yan, G.; Yuan, Y.; Patel, B.K.; Cahen, D.; Nayak, P.K. The state of the art in photovoltaic materials and device research. Nat. Rev. Mater. 2025, 10, 335–354. [Google Scholar] [CrossRef]
  234. Melskens, J.; Theeuwes, R.J.; Black, L.E.; Berghuis, W.-J.H.; Macco, B.; Bronsveld, P.C.P.; Kessels, W.M.M. Excellent Passivation of n-Type Silicon Surfaces Enabled by Pulsed-Flow Plasma-Enhanced Chemical Vapor Deposition of Phosphorus Oxide Capped by Aluminum Oxide. Phys. Status Solidi (RRL)—Rapid Res. Lett. 2020, 15, 2000399. [Google Scholar] [CrossRef]
  235. Banerjee, S.; Das, M.K. A review of Al2O3 as surface passivation material with relevant process technologies on c-Si solar cell. Opt. Quantum Electron. 2021, 53, 60. [Google Scholar] [CrossRef]
  236. Madani, K.; Rohatgi, A.; Min, K.H.; Song, H.-E.; Huang, Y.-Y.; Upadhyaya, A.D.; Upadhyaya, V.; Rounsaville, B.; Ok, Y.-W. Comparison of passivation properties of plasma-assisted ALD and APCVD deposited Al2O3 with SiNx capping. Sol. Energy Mater. Sol. Cells 2020, 218, 110718. [Google Scholar] [CrossRef]
  237. Grant, N.; Pain, S.; Khorani, E.; Jefferies, R.; Wratten, A.; McNab, S.; Walker, D.; Han, Y.; Beanland, R.; Bonilla, R.; et al. Activation of Al2O3 surface passivation of silicon: Separating bulk and surface effects. Appl. Surf. Sci. 2023, 645, 158786. [Google Scholar] [CrossRef]
  238. Berghuis, W.J.H.-J.; Melskens, J.; Macco, B.; Theeuwes, R.J.; Black, L.E.; Verheijen, M.A.; Kessels, W.M.M. Excellent surface passivation of germanium by a-Si:H/Al2O3 stacks. J. Appl. Phys. 2021, 130, 135303. [Google Scholar] [CrossRef]
  239. Ramos-Carrazco, A.; Berman-Mendoza, D.; Ramirez-Espinoza, R.; Gutierrez, R.G.; Vazquez-Arce, J.L.; Rangel, R.; Melendrez-Amavizca, R.; Bartolo-Pérez, P. Al2O3/Si NPs multilayered antireflective coating to enhance the photovoltaic performance of solar cells. J. Mater. Sci. Mater. Electron. 2023, 34, 2328. [Google Scholar] [CrossRef]
  240. Huang, Y.-C.; Chuang, R.W. Study on Annealing Process of Aluminum Oxide Passivation Layer for PERC Solar Cells. Coatings 2021, 11, 1052. [Google Scholar] [CrossRef]
  241. Uzum, A.; Kanmaz, I. Passivation properties of HfO2-SiO2 mixed metal oxide thin films with low reflectivity on silicon substrates for semiconductor devices. Thin Solid Film. 2021, 738, 138965. [Google Scholar] [CrossRef]
  242. Prishya, A.A.; Chopra, L. Comprehensive review on uses of silicon dioxide in solar cell. Mater. Today Proc. 2022, 72, 1471–1478. [Google Scholar] [CrossRef]
  243. Yang, Y.; Chen, D.; Li, D.; Zhao, T.; Zhang, W.; Qiao, W.; Liang, Y.; Yang, Y. Research on the Evolution of Defects Initiation and the Diffusion of Dopant on the Si/SiO2 Interface. Adv. Mater. Interfaces 2024, 12, 2400751. [Google Scholar] [CrossRef]
  244. van de Loo, B.; Knoops, H.; Dingemans, G.; Janssen, G.; Lamers, M.; Romijn, I.; Weeber, A.; Kessels, W. “Zero-charge” SiO2/Al2O3 stacks for the simultaneous passivation of n+ and p+ doped silicon surfaces by atomic layer deposition. Sol. Energy Mater. Sol. Cells 2015, 143, 450–456. [Google Scholar] [CrossRef]
  245. Hoex, B.; Schmidt, J.; Bock, R.; Altermatt, P.P.; van de Sanden, M.C.M.; Kessels, W.M.M. Excellent passivation of highly doped p-type Si surfaces by the negative-charge-dielectric Al2O3. Appl. Phys. Lett. 2007, 91, 112107. [Google Scholar] [CrossRef]
  246. Smirnov, M.; Roginskii, E.; Savin, A.; Mazhenov, N.; Pankin, D. Density-Functional Study of the Si/SiO2 Interfaces in Short-Period Superlattices: Structures and Energies. Coatings 2023, 13, 1231. [Google Scholar] [CrossRef]
  247. Glunz, S.W.; Steinhauser, B.; Polzin, J.; Luderer, C.; Grübel, B.; Niewelt, T.; Okasha, A.M.O.M.; Bories, M.; Nagel, H.; Krieg, K.; et al. Silicon-based passivating contacts: The TOPCon route. Prog. Photovolt. Res. Appl. 2021, 31, 341–359. [Google Scholar] [CrossRef]
  248. Yousuf, H.; Khokhar, M.Q.; Zahid, M.A.; Rabelo, M.; Kim, S.; Pham, D.P.; Kim, Y.; Yi, J. Tunnel Oxide Deposition Techniques and Their Parametric Influence on Nano-Scaled SiOx Layer of TOPCon Solar Cell: A Review. Energies 2022, 15, 5753. [Google Scholar] [CrossRef]
  249. Rahman, R.U.; Khokhar, M.Q.; Hussain, S.Q.; Mehmood, H.; Yousuf, H.; Jony, J.A.; Park, S.; Yi, J. Progress in TOPCon solar cell technology: Investigating hafnium oxide through simulation. Curr. Appl. Phys. 2024, 63, 96–104. [Google Scholar] [CrossRef]
  250. Li, S.; Liu, C.; Zhu, T.; Wang, Y.; He, J.; Yang, G.; Sun, P.; Li, H.; Liu, H.; Jiang, N. Effects of the thermal treatments on the optical properties of SiO2 anti-reflective coatings on sapphire windows. Infrared Phys. Technol. 2024, 137, 105151. [Google Scholar] [CrossRef]
  251. Jimenéz-Vivanco, M.R.; García, G.; Carrillo, J.; Agarwal, V.; Díaz-Becerril, T.; Doti, R.; Faubert, J.; Lugo, J.E. Porous Si-SiO2 based UV Microcavities. Sci. Rep. 2020, 10, 2220. [Google Scholar] [CrossRef]
  252. Pliskin, W.A.; Esch, R.P. Refractive Index of SiO2 Films Grown on Silicon. J. Appl. Phys. 1965, 36, 2011–2013. [Google Scholar] [CrossRef]
  253. Ali, K.; Khan, S.A.; Jafri, M.M. Effect of Double Layer (SiO2/TiO2) Anti-reflective Coating on Silicon Solar Cells. Int. J. Electrochem. Sci. 2014, 9, 7865–7874. [Google Scholar] [CrossRef]
  254. Rad, A.S.; Afshar, A.; Azadeh, M. Anti-reflection and self-cleaning meso-porous TiO2 coatings as solar systems protective layer: Investigation of effect of porosity and roughness. Opt. Mater. 2020, 107, 110027. [Google Scholar] [CrossRef]
  255. Richards, B.S. Comparison of TiO2and other dielectric coatings for buried-contact solar cells: A review. Prog. Photovolt. Res. Appl. 2004, 12, 253–281. [Google Scholar] [CrossRef]
  256. Elrashidi, A.; Elleithy, K. High-Efficiency Crystalline Silicon-Based Solar Cells Using Textured TiO2 Layer and Plasmonic Nanoparticles. Nanomaterials 2022, 12, 1589. [Google Scholar] [CrossRef]
  257. Richards, B.S.; Cotter, J.E.; Honsberg, C.B. Enhancing the surface passivation of TiO2 coated silicon wafers. Appl. Phys. Lett. 2002, 80, 1123–1125. [Google Scholar] [CrossRef]
  258. Yadav, C.; Kumar, S. Numerical Simulation for Optimization of Ultra-thin n-type AZO and TiO2 Based Textured p-type c-Si Heterojunction Solar Cells. Silicon 2021, 14, 4291–4299. [Google Scholar] [CrossRef]
  259. Lettieri, S.; Pavone, M.; Fioravanti, A.; Amato, L.S.; Maddalena, P. Charge Carrier Processes and Optical Properties in TiO2 and TiO2-Based Heterojunction Photocatalysts: A Review. Materials 2021, 14, 1645. [Google Scholar] [CrossRef] [PubMed]
  260. Singh, M.; Singh, D.; Mishra, V.; Singh, V.; Singh, S.; Dev, R.; Singh, D. Efficiency and environmental stability of TiO2 based solar cells for green electricity production. Int. J. Chem. React. Eng. 2023, 22, 69–77. [Google Scholar] [CrossRef]
  261. Park, G.S.; Lee, S.; Kim, D.; Park, S.Y.; Koh, J.H.; Won, D.H.; Lee, P.; Do, Y.R.; Min, B.K. Amorphous TiO2 Passivating Contacts for Cu(In,Ga)(S,Se)2 Ultrathin Solar Cells: Defect-State-Mediated Hole Conduction. Adv. Energy Mater. 2023, 13, 2203183. [Google Scholar] [CrossRef]
  262. Khir, H.; Pandey, A.; Saidur, R.; Ahmad, M.S.; Rahim, N.A.; Dewika, M.; Samykano, M. Recent advancements and challenges in flexible low temperature dye sensitised solar cells. Sustain. Energy Technol. Assess. 2022, 53, 102745. [Google Scholar] [CrossRef]
