Home Physical Sciences Harnessing 1D TiO2 nanostructures: insights into synthesis, properties and photocatalytic applications for pollution control
Article Open Access

Harnessing 1D TiO2 nanostructures: insights into synthesis, properties and photocatalytic applications for pollution control

  • Jyoti Rawat ORCID logo EMAIL logo , Charu Dwivedi ORCID logo EMAIL logo and Pankaj Sharma ORCID logo EMAIL logo
Published/Copyright: April 22, 2025

Abstract

This review explores the development of one dimensional TiO2 nanostructures, which have garnered significant attention due to their exceptional structural and functional properties. Additionally, it provides valuable insights into the synthesis of one dimensional TiO2 nanostructures with controlled size, morphology, composition, and properties, as well as their photocatalytic applications for pollution control, compared to zero dimensional and two dimensional structures. Furthermore, the review places particular emphasis on their role in environmental remediation, especially in the photocatalytic degradation of both air and aqueous pollutants. It highlights the transformative potential of these nanomaterials in addressing pressing environmental challenges and advancing future technologies.

1 Introduction

The rapid transformation of industries and the global economy over the past century has introduced significant challenges. The increasing reliance on fossil fuels has accelerated global warming and climate change, while also contributing to environmental pollution. 1 This situation highlights the urgent need for clean energy sourced from renewables and effective methods for pollutant degradation. 2 Solar energy, particularly through photocatalysis, presents a compelling opportunity for breaking down organic pollutants, generating hydrogen via water splitting, and converting CO2 into renewable hydrocarbon fuels. 3 Furthermore, innovations like supercapacitors and solar cells are emerging as essential tools in combating resource depletion, offering sustainable solutions to both environmental degradation and the energy crisis. 4

TiO2 is a widely studied semiconductor known for its high stability, non-toxicity, and strong photocatalytic activity under ultraviolet (UV) light. Its three primary crystalline forms anatase, rutile, and brookite-exhibit different electronic and optical properties. 5 Among these, anatase TiO2 is particularly promising for photocatalytic applications because of its superior electron-hole pair separation, which is essential for efficient photocatalytic reactions. 6 The dimensionality of TiO2 nanostructures significantly influences their electronic and optical behaviors; thus, fabricating them in a one dimensional (1D) format can lead to enhanced properties compared to their bulk counterparts. 7

1D TiO2 nanostructures, such as nanorods (NRs), nanotubes (NTs), nanobelts (NBs), nanofiber (NFs) and nanowires (NWs), exhibit superior properties compared to their bulk materials due to their increased surface area-to-volume ratio and unique optical and electronic characteristics. 8 The 1D nanostructure also promotes better charge carrier mobility compared to bulk TiO2, making it especially beneficial for applications such as photocatalysis, sensors, and energy storage. 9 Additionally, 1D TiO2 often shows increased photocatalytic activity because of its ability to facilitate faster electron-hole pair separation, which is crucial for efficient light absorption and reaction processes. The high surface area of 1D TiO2 materials can also enhance interactions with molecules, contributing to their superior catalytic performance. 10

These 1D nanostructures of TiO2 offer a higher surface area and enhanced light absorption compared to their bulk counterparts, making them highly efficient for environmental remediation processes. One of the most notable applications of 1D TiO2 is in the degradation of organic pollutants in water. When exposed to UV light, TiO2 generates electron-hole pairs that can break down harmful contaminants such as dyes, pesticides, and pharmaceuticals, converting them into less toxic substances. 11 Additionally, the high surface-to-volume ratio of 1D TiO2 structures facilitates a more efficient interaction with pollutants, enhancing the degradation rate and ensuring a more sustainable and effective treatment process. 12 Beyond water purification, 1D TiO2 has also been applied in air pollution control. TiO2 based photocatalysts can be used to remove harmful gaseous compounds like nitrogen oxides (NOx) and volatile organic compounds (VOCs) from the atmosphere. The photocatalytic oxidation of these pollutants under UV light not only breaks them down but also converts them into less harmful substances like carbon dioxide and water vapor. 13 Furthermore, the robust stability and recyclability of 1D TiO2 materials allow for long-term environmental applications with minimal degradation over time. As a result, 1D TiO2 structures are emerging as promising materials for sustainable environmental technologies, offering a solution to some of the most pressing challenges related to pollution control and resource management. 14

In contrast, two dimensional (2D) and three dimensional (3D) forms of TiO2 exhibit different properties due to their varying dimensionality. 2D TiO2, such as nanosheets (NSs) or thin films, offers excellent charge transport and mechanical flexibility, making it suitable for applications in flexible electronics and optoelectronics. 15 The reduced dimensionality in 2D materials leads to distinct quantum effects and can also enhance photocatalytic performance, especially in terms of light absorption and surface interaction. 16 Meanwhile, 3D TiO2, typically in the form of bulk particles or crystals, possesses the most conventional properties in terms of mechanical strength and stability but lacks the enhanced surface area or charge transport capabilities seen in the 1D and 2D counterparts. The increased dimensionality in 3D TiO2 can also reduce its efficiency in photocatalytic and energy conversion applications due to slower electron transfer rates and the difficulty of light penetration. 17

In recent years, significant advancements have been made in the synthesis, modification, and applications of 1D TiO2 nanostructured materials. Ratanatawanate et al. successfully synthesized TiO2 NTs integrated with PbS quantum dots (QDs) using thiolactic acid. The process involved applying an initial layer of a double-chain cationic surfactant to the TiO2 NTs, and the resulting PbS QDs proved effective for photocatalytic degradation. 11 Zhang et al. synthesized a ZnO NRs based platform that synergizes gas sensing, optoelectronic, and photocatalytic capabilities. Their experiments, combined with first-principles DFT calculations, revealed the impact of surface structure and oxygen vacancies on the (100) facets. The small diameter and high density of the ZnO array enhanced surface area and adsorption, boosting visible light absorption. 18 Bai and his team successfully recycled graphite target waste into 2D few-layer graphite for creating nanoscale photocatalysts. The resulting photocatalyst showed a 19 % improvement in efficiency and enhanced photostability for sunlight-driven degradation of organic pollutants. They also examined how temperature and pH affect photocatalytic degradation performance. 19 Su et al. developed a sensing-transducing strategy by incorporating high-piezoresponse Sm-PMNPT ceramic (d33 = 1,500 pC/N) into a moisture-sensitive polyetherimide (PEI) matrix through electrospinning. This enabled simultaneous humidity sensing and signal transduction. The resulting STP textile can be worn on the body for emotion detection, exercise monitoring, and stress identification. 20 Huang et al. developed a degradable multilayer fabric (DMF) made of ellipsoidal carbon nanotubes (ECNT) and polyvinylpyrrolidone/cellulose acetate electrospun fibers (PEF). The optimized device showed a sensitivity of 3.38 kPa−1 across a range of 0.1–500 kPa and maintained mechanical stability over 2000 cycles. This work deepens the understanding of mechanical interfacial coupling in piezoresistive materials, offering new opportunities for wearable electronics design. 21 Luo et al. developed a PZT-CNT/PVDF composite for infrared sensing and noncontact human-machine interfacing. The 3 wt% CNT doping improved proximity sensitivity by 55.78 % and infrared detection by 13.97 %, offering new potential for high-performance, self-powered sensors. 22 Huang’s team developed a PZT-CNT/PVDF composite for infrared sensing and noncontact HMI. With 3 wt% CNT doping, it achieved optimal sensitivity and infrared detection, improving performance by 13.97 % over the undoped version. This work advances high-performance, self-powered sensors. 23

The unique characteristics of 1D TiO2 nanostructures contribute to their diverse range of applications. In photocatalysis, they are utilized for the degradation of organic pollutants, water splitting for hydrogen production, and carbon dioxide reduction, all of which are critical for environmental remediation and energy conversion. 14 Additionally, their high surface area and conductivity make them suitable for energy storage applications, such as in supercapacitors and lithium-ion batteries, where fast charge-discharge rates are essential. Moreover, 1D TiO2 nanostructures have shown potential in sensing applications, where their sensitivity to changes in environmental conditions can be harnessed for detecting gases and other analytes. 15 Continued research and innovation in synthesis techniques and material design are crucial to overcoming these hurdles and fully realizing the potential of 1D TiO2 nanostructures in practical applications. 14

This review provides a comprehensive discussion on 1D TiO2 nanostructures, which represent a fascinating area of study in nanotechnology, combining unique properties with a broad spectrum of applications. Their potential to revolutionize fields such as environmental science, energy, and electronics makes them a key focus for researchers and industry alike. As advancements continue in their synthesis and application, 1D TiO2 nanostructures are poised to play a significant role in addressing some of the most pressing challenges of our time, paving the way for a more sustainable future.

2 Synthesis of 1D TiO2 nanostructures with various approaches

Nanomaterial synthesis methods are divided into two main approaches: top-down and bottom-up (Figure 1). This section provides a concise overview of key preparation methods, including top-down approaches and bottom-up methods such as the hydrothermal method, chemical vapor deposition (CVD), and the sol-gel method, which will be discussed in more detail in the following section (Tables 1 and 2).

Figure 1: 
A systematic diagram of top-down and bottom-up techniques for nanostructures.
Figure 1:

A systematic diagram of top-down and bottom-up techniques for nanostructures.

Table 1:

Top-down approaches for the synthesis of 1D nanostructures.

S.No. 1D nanostructures Synthesis method Size (nm) Reference
1. TiO2 nanotubes Photolithography method 38.5–53.9 [ 24
2. TiO2 nanotubes Photolithography method 14–17 [ 25
3. TiO2 nanowires Electron Beam lithography 100 [ 26
4. TiO2 nanotubes Thermal evaporation 90 [ 27
5. TiO2 nanowires Thermal evaporation 60–90 [ 28
6. ZnO nanorods Thermal evaporation 50 [ 29
7. ZnS nanobelts Thermal evaporation 100 [ 30
8.  TiO2 nanotubes Lithography method 32 [ 31
9. TiO2 nanowires Photolithography method 90 to 180 [ 32
10. ZnO nanowires Electron-beam-lithography 10 [ 33
11. TiO2 nanorods Thermal evaporation 70–150 [ 34
12. TiO2 nanotubes Photolithography method 30 [ 35
13. TiO2 nanowires Thermal evaporation 70–150 [ 36
14. ZnO nanorods Thermal evaporation 50 [ 37
15. ZnO nanorods Thermal evaporation 300–350 [ 38
Table 2:

Bottom-up approaches for the synthesis of 1D nanostructures.