  263. Shukla, G.; Angappane, S. Dimensional constraints favour high temperature anatase phase stability in TiO2 nanorods. Appl. Surf. Sci. 2022, 577, 151874. [Google Scholar] [CrossRef]
  264. Cui, W.; Chen, F.; Li, Y.; Su, X.; Sun, B. Status and perspectives of transparent conductive oxide films for silicon heterojunction solar cells. Mater. Today Nano 2023, 22, 100329. [Google Scholar] [CrossRef]
  265. Dong, G.; Li, J.; Zhao, Y.; Ran, X.; Peng, C.; He, D.; Jin, C.; Wang, Q.; Jiang, H.; Zhang, Y.; et al. Highly efficient silicon heterojunction solar cells with ZnO:Al transparent electrode and transition metal doped indium oxide interfacial layer. Prog. Photovolt. Res. Appl. 2023, 31, 931–938. [Google Scholar] [CrossRef]
  266. Wang, J.; Meng, C.; Liu, H.; Hu, Y.; Zhao, L.; Wang, W.; Xu, X.; Zhang, Y.; Yan, H. Application of Indium Tin Oxide/Aluminum-Doped Zinc Oxide Transparent Conductive Oxide Stack Films in Silicon Heterojunction Solar Cells. ACS Appl. Energy Mater. 2021, 4, 13586–13592. [Google Scholar] [CrossRef]
  267. Kumari, P.; Srivastava, A.; Sharma, R.K.; Sharma, D.; Srivastava, S.K. Zinc Oxide: A Fascinating Material for Photovoltaic Applications. In Materials Horizons from Nature to Nanomaterials; Springer Nature: Singapore, 2022; pp. 173–241. [Google Scholar] [CrossRef]
  268. Üzar, N.; Abdulaziz, U. Investigation of the effects of coating numbers of thin films and metal contact type on physical properties of undoped ZnO, Fe-doped ZnO, and Fe–B co-doped ZnO thin films. J. Mater. Sci. Mater. Electron. 2024, 35, 1136. [Google Scholar] [CrossRef]
  269. Aberle, A.G. Overview on SiN surface passivation of crystalline silicon solar cells. Sol. Energy Mater. Sol. Cells 2001, 65, 239–248. [Google Scholar] [CrossRef]
  270. Tepner, S.; Lorenz, A. Printing technologies for silicon solar cell metallization: A comprehensive review. Prog. Photovolt. Res. Appl. 2023, 31, 557–590. [Google Scholar] [CrossRef]
  271. Suchikova, Y.; Kovachov, S.; Bohdanov, I.; Karipbayev, Z.T.; Zhydachevskyy, Y.; Lysak, A.; Pankratov, V.; Popov, A.I. Advanced Synthesis and Characterization of CdO/CdS/ZnO Heterostructures for Solar Energy Applications. Materials 2024, 17, 1566. [Google Scholar] [CrossRef]
  272. Suchikova, Y.; Kovachov, S.; Bohdanov, I.; Karipbaev, Z.T.; Pankratov, V.; Popov, A.I. Study of the structural and morphological characteristics of the CdxTeyOz nanocomposite obtained on the surface of the CdS/ZnO heterostructure by the SILAR method. Appl. Phys. A 2023, 129, 116625. [Google Scholar] [CrossRef]
  273. Balasubramani, V.; Akila, T.; Suresh, R.; Alodhay, A.N.; Muthuramamoorthy, M.; Sasikumar, K. Optimizing Oxygen Vacancy Concentrations in CeO2 Thin Films for Enhanced Photodetector Sensitivity. Opt. Mater. 2025, 159, 116625. [Google Scholar] [CrossRef]
  274. Ali, Z.; Ali, K. Modeling of improved efficiency and spectral response of a Si-based heterojunction solar cell by using CeO2 as a buffer layer. J. Comput. Electron. 2022, 21, 1320–1328. [Google Scholar] [CrossRef]
  275. Kanneboina, V. The simulated performance of c-Si/a-Si:H heterojunction solar cells with nc-Si:H, µc-Si:H, a-SiC:H, and a-SiGe:H emitter layers. J. Comput. Electron. 2020, 20, 344–352. [Google Scholar] [CrossRef]
  276. Vikhrov, S.P.; Vishnyakov, N.V.; Gudzev, V.V.; Ermachikhin, A.V.; Shilina, D.V.; Litvinov, V.G.; Maslov, A.D.; Mishustin, V.G.; Terukov, E.I.; Titov, A.S. Study of Deep Levels in a HIT Solar Cell. Semiconductors 2018, 52, 926–930. [Google Scholar] [CrossRef]
  277. Crovetto, A.; Yan, C.; Iandolo, B.; Zhou, F.; Stride, J.; Schou, J.; Hao, X.; Hansen, O. Lattice-matched Cu2ZnSnS4/CeO2 solar cell with open circuit voltage boost. Appl. Phys. Lett. 2016, 109, 233904. [Google Scholar] [CrossRef]
  278. Maniglia, R.; Reed, K.J.; Texter, J. Reactive CeO2 nanofluids for UV protective films. J. Colloid Interface Sci. 2017, 506, 346–354. [Google Scholar] [CrossRef]
  279. Caputo, F.; De Nicola, M.; Sienkiewicz, A.; Giovanetti, A.; Bejarano, I.; Licoccia, S.; Traversa, E.; Ghibelli, L. Cerium oxide nanoparticles, combining antioxidant and UV shielding properties, prevent UV-induced cell damage and mutagenesis. Nanoscale 2015, 7, 15643–15656. [Google Scholar] [CrossRef]
  280. Han, J.; Park, K.; Tan, S.; Vaynzof, Y.; Xue, J.; Diau, E.W.-G.; Bawendi, M.G.; Lee, J.-W.; Jeon, I. Perovskite solar cells. Nat. Rev. Methods Prim. 2025, 5, 3. [Google Scholar] [CrossRef]
  281. Prince, K.J.; Mirletz, H.M.; Gaulding, E.A.; Wheeler, L.M.; Kerner, R.A.; Zheng, X.; Schelhas, L.T.; Tracy, P.; Wolden, C.A.; Berry, J.J.; et al. Sustainability pathways for perovskite photovoltaics. Nat. Mater. 2024, 24, 22–33. [Google Scholar] [CrossRef]
  282. Elangovan, N.K.; Kannadasan, R.; Beenarani, B.; Alsharif, M.H.; Kim, M.-K.; Inamul, Z.H. Recent developments in perovskite materials, fabrication techniques, band gap engineering, and the stability of perovskite solar cells. Energy Rep. 2024, 11, 1171–1190. [Google Scholar] [CrossRef]
  283. Sewela, T.; Ocaya, R.O.; Malevu, T.D. Recent insights into the transformative role of Graphene-based/TiO2 electron transport layers for perovskite solar cells. Energy Sci. Eng. 2024, 13, 4–26. [Google Scholar] [CrossRef]
  284. Liao, Y.-H.; Chang, Y.-H.; Lin, T.-H.; Lee, K.-M.; Wu, M.-C. Recent Advances in Metal Oxide Electron Transport Layers for Enhancing the Performance of Perovskite Solar Cells. Materials 2024, 17, 2722. [Google Scholar] [CrossRef]
  285. Chen, X.; Yue, Z.; Yang, H.; Xu, B.; Cheng, Y. N-Type Self-Assembled Monolayers (SAMs): The Next Star Materials in the Perovskite Photovoltaic Field. Small 2025, 21, e2411312. [Google Scholar] [CrossRef] [PubMed]
  286. Samantaray, M.R.; Wang, Z.; Hu, D.; Yuan, M.; Song, H.; Li, F.; Jia, G.; Ji, L.; Zou, X.; Shen, H.; et al. Scalable Fabrication Methods of Large-Area (n-i-p) Perovskite Solar Panels. Sol. RRL 2024, 8, 2400235. [Google Scholar] [CrossRef]
  287. Kothandaraman, R.K.; Jiang, Y.; Feurer, T.; Tiwari, A.N.; Fu, F. Near-Infrared-Transparent Perovskite Solar Cells and Perovskite-Based Tandem Photovoltaics. Small Methods 2020, 4, 2000395. [Google Scholar] [CrossRef]
  288. Lin, L.; Jones, T.W.; Yang, T.C.; Duffy, N.W.; Li, J.; Zhao, L.; Chi, B.; Wang, X.; Wilson, G.J. Inorganic Electron Transport Materials in Perovskite Solar Cells. Adv. Funct. Mater. 2020, 31, 2008300. [Google Scholar] [CrossRef]
  289. Arora, I.; Chawla, H.; Chandra, A.; Sagadevan, S.; Garg, S. Advances in the strategies for enhancing the photocatalytic activity of TiO2: Conversion from UV-light active to visible-light active photocatalyst. Inorg. Chem. Commun. 2022, 143, 109700. [Google Scholar] [CrossRef]
  290. Gopinath, V.M.; Arulvel, S. A review on the steels, alloys/high entropy alloys, composites and coatings used in high temperature wear applications. Mater. Today Proc. 2021, 43, 817–823. [Google Scholar] [CrossRef]
  291. Manabeng, M.; Mwankemwa, B.S.; Ocaya, R.O.; Motaung, T.E.; Malevu, T.D. A Review of the Impact of Zinc Oxide Nanostructure Morphology on Perovskite Solar Cell Performance. Processes 2022, 10, 1803. [Google Scholar] [CrossRef]
  292. Abidin, N.A.Z.; Arith, F.; Noorasid, N.S.; Sarkawi, H.; Mustafa, A.N.; Safie, N.E.; Shah, A.S.M.; Azam, M.A.; Chelvanathan, P.; Amin, N. Dopant engineering for ZnO electron transport layer towards efficient perovskite solar cells. RSC Adv. 2023, 13, 33797–33819. [Google Scholar] [CrossRef]
  293. Apergi, S.; Brocks, G.; Tao, S.; Olthof, S. Probing the Reactivity of ZnO with Perovskite Precursors. ACS Appl. Mater. Interfaces 2024, 16, 14984–14994. [Google Scholar] [CrossRef] [PubMed]
  294. Znaidi, L. Sol–gel-deposited ZnO thin films: A review. Mater. Sci. Eng. B 2010, 174, 18–30. [Google Scholar] [CrossRef]
  295. Han, Y.; Guo, J.; Luo, Q.; Ma, C.-Q. Solution-Processable Zinc Oxide for Printed Photovoltaics: Progress, Challenges, and Prospect. Adv. Energy Sustain. Res. 2023, 4, 2200179. [Google Scholar] [CrossRef]
  296. Luo, J.; Wang, Y.; Zhang, Q. Progress in perovskite solar cells based on ZnO nanostructures. Sol. Energy 2018, 163, 289–306. [Google Scholar] [CrossRef]
  297. Qiu, C.; Wu, Y.; Song, J.; Wang, W.; Li, Z. Efficient Planar Perovskite Solar Cells with ZnO Electron Transport Layer. Coatings 2022, 12, 1981. [Google Scholar] [CrossRef]
  298. Otalora, C.; Botero, M.A.; Ordoñez, G. ZnO compact layers used in third-generation photovoltaic devices: A review. J. Mater. Sci. 2021, 56, 15538–15571. [Google Scholar] [CrossRef]
  299. Kim, S.Y.; Cho, S.J.; Byeon, S.E.; He, X.; Yoon, H.J. Self-Assembled Monolayers as Interface Engineering Nanomaterials in Perovskite Solar Cells. Adv. Energy Mater. 2020, 10, 2002606. [Google Scholar] [CrossRef]
  300. Sekar, K.; Doineau, R.; Mayarambakam, S.; Schmaltz, B.; Poulin-Vittrant, G. Control of ZnO nanowires growth in flexible perovskite solar cells: A mini-review. Heliyon 2024, 10, e24706. [Google Scholar] [CrossRef] [PubMed]
  301. Jiang, Q.; Zhang, X.; You, J. SnO2: A Wonderful Electron Transport Layer for Perovskite Solar Cells. Small 2018, 14, e1801154. [Google Scholar] [CrossRef]
  302. Deng, K.; Chen, Q.; Li, L. Modification Engineering in SnO2 Electron Transport Layer toward Perovskite Solar Cells: Efficiency and Stability. Adv. Funct. Mater. 2020, 30, 2004209. [Google Scholar] [CrossRef]
  303. Chen, Y.; Meng, Q.; Zhang, L.; Han, C.; Gao, H.; Zhang, Y.; Yan, H. SnO2-based electron transporting layer materials for perovskite solar cells: A review of recent progress. J. Energy Chem. 2019, 35, 144–167. [Google Scholar] [CrossRef]
  304. Ahmmed, S.; He, Y.; Kayesh, E.; Karim, A.; Matsuishi, K.; Islam, A. Ce-Doped SnO2 Electron Transport Layer for Minimizing Open Circuit Voltage Loss in Lead Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2024, 16, 32282–32290. [Google Scholar] [CrossRef]
  305. Lee, M.M.; Teuscher, J.; Miyasaka, T.; Murakami, T.N.; Snaith, H.J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643–647. [Google Scholar] [CrossRef]
  306. Noh, M.F.M.; Teh, C.H.; Daik, R.; Lim, E.L.; Yap, C.C.; Ibrahim, M.A.; Ludin, N.A.; Yusoff, A.R.b.M.; Jang, J.; Teridi, M.A.M. The architecture of the electron transport layer for a perovskite solar cell. J. Mater. Chem. C 2017, 6, 682–712. [Google Scholar] [CrossRef]
  307. Al-Shujaa, S.; Zhao, P.; He, D.; Al-Anesi, B.; Feng, Y.; Xia, J.; Zhang, B.; Zhang, Y. Improving the Efficiency and Stability of Perovskite Solar Cells by Refining the Perovskite-Electron Transport Layer Interface and Shielding the Absorber from UV Effects. ACS Appl. Mater. Interfaces 2024, 16, 28493–28504. [Google Scholar] [CrossRef]
  308. Bhattarai, S.; Borah, D.; Rout, J.; Pandey, R.; Madan, J.; Hossain, I.; Handique, P.; Ansari, M.Z.; Hossain, M.K.; Rahman, F. Designing an efficient lead-free perovskite solar cell with green-synthesized CuCrO2 and CeO2 as carrier transport materials. RSC Adv. 2023, 13, 34693–34702. [Google Scholar] [CrossRef]
  309. Hu, T.; Xiao, S.; Yang, H.; Chen, L.; Chen, Y. Cerium oxide as an efficient electron extraction layer for p–i–n structured perovskite solar cells. Chem. Commun. 2017, 54, 471–474. [Google Scholar] [CrossRef]
  310. Chen, W.; Luo, Q.; Zhang, C.; Shi, J.; Deng, X.; Yue, L.; Wang, Z.; Chen, X.; Huang, S. Effects of down-conversion CeO2:Eu3+ nanophosphors in perovskite solar cells. J. Mater. Sci. Mater. Electron. 2017, 28, 11346–11357. [Google Scholar] [CrossRef]
  311. Abdulrahman, A.; Basem, A.; Algarni, Z.; Soliman, N.F.; Amari, A.; Saydullaev, B.; Abduvokhidov, A.; Khudoynazarov, E.; Mahariq, I. Sustainable Synthesis of a Novel α-Fe2O3/ZIF-8/PANI Composite Electrodes for High-Energy Hybrid Supercapacitors. J. Alloys Compd. 2025, 1040, 183422. [Google Scholar] [CrossRef]
  312. Dissanayake, D.M.S.N.; Mantilaka, M.M.M.G.P.G.; Palihawadana, T.C.; Chandrakumara, G.T.D.; De Silva, R.T.; Pitawala, H.M.T.G.A.; de Silva, K.M.N.; Amaratunga, G.A.J. Facile and low-cost synthesis of pure hematite (α-Fe2O3) nanoparticles from naturally occurring laterites and their superior adsorption capability towards acid-dyes. RSC Adv. 2019, 9, 21249–21257. [Google Scholar] [CrossRef] [PubMed]
  313. Qureshi, A.A.; Javed, S.; Akram, M.A.; Schmidt-Mende, L.; Fakharuddin, A. Solvent-Assisted Crystallization of an α-Fe2O3 Electron Transport Layer for Efficient and Stable Perovskite Solar Cells Featuring Negligible Hysteresis. ACS Omega 2023, 8, 18106–18115. [Google Scholar] [CrossRef] [PubMed]
  314. Piccinin, S. The band structure and optical absorption of hematite (α-Fe2O3): A first-principles GW-BSE study. Phys. Chem. Chem. Phys. 2019, 21, 2957–2967. [Google Scholar] [CrossRef]
  315. Fan, X.; Zhu, F.; Wang, Z.; Wang, X.; Zou, Y.; Gao, B.; Song, L.; He, J.; Wang, T. Preparation of p-type Fe2O3 nanoarray and its performance as photocathode for photoelectrochemical water splitting. Front. Chem. 2025, 13, 1526745. [Google Scholar] [CrossRef]
  316. Rahman, A.U.; El Astal-Quirós, A.; Susanna, G.; Javanbakht, H.; Calabrò, E.; Polino, G.; Paci, B.; Generosi, A.; Riva, F.R.; Brunetti, F.; et al. Scaling-Up of Solution-Processable Tungsten Trioxide (WO3) Nanoparticles as a Hole Transport Layer in Inverted Organic Photovoltaics. Energies 2024, 17, 814. [Google Scholar] [CrossRef]
  317. Ávila-López, A.; Cruz, J.C.; Díaz-Real, J.A.; García-Uitz, K.; Cante-Góngora, D.; Rodríguez-May, G. A Review of Perovskite-Based Solar Cells over the Last Decade: The Evolution of the Hole Transport Layer and the Use of WO3 as an Electron Transport Layer. Coatings 2025, 15, 132. [Google Scholar] [CrossRef]
  318. Haque, M.M.; Mahjabin, S.; Abdullah, H.B.; Akhtaruzzaman, M.; Almohamadi, H.; Islam, M.A.; Hossain, M.I.; Ibrahim, M.A.; Chelvanathan, P. Exploring the Theoretical Potential of Tungsten Oxide (WOx) as a Universal Electron Transport Layer (ETL) for Various Perovskite Solar Cells through Interfacial Energy Band Alignment Modulation. J. Phys. Chem. Solids 2024, 196, 112324. [Google Scholar] [CrossRef]
  319. Roy, A.; Bhandari, S.; Ghosh, A.; Sundaram, S.; Mallick, T.K. Incorporating Solution-Processed Mesoporous WO3 as an Interfacial Cathode Buffer Layer for Photovoltaic Applications. J. Phys. Chem. A 2020, 124, 5709–5719. [Google Scholar] [CrossRef]
  320. Sengupta, D.; Das, P.; Mondal, B.; Mukherjee, K. Effects of doping, morphology and film-thickness of photo-anode materials for dye sensitized solar cell application—A review. Renew. Sustain. Energy Rev. 2016, 60, 356–376. [Google Scholar] [CrossRef]
  321. Grätzel, M. Mesoporous oxide junctions and nanostructured solar cells. Curr. Opin. Colloid Interface Sci. 1999, 4, 314–321. [Google Scholar] [CrossRef]