S.No. 1D nanostructures Synthesis method Size (nm) Reference
1. TiO2 nanobelts Hydrothermal method 7–8 [ 39
2. TiO2 spherical Sol-gel method 10 [ 40
3. TiO2 nanotubes Hydrothermal method 6 [ 41
4. TiO2 nanorods Chemical vapor deposition 10 [ 42
5. TiO2 nanobelts Chemical vapor deposition 10–20 [ 43
6. SnO2 nanorods Hydrothermal method 4–15 [ 44
7. MnV2O6 nanorods Hydrothermal method 20–30 [ 45
8. Si nanowires Chemical vapor deposition 3 [ 46
9. TiO2 nanowires Hydrothermal process 50 [ 47
10. TiO2 nanowire Chemical vapor deposition 30 [ 48
11. TiO2 nanowires Sol-gel method 20 [ 49
12. Ag/TiO2 nanowires Sol-gel method 70 [ 50
13. Ag2S QDs decorated TiO2 nanotubes Hydrothermal method 0.53 [ 51
14. CdSe incorporated TiO2 nanotubes Hydrothermal method 0.35 [ 52
15. TiO2 nanotubes Chemical vapor deposition 25 [ 53

Top-down approaches for fabricating 1D TiO2 nanostructures involve the sequential reduction of bulk materials to achieve desired nanoscale features. Common techniques include lithography, where patterns are etched onto a substrate to define the nanostructures, followed by processes like etching or milling to remove unwanted material. 54 Another method is the use of templates, where bulk TiO2 is shaped into NFs or NRs through methods like electrospinning or sol-gel processes, which create porous structures that can be further refined (Table 1). Additionally, mechanical methods such as grinding or ball milling can be employed to break down larger TiO2 NPs into nanostructured forms. These approaches allow for precise control over the dimensions and morphology of the TiO2 nanostructures, enabling their integration into various applications such as photocatalysis, sensors, and energy storage devices. 55

Bottom-up approaches in nanotechnology focus on the assembly or synthesis of nanoscale structures from smaller components, often at the molecular or atomic level. A common method is sol-gel synthesis, where titanium precursors are transformed into a gel and subsequently heated to produce TiO2 NTs. Hydrothermal and solvothermal methods also allow for the growth of TiO2 nanostructures by controlling temperature and pressure in a solvent, facilitating the formation of well-defined NWs or NRs. 56 Another technique is CVD, which enables the deposition of TiO2 films that can be further processed into 1D nanostructures (Table 2). These bottom-up methods provide advantages such as greater uniformity, the ability to control composition and morphology, and the potential for producing complex nanostructures. Bottom-up strategies are essential in advancing nanotechnology, facilitating innovations in nanoelectronics, nanomedicine, and other fields by leveraging the inherent characteristics of small-scale components to create functional and customized nano systems. 18 , 19

Wang et al. focus on 1D TiO2 nanostructures, specifically NWs, NRs, and NTs, examining their structure, growth mechanisms, synthesis methods, properties, and applications, particularly in energy-related fields. This review covers the structural features of TiO2 polymorphs and the 1D growth mechanisms, such as oriented attachment and surface-reaction-limited growth. It also discusses various synthetic methods, including solution-based, vapor deposition, templated growth, and top-down fabrication. The strengths and limitations of each method are highlighted, with a focus on how TiO2 NWs can improve solar energy conversion efficiency in photovoltaic and photoelectrochemical systems. Additionally, the review explores their potential in energy storage, electrochromism, sensing, and mechanical applications. The article emphasizes the need for reliable, scalable synthesis methods and a deeper understanding of the relationship between 1D morphology, crystal structure, and material properties to enable better integration with other functional materials in advanced device systems. 57

2.1 Top-down approach

Top-down approaches in nanotechnology involve the fabrication or manipulation of nanoscale structures by breaking down or carving out larger materials. These methods start with bulk materials and use various techniques to reduce them to the desired size Lithography techniques are essential for the precise fabrication of 1D TiO2 nanostructures, which are crucial for applications in fields such as electronics, sensors, and catalysis. 18 Traditional photolithography, although widely used, typically faces limitations when creating ultra-small features due to the diffraction limits of light. However, advancements in electron beam lithography (EBL) and X-ray lithography (XRL) have enabled the creation of high-resolution 1D TiO2 structures. These methods allow for the direct writing of intricate patterns, making them suitable for fabricating nanostructures with superior control over dimensions and alignment. 20

EBL offers a significant advantage in terms of resolution, achieving feature sizes down to the nanometer scale. In this process, a focused electron beam is used to expose a resist material, allowing for the precise patterning of 1D TiO2 nanostructures. The high-resolution capabilities of EBL are particularly beneficial when fabricating TiO2 NWs and NTs, which require precise control over their diameter and length. 54 Additionally, EBL enables the production of complex and custom geometries, which is critical for advancing applications in nanoelectronics and photonic devices. However, it is a slower process compared to photolithography, making it more suitable for research and small-scale production. 55

XRL, on the other hand, provides several advantages for large-scale manufacturing of 1D TiO2 structures due to its ability to pattern with finer details than conventional optical lithography. XRL employs high-energy X-rays to expose a photosensitive resist material, enabling the creation of highly intricate and deep structures. 58 This method is particularly advantageous for creating deep, vertical TiO2 nanostructures, as the X-rays can penetrate thicker resist layers without significant loss of resolution. The high aspect ratios achievable with XRL make it an ideal technique for producing 1D TiO2 nanostructures with enhanced mechanical strength and stability, which are important for various industrial applications such as photocatalysis and energy storage. 59 Despite its advantages, XRL requires specialized equipment and is typically more costly than other lithographic methods. Nonetheless, it remains a powerful tool for the fabrication of high-precision 1D TiO2 nanostructures, especially when scalability and feature depth are key considerations. 54

2.1.1 Photolithography

Photolithography is a crucial process used in semiconductor manufacturing and microfabrication, where patterns are transferred onto a substrate, typically silicon, to create intricate circuit designs. The process begins with the application of a light-sensitive photoresist material on the substrate. A patterned mask is then used to expose specific areas of the photoresist to UV light, causing chemical changes in the exposed regions. 58 After developing the photoresist, unwanted areas are washed away, leaving a precise pattern that can be further processed through etching or deposition methods. This technique enables the fabrication of complex electronic components at micro and nanoscale levels, driving advancements in technology and the miniaturization of devices. 59

Y.S. Kim et al. have proposed a novel soft lithography technique that does not require polydimethylsiloxane (PDMS). This approach utilizes a mold made from UV-curable polyurethane acrylate and incorporates water-soluble sulfonated polystyrene (SPS) in the soft molding process. To enhance adhesion, a multilayer deposition is performed between the substrate and the polymer layer. This method effectively addresses the issues related to PDMS molds, particularly their tendency to collapse laterally, enabling the successful transfer of well-organized polymeric nanostructures (80 nm wide and 400 nm tall) onto diverse substrates, including glass and flexible polymer films. 60 Liau et al. developed a photolithography technique using a polyvinyl butyral (PVB) and TiO2 composite, which involves UV exposure. This process facilitates the photodegradation of the robust PVB/TiO2 photoresist, allowing for the easy removal of TiO2 particles due to their lower mechanical strength. The successful formation of patterns on various substrates was demonstrated, and the study also analysed the surface morphology of the films with different exposure times. Additionally, they examined the dynamic intensities of functional groups during the photodegradation of PVB/TiO2. 61

2.1.2 Electron beam lithography

EBL is a high-resolution method for creating detailed micro and nanoscale patterns on a substrate using a focused electron beam. In this process, a substrate coated with an electron-sensitive resist material is exposed to the electron beam, which alters the chemical structure of the resist in the exposed areas. After exposure, the resist is developed, revealing precise patterns that can then be used for further fabrication processes, such as etching or metal deposition. 62 EBL is particularly valued for its ability to achieve feature sizes below 10 nm, making it essential in fields such as semiconductor manufacturing, nanotechnology, and materials research. Although EBL is slower compared to other lithographic methods, its unmatched resolution and flexibility in pattern design make it a powerful tool for creating complex devices and structures at the nanoscale. 63

F. Bretagnol et al. have created a two-step nanoscale writing method that enables the formation of bio-adhesive patterns on a non-bio-adhesive substrate. This nanofabrication process begins with the development of a plasma-derived polyethylene oxide film, which exhibits protein-repelling properties. Following this, electron beam lithography is used to construct nanoscale bio-adhesive patterns by modifying the density of ether bonds in the coating. 64 Burek et al. successfully fabricated arrays of vertical gold and copper nanopillars with diameters as small as 25 nm using electron beam lithography on polymethylmethacrylate. These nanopillars have a broad spectrum of potential applications, including sensing, field emission, and nanomedicine. 65

2.1.3 Thermal evaporation

Thermal evaporation is an endothermic process that relies on heat to induce chemical breakdown within materials, specifically by breaking chemical bonds in molecules. This technique is particularly favoured for producing stable, monodisperse suspensions capable of self-assembly, making it one of the most effective methods for synthesizing inorganic nanoparticles. The process involves depositing thin films onto a wide range of substrates, providing versatility in applications across various fields. 66 In a typical thermal evaporation setup, as illustrated in Figure 2, an alumina crucible containing the source material is placed centrally within the heating zone, with the untreated substrate positioned directly opposite the powder. The source material and substrate are heated using a resistive heating method, and a consistent distance of 20 cm is maintained between them to ensure uniform deposition. To facilitate precise control over deposition time, the entrance of the crucible is equipped with a shutter. Prior to the deposition process, glass slides are meticulously cleaned using a soap solution and then subjected to ultrasonic washing in acetone and de-ionized water to remove any contaminants. For the source materials, stoichiometric powders of each chemical, with a purity of 99.99 %, are employed individually within the crucibles to achieve optimal results. 66 , 67

Figure 2: 
A schematic drawing of thermal evaporation chamber setup.
Figure 2:

A schematic drawing of thermal evaporation chamber setup.

2.2 Bottom-up approach

2.2.1 Chemical vapor deposition

CVD is a versatile and widely used technique for producing thin films and coatings on various substrates through the chemical reaction of gaseous precursors. In this process, gases are added to a chamber, where they undergo chemical reactions that deposit solid material onto the substrate’s surface. 68 , 69 CVD is particularly valued in the semiconductor industry for its ability to create uniform, high-purity films with excellent conformality, making it ideal for applications like integrated circuits, photovoltaics, and protective coatings. 70 The ability to precisely control the film’s composition, thickness, and microstructure allows for the engineering of materials with specific properties tailored to diverse technological needs. 71 , 72

Chen et al. synthesized rutile TiO2 NRs using the CVD technique, with titanium tetraisopropoxide (TTIP) as the precursor. The deposition was carried out at 550 °C under oxygen pressures of 1.5 and 5 mbar. The resulting NRs exhibited increased density and uniform height. However, the high cost of the required equipment limits the practical application of this method, resulting in steep production expenses. 73 Du et al. successfully produced TiO2 NWs through the CVD technique, utilizing TiCl4 as the precursor (Figure 3). 74

Figure 3: 
Shows SEM images of nanowires (a), nanorods (b), and nanobelts (c) fabricated using the CVD method.
74
 Copyright (2015), Journal of Crystal Growth.
Figure 3:

Shows SEM images of nanowires (a), nanorods (b), and nanobelts (c) fabricated using the CVD method. 74 Copyright (2015), Journal of Crystal Growth.