  322. Zhang, Y.; Zhang, B.; Peng, X.; Liu, L.; Dong, S.; Lin, L.; Chen, S.; Meng, S.; Feng, Y. Preparation of dye-sensitized solar cells with high photocurrent and photovoltage by using mesoporous titanium dioxide particles as photoanode material. Nano Res. 2015, 8, 3830–3841. [Google Scholar] [CrossRef]
  323. Di Carlo, G.; Calogero, G.; Brucale, M.; Caschera, D.; de Caro, T.; Di Marco, G.; Ingo, G.M. Insights into meso-structured photoanodes based on titanium oxide thin film with high dye adsorption ability. J. Alloys Compd. 2014, 609, 116–124. [Google Scholar] [CrossRef]
  324. Agarwala, S.; Kevin, M.; Wong, A.S.W.; Peh, C.K.N.; Thavasi, V.; Ho, G.W. Mesophase Ordering of TiO2 Film with High Surface Area and Strong Light Harvesting for Dye-Sensitized Solar Cell. ACS Appl. Mater. Interfaces 2010, 2, 1844–1850. [Google Scholar] [CrossRef]
  325. Li, W.; Elzatahry, A.; Aldhayan, D.; Zhao, D. Core–shell structured titanium dioxide nanomaterials for solar energy utilization. Chem. Soc. Rev. 2018, 47, 8203–8237. [Google Scholar] [CrossRef] [PubMed]
  326. Korir, B.K.; Kibet, J.K.; Ngari, S.M. A review on the current status of dye-sensitized solar cells: Toward sustainable energy. Energy Sci. Eng. 2024, 12, 3188–3226. [Google Scholar] [CrossRef]
  327. Kumar, D. A short review on the advancement in the development of TiO2 and ZnO based photo-anodes for the application of Dye-Sensitized Solar Cells (DSSCs). Eng. Res. Express 2021, 3, 042004. [Google Scholar] [CrossRef]
  328. Clarke, B.; Ghandi, K. The Interplay of Growth Mechanism and Properties of ZnO Nanostructures for Different Applications. Small 2023, 19, e2302864. [Google Scholar] [CrossRef] [PubMed]
  329. Ali, J.; Bibi, S.; Jatoi, W.B.; Tuzen, M.; Jakhrani, M.A.; Feng, X.; Saleh, T.A. Green synthesized zinc oxide nanostructures and their applications in dye-sensitized solar cells and photocatalysis: A review. Mater. Today Commun. 2023, 36, 106840. [Google Scholar] [CrossRef]
  330. Vittal, R.; Ho, K.-C. Zinc oxide based dye-sensitized solar cells: A review. Renew. Sustain. Energy Rev. 2017, 70, 920–935. [Google Scholar] [CrossRef]
  331. Boro, B.; Gogoi, B.; Rajbongshi, B.; Ramchiary, A. Nano-structured TiO2/ZnO nanocomposite for dye-sensitized solar cells application: A review. Renew. Sustain. Energy Rev. 2018, 81, 2264–2270. [Google Scholar] [CrossRef]
  332. Kumar, R.; Umar, A.; Kumar, G.; Nalwa, H.S.; Kumar, A.; Akhtar, M.S. Zinc oxide nanostructure-based dye-sensitized solar cells. J. Mater. Sci. 2017, 52, 4743–4795. [Google Scholar] [CrossRef]
  333. Abdellah, I.M. Molecular engineering and electrolyte optimization strategies for enhanced performance of Ru(ii) polypyridyl-sensitized DSSCs. RSC Adv. 2025, 15, 9763–9786. [Google Scholar] [CrossRef]
  334. Bouclé, J.; Ackermann, J. Solid-state dye-sensitized and bulk heterojunction solar cells using TiO2 and ZnO nanostructures: Recent progress and new concepts at the borderline. Polym. Int. 2011, 61, 355–373. [Google Scholar] [CrossRef]
  335. Prabavathy, N.; Shalini, S.; Balasundaraprabhu, R.; Velauthapillai, D.; Prasanna, S.; Muthukumarasamy, N. Enhancement in the photostability of natural dyes for dye-sensitized solar cell (DSSC) applications: A review. Int. J. Energy Res. 2017, 41, 1372–1396. [Google Scholar] [CrossRef]
  336. Zheng, H.; Tachibana, Y.; Kalantar-Zadeh, K. Dye-Sensitized Solar Cells Based on WO3. Langmuir 2010, 26, 19148–19152. [Google Scholar] [CrossRef]
  337. Chen, W.; Qiu, Y.; Zhong, Y.; Wong, K.S.; Yang, S. High-Efficiency Dye-Sensitized Solar Cells Based on the Composite Photoanodes of SnO2 Nanoparticles/ZnO Nanotetrapods. J. Phys. Chem. A 2010, 114, 3127–3138. [Google Scholar] [CrossRef]
  338. Memar, A.; Phan, C.M.; Tade, M.O. Photocatalytic activity of WO3/Fe2O3 nanocomposite photoanode. Int. J. Hydrogen Energy 2015, 40, 8642–8649. [Google Scholar] [CrossRef]
  339. Younas, M.; Gondal, M.; Dastageer, M.; Baig, U. Fabrication of cost effective and efficient dye sensitized solar cells with WO3-TiO2 nanocomposites as photoanode and MWCNT as Pt-free counter electrode. Ceram. Int. 2019, 45, 936–947. [Google Scholar] [CrossRef]
  340. Kim, M.-H.; Kwon, Y.-U. Semiconducting Divalent Metal Oxides as Blocking Layer Material for SnO2-Based Dye-Sensitized Solar Cells. J. Phys. Chem. C 2011, 115, 23120–23125. [Google Scholar] [CrossRef]
  341. Shahpari, M.; Behjat, A.; Khajaminian, M.; Torabi, N. The influence of morphology of hematite (α-Fe2O3) counter electrodes on the efficiency of dye-sensitized solar cells. Sol. Energy 2015, 119, 45–53. [Google Scholar] [CrossRef]
  342. Suh, D. Status of Al2O3/TiO2-Based Antireflection and Surface Passivation for Silicon Solar Cells. Phys. Status Solidi (RRL)—Rapid Res. Lett. 2021, 15, 2100236. [Google Scholar] [CrossRef]
  343. Lee, M.-J.; Park, J.-Y.; Kim, C.-S.; Okuyama, K.; Lee, S.-E.; Kim, T.-O. Improvement of light scattering capacity in dye-sensitized solar cells by doping with SiO2 nanoparticles. J. Power Sources 2016, 327, 96–103. [Google Scholar] [CrossRef]
  344. Ramanujam, J.; Bishop, D.M.; Todorov, T.K.; Gunawan, O.; Rath, J.; Nekovei, R.; Artegiani, E.; Romeo, A. Flexible CIGS, CdTe and a-Si:H based thin film solar cells: A review. Prog. Mater. Sci. 2020, 110, 100619. [Google Scholar] [CrossRef]
  345. Luque, A.; Hegedus, S. (Eds.) Handbook of Photovoltaic Science and Engineering; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2011. [Google Scholar] [CrossRef]
  346. Vijayan, K.; Vijayachamundeeswari, S.; Sivaperuman, K.; Ahsan, N.; Logu, T.; Okada, Y. A review on advancements, challenges, and prospective of copper and non-copper based thin-film solar cells using facile spray pyrolysis technique. Sol. Energy 2022, 234, 81–102. [Google Scholar] [CrossRef]
  347. Ishizuka, S.; Sakurai, K.; Yamada, A.; Matsubara, K.; Fons, P.; Iwata, K.; Nakamura, S.; Kimura, Y.; Baba, T.; Nakanishi, H.; et al. Fabrication of wide-gap Cu(In1−xGax)Se2 thin film solar cells: A study on the correlation of cell performance with highly resistive i-ZnO layer thickness. Sol. Energy Mater. Sol. Cells 2005, 87, 541–548. [Google Scholar] [CrossRef]
  348. Efaz, E.T.; Rhaman, M.; Al Imam, S.; Bashar, K.L.; Kabir, F.; Mourtaza, E.; Sakib, S.N.; Mozahid, F.A. A review of major technologies of thin-film solar cells. Eng. Res. Express 2021, 3, 032001. [Google Scholar] [CrossRef]
  349. Romeo, A.; Artegiani, E. CdTe-Based Thin Film Solar Cells: Past, Present and Future. Energies 2021, 14, 1684. [Google Scholar] [CrossRef]
  350. Salhi, B. The Photovoltaic Cell Based on CIGS: Principles and Technologies. Materials 2022, 15, 1908. [Google Scholar] [CrossRef]
  351. Rahman, M.A. Enhancing the photovoltaic performance of Cd-free Cu2ZnSnS4 heterojunction solar cells using SnS HTL and TiO2 ETL. Sol. Energy 2021, 215, 64–76. [Google Scholar] [CrossRef]
  352. Qin, D.; Yang, P.; Pan, Y.; Wang, Y.; Pan, Y.; Weng, G.; Hu, X.; Tao, J.; Chu, J.; Akiyama, H.; et al. Non-Cadmium TiO2/Sb2(Se, S)3 Heterojunction Solar Cells with Improved Efficiency by NaCl-Treated Interface Engineering. ACS Appl. Mater. Interfaces 2025, 17, 22050–22059. [Google Scholar] [CrossRef] [PubMed]
  353. Werner, F.; Veith-Wolf, B.; Spindler, C.; Barget, M.R.; Babbe, F.; Guillot, J.; Schmidt, J.; Siebentritt, S. Oxidation as Key Mechanism for Efficient Interface Passivation in Cu(In,Ga)Se2 Thin-Film Solar Cells. Phys. Rev. Appl. 2020, 13, 054004. [Google Scholar] [CrossRef]
  354. Cunha, J.M.V.; Fernandes, P.A.; Hultqvist, A.; Teixeira, J.P.; Bose, S.; Vermang, B.; Garud, S.; Buldu, D.; Gaspar, J.; Edoff, M.; et al. Insulator Materials for Interface Passivation of Cu(In,Ga)Se2Thin Films. IEEE J. Photovolt. 2018, 8, 1313–1319. [Google Scholar] [CrossRef]
  355. Abbas, K.J.; Bahrami, A. Investigating the potential of WO3 and WS2 as Cd-free buffer layers in Sb2Se3-Based thin-film solar cells: A numerical study with SCAPS-1D software. Sol. Energy Mater. Sol. Cells 2024, 272, 112891. [Google Scholar] [CrossRef]
  356. Ouyang, D.; Huang, Z.; Choy, W.C.H. Solution-Processed Metal Oxide Nanocrystals as Carrier Transport Layers in Organic and Perovskite Solar Cells. Adv. Funct. Mater. 2018, 29, 1804660. [Google Scholar] [CrossRef]
  357. Concina, I.; Vomiero, A. Metal Oxide Semiconductors for Dye- and Quantum-Dot-Sensitized Solar Cells. Small 2014, 11, 1744–1774. [Google Scholar] [CrossRef]
  358. Pandi, K.; Kumar, T.R.N.; Lakhera, S.K.; Neppolian, B. Simultaneous Passivation of Surface Vacancies and Enhancement in Charge Transfer Property of ZnO Electron Transport Layer for Inverted Organic Solar Cells. Energy Technol. 2020, 8, 2000481. [Google Scholar] [CrossRef]
  359. Chaudhary, D.K.; Dhawan, P.K.; Patel, S.P.; Bhasker, H. Large area semitransparent inverted organic solar cells with enhanced operational stability using TiO2 electron transport layer for building integrated photovoltaic devices. Mater. Lett. 2021, 283, 128725. [Google Scholar] [CrossRef]
  360. Bivour, M.; Temmler, J.; Steinkemper, H.; Hermle, M. Molybdenum and tungsten oxide: High work function wide band gap contact materials for hole selective contacts of silicon solar cells. Sol. Energy Mater. Sol. Cells 2015, 142, 34–41. [Google Scholar] [CrossRef]
  361. Tan, Z.; Li, L.; Cui, C.; Ding, Y.; Xu, Q.; Li, S.; Qian, D.; Li, Y. Solution-Processed Tungsten Oxide as an Effective Anode Buffer Layer for High-Performance Polymer Solar Cells. J. Phys. Chem. C 2012, 116, 18626–18632. [Google Scholar] [CrossRef]
  362. United Nations Framework Convention on Climate Change (UNFCCC). Paris Agreement; UNFCCC: Bonn, Germany, 2015; Available online: https://unfccc.int/sites/default/files/english_paris_agreement.pdf (accessed on 5 September 2025).
  363. European Union. Directive (EU) 2018/2001 of the European Parliament and of the Council of 11 December 2018 on the Pro-motion of the Use of Energy from Renewable Sources (Recast). Official Journal of the European Union L 328 (21 December 2018), 82–209. Available online: https://eur-lex.europa.eu/legal-content/en/ALL/?uri=CELEX:32018L2001 (accessed on 5 September 2025).
  364. International Energy Agency (IEA). Solar PV Feed-in Tariff (China)—IEA Policies Database, 2011 (Notice No. 1594). Available online: https://www.iea.org/policies/5100-solar-pv-feed-in-tariff (accessed on 5 September 2025).
  365. U.S. Department of Energy (DOE). Recovery Act (American Recovery and Reinvestment Act of 2009). Available online: https://www.energy.gov/recovery-act (accessed on 5 September 2025).
  366. Alternative Fuels Data Center (AFDC). American Recovery and Reinvestment Act of 2009 (Public Law 111-5)—Energy-Related Provisions. Available online: https://afdc.energy.gov/laws/arra.html (accessed on 5 September 2025).
  367. International Energy Agency (IEA). Solar PV—Analysis (Overview of Global Trends and Market Dynamics). IEA: Paris, France, 2024–2025 (Living Page). Available online: https://www.iea.org/energy-system/renewables/solar-pv (accessed on 5 September 2025).
  368. IEA Photovoltaic Power Systems Programme (IEA-PVPS). A Snapshot of Global PV—1992–2016, 5th ed.; IEA-PVPS: Paris, France, 2017; Available online: https://iea-pvps.org/wp-content/uploads/2020/01/IEA-PVPS_-_A_Snapshot_of_Global_PV_-_1992-2016__1_.pdf (accessed on 5 September 2025).
  369. Suchikova, Y.; Kovachov, S.; Bohdanov, I.; Konuhova, M.; Zhydachevskyy, Y.; Kumarbekov, K.; Pankratov, V.; Popov, A.I. Wet Chemical Synthesis of AlxGa1−xAs Nanostructures: Investigation of Properties and Growth Mechanisms. Crystals 2024, 14, 633. [Google Scholar] [CrossRef]
  370. Suchikova, Y.; Kovachov, S.; Bohdanov, I.; Popova, E.; Moskina, A.; Popov, A. Characterization of CdxTeyOz/CdS/ZnO Heterostructures Synthesized by the SILAR Method. Coatings 2023, 13, 639. [Google Scholar] [CrossRef]
  371. Kovachov, S.; Bohdanov, I.; Karipbayev, Z.; Suchikova, Y.; Tsebriienko, T.; Popov, A.I. Layer-by-Layer Synthesis and Analysis of the the Phase Composition of Cdx TeyOz/CdS/por-ZnO/ZnO Heterostructure. In Proceedings of the 2022 IEEE 3rd KhPI Week on Advanced Technology (KhPIWeek), Kharkiv, Ukraine, 3–7 October 2022; pp. 1–6. [Google Scholar] [CrossRef]
  372. Xu, P.; Wang, H.; Ren, L.; Tu, B.; Wang, W.; Fu, Z. Theoretical study on composition-dependent properties of ZnO·nAl2O3 spinels. Part I: Optical and dielectric. J. Am. Ceram. Soc. 2021, 104, 5099–5109. [Google Scholar] [CrossRef]
  373. Aryal, S.; Rulis, P.; Ching, W.; Green, D.J. Mechanical Properties and Electronic Structure of Mullite Phases Using First-Principles Modeling. J. Am. Ceram. Soc. 2012, 95, 2075–2088. [Google Scholar] [CrossRef]
  374. Edan, T.S.; Haider, A.J.; Sultan, F.I. Core–Shell Nanostructures in Self-Cleaning Surface Technology: A Mini Review. Plasmonics 2025, 1–20. [Google Scholar] [CrossRef]
  375. Arulkumar, E.; Manivannan, R.K.; Dhamodaran, G.; Krishnan, R. CuO@α-Fe2O3 nanocomposite via one-pot synthesis for multifunctional photocatalytic, optoelectronic and antimicrobial applications. J. Sol-Gel Sci. Technol. 2025, 1–33. [Google Scholar] [CrossRef]
  376. de Freitas Mendes, M.V.; Pedroti, L.G.; de Carvalho, J.M.F.; Ferreira, F.A.; Mendes, B.C.; Xavier, L.M.; Fernandes, W.E.H.; Oliveira, N.M.A. Characterization of Yellow and Red Clays and Bauxite Residue: Analysis of Properties and Potential for Ceramic Products. In Characterization of Minerals, Metals, and Materials 2025; Peng, Z., Xie, K.Y., Zhang, M., Li, J., Li, B., Monteiro, S.N., Soman, R., Hwang, J.-Y., Kalay, Y.E., Escobedo-Diaz, J.P., et al., Eds.; TMS 2025. The Minerals, Metals & Materials Series; Springer: Cham, Switzerland, 2025. [Google Scholar] [CrossRef]
  377. European Chemicals Agency (ECHA). Hazardous to the Aquatic Environment (Long-Term). ECHA Registration Dossier; ECHA: Helsinki, Finland, [Online]. Available online: https://echa.europa.eu/registration-dossier/-/registered-dossier/16139/2/1 (accessed on 7 September 2025).
  378. RMIS (Raw Materials Information System—Tungsten): European Commission, Joint Research Centre (JRC). RMIS—Raw Materials Information System: Tungsten. European Commission: Brussels, Belgium, [Online]. Available online: https://rmis.jrc.ec.europa.eu/rmp/Tungsten (accessed on 7 September 2025).
  379. European Commission, Directorate-General for Internal Market, Industry, Entrepreneurship and SMEs (DG GROW). Critical Raw Materials. European Commission: Brussels, Belgium, [Online]. Available online: https://single-market-economy.ec.europa.eu/sectors/raw-materials/areas-specific-interest/critical-raw-materials_en (accessed on 7 September 2025).