2.2.2 Solvothermal

The solvothermal method a widely used synthetic technique for producing nanostructured materials, involves the reaction of precursors in a solvent under elevated temperature and pressure conditions. This approach allows the controlled growth of materials such as nanoparticles, NWs, and thin films, often resulting in high purity and uniformity. 75 By adjusting parameters like pressure, solvent type, and reaction time, researchers can tailor the size, shape, and crystalline structure of the resulting materials, making it a versatile option for different applications in fields including catalysis, electronics, and photonics. The solvothermal method is particularly beneficial for synthesizing complex oxides and composite materials, enabling the incorporation of different elements and phases to enhance material properties. 76 , 77

Chen et al. synthesized a porous anatase TiO2 nanostructure using a solvothermal method with rod-like titanyl sulphate (Figure 4(a)). The resulting morphology was dependent on reaction time, the form of titanyl sulphate, and the type of solvent. 78 He et al. developed nitrogen-fluorine co-doped TiO2 NBs (Figure 4(b)) from amorphous titania microspheres, showing improved photocatalytic degradation of methyl orange. 79 Zhao and colleagues created TiO2 NRs arrays on FTO glass (Figure 4(c)) through solvothermal methods and thermal treatments, observing a surface transformation that increased photoelectric conversion efficiency by 39 % in dye-sensitized solar cells (DSSC), indicating the potential of solvothermal method in the photoelectrical applications. 80

Figure 4: 
SEM images of (a) nanosheets 
78
 copyright (2010), John Wiley and Sons, (b) nanobelts 
79
 copyright (2012), ACS applied materials & interfaces, and (c) nanorods 
80
 copyright (2014), Journal of Power Sources.
Figure 4:

SEM images of (a) nanosheets 78 copyright (2010), John Wiley and Sons, (b) nanobelts 79 copyright (2012), ACS applied materials & interfaces, and (c) nanorods 80 copyright (2014), Journal of Power Sources.

2.2.3 Hydrothermal method

The hydrothermal method is a synthesis technique used to produce nanostructured materials under high temperature and pressure in a liquid phase, typically water. This method allows for the controlled growth of crystalline materials, often resulting in high purity and well-defined structures. By adjusting parameters such as temperature, pressure, and reaction time, researchers can influence the size, shape, and morphology of the resulting products, making it suitable for a variety of applications, including ceramics, nanomaterials, and catalysts. The hydrothermal approach is particularly advantageous for synthesizing complex oxides and other inorganic compounds, enabling the formation of diverse structures like NTs, NWs, and NRs, while also facilitating the incorporation of different elements and dopants to enhance material properties. 81 Furthermore, Tang et al. developed elongated titanate NTs up to several micrometres long using a stirring hydrothermal method, representing a significant advancement in TiO2 NTs technology. By optimizing the stirring rate, they produced NTs with a high aspect ratio, enabling the creation of 1D TiO2 based materials suitable for lithium-ion batteries. 82 Morgan et al. explored how the amount of NaOH solution and temperature affect nanostructure formation from Degussa P25 using hydrothermal techniques. As depicted in Figure 5, varying hydrothermal conditions can lead to different TiO2 morphologies and structures. 84 Furthermore, Tanaka and Peng et al. examined the impact of NaOH solution amount, temperature, and reaction duration on the formation of nanostructures, using a titanium substrate as the precursor in hydrothermal method. 83 , 85

Figure 5: 
A morphological phase diagram showing the evolution of Degussa P25, highlighting the regions where nanostructures form following 20 h of hydrothermal treatment.
83
 Copyright (2008), chemistry of materials.
Figure 5:

A morphological phase diagram showing the evolution of Degussa P25, highlighting the regions where nanostructures form following 20 h of hydrothermal treatment. 83 Copyright (2008), chemistry of materials.

3 Modification strategy

To enhance the performance of 1D TiO2 nanostructures, such as NRs, NTs, NWs, NSs, and NFs, various modification techniques have been employed. These methods, including thermal treatment, electrochemical deposition, sputtering, sol-gel synthesis, and hydrothermal processing, aim to reduce charge recombination and improve charge transfer efficiency. By integrating additional materials with TiO2, the modified structures can achieve better photocatalytic activity and increased stability, ultimately broadening their applications in fields like solar energy conversion and environmental remediation. The strategic combination of materials and innovative fabrication techniques is essential for optimizing the electronic and optical properties of these nanostructures, making them more effective for practical applications. 86

3.1 Metal and non-metal dopants into 1D TiO2

The strategic incorporation of selectively-doped metal ions into 1D TiO2 nanostructures effectively enhances visible light absorption and reduces charge carrier recombination rates. Introducing transition metal ions into these nanostructures leads to an increase in Ti3+ ions, which boosts photocatalytic activity. 87 This enhancement is linked to a h1igher concentration of oxygen defects that promote better oxygen adsorption on the surface of titania. Additionally, the presence of metal ions creates intraband states close to the valence band (VB) or conduction band (CB) edges, facilitating charge transfer transitions between the TiO2 nanostructures’ VB (or CB) and the d-electrons of the dopants, ultimately improving the absorption of visible light. 88

Asahi et al. incorporated nitrogen into TiO2 through sputtering in a nitrogen-rich gas blend, which broadened its absorption spectrum from the UV to visible light. This modification improved the photocatalytic degradation of methylene orange (MO) when exposed to visible light. Xing et al. fabricated TiO2/graphene composites using a hydrothermal method, in which Ti3+ self-doped TiO2 NRs were integrated onto boron-doped graphene sheets. Sodium borohydride (NaBH4) was employed as the boron doping source and as a reducing agent for the graphene. 89 Hou’s group systematically studied nitrogen-doped TiO2 NTs treated with hot ammonia via a hydrothermal method, discovering that the ammonia-to-deionized water ratio affected the NTs morphology (Figure 6). Nitrogen doping narrows the band gap and improves photogenerated carrier transfer, significantly enhancing photocatalytic activity for MO. Additionally, non-metals like carbon and fluorine have been incorporated into the TiO2 lattice using different methods, with carbon being the second most common element for modifying 1D TiO2 nanostructures. 90

Figure 6: 
Shows SEM images of TiO2 NTs arrays after immersion in water (a) ammonia solutions (b–f) with varying concentrations (vol (A): Vol (DI) = 1:10 to 1:1), with a scale bar of 200 nm. It also includes of the band structure for both pure and nitrogen-doped TiO2 NTAs (g), absorption spectra with estimated bandgaps (h), and the photocatalytic degradation of MO under visible light (i), where A represents ammonia solution and deionized water.
90
 Copyright (2014), solid state Sciences.
Figure 6:

Shows SEM images of TiO2 NTs arrays after immersion in water (a) ammonia solutions (b–f) with varying concentrations (vol (A): Vol (DI) = 1:10 to 1:1), with a scale bar of 200 nm. It also includes of the band structure for both pure and nitrogen-doped TiO2 NTAs (g), absorption spectra with estimated bandgaps (h), and the photocatalytic degradation of MO under visible light (i), where A represents ammonia solution and deionized water. 90 Copyright (2014), solid state Sciences.

3.2 Combining with a semiconductor or conductor

Combining TiO2 with a semiconductor or conductor can significantly enhance its photocatalytic efficiency and overall performance in various applications. By integrating TiO2 with materials such as graphene, ZnO, or metal NPs, the resulting composites can improve charge separation and reduce electron-hole recombination, leading to greater photocatalytic activity. For instance, the presence of conductive materials can facilitate faster electron transport, while semiconductors can provide additional energy levels that help to harness more of the visible light spectrum. These hybrid structures not only enhance the photocatalytic degradation of pollutants but also improve the performance of solar cells, making them highly versatile for environmental applications. The synergistic effects of these combinations enable the development of advanced materials with tailored properties for specific applications. 91

Kim et al. used the S-CBD method to fabricate CdS/TiO2 NWs. They first synthesized TiO2 NWs through a hydrothermal process, then immersed them in a Cd2+ solution to uniformly deposit 8 nm CdS NPs on the surface. 92 Wang et al. created Cu2O/TiO2 NTs array (NTAs) p-n heterojunction using the chemical bath deposition method. As shown in Figure 7(a–e), the Cu2O NPs were uniformly coated on the TiO2 NTs. The quantity of Cu2O could be controlled by adjusting the deposition time. This p-n heterojunction significantly better the separation of photogenerated electrons and holes and enhanced visible light absorption. 94 Ye et al. enhanced n-type TiO2 NTs by decorating them with n-type TiO2 NPs through the hydrolysis of TiCl4 solution, which significantly increased surface area and improved solar cell efficiency (Figure 7(c and g)). 93 Meanwhile, Xu’s group employed an n-n type TiO2/CuInS2 NRs nanostructure solar cell using a solvothermal method (Figure 7(d and h). The CuInS2/TiO2 NRs exhibited higher power transformation efficiency under solar light than TiO2 NTs, owing to better electron-hole separation and enhanced light harvesting. 95

Figure 7: 
SEM images: (a) Cu2O/TiO2 NTAs, (b) CuO/TiO2 NFs, (c) TiO2 NPs/TiO2 NTAs, (d), (e), (f), (g), and (h) TEM images of CuInS2/TiO2 NRs.
93
 Copyright (2011), nano letters.
Figure 7:

SEM images: (a) Cu2O/TiO2 NTAs, (b) CuO/TiO2 NFs, (c) TiO2 NPs/TiO2 NTAs, (d), (e), (f), (g), and (h) TEM images of CuInS2/TiO2 NRs. 93 Copyright (2011), nano letters.

4 Applications of 1D nanostructures

Diverse and effective methodologies have been employed for the fabrication of 1D TiO2 nanostructures. The intrinsic excellence of these nanostructures lies in their exceptional ion-exchange/intercalation capabilities, coupled with their photocatalytic and adsorption characteristics. Consequently, 1D TiO2 nanostructured materials have garnered significant attention, particularly for their roles in the photocatalytic degradation of airborne and aqueous pollutants, the augmentation of supercapacitors, the advancement of solar cells, and the progression of lithium-ion batteries. A comprehensive discussion on the applications of 1D TiO2 nanostructures is provided in this section and Table 3.