  380. Jungbluth, N. Life cycle assessment of crystalline photovoltaics in the Swiss ecoinvent database. Prog. Photovolt. Res. Appl. 2005, 13, 429–446. [Google Scholar] [CrossRef]
Figure 1. Design of the Bibliometric Analysis.
Figure 1. Design of the Bibliometric Analysis.
Ceramics 08 00119 g001
Figure 2. Dynamics of publication numbers for selected binary oxides in the context of solar energy applications (Web of Science Core Collection data, 1970–2024).
Figure 2. Dynamics of publication numbers for selected binary oxides in the context of solar energy applications (Web of Science Core Collection data, 1970–2024).
Ceramics 08 00119 g002
Figure 3. Sankey diagram illustrating the distribution of research focus among binary oxide ceramics across four disciplines: Energy, Materials Science, Environmental Science, and Physics.
Figure 3. Sankey diagram illustrating the distribution of research focus among binary oxide ceramics across four disciplines: Energy, Materials Science, Environmental Science, and Physics.
Ceramics 08 00119 g003
Figure 4. Bibliometric analysis in VOSviewer for TiO2: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/245z4qkv, https://tinyurl.com/29ecu8nc (accessed on 7 September 2025).
Figure 4. Bibliometric analysis in VOSviewer for TiO2: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/245z4qkv, https://tinyurl.com/29ecu8nc (accessed on 7 September 2025).
Ceramics 08 00119 g004
Figure 5. Bibliometric analysis in VOSviewer for ZnO: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/2cjolpkj, https://tinyurl.com/29k59moj (accessed on 7 September 2025).
Figure 5. Bibliometric analysis in VOSviewer for ZnO: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/2cjolpkj, https://tinyurl.com/29k59moj (accessed on 7 September 2025).
Ceramics 08 00119 g005
Figure 6. Bibliometric analysis in VOSviewer for SiO2: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/23ktygy6, https://tinyurl.com/224hqtyt (accessed on 7 September 2025).
Figure 6. Bibliometric analysis in VOSviewer for SiO2: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/23ktygy6, https://tinyurl.com/224hqtyt (accessed on 7 September 2025).
Ceramics 08 00119 g006
Figure 7. Bibliometric analysis in VOSviewer for Al2O3: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/2dceubry, https://tinyurl.com/22qy7qzs (accessed on 7 September 2025).
Figure 7. Bibliometric analysis in VOSviewer for Al2O3: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/2dceubry, https://tinyurl.com/22qy7qzs (accessed on 7 September 2025).
Ceramics 08 00119 g007
Figure 8. Bibliometric analysis in VOSviewer for CeO2: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/25ba3qky, https://tinyurl.com/26wxuydq (accessed on 7 September 2025).
Figure 8. Bibliometric analysis in VOSviewer for CeO2: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/25ba3qky, https://tinyurl.com/26wxuydq (accessed on 7 September 2025).
Ceramics 08 00119 g008
Figure 9. Bibliometric analysis in VOSviewer for Fe2O3: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/26xze96n, https://tinyurl.com/28kszl94 (accessed on 7 September 2025).
Figure 9. Bibliometric analysis in VOSviewer for Fe2O3: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/26xze96n, https://tinyurl.com/28kszl94 (accessed on 7 September 2025).
Ceramics 08 00119 g009
Figure 10. Bibliometric analysis in VOSviewer for WO3: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/22smuu6m, https://tinyurl.com/26s56kfs (accessed on 7 September 2025).
Figure 10. Bibliometric analysis in VOSviewer for WO3: (a) visualization of international collaboration; (b) visualization of keyword clustering. The interactive visualization is available at https://tinyurl.com/22smuu6m, https://tinyurl.com/26s56kfs (accessed on 7 September 2025).
Ceramics 08 00119 g010
Figure 11. Major types of solar cells.
Figure 11. Major types of solar cells.
Ceramics 08 00119 g011
Figure 12. Matrix of Oxide Ceramic Applications in Solar Cells.
Figure 12. Matrix of Oxide Ceramic Applications in Solar Cells.
Ceramics 08 00119 g012
Table 2. Description of Statistical Indicators.
Table 2. Description of Statistical Indicators.
Statistical IndicatorDescription
MedianThe central value of an ordered dataset, less affected by outliers.
MeanThe arithmetic average, indicating the overall level of research activity.
Standard DeviationA measure of data variability relative to the mean.
Coefficient of Variation (CV)The standard deviation, expressed as a percentage of the mean, reflecting relative variability.
Maximum ValueThe highest recorded number of publications.
First Quartile (Q1)The value below which 25% of the data fall, representing the lower range of activity.
Third Quartile (Q3)The value below which 75% of the data fall, representing the upper range of activity.
Interquartile Range (IQR)The range containing the central 50% of the data, enabling assessment of variability without outliers.
Table 3. Number of publications identified for each oxide material in the context of solar energy applications (Web of Science Core Collection).
Table 3. Number of publications identified for each oxide material in the context of solar energy applications (Web of Science Core Collection).
MaterialsResults from WoS CC
TiO2 (titanium dioxide)22,898
ZnO (zinc oxide)19,092
SiO2 (silicon dioxide)4140
Al2O3 (aluminum oxide)3268
Fe2O3 (iron oxide)2633
WO3 (tungsten trioxide)2062
CeO2 (cerium dioxide)491
Table 4. Statistical analysis of publications on binary oxide ceramics *.
Table 4. Statistical analysis of publications on binary oxide ceramics *.
MaterialMedianStandard DeviationCoefficient of Variation (%)MeanMaximum1st Quartile (Q1)3rd Quartile (Q3)Interquartile Range (IQR)
TiO2119.0652.44125.37520.41166513.751180.01166.25
ZnO126.5529.57116.50454.57142215.75977.25961.5
SiO225.0110.43117.3394.112995.5231.5226.0
Al2O320.0100.99108.1693.372626.5213.5207.0
Fe2O39.098.82131.3675.232712.5147.0144.5
WO316.071.78118.3660.651924.25119.5115.25
CeO25.019.42114.7216.93632.033.031.0
* The mean values are calculated over active years only, i.e., years in which at least one publication on the given oxide was recorded. Years with zero publications were excluded from the denominator in order to provide a meaningful average for periods of active research.
Table 5. Top 5 Most Cited Publications on Binary Oxide Ceramics *.
Table 5. Top 5 Most Cited Publications on Binary Oxide Ceramics *.
TiO2 (Titanium Dioxide)Citations
O’Regan, B., & Grätzel, M. (1991). A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature, 353(6346), 737–740. https://doi.org/10.1038/353737a0 [191]25,829
Kojima, A., Teshima, K., Shirai, Y., & Miyasaka, T. (2009). Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. Journal of the American Chemical Society, 131(17), 6050–6051. https://doi.org/10.1021/ja809598r [192]18,169
Grätzel, M. (2001). Photoelectrochemical cells. Nature, 414(6861), 338–344. https://doi.org/10.1038/35104607 [193]11,772
Kim, H.-S., Lee, C.-R., Im, J.-H., Lee, K.-B., Moehl, T., Marchioro, A., et al. (2012). Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Scientific Reports, 2(1), 591. https://doi.org/10.1038/srep00591 [194]7149
Nazeeruddin, M. K., Kay, A., Rodicio, I., Humphry-Baker, R., Mueller, E., Liska, P., et al. (1993). Conversion of light to electricity by cis-X2bis(2,2′-bipyridyl-4,4′-dicarboxylate)ruthenium(II) charge-transfer sensitizers (X = Cl-, Br-, I-, CN-, and SCN-) on nanocrystalline titanium dioxide electrodes. Journal of the American Chemical Society, 115(14), 6382–6390. https://doi.org/10.1021/ja00067a063 [195]5854
ZnO (zinc oxide)Citations
Grätzel, M. (2001). Photoelectrochemical cells. Nature, 414(6861), 338–344. https://doi.org/10.1038/35104607 [193]11,772
Law, M., Greene, L. E., Johnson, J. C., Saykally, R., & Yang, P. (2005). Nanowire dye-sensitized solar cells. Nature Materials, 4(6), 455–459. https://doi.org/10.1038/nmat1387 [196]5135
Grätzel, M. (2003). Dye-sensitized solar cells. Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 4(2), 145–153. https://doi.org/10.1016/S1389-5567(03)00026-1 [197]4640
Liu, D., & Kelly, T. L. (2014). Perovskite solar cells with a planar heterojunction structure prepared using room-temperature solution processing techniques. Nature Photonics, 8(2), 133–138. https://doi.org/10.1038/nphoton.2013.342 [198]2383
Liu, B., & Aydil, E. S. (2009). Growth of Oriented Single-Crystalline Rutile TiO 2 Nanorods on Transparent Conducting Substrates for Dye-Sensitized Solar Cells. Journal of the American Chemical Society, 131(11), 3985–3990. https://doi.org/10.1021/ja8078972 [199]2215
SiO2 (silicon dioxide)Citations
Kay, A., Cesar, I., & Grätzel, M. (2006). New Benchmark for Water Photooxidation by Nanostructured α-Fe 2 O 3 Films. Journal of the American Chemical Society, 128(49), 15714–15721. https://doi.org/10.1021/ja064380l [200]1437
Cushing, S. K., Li, J., Meng, F., Senty, T. R., Suri, S., Zhi, M., et al. (2012). Photocatalytic Activity Enhanced by Plasmonic Resonant Energy Transfer from Metal to Semiconductor. Journal of the American Chemical Society, 134(36), 15033–15041. https://doi.org/10.1021/ja305603t [201]1032
Palomares, E., Clifford, J. N., Haque, S. A., Lutz, T., & Durrant, J. R. (2003). Control of Charge Recombination Dynamics in Dye Sensitized Solar Cells by the Use of Conformally Deposited Metal Oxide Blocking Layers. Journal of the American Chemical Society, 125(2), 475–482. https://doi.org/10.1021/ja027945w [202]1029
Zou, S., Liu, Y., Li, J., Liu, C., Feng, R., Jiang, F., et al. (2017). Stabilizing Cesium Lead Halide Perovskite Lattice through Mn(II) Substitution for Air-Stable Light-Emitting Diodes. Journal of the American Chemical Society, 139(33), 11443–11450. https://doi.org/10.1021/jacs.7b04000 [203]728
Aberle, A. G. (2000). Surface passivation of crystalline silicon solar cells: a review. Progress in Photovoltaics: Research and Applications, 8(5), 473–487. https://doi.org/10.1002/1099-159X(200009/10)8:5%3C473::AID-PIP337%3E3.0.CO;2-D [204]625
Al2O3 (aluminum oxide)Citations
Palomares, E., Clifford, J. N., Haque, S. A., Lutz, T., & Durrant, J. R. (2003). Control of Charge Recombination Dynamics in Dye Sensitized Solar Cells by the Use of Conformally Deposited Metal Oxide Blocking Layers. Journal of the American Chemical Society, 125(2), 475–482. https://doi.org/10.1021/ja027945w [202]1886
Mor, G. K., Varghese, O. K., Paulose, M., Shankar, K., & Grimes, C. A. (2006). A review on highly ordered, vertically oriented TiO2 nanotube arrays: Fabrication, material properties, and solar energy applications. Solar Energy Materials and Solar Cells, 90(14), 2011–2075. https://doi.org/10.1016/j.solmat.2006.04.007 [205]1605
Huang, Z., Geyer, N., Werner, P., de Boor, J., & Gösele, U. (2011). Metal-Assisted Chemical Etching of Silicon: A Review. Advanced Materials, 23(2), 285–308. https://doi.org/10.1002/adma.201001784 [206]1285
Malinkiewicz, O., Yella, A., Lee, Y. H., Espallargas, G. M., Graetzel, M., Nazeeruddin, M. K., & Bolink, H. J. (2014). Perovskite solar cells employing organic charge-transport layers. Nature Photonics, 8(2), 128–132. https://doi.org/10.1038/nphoton.2013.341 [207]1029
Niu, G., Li, W., Meng, F., Wang, L., Dong, H., & Qiu, Y. (2014). Study on the stability of CH 3 NH 3 PbI 3 films and the effect of post-modification by aluminum oxide in all-solid-state hybrid solar cells. J. Mater. Chem. A, 2(3), 705–710. https://doi.org/10.1039/C3TA13606J [208]955
Fe2O3 (iron oxide)Citations
Sivula, K., Le Formal, F., & Grätzel, M. (2011). Solar Water Splitting: Progress Using Hematite (α-Fe2O3) Photoelectrodes. ChemSusChem, 4(4), 432–449. https://doi.org/10.1002/cssc.201000416 [209]2332
Osterloh, F. E. (2013). Inorganic nanostructures for photoelectrochemical and photocatalytic water splitting. Chem. Soc. Rev., 42(6), 2294–2320. https://doi.org/10.1039/C2CS35266D [210]1776
Kango, S., Kalia, S., Celli, A., Njuguna, J., Habibi, Y., & Kumar, R. (2013). Surface modification of inorganic nanoparticles for development of organic–inorganic nanocomposites—A review. Progress in Polymer Science, 38(8), 1232–1261. https://doi.org/10.1016/j.progpolymsci.2013.02.003 [211]1685
Park, J. H., Kim, S., & Bard, A. J. (2006). Novel Carbon-Doped TiO 2 Nanotube Arrays with High Aspect Ratios for Efficient Solar Water Splitting. Nano Letters, 6(1), 24–28. https://doi.org/10.1021/nl051807y [212]1647
Wang, C.-C., Li, J.-R., Lv, X.-L., Zhang, Y.-Q., & Guo, G. (2014). Photocatalytic organic pollutants degradation in metal–organic frameworks. Energy Environ. Sci., 7(9), 2831–2867. https://doi.org/10.1039/C4EE01299B [213]1444
WO3 (tungsten trioxide)Citations
Park, J. H., Kim, S., & Bard, A. J. (2006). Novel Carbon-Doped TiO 2 Nanotube Arrays with High Aspect Ratios for Efficient Solar Water Splitting. Nano Letters, 6(1), 24–28. https://doi.org/10.1021/nl051807y [212]1647
Bak, T., Nowotny, J., Rekas, M., & Sorrell, C. (2002). Photo-electrochemical hydrogen generation from water using solar energy. Materials-related aspects. International Journal of Hydrogen Energy, 27(10), 991–1022. https://doi.org/10.1016/S0360-3199(02)00022-8 [214]1346
Granqvist, C. (2000). Electrochromic tungsten oxide films: Review of progress 1993–1998. Solar Energy Materials and Solar Cells, 60(3), 201–262. https://doi.org/10.1016/S0927-0248(99)00088-4 [215]1324
Meyer, J., Hamwi, S., Kröger, M., Kowalsky, W., Riedl, T., & Kahn, A. (2012). Transition Metal Oxides for Organic Electronics: Energetics, Device Physics and Applications. Advanced Materials, 24(40), 5408–5427. https://doi.org/10.1002/adma.201201630 [216]1049
Baetens, R., Jelle, B. P., & Gustavsen, A. (2010). Properties, requirements and possibilities of smart windows for dynamic daylight and solar energy control in buildings: A state-of-the-art review. Solar Energy Materials and Solar Cells, 94(2), 87–105. https://doi.org/10.1016/j.solmat.2009.08.021 [217]1047
CeO2 (cerium dioxide)Citations
Liu, X., Iocozzia, J., Wang, Y., Cui, X., Chen, Y., Zhao, S., et al. (2017). Noble metal–metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy & Environmental Science, 10(2), 402–434. https://doi.org/10.1039/C6EE02265K [218]832
Corma, A., Atienzar, P., García, H., & Chane-Ching, J.-Y. (2004). Hierarchically mesostructured doped CeO2 with potential for solar-cell use. Nature Materials, 3(6), 394–397. https://doi.org/10.1038/nmat1129 [219]728
Ou, G., Xu, Y., Wen, B., Lin, R., Ge, B., Tang, Y., et al. (2018). Tuning defects in oxides at room temperature by lithium reduction. Nature Communications, 9(1), 1302. https://doi.org/10.1038/s41467-018-03765-0 [220]502
Abanades, S., & Flamant, G. (2006). Thermochemical hydrogen production from a two-step solar-driven water-splitting cycle based on cerium oxides. Solar Energy, 80(12), 1611–1623. https://doi.org/10.1016/j.solener.2005.12.005 [221]491
Boyjoo, Y., Sun, H., Liu, J., Pareek, V. K., & Wang, S. (2017). A review on photocatalysis for air treatment: From catalyst development to reactor design. Chemical Engineering Journal, 310(2, SI), 537–559. https://doi.org/10.1016/j.cej.2016.06.090 [222]435
* Some publications appear in multiple oxide lists because they reference or use several materials simultaneously. The selection was based on keywords and abstracts, so even if the main focus of the publication is not exclusively on the respective oxide, its role in the research context is confirmed.
Table 6. Distribution of Top 5 Publications by Year for Each Oxide.
Table 6. Distribution of Top 5 Publications by Year for Each Oxide.
YearTiO2ZnOSiO2Al2O3Fe2O3WO3CeO2
1991O’Regan, B., & Grätzel, M.
[191]
(25,829)
1993Nazeeruddin et al.
[195]
(5854)
2000--Aberle
[204]
(2000)
--Granqvist
[215]
(1324)
-
2001Grätzel
[193]
(11,772)
Grätzel
[193]
(11,772)
2002-----Bak et al.
[214]
(1346)
-
2003 Grätzel
[197]
(4640)
Palomares et al.
[202]
(1029)
Palomares et al.
[202]
(1886)
2004------Corma et al.
[219]
(728)
2005Law et al.
[196]
(5135)
2006Kay et al.
[200] (1437)
Mor et al.
[205]
(1605)
Park et al.
[212]
(1647)
Park et al.
[212]
(1647)
Abanades et al.
[221]
(491)
2009Kojima et al.
[192]
(18,169)
Liu, & Aydil,
[199]
(2215)
2010-----Baetens et al.
[217]
(1047)
2011Huang et al.
[206]
(1285)
Sivula et al.
[209]
(2332)
2012Kim et al.
[194]
(7149)
Cushing et al.
[201]
(1032)
Meyer et al.
[216]
(1049)
2013Osterloh
[210]
(1776), Kango
[211]
(1685)
2014Liu & Kelly
[198]
(2383)
Malinkiewicz
[207]
(1029),
Niu et al.
[208]
(955)
Wang et al.
[213]
(1444)
2017Zou et al.
[203]
(728)
Liu et al.
[218]
(832),
Boyjoo et al.
[222]
(435)
2018Ou et al.
[220]
(502)
Table 7. Comparison of key oxide materials used in crystalline silicon (c-Si) solar cells.
Table 7. Comparison of key oxide materials used in crystalline silicon (c-Si) solar cells.