Table 3:

Comparison of the degradation capabilities of 1D TiO2 nanostructures against different organic pollutants.

S. No. Catalyst Synthesis

Process
Photocatalytic applications Degradation capabilities Degradation

Time (min)
References
1. Carbon/TiO2 nanotubes Hydrothermal method Rhodamine 6G (RhB-6G) dye 89 % 60 [ 96
2. Nitrogen doped

TiO2 nanotubes
Anodization method Methylene orange 98.5 % 40 [ 97
3. TiO2 nanowires Hydrothermal approach Resorcinol photodegradation 99 % 60 [ 98
4. Fe2O3-TiO2 nanorods Electron beam deposition Methylene blue 85 % 60 [ 99
5. Ag/AgBr/TiO2 nanofibres Electrospinning method Methylene blue 100 % 20 [ 100
6. BiVO4/TiO2 nanofibers Solvothermal method Rhodamine B 98 % 30 [ 101
7. 2H-MoS2/1D-TNTs Anodization method Rhodamine B 99 % 120 [ 102
8. C-TiO2 nanorods Hydrothermal method Methylene blue, Rhodamine B (RhB) and p-nitrophenol (PNP) 63 % 80 [ 103
9. TiO2 nanotubes Liquid phase deposition Tetracycline hydrochloride 60 % 50 [ 104
10. Ag decorated TiO2 nanoneedles Hydrothermal method Methylene blue 98.7 % 50 [ 105

4.1 Photocatalytic applications

The widespread use of TiO2 nanostructured materials as photocatalysts is due to their impressive ability to facilitate both oxidation and reduction reactions. TiO2 reliably exhibits strong photocatalytic activity in tackling both airborne and aqueous pollutants.

4.1.1 Degradation of air pollutants

In recent years the air pollution has been an important and trouble for the human health therefore, a number of efforts have been made to remove these pollutants and purify the air. Millions of people throughout the world suffer from the consequences of air pollution. VOCs are chemical pollutants that cause many diseases.

The degradation of air pollutants is a critical environmental challenge that can be effectively addressed through advanced photocatalytic processes, particularly using materials like TiO2. When exposed to UV light, TiO2 generates reactive oxygen species that can break down harmful compounds such as VOCs, nitrogen oxides, and particulate matter. 95 The advantage of photocatalytic oxidation (PCO) processes is that it occurs at low temperature and pressure. 95 Additionally, the effectiveness of TiO2 in degrading air pollutants can be enhanced by optimizing its nanostructure and incorporating co-catalysts or doping elements to extend its activity into the visible light spectrum. As urban air quality continues to decline, leveraging such photocatalytic technologies offers a promising approach to mitigate pollution and improve public health. 106 Yang et al. synthesized an Ag/TiO2 NTs composite using a straightforward method, with Ag nanocrystals around 3.8 nm uniformly distributed on the TiO2 surface (Figure 8(a and b)). This composite showed exceptional visible-light photocatalytic activity, degrading nearly 100 % of rhodamine B (RhB) in 2 h, outperforming P25 and TiO2 NTs (Figure 8(c, d)). Additionally, TiO2@carbon core/shell NFs were fabricated, allowing for controlled carbon layer thickness, which enhanced photocatalytic efficiency. The optimal performance was observed with a 2 nm thick carbon layer, benefiting from improved charge carrier separation and the synergistic effects of the carbon sensitizer. 107

Figure 8: 
Ag/TiO2 NTs (a) TEM images of Ag/TiO2 NTs (20nm), (b) TEM images of Ag/TiO2 NTs (2nm), (c) photocatalytic degradation of RhB over P25, TiO2 NTs, and Ag/TiO2 NTs, and (d) mechanism of RhB degradation over Ag/TiO2 NTs under visible light.
107
 Copyright (2015), The Journal of Physical chemistry C.
Figure 8:

Ag/TiO2 NTs (a) TEM images of Ag/TiO2 NTs (20nm), (b) TEM images of Ag/TiO2 NTs (2nm), (c) photocatalytic degradation of RhB over P25, TiO2 NTs, and Ag/TiO2 NTs, and (d) mechanism of RhB degradation over Ag/TiO2 NTs under visible light. 107 Copyright (2015), The Journal of Physical chemistry C.

Sopyan et al. examined the kinetics of acetaldehyde, ammonia, and hydrogen sulfide photocatalytic oxidation (PCO) in a batch photoreactor, analysing the data based on the Langmuir-Hinshelwood model. 108 Khalilzadeh et al. designed a photoreactor using N-TiO2 and TiO2 NPs as photocatalysts, observing that the photocatalytic activity under visible light was greater for N-TiO2 than for TiO2. 109 Montalvo et al. developed a pseudo-first-order rate for formaldehyde degradation and found that sound-magnetic assisted fluidized bed (SMFB) outperformed multi-stage conversion fluidized bed (MFB) in enhancing degradation efficiency. 110 Qi et al. noted that while many studies have examined acetaldehyde degradation kinetics, literature reviews reveal a gap in research on acetaldehyde degradation in fluidized bed photoreactors and the various parameters affecting degradation efficiency in these systems. 111

4.1.2 Degradation of aqueous pollutants

The degradation of aqueous pollutants is a critical focus in environmental remediation, and photocatalysis has emerged as a promising approach to address this issue. Utilizing photocatalytic materials like TiO2, researchers can effectively break down harmful substances in water, including organic dyes, heavy metals, and pharmaceuticals. When TiO2 is exposed to UV light, photon energy excites electrons from the VB to the CB. 8 The electrons in the CB can react with oxygen to generate peroxide anions, while the holes in the VB create hydroxyl radicals by interacting with water, facilitating the degradation of pollutants (Figure 9(a)). Although traditional photocatalysis effectively degrades pollutants using various forms of TiO2, photoelectrocatalysis has shown enhanced efficiency, especially with 1D TiO2 nanostructures (NTs, NRs, NWs, NBs, and NFs) grown on FTO or titanium substrates (Figure 9 (b)). 8

Figure 9: 
A systematic representation of (a) photoelectrocatalysis and (b) degradation of pollutants for 1D TiO2.
Figure 9:

A systematic representation of (a) photoelectrocatalysis and (b) degradation of pollutants for 1D TiO2.

Wang et al. utilized electrospinning to create ultrafine fibers of cellulose acetate/TiO2 under UV light, aiming to improve dye wastewater treatment. To address TiO2’s low solar light absorption and large band gap, it is essential to couple 1D TiO2 with semiconductors, nonmetals, and metals to enhance visible light absorption and reduce the band gap, thereby improving both photoelectrocatalytic and photocatalytic activity. 112 Zhang et al. used an electrodeposition method to create Cu2O/TiO2 NTA p-n heterojunction photoelectrodes, with Cu2O NPs evenly distributed on TiO2 NTs. This configuration enhanced the separation of photogenerated electrons and holes, as well as visible light absorption. Compared to TiO2 NTs, the Cu2O/TiO2 NTA photoelectrodes exhibited improved photoconversion efficiency and greater stability in degrading rhodamine B (RhB), demonstrating the synergistic benefits of combining visible light and electricity. 113

4.2 Solar cells

Solar cells are devices that transform sunlight into electricity directly using the photovoltaic effect. These cells harness the abundant energy from the sun in a reliable and cost-effective means, making them a cornerstone of renewable energy technology. When sunlight strikes the surface of a solar cell, it excites electrons in a semiconductor material, typically silicon, creating electron-hole pairs. 114 This process is facilitated by a p-n junction, which generates an internal electric field that directs the flow of these charges, resulting in a direct current (DC) of electricity. With advances in technology, various types of solar cells have emerged, including monocrystalline, polycrystalline, thin-film, and perovskite cells, each with unique characteristics and efficiencies tailored for different applications. 115 A comparison of various applications of 1D TiO2 nanostructures is described in Table 4.

Table 4:

Comparison of the different applications and synthesis of 1D TiO2 nanostructures.

S. No. Precursor Product Synthesis method Phase Applications References
1. Titanium tetrachloride, ethanol, distilled water TiO2 nanowires Hydrothermal Rutile Dye-sensitized solar cells [ 116
2. Poly (vinyl pyrrolidone), tetrabutyl titanate, ethanol TiO2 nanowires Electrospinning and annealing Rutile Gas sensing [ 117
3. Titanium tetraisopropoxide, poly(vinyl acetate), acetic acid, DMSO TiO2 nanofibers Electrospinning Rutile Dye-sensitized solar cells [ 118
4. Ethylene glycol, ammonium fluoride, ethanol TiO2 nanotubes Anodization Anatase Photo-electrochemical and gas sensing [ 119
5. Titanium (Ti) sheets, glycerol, boron-source gas, ammonium fluoride, ethanol, TiO2 nanotubes Chemical vapour deposition Anatase Lithium-ion capacitor [ 120
6. Tetrabutyl titanate, hydrochloric acid, ethanol TiO2 nanorods Solvothermal Anatase Lithium-ion batteries [ 121
7. Titanium sulfate, sodium hydroxide, potassium hydroxide, hydrochloric acid TiO2 nanoflakes Hydrothermal Anatase Lithium-ion batteries [ 122
8. Sodium hydroxide, titanium sulfate, hydrochloric acid TiO2 nanoflakes Hydrothermal precipitation Anatase Lithium-ion battery [ 123
9. Acetone, methanol, isopropanol, tetrabutyl titanate, titanium(IV) butoxide TiO2 nanosheets Hydrothermal Anatase Sensors [ 124
10. Tetrabutyl titantate, hydrogen fluoride TiO2 nanosheets Hydrothermal Anatase Photocatalytic organic conversion [ 125
11. Ti foil samples, acetone, ethanol, and deionized (DI) TiO2 micro-flowers Anodic oxidation Anatase Dye-sensitized solar cells [ 126
12. Titanium n butoxide, acetic acid, TiO2 spheres Acid thermal method Anatase Dye-sensitized solar cells [ 127
13. Titanium (IV) isopropoxide, polyvinyl pyrrolidine TiO2 nanospheres Hydrothermal Anatase Li-ion battery [ 128
14. Oleic acid, ethanol, tetrabutyl titanate TiO2 nanospheres Solvothermal Anatase Photodegradation [ 129
15. Tetrabutyl titanate, hydrochloric acid, deionized water TiO2 nanoflowers Hydrothermal Anatase Photocatalytic degradation of organic pollutants [ 130

The applications of solar cells are vast and varied, ranging from small scale uses in calculators and garden lights to large solar power plants that can provide electricity to thousands of homes. They are also increasingly integrated into building materials, a trend known as building-integrated photovoltaics (BIPV), allowing structures to generate energy while maintaining aesthetic appeal. Despite their significant benefits, such as lowering greenhouse gas emissions and enhancing energy independence, solar cells face challenges, including efficiency limitations and environmental impacts from production and disposal. Ongoing research is focused on improving efficiency, developing new materials, and enhancing recycling methods, ensuring that solar energy can play an increasingly vital role in a sustainable energy future. 115

Recent studies on solar cells using II-VI semiconductor nanostructures have garnered global attention and yielded significant results. Wu et al. developed CdS-based solar systems with Ga nanoribbons and Si heterojunctions. Under UV activation, the transistors demonstrated photovoltaic activity, with an open-circuit voltage of 0.45 V and a short-circuit current of 3.49 nA, achieving a throughput of 44.1 % and energy transfer efficiency of 1.2 %. 131 Britt et al. successfully fabricated 1 cm2 thin-film CdS/CdTe solar cells, achieving an efficiency of 15.8 %. 132 Zhou et al. created a ZnO/ZnSe type II core-shell NWs array solar cell, enhancing light absorption by increasing junction area and light trapping. Their work offers new possibilities for selecting absorber materials in solar cells. 133 Zang et al. developed CdS/Cu2S core-shell nanowire solar cells, which demonstrated improved open-circuit voltage and fill factor compared to planar cells, achieving an efficiency of approximately 5.4 %, despite having limited light absorption. They also connected multiple cells along single nanowires in both series and parallel arrangements to increase output. The enhanced performance is attributed to a low-temperature cation exchange process, which creates a heteroepitaxial junction between the CdS core and Cu2S shell (Figure 10). 135

Figure 10: 
SEM and I-V images: (a) SEM image of the unit, with CdS and Cu2S in yellow and brown, (b) I-V characteristic of the NWs under 1 sun (AM 1.5 G) illumination, (c) SEM image showing three PV units from a single NWs connected in series, (d) I-V characteristic of the series-connected units, showing an increase in voltage while the current stays the same, (e) SEM image of four PV units from a single nanowire connected in parallel, (f) I-V characteristic of the parallel units, showing an increase in current while the voltage remains constant.
134
 Copyright (2013) crystal (MDPI).
Figure 10:

SEM and I-V images: (a) SEM image of the unit, with CdS and Cu2S in yellow and brown, (b) I-V characteristic of the NWs under 1 sun (AM 1.5 G) illumination, (c) SEM image showing three PV units from a single NWs connected in series, (d) I-V characteristic of the series-connected units, showing an increase in voltage while the current stays the same, (e) SEM image of four PV units from a single nanowire connected in parallel, (f) I-V characteristic of the parallel units, showing an increase in current while the voltage remains constant. 134 Copyright (2013) crystal (MDPI).

4.2.1 Dye-sensitized solar cells

DSSCs offer an innovative approach to solar energy conversion, characterized by their use of a photo-sensitizing dye to absorb sunlight and generate electricity. 134 Invented by Michael Gratzel in the early 1990s, these cells consist of a semiconductor, typically TiO2, which is coated with a dye that captures sunlight, exciting electrons and injecting them into the TiO2’s conduction band. This process allows for efficient electron transport to the external circuit, while an electrolyte regenerates the dye by transferring electrons back to it. The advantages of DSSCs include their versatility in design, cost-effectiveness, and good performance under low light conditions, making them suitable for various applications, including integration into building materials. 136 However, challenges such as long-term stability and efficiency improvements remain, prompting ongoing research into robust materials and advanced manufacturing techniques to enhance their commercial viability. 134 Ohsaki et al. pioneered the synthesis of TiO2 NTs for DSSCs applications using the hydrothermal process. Their research examined how various fabrication conditions, such as electrolyte composition, sintering temperature, and paste pH, significantly influence the performance of TiO2 NTs electrodes in DSSCs. Shankar et al. achieved a power conversion efficiency of 6.89 % for DSSCs by fabricating vertically aligned TiO2 NTAs with lengths ranging from 10 to 220 mm through the electrochemical anodization technique. 136

Wang et al. used rutile TiO2 NRs in solid-state DSSCs using C106 dye and spiro-MeOTAD as the hole-transporting material (HTM). The J-V curves shown in Figure 11(a) indicated a 2.9 % efficiency under full sunlight. However, transient photovoltage decay analysis in Figure 11(b and c) revealed that the recombination time for photogenerated electrons and holes in the rutile TiO2 NRs DSSCs was shorter than in the TiO2 NPs at the same charge density. This highlights the differing charge dynamics between the two materials. 57

Figure 11: 
Shows the J-V curves for DSSCs: (a) TiO2 NRs of varying lengths, (b) displays the charge recombination lifetime plots for device A and device D, while (c) illustrates the electron diffusion coefficients for the same devices.
57
 Copyright (2014), American Chemical Society.
Figure 11:

Shows the J-V curves for DSSCs: (a) TiO2 NRs of varying lengths, (b) displays the charge recombination lifetime plots for device A and device D, while (c) illustrates the electron diffusion coefficients for the same devices. 57 Copyright (2014), American Chemical Society.

4.2.2 Lithium-ion batteries

Lithium-ion batteries (LIBs) are commonly used in portable electronics and the automotive sector due to their high energy density. They play a key role in reducing CO2 emissions and mitigating environmental pollution. Traditional LIB cells consist of graphite anodes, LiCoO2 cathodes, and a porous membrane separator soaked in electrolytes, which allow lithium ions to move between the electrodes while electrons flow through external circuits during charging and discharging. 137 To meet diverse application needs, LIBs require high power/energy density and long cycle life. However, graphite anodes face potential safety risks, such as thermal runaway, due to the formation of lithium dendrites and the solid-electrolyte interphase (SEI) layer, which arise from their relatively low lithiation potential. These issues are particularly problematic when LIBs are used at high current rates. As a result, researchers continue to search for alternative electrode materials that can address these challenges. 138

The desired battery performances include energy storage and portable power applications, thanks to their high energy density, and relatively low self-discharge rates. These batteries include a cathode, an anode, an electrolyte, and a separator, which enable the movement of lithium ions between the anode and cathode during charging and discharging cycles. 134 The performance of LIBs is often illustrated in graphs that display key metrics such as voltage, capacity, and charge-discharge cycles. For instance, capacity fade can be depicted over multiple cycles, showing how the battery’s ability to hold charge diminishes over time due to factors like electrolyte degradation and electrode material fatigue. Understanding these graphs is crucial for improving battery design and efficiency, guiding advancements in technologies including electric transports and renewable energy storage systems, and ultimately enabling a more sustainable energy future. 138

In 2005, Li et al. utilized a hydrothermal method to synthesize hydrogen titanate NTs. However, the drawbacks of the irregular and short TiO2 NTs became apparent, particularly in terms of their low discharge capacity. 139 Wang group created MoS2 nanosheet@TiO2 NTs nanostructures using a two-step process. They first synthesized porous TiO2 NTs through the sol-gel method and then added ultrathin MoS2 nanoclusters via solvothermal assembly. The composites achieved an initial discharge capacity of 931 mAh g−1 at 100 mA g−1, maintaining high capacity after 50 cycles. Additionally, a TiO2/MoO2 core-shell NWs array anode was developed using a hydrothermal method subsequent to controlled electrodeposition. 140

Tammawat and colleagues demonstrated that anatase TiO2 NFs anodes could be used directly as the active anode material in LIBs, without the need for additives or binders. These NFs showed excellent lithium storage capacity, stable cycling performance, and strong rate capability. 141 Wang et al. developed a hybrid lithium-ion capacitor using a TiO2 NBs array for the anode and graphene hydrogel for the cathode. Their findings revealed that the capacitor achieved an energy density of 21 Wh/kg and a high-power density of 19 kW/kg. 142 Zhu et al. successfully synthesized mesoporous single-grain layer anatase TiO2 NSs through a straightforward and easily reproducible method. These TiO2 NSs demonstrated a discharge capacity of 73 mAh/g and maintained distinct voltage plateaus over 4,000 cycles, making them a promising anode material for long-lasting LIBs. Wu et al. presented an efficient and eco-friendly method for synthesizing anatase petal-like TiO2 NSs. The distinctive structure exhibited high capacity and excellent cycling stability. This performance can be attributed to the relatively large surface area of 28.4 m2/g, which facilitates shorter lithium-ion diffusion distances and enhances the electrochemical performance of the electrode. 143

Xu’s group created MoS2 NSs@TiO2 NTs hybrids using a two-step method (Figure 12). The TiO2 NTs were first formed via sol-gel, then MoS2 nanoclusters were added through solvothermal treatment. The composites achieved 931 mAh/g at 100 mA/g, maintaining high capacity after 50 cycles. The large surface area and efficient Li-ion transfer contributed to this performance. MoS2@TiO2 hybrids, including NBs and NSs forms, showed over 700 mAh/g after 100 cycles (Figure 12). A TiO2/MoO2 core-shell NWs anode also demonstrated high capacity (∼670 mAh/g), excellent cycling (>200 cycles), and good rate performance (up to 2000 mA/g). 144

Figure 12: 
Shows SEM (a) TEM images (Inset (a)) of MoS2@TiO2 hybrids, (b) charge-discharge profiles at 100 mA/g, (c) compares the cycling performance of MoS2 flakes (I) and MoS2@TiO2 hybrids (II) at varying current densities, (d) show SEM images of TiO2/MoO2 core-shell NWs arrays, (e) shows the cycling performance and coulombic efficiency of TiO2/MoO2 at 50 mA/g cycling performance at 50 mA/g, (f) response at increasing current densities.
144
 Copyright (2014), Royal Society of Chemistry.
Figure 12:

Shows SEM (a) TEM images (Inset (a)) of MoS2@TiO2 hybrids, (b) charge-discharge profiles at 100 mA/g, (c) compares the cycling performance of MoS2 flakes (I) and MoS2@TiO2 hybrids (II) at varying current densities, (d) show SEM images of TiO2/MoO2 core-shell NWs arrays, (e) shows the cycling performance and coulombic efficiency of TiO2/MoO2 at 50 mA/g cycling performance at 50 mA/g, (f) response at increasing current densities. 144 Copyright (2014), Royal Society of Chemistry.