Oxide MaterialMain Functions in c-Si CellsAdvantagesDisadvantages and Limitations
Al2O3Surface passivation, insulationExcellent chemical passivation, negative charge (effective field-effect passivation), high thermal stability, chemical inertnessDoes not conduct electrons (only a passivation layer), often requires an additional protective layer (e.g., SiNx)
SiO2Surface passivation, antireflection coatings, tunneling layerExceptional chemical passivation, very stable interface with Si, excellent antireflection properties (low refractive index, ~1.45)Lack of effective field-effect passivation (neutral/weakly positive charge), primarily used as a passive layer
TiO2Antireflection coatings, surface passivation of n-SiHigh transparency, high refractive index (2.0–2.5), good chemical stability, low costUnintended shunting due to n-type conductivity, photocatalytic activity under UV, requires precise film quality control
ZnOTransparent conductive oxide (TCO), buffer layer in heterojunctionsHigh transparency, high conductivity when doped (e.g., AZO), low cost, application flexibilityLower chemical stability, sensitive to humidity and acidic environments, requires additional passivation
CeO2Experimental window layer, UV filter, potential passivation layerGood chemical stability, UV absorption capability (Si protection), promising CeO2/n-Si heterojunctionLow electron mobility, further research needed for deposition optimization
Table 8. Comparison of key characteristics of oxide materials used in perovskite solar cells.
Table 8. Comparison of key characteristics of oxide materials used in perovskite solar cells.
Oxide MaterialLayer TypeMain Advantages and FunctionsDrawbacks and Technological Features
TiO2ETLHigh transparency; favorable band alignment with perovskite; thermal stability; efficiency > 20%Photocatalytic activity (UV-induced perovskite degradation); requires high-temperature processing (>450 °C)
ZnOETLHigh electron mobility; low-temperature deposition; compatible with solution-based methodsChemical instability in contact with MA+-based perovskites; requires interfacial protection or surface modification
CeO2ETLUV absorption; chemical inertness; interface passivation; potential for enhanced stabilityLower electron mobility; difficulty in forming high-quality films without thermal treatment
Fe2O3ETL, also studied as experimental absorberLow cost; environmental friendliness; high resistance to UV and moistureLow electron mobility; high charge recombination; partial visible light absorption; lower efficiency (~13%)
WO3HTL/ETLHigh work function (HTL); resistance to moisture and temperature; solution-processable; UV protectionSuboptimal band alignment when used as ETL; property variation depending on stoichiometry level
Table 9. Characteristics of Oxide Materials for DSSC.
Table 9. Characteristics of Oxide Materials for DSSC.
Oxide MaterialRole in DSSCAdvantagesLimitations
TiO2PhotoanodeIdeal energy alignment with dyes; high chemical stability; large surface area for dye adsorptionSlow electron transport; recombination with oxidized electrolyte species
ZnOPhotoanodeHigh electron mobility; easy nanostructuring (nanorods, nanoparticles); low-temperature depositionChemical instability in the presence of some dyes (especially acidic); risk of defect formation
WO3Photoanode/AdditiveUV absorption; chemical stability; electron conductivityLess favorable energy alignment; high recombination; low efficiency
Fe2O3 (Hematite)Experimental PhotoanodeVisible light absorption; non-toxicity; UV stabilityVery short hole diffusion length (~2–4 nm); intense recombination; low photovoltage
Al2O3Passivating BarrierDefect passivation; reduced recombination; increased V_OCInsulator—does not conduct electrons; requires precise thickness control
SiO2Optical Additive/BarrierEnhanced light scattering; structural stabilization; chemical inertnessNon-conductive; indirect effect via morphology and optics
Table 10. Role of Oxides in Thin-Film Chalcogenide and Inorganic Solar Cells.
Table 10. Role of Oxides in Thin-Film Chalcogenide and Inorganic Solar Cells.
Oxide MaterialRole in the DeviceAdvantagesLimitations or Application Conditions
ZnOTransparent contact, buffer, textured layerHigh transparency, good conductivity when doped, texturing capabilityMay require protection during deposition, vulnerable to acids
TiO2Buffer layer, grain boundary passivationCd-free replacement for CdS, visible-range transparency, thermal stabilityRequires interface control due to risk of recombination
Al2O3Passivating layer, dielectric barrierReduces recombination, improves V_OC, used in nanopatterned structuresInsulator—does not conduct charge, precise thickness critical
SiO2Dielectric layer, diffusion barrierOptical transparency, thermal stability, interlayer diffusion barrierDoes not contribute to charge transport, auxiliary function
WO3Back contact buffer (CdTe, CIGS)High work function, transparency, improved hole extractionRequires thin deposition (a few nm), critical energy level alignment
CeO2Experimental buffer/window layer between absorber; surface passivation; UV barrierWide band gap, high transparency; chemical inertness; Cd-free; UV absorption and surface recombination reductionLow electron mobility increases series resistance; electrical properties sensitive to oxygen vacancies; requires optimized deposition methods (ALD, solution processes) and post-treatment; efficiency demonstrated only on lab-scale samples
Table 11. Oxide Materials in Organic, Quantum Dot, and “All-Oxide” Solar Cells: Functions, Advantages, and Technological Limitations.
Table 11. Oxide Materials in Organic, Quantum Dot, and “All-Oxide” Solar Cells: Functions, Advantages, and Technological Limitations.
Oxide MaterialRole in the DeviceAdvantagesLimitations or Application Conditions
ZnOETL (OSC, QD)High transparency; solution-processable; high electron mobility; chemical stabilityGenerates reactive radicals under UV; requires surface modification or encapsulation
TiO2ETL (OSC, QD); contact in Cu2O cellsWide band gap; stability; solution-processableLow electron mobility; interface quality is critical
WO3HTL (OSC); rear contact (QD)High work function; transparency; thermal stability; UV protectionLower work function than MoO3; sensitive to stoichiometry and thickness
Fe2O3Absorber (experimental)Low cost; non-toxic; stableRequires cascade/tandem architecture; limited spectral absorption; low carrier mobility; low efficiency
CeO2ETL or protective interlayer/UV filter in all-oxide cellsUV absorption prevents degradation of the active layer; chemically inert; compatible with low-temperature depositionLow electron mobility; properties sensitive to oxygen vacancies; large-scale solution processing not yet optimized
Al2O3Inert encapsulation, passivating/optical spacerReduces surface recombination; stabilizes morphology; chemically/thermally inert; may enhance V_OCDoes not conduct charge; thickness must be <3 nm to avoid adding series resistance
SiO2Anti-reflective front (AR) coating or dielectric stabilizing barrierLow refractive index (~1.45) reduces reflection; barrier to oxygen/moisture diffusion; low-T compatibleNot charge-selective; effect is purely optical/encapsulation-related, requiring careful integration with ETL/HTL
Table 12. Applications of Key Binary Oxides in Different Types of Solar Cells.
Table 12. Applications of Key Binary Oxides in Different Types of Solar Cells.
Oxide Materialc-SiPSCDSSCThin-Film (CIGS, CdTe, CZTS, a-Si:H)OSC, QD, All-Oxide
TiO2Passivation, anti-reflectionETL, barrier, mesoporous scaffoldPhotoanode (ETL)Buffer, grain boundary passivationETL, contact with Cu2O
ZnOTCO, bufferETLPhotoanode (ETL)TCO, buffer, textured layerETL, all-oxide component
Al2O3Passivation, dielectricPassivation, inert insulatorBarrier, passivation of TiO2Passivation, dielectric barrierOptical spacer, inert interlayer
SiO2Anti-reflective, tunnel layerAnti-reflective, optical layerOptical additive, light scattererDiffusion barrier, optical stabilizationDielectric, optical substrate
CeO2UV protection, passivationETL, UV filter, stabilizationPotential passivation, buffer (experimental)ETL, absorber, protective layer
Fe2O3Absorber (experimental)Photoanode (low efficiency)Absorber in all-oxide architectures
WO3HTL, ETL (investigated)Photoanode/additiveRear buffer contact (HTL)HTL, absorber, rear contact
ETL—electron transport layer; HTL—hole transport layer; TCO—transparent conductive oxide; Barrier/buffer—for energetic or chemical alignment; Passivation—reduction in recombination; Absorber—light-absorbing layer.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Suchikova, Y.; Nazarovets, S.; Konuhova, M.; Popov, A.I. Binary Oxide Ceramics (TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3) for Solar Cell Applications: A Comparative and Bibliometric Analysis. Ceramics 2025, 8, 119. https://doi.org/10.3390/ceramics8040119

AMA Style

Suchikova Y, Nazarovets S, Konuhova M, Popov AI. Binary Oxide Ceramics (TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3) for Solar Cell Applications: A Comparative and Bibliometric Analysis. Ceramics. 2025; 8(4):119. https://doi.org/10.3390/ceramics8040119

Chicago/Turabian Style

Suchikova, Yana, Serhii Nazarovets, Marina Konuhova, and Anatoli I. Popov. 2025. "Binary Oxide Ceramics (TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3) for Solar Cell Applications: A Comparative and Bibliometric Analysis" Ceramics 8, no. 4: 119. https://doi.org/10.3390/ceramics8040119

APA Style

Suchikova, Y., Nazarovets, S., Konuhova, M., & Popov, A. I. (2025). Binary Oxide Ceramics (TiO2, ZnO, Al2O3, SiO2, CeO2, Fe2O3, and WO3) for Solar Cell Applications: A Comparative and Bibliometric Analysis. Ceramics, 8(4), 119. https://doi.org/10.3390/ceramics8040119

Article Metrics

Back to TopTop