5 Future challenges and perspectives

This review highlights the promise of nanostructured titanium dioxide as an effective and appealing alternative for developing flexible devices across various applications. However, significant challenges remain, particularly in enhancing the spectral sensitivity of these structures to visible and near-infrared light, as well as improving the biocompatibility of TiO2 nanostructures. Future research must prioritize long-term, consistent photoactivity, potentially achieved by exploring alternative synthesis routes. Non-metal doped TiO2 nanostructures currently show limited photocatalytic activity under visible light, suggesting a need for innovative materials such as polymers, glasses, ceramics, and metals, which can facilitate cost-effective and environmentally friendly applications. Additionally, advancing the development of new synthetic methods and nanostructures with increased surface area will be crucial. Techniques compatible with non-lithographic complementary metal oxide semiconductor (CMOS) processes may offer promising avenues for integrating new dopant materials into TiO2 nanostructures. Moreover, these advancements could open doors to applications in environmental remediation and alternative energy solutions. To address these challenges, interdisciplinary knowledge in chemistry, physics, and computational modeling will be essential, enabling researchers to optimize the structure and properties of TiO2 based materials for enhanced performance and broader applicability. Continued collaboration between materials scientists, chemists, and engineers will be vital in overcoming these hurdles and advancing the field.

6 Conclusions

In conclusion, advancements in the synthesis of 1D TiO2 nanostructures have significantly enhanced their photocatalytic applications for organic degradation. With the progress of nanotechnology, researchers have achieved controllable synthesis of 1D TiO2 materials, including solid NWs, NRs, and NBs, each with varying aspect ratios and internal structures. These innovations have spurred investigations into their photocatalytic properties and potential environmental applications. Solar photocatalysis using 1D TiO2 nanostructures presents a promising method for degrading organic pollutants, although successful industrial application hinges on the utilization of solar energy. Future efforts should focus on creating solar-light-active TiO2 nanomaterials and exploring the assembly of 1D nanostructures into integrated nanosystems, such as membranes or films. Rapid advancements are anticipated in developing cost-effective and scalable synthesis strategies for 1D TiO2 nanostructures, particularly for large-scale water treatment applications.


Jyoti Rawat and Pankaj Sharma, Instituto de Quimica, UNAM, Circuito Exterior, Coyoacan CDMX 04510, Mexico, E-mail: (J. Rawat), (P. Sharma); and Charu Dwivedi, Department of Chemistry, School of Physical Sciences, Doon University, Dehradun 248001, India, E-mail:

Acknowledgments

I would also like to thank Direccion General de Asuntosdel Personal Academico (DGAPA), UNAM.

  1. Research ethics: Neither people nor animals were used as subjects in this investigation.

  2. Informed consent: All the authors are agreed in the manuscript.

  3. Author contributions: Jyoti Rawat: Writing-Original Draft, Review & Editing. Charu Dwivedi: Supervision, Review & Editing. Pankaj Sharma: Supervision, Review & Editing.

  4. Use of Large Language Models, AI and Machine Learning Tools: No ChatGPT, Artificial Intelligence (AI) and Machine Learning Tools (MLT) have been used for the preparation of manuscript.

  5. Conflict of interest: There is no conflict of interest.

  6. Research funding: Dirección General de Asuntos del Personal Académico, Universidad Nacional Autónoma de México. http://dx.doi.org/10.13039/501100006087 (posdoc position IQ).

  7. Data availability: Not appilicable.

References

1. Snyder, E. G.; Watkins, T. H.; Solomon, P. A.; Thoma, E. D.; Williams, R. W.; Hagler, G. S.; Shelow, D.; Hindin, D. A.; Kilaru, V. J.; Preuss, P. W. The Changing Paradigm of Air Pollution Monitoring. Environ. Sci. Technol. 2013, 47 (20), 11369–11377; https://doi.org/10.1021/es4022602.Search in Google Scholar PubMed

2. Xie, G.; Zhang, K.; Guo, B.; Liu, Q.; Fang, L.; Gong, J. R. Graphene based Materials for Hydrogen Generation from Light driven Water Splitting. Adv. Mater. 2013, 25 (28), 3820–3839; https://doi.org/10.1002/adma.201301207.Search in Google Scholar PubMed

3. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S. Z. Earth-abundant Cocatalysts for Semiconductor-Based Photocatalytic Water Splitting. Chem. Soc. Rev. 2014, 43 (22), 7787–7812; https://doi.org/10.1039/c3cs60425j.Search in Google Scholar PubMed

4. Zhang, Q.; Suresh, L.; Liang, Q.; Zhang, Y.; Yang, L.; Paul, N.; Tan, S. C. Emerging Technologies for Green Energy Conversion and Storage. Adv. Sustain. Syst. 2021, 5 (3), 2000152; https://doi.org/10.1002/adsu.202000152.Search in Google Scholar

5. Rawat, J.; Sharma, H.; Dwivedi, C. One Dimensional Semiconducting Hybrid Nanostructure: Gas Sensing and Optoelectronic Applications; Wiley, 2023; pp 1–26.10.1002/9783527837649.ch1Search in Google Scholar

6. Katal, R.; Masudy-Panah, S.; Tanhaei, M.; Farahani, M. H. D. A.; Jiangyong, H. A Review on the Synthesis of the Various Types of Anatase TiO2 Facets and Their Applications for Photocatalysis. Chem. Eng. J. 2020, 384, 1–77; https://doi.org/10.1016/j.cej.2019.123384.Search in Google Scholar

7. Nikoobakht, B.; Wang, X.; Herzing, A.; Shi, J. Scalable Synthesis and Device Integration of Self-Registered One-Dimensional Zinc Oxide Nanostructures and Related Materials. Chem. Soc. Rev. 2013, 42 (1), 342–365; https://doi.org/10.1039/c2cs35164a.Search in Google Scholar PubMed

8. Ge, M.; Cao, C.; Huang, J.; Li, S.; Chen, Z.; Zhang, K. Q.; Deyab, S. S. Al.; Lai, Y. A Review of One-Dimensional TiO2 Nanostructured Materials for Environmental and Energy Applications. J. Mater. Chem. A. 2016, 4 (18), 6772–6801; https://doi.org/10.1039/c5ta09323f.Search in Google Scholar

9. Prakash, J.; Swart, H. Plasmonic Photocatalysts as Emerging Multifunctional Nanomaterials for Energy and Environmental Applications. Phys. Rev. B Condens. 2023, 699, 415297; https://doi.org/10.1016/j.physb.2023.415297.Search in Google Scholar

10. Cabrera, J.; Alarcón, H.; López, A.; Candal, R.; Acosta, D.; Rodriguez, J. Synthesis, Characterization and Photocatalytic Activity of 1D TiO2 Nanostructures. Water Sci. Technol. 2014, 70 (6), 972–979; https://doi.org/10.2166/wst.2014.312.Search in Google Scholar PubMed

11. Ratanatawanate, C.; Xiong, C.; Balkus Jr, K. J. Fabrication of PbS Quantum Dot Doped TiO2 Nanotubes. ACS Nano 2008, 2 (8), 1682–1688; https://doi.org/10.1021/nn800141e.Search in Google Scholar PubMed

12. Tian, J.; Zhao, Z.; Kumar, A.; Boughton, R. I.; Liu, H. Recent Progress in Design, Synthesis, and Applications of One-Dimensional TiO2 Nanostructured Surface Heterostructures: A Review. Chem. Soc. Rev. 2014, 43 (20), 6920–6937; https://doi.org/10.1039/c4cs00180j.Search in Google Scholar PubMed

13. Tobaldi, D. M.; Dvoranová, D.; Lajaunie, L.; Rozman, N.; Figueiredo, B.; Seabra, M. P.; Škapin, A. S.; Calvino, J. J.; Brezová, V.; Labrincha, J. A. Graphene-TiO2 Hybrids for Photocatalytic Aided Removal of VOCs and Nitrogen Oxides from Outdoor Environment. Chem. Eng. J 2021, 405, 126651; https://doi.org/10.1016/j.cej.2020.126651.Search in Google Scholar PubMed PubMed Central

14. Zhang, W.; Tian, Y.; He, H.; Xu, L.; Li, W.; Zhao, D. Recent Advances in the Synthesis of Hierarchically Mesoporous TiO2 Materials for Energy and Environmental Applications. Natl. Sci. Rev. 2020, 7 (11), 1702–1725; https://doi.org/10.1093/nsr/nwaa021.Search in Google Scholar PubMed PubMed Central

15. Joshi, R. K.; Schneider, J. J. Assembly of One-Dimensional Inorganic Nanostructures into Functional 2D and 3D Architectures. Synthesis, Arrangement and Functionality. Chem. Soc. Rev. 2012, 41 (15), 5285–5312; https://doi.org/10.1039/c2cs35089k.Search in Google Scholar PubMed

16. Liu, G.; Zhen, C.; Kang, Y.; Wang, L.; Cheng, H. M. Unique Physicochemical Properties of Two-Dimensional Light Absorbers Facilitating Photocatalysis. Chem. Soc. Rev. 2018, 47 (16), 6410–6444; https://doi.org/10.1039/c8cs00396c.Search in Google Scholar PubMed

17. Pang, Y. L.; Lim, S.; Ong, H. C.; Chong, W. T. A Critical Review on the Recent Progress of Synthesizing Techniques and Fabrication of TiO2-Based Nanotubes Photocatalysts. Appl. Catal. Gen. 2014, 481, 127–142; https://doi.org/10.1016/j.apcata.2014.05.007.Search in Google Scholar

18. Zhang, Q.; Xie, G.; Duan, M.; Liu, Y.; Cai, Y.; Xu, M.; Zhao, K.; Tai, H.; Jiang, Y.; Su, Y. Zinc Oxide Nanorods for Light-Activated Gas Sensing and Photocatalytic Applications. ACS Appl. Nano Mater. 2023, 6 (19), 17445–17456; https://doi.org/10.1021/acsanm.3c02403.Search in Google Scholar

19. Bai, M.; Chen, R.; Liu, X.; Li, H.; Li, J.; Huang, H.; Song, M.; Zhang, Q.; Su, Y.; Wang, H.; Xu, M.; Xie, G. An Effective Strategy for Synthesizing High-Performance Photocatalyst by Recycling the Graphite Target Wastes. J. Environ. Chem. Eng. 2024, 12 (5), 113872; https://doi.org/10.1016/j.jece.2024.113872.Search in Google Scholar

20. Su, Y.; Liu, Y.; Li, W.; Xiao, X.; Chen, C.; Lu, H.; Yuan, Z.; Tai, H.; Jiang, Y.; Zou, J.; Xie, G.; Chen, J. Sensing-transducing Coupled Piezoelectric Textiles for Self-Powered Humidity Detection and Wearable Biomonitoring. Mater. Horiz. 2023, 10 (3), 842–851; https://doi.org/10.1039/d2mh01466a.Search in Google Scholar PubMed

21. Huang, J.; Xie, G.; Xu, X.; Geng, Z.; Su, Y. Degradable Multilayer Fabric Sensor with Wide Detection Range and High Linearity. ACS Appl. Mater. Interfaces 2024, 16 (43), 58838–58847; https://doi.org/10.1021/acsami.4c12066.Search in Google Scholar PubMed

22. Luo, X.; Li, W.; Yuan, L.; Xie, G.; Su, Y. Self-powered Infrared Detector Enabled by Interfacial Anchoring and Thermal Reinforcement. Nano Trends 2024, 8, 100061; https://doi.org/10.1016/j.nwnano.2024.100061.Search in Google Scholar

23. Huang, J.; Cai, Y.; Xie, G.; Xu, X.; Geng, Z.; Jiang, Y.; Su, Y. Hierarchical Carbon Nanotube-Decorated Polyacrylonitrile Smart Textiles for Wearable Biomonitoring. Wearables 2024, 1, 180–188; https://doi.org/10.1016/j.wees.2024.07.002.Search in Google Scholar

24. Fraga, T. M.; Albertin, K. F.; Pereyra, I. TiO2 Nanotubes Production and Characterization. ECS Trans. 2012, 49 (1), 199; https://doi.org/10.1149/04901.0199ecst.Search in Google Scholar

25. Berger, S.; Kunze, J.; Schmuki, P.; LeClere, D.; Valota, A. T.; Skeldon, P.; Thompson, G. E. A Lithographic Approach to Determine Volume Expansion Factors during Anodization: Using the Example of Initiation and Growth of TiO2 Nanotubes. Electrochim. Acta. 2009, 54 (24), 5942–5948; https://doi.org/10.1016/j.electacta.2009.05.064.Search in Google Scholar

26. Tian, W. C.; Ho, Y. H.; Chen, C. H.; Kuo, C. Y. Sensing Performance of Precisely Ordered TiO2 Nanowire Gas Sensors Fabricated by Electron-Beam Lithography. Sens 2013, 13 (1), 865–874; https://doi.org/10.3390/s130100865.Search in Google Scholar PubMed PubMed Central

27. Lai, C. W.; Sreekantan, S. WO3 Deposited TiO2 Nanotube Arrays by Thermal Evaporation Method for Effective Photocatalytic Activity. J. Eng. Sci. 2012, 8, 39–50.Search in Google Scholar

28. Wu, J. M.; Wu, W. T.; Shih, H. C. Characterization of Single-Crystalline TiO2 Nanowires Grown by Thermal Evaporation. J. Electrochem. Soc. 2005, 152 (8), 613; https://doi.org/10.1149/1.1949087.Search in Google Scholar

29. Wang, B.; Jin, X.; Ouyang, Z. B.; Xu, P. Photoluminescence and Field Emission of 1D ZnO Nanorods Fabricated by Thermal Evaporation. Appl. Phys. A. 2012, 108, 195–200; https://doi.org/10.1007/s00339-012-6870-1.Search in Google Scholar

30. Li, Q.; Wang, C. Fabrication of Wurtzite ZnS Nanobelts via Simple Thermal Evaporation. Appl. Phys. Lett. 2003, 83 (2), 359–361; https://doi.org/10.1063/1.1591999.Search in Google Scholar

31. Vega, V.; Montero-Moreno, J. M.; García, J.; Prida, V. M.; Rahimi, W.; Waleczek, M.; Bae, C.; Zierold, R.; Nielsch, K. Long-range Hexagonal Arrangement of TiO2 Nanotubes by Soft Lithography-Guided Anodization. Electrochim. Acta. 2016, 203, 51–58.10.1016/j.electacta.2016.04.016Search in Google Scholar

32. Francioso, L.; Siciliano, P. Top-down Contact Lithography Fabrication of aTiO2 Nanowirearray over a SiO2 Mesa. Nanotechnology 2006, 17 (15), 3761; https://doi.org/10.1088/0957-4484/17/15/025.Search in Google Scholar

33. Nicaise, S. M.; Cheng, J. J.; Kiani, A.; Gradečak, S.; Berggren, K. Control of Zinc Oxide Nanowire Array Properties with Electron Beam Lithography Templating for PV Applications. Nanotechnology 2015, 26 (7), 075303, https://doi.org/10.1088/0957-4484/26/7/075303.10.1088/0957-4484/26/7/075303Search in Google Scholar PubMed

34. Wu, J. M.; Shih, H. C.; Wu, W. T. Growth of TiO2 Nanorods by Two-step Thermal Evaporation. J. Vac. Sci. Technol. B. 2005, 23 (5), 2122–2126; https://doi.org/10.1116/1.2038067.Search in Google Scholar

35. Chappanda, K. N.; Smith, Y. R.; Misra, M.; Mohanty, S. K. Site-specific and Patterned Growth of TiO2 Nanotube Arrays from E-Beam Evaporated Thin Titanium Film on Si Wafer. Nanotechnology 2012, 23 (38), 385601; https://doi.org/10.1088/0957-4484/23/38/385601.Search in Google Scholar PubMed

36. Shang, Z. G.; Liu, Z. Q.; Shang, P. J.; Shang, J. K. Synthesis of Single-Crystal TiO2 Nanowire Using Titanium Monoxide Powder by Thermal Evaporation. J. Mater. Sci. Technol. 2012, 28 (5), 385–390; https://doi.org/10.1016/s1005-0302(12)60072-3.Search in Google Scholar

37. Abdallah, B.; Kakhia, M.; Obaide, A. Morphological and Structural Studies of ZnO Nanotube Films Using Thermal Evaporation Technique. Plasmonics 2021, 16 (5), 1549–1556; https://doi.org/10.1007/s11468-021-01420-x.Search in Google Scholar

38. Umar, A.; Karunagaran, B.; Suh, E. K.; Hahn, Y. B. Structural and Optical Properties of Single-Crystalline ZnO Nanorods Grown on Silicon by Thermal Evaporation. Nanotechnology 2006, 17 (16), 4072; https://doi.org/10.1088/0957-4484/17/16/013.Search in Google Scholar PubMed

39. Luo, W.; Taleb, A. Large-scale Synthesis Route of TiO2 Nanomaterials with Controlled Morphologies Using Hydrothermal Method and TiO2 Aggregates as Precursor. Nanomater 2021, 11 (2), 365; https://doi.org/10.3390/nano11020365.Search in Google Scholar PubMed PubMed Central

40. Vijayalakshmi, R.; Rajendran, V. Synthesis and Characterization of Nano-TiO2 via Different Methods. Arch. Appl. Sci. Res. 2012, 4 (2), 1183–1190.Search in Google Scholar

41. Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. Formation of Titanium Oxide Nanotube. Langmuir 1998, 14 (12), 3160–3163; https://doi.org/10.1021/la9713816.Search in Google Scholar

42. Feng, S.; Yang, J.; Liu, M.; Zhu, H.; Zhang, J.; Li, G.; Peng, J.; Liu, Q. CdS Quantum Dots Sensitized TiO2 Nanorod-Array-Film Photoelectrode on FTO Substrate by Electrochemical Atomic Layer Epitaxy Method. Electrochim. Acta 2012, 83, 321–326; https://doi.org/10.1016/j.electacta.2012.07.130.Search in Google Scholar

43. Quynh Hoa, N. T.; Lee, Z.; Kang, S. H.; Radmilovic, V.; Kim, E. T. Synthesis and Ferromagnetism of Co-doped TiO2-δ Nanobelts by Metallorganic Chemical Vapor Deposition. Appl. Phys. Lett. 2008, 92 (12); https://doi.org/10.1063/1.2904648.Search in Google Scholar

44. Chen, Y. J.; Xue, X. Y.; Wang, Y. G.; Wang, T. H. Synthesis and Ethanol Sensing Characteristics of Single Crystalline SnO2 Nanorods. Appl. Phys. Lett. 2005, 87 (23); https://doi.org/10.1063/1.2140091.Search in Google Scholar

45. Lei, S.; Tang, K.; Jin, Y.; Chen, C. Preparation of Aligned MnV2O6 Nanorods and Their Anodic Performance for Lithium Secondary Battery Use. Nanotechnology 2007, 18 (17), 175605; https://doi.org/10.1088/0957-4484/18/17/175605.Search in Google Scholar

46. Wu, Y.; Cui, Y.; Huynh, L.; Barrelet, C. J.; Bell, D. C.; Lieber, C. M. Controlled Growth and Structures of Molecular-Scale Silicon Nanowires. Nano Lett. 2004, 4 (3), 433–436; https://doi.org/10.1021/nl035162i.Search in Google Scholar

47. Wei, M.; Qi, Z. M.; Ichihara, M.; Honma, I.; Zhou, H. Ultralong Single-Crystal TiO2-B Nanowires: Synthesis and Electrochemical Measurements. Chem. Phys. Lett. 2006, 424 (4-6), 316–320; https://doi.org/10.1016/j.cplett.2006.04.066.Search in Google Scholar

48. Shi, J.; Wang, X. Growth of Rutile Titanium Dioxide Nanowires by Pulsed Chemical Vapor Deposition. Cryst. Growth Des. 2011, 11 (4), 949–954; https://doi.org/10.1021/cg200140k.Search in Google Scholar

49. Liu, Z.; Liu, Z.; Cui, T.; Dong, L.; Zhang, J.; Han, Li.; Guangmin, Li.; Liu, C. Photocatalyst from One-Dimensional TiO2 Nanowires/synthetic Zeolite Composites. Mater. Express 2014, 4 (6), 465–474; https://doi.org/10.1166/mex.2014.1196.Search in Google Scholar

50. Chin, S. F.; Pang, S. C.; Dom, F. E. I. Sol-gel Synthesis of Silver/titanium Dioxide (Ag/TiO2) Core-Shell Nanowires for Photocatalytic Applications. Mater. Lett. 2011, 65 (17–18), 2673–2675; https://doi.org/10.1016/j.matlet.2011.05.076.Search in Google Scholar

51. Rawat, J.; Sharma, H.; Dwivedi, C. Improved Photocatalytic Activity of Ag2S Quantum Dots Decorated TiO2 Nanotubes under Visible Light. J. Mater. Res. 2024, 39 (9), 1344–1357; https://doi.org/10.1557/s43578-024-01313-9.Search in Google Scholar

52. Rawat, J.; Rawat, S.; Juyal, A.; Sharma, H.; Dwivedi, C. Enhancing Electrochemical Properties of TiO2 Nanotube by Incorporation of CdSe Quantum Dots. J. Nanopart. Res. 2023, 25 (4), 82; https://doi.org/10.1007/s11051-023-05731-4.Search in Google Scholar

53. Karaman, M.; Sarıipek, F.; Köysüren, Ö.; Yıldız, H. B. Template Assisted Synthesis of Photocatalytic Titanium Dioxide Nanotubes by Hot Filament Chemical Vapor Deposition Method. Appl. Surf. Sci. 2013, 283, 993–998.10.1016/j.apsusc.2013.07.058Search in Google Scholar

54. Abid, N.; Khan, A. M.; Shujait, S.; Chaudhary, K.; Ikram, M.; Imran, M.; Haidar, J.; Khan, M.; Khan, Q.; Maqbool, M. Synthesis of Nanomaterials Using Various Top-Down and Bottom-Up Approaches, Influencing Factors, Advantages, and Disadvantages: A Review. Adv. Colloid Interface Sci. 2022, 300, 102597; https://doi.org/10.1016/j.cis.2021.102597.Search in Google Scholar PubMed

55. Lee, H. K.; Okada, T.; Fujiwara, T.; Lee, S. W. Top-down Synthesis and Deposition of Highly Porous TiO2 Nanoparticles from NH4TiOF3 Single Crystals on Multi-Walled Carbon Nanotubes and Graphene Oxides. Mater. Des. 2016, 108, 269–276; https://doi.org/10.1016/j.matdes.2016.06.101.Search in Google Scholar

56. Oh, D. K.; Jeong, H.; Kim, J.; Kim, Y.; Kim, I.; Ok, J. G.; Rho, J. Top-down Nanofabrication Approaches toward Single-Digit-Nanometer Scale Structures. J. Mech. Sci. Techn. 2021, 35, 837–859; https://doi.org/10.1007/s12206-021-0243-7.Search in Google Scholar

57. Wang, X.; Li, Z.; Shi, J.; Yu, Y. One-dimensional Titanium Dioxide Nanomaterials: Nanowires, Nanorods, and Nanobelts. Chem. Rev. 2014, 114 (19), 9346–9384; https://doi.org/10.1021/cr400633s.Search in Google Scholar PubMed

58. Rawat, J.; Sajwan, D.; Garimella, S. V.; Sharma, H.; Dwivedi, C. Boron Nitride Quantum Dots: A Rising Star in Sensing Applications. Nano Trends 2023, 2, 100008; https://doi.org/10.1016/j.nwnano.2023.100008.Search in Google Scholar

59. Vandenbroucke, S. S.; Mattelaer, F.; Jans, K.; Detavernier, C.; Stakenborg, T.; Vos, R. Photocatalytic Lithography with Atomic Layer Deposited TiO2 Films to Tailor Biointerface Properties. Adv. Mater. Interfaces 2019, 6 (9), 1900035; https://doi.org/10.1002/admi.201900035.Search in Google Scholar

60. Kim, Y. S.; Lee, H. H.; Hammond, P. T. High Density Nanostructure Transfer in Soft Molding Using Polyurethane Acrylate Molds and Polyelectrolyte Multilayers. Nanotechnology 2003, 14 (10), 1140; https://doi.org/10.1088/0957-4484/14/10/312.Search in Google Scholar

61. Liau, L. C. K.; Chou, W. W.; Wu, R. K. Photocatalytic Lithography Processing via Poly (Vinyl butyral)/TiO2 Photoresists by Ultraviolet (UV) Exposure. Ind. Eng. Chem. Res. 2008, 47 (7), 2273–2278; https://doi.org/10.1021/ie071331o.Search in Google Scholar

62. Tseng, A. A.; Chen, K.; Chen, C. D.; Ma, K. J. Electron Beam Lithography in Nanoscale Fabrication: Recent Development. IEEE Trans. Compon. Packag. Manuf. 2003, 26 (2), 141–149.10.1109/TEPM.2003.817714Search in Google Scholar

63. Chen, Y. Nanofabrication by Electron Beam Lithography and its Applications: A Review. Microelectron. Eng. 2015, 135, 57–72; https://doi.org/10.1016/j.mee.2015.02.042.Search in Google Scholar

64. Brétagnol, F.; Sirghi, L.; Mornet, S.; Sasaki, T.; Gilliland, D.; Colpo, P.; Rossi, F. Direct Fabrication of Nanoscale Bio-Adhesive Patterns by Electron Beam Surface Modification of Plasma Polymerized Poly Ethylene Oxide-like Coatings. Nanotech 2008, 19 (12), 125306; https://doi.org/10.1088/0957-4484/19/12/125306.Search in Google Scholar PubMed

65. Burek, M. J.; Greer, J. R. Fabrication and Microstructure Control of Nanoscale Mechanical Testing Specimens via Electron Beam Lithography and Electroplating. Nano Lett. 2010, 10 (1), 69–76; https://doi.org/10.1021/nl902872w.Search in Google Scholar PubMed

66. Wu, J. M.; Shih, H. C.; Wu, W. T.; Tseng, Y. K.; Chen, I. C. Thermal Evaporation Growth and the Luminescence Property of TiO2 Nanowires. J. Cryst. Growth 2005, 281 (2-4), 384–390; https://doi.org/10.1016/j.jcrysgro.2005.04.018.Search in Google Scholar

67. Hassan, N. K.; Hashim, M. R.; Bououdina, M. One-dimensional ZnO Nanostructure Growth Prepared by Thermal Evaporation on Different Substrates: Ultraviolet Emission as a Function of Size and Dimensionality. Ceram. Int. 2013, 39 (7), 7439–7444; https://doi.org/10.1016/j.ceramint.2013.02.088.Search in Google Scholar

68. Rawat, J.; Sharma, H.; Dwivedi, C. Enhanced Photocatalytic Activity and Thermal Conductivity of Boron Nitride Quantum Dots Decorated TiO2 Nanograss Composite. J. Mater. Sci. 2023, 58 (37), 14788–14806; https://doi.org/10.1007/s10853-023-08939-w.Search in Google Scholar

69. Arole, V. M.; Munde, S. V. Fabrication of Nanomaterials by Top-Down and Bottom-Up Approaches-An Overview. J. Mater. Sci. 2014, 89–93.Search in Google Scholar

70. Kumari, S.; Raturi, S.; Kulshrestha, S.; Chauhan, K.; Dhingra, S.; András, K.; Thu, K.; Khargotra, R.; Singh, T. A Comprehensive Review on Various Techniques Used for Synthesizing Nanoparticles. J. Mater. Res. Technol. 2023, 27, 1739–1763; https://doi.org/10.1016/j.jmrt.2023.09.291.Search in Google Scholar

71. Choy, K. L. Chemical Vapour Deposition of Coatings. Mater. Res. Technol. 2003, 48 (2), 57–170; https://doi.org/10.1016/s0079-6425(01)00009-3.Search in Google Scholar

72. Katsui, H.; Goto, T. Chemical Vapor Deposition. Multi-Dimen. Addit. Manuf. 2021, 75–95.10.1007/978-981-15-7910-3_6Search in Google Scholar

73. Chen, C. A.; Chen, Y. M.; Korotcov, A.; Huang, Y. S.; Tsai, D. S.; Tiong, K. K. Growth and Characterization of Well-Aligned Densely-Packed Rutile TiO2 Nanocrystals on Sapphire Substrates via Metal–Organic Chemical Vapor Deposition. Nanotechnology 2008, 19 (7), 075611; https://doi.org/10.1088/0957-4484/19/7/075611.Search in Google Scholar PubMed

74. Du, J.; Gu, X.; Guo, H.; Liu, J.; Wu, Q.; Zou, J. Self-induced Preparation of TiO2 Nanowires by Chemical Vapor Deposition. J. Cryst. Growth 2015, 427, 54–59; https://doi.org/10.1016/j.jcrysgro.2015.07.004.Search in Google Scholar

75. Nakahira, A.; Kubo, T.; Numako, C. Formation Mechanism of TiO2-Derived Titanate Nanotubes Prepared by the Hydrothermal Process. Inorg. Chem. 2010, 49 (13), 5845–5852; https://doi.org/10.1021/ic9025816.Search in Google Scholar PubMed

76. Nyabadza, A.; McCarthy, É.; Makhesana, M.; Heidarinassab, S.; Plouze, A.; Vazquez, M.; Brabazon, D. A Review of Physical, Chemical and Biological Synthesis Methods of Bimetallic Nanoparticles and Applications in Sensing, Water Treatment, Biomedicine, Catalysis and Hydrogen Storage. Adv. Colloid Interface Sci. 2023, 103010; https://doi.org/10.1016/j.cis.2023.103010.Search in Google Scholar PubMed

77. Hendi, A. A.; Alanazi, M. M.; Alharbi, W.; Ali, T.; Awad, M. A.; Ortashi, K. M.; Aldosari, H.; Alfaifi, F. H.; Qindeel, R.; Naz, G.; Alsheddi, T. H. Dye-sensitized Solar Cells Constructed Using Titanium Oxide Nanoparticles and Green Dyes as Photosensitizers. J. King Saud Univ. Sci. 2023, 35 (3), 102555; https://doi.org/10.1016/j.jksus.2023.102555.Search in Google Scholar

78. Chen, Y.; Tian, G.; Ren, Z.; Tian, C.; Pan, K.; Zhou, W.; Fu, H. Solvothermal Synthesis, Characterization, and Formation Mechanism of a Single Layer Anatase TiO2 Nanosheet with a Porous Structure. Eur. J. Inorg. Chem. 2011 (5), 754–770; https://doi.org/10.1002/ejic.201000999.Search in Google Scholar

79. He, Z.; Que, W.; Chen, J.; Yin, X.; He, Y.; Ren, J. Photocatalytic Degradation of Methyl Orange over Nitrogen–Fluorine Codoped TiO2 Nanobelts Prepared by Solvothermal Synthesis. ACS Appl. Mater. Interfaces 2012, 4 (12), 6816–6826; https://doi.org/10.1021/am3019965.Search in Google Scholar PubMed

80. Zhao, J.; Yao, J.; Zhang, Y.; Guli, M.; Xiao, L. Effect of Thermal Treatment on TiO2 Nanorod Electrodes Prepared by the Solvothermal Method for Dye-Sensitized Solar Cells: Surface Reconfiguration and Improved Electron Transport. J. Power Sources 2014, 255, 16–23; https://doi.org/10.1016/j.jpowsour.2013.12.127.Search in Google Scholar

81. Gan, Y. X.; Jayatissa, A. H.; Yu, Z.; Chen, X.; Li, M. Hydrothermal Synthesis of Nanomaterials. J. Nanomater. 2020, 1–3; https://doi.org/10.1155/2020/8917013.Search in Google Scholar

82. Tang, Y.; Zhang, Y.; Deng, J.; Qi, D.; Leow, W. R.; Wei, J.; Yin, S.; Dong, Z.; Yazami, R.; Chen, X. Unravelling the Correlation between the Aspect Ratio of Nanotubular Structures and Their Electrochemical Performance to Achieve High rate and Long life Lithium ion Batteries. Angew. Chem. 2014, 126 (49), 13706–13710; https://doi.org/10.1002/ange.201406719.Search in Google Scholar

83. Tanaka, S. I.; Hirose, N.; Tanaki, T. Effect of the Temperature and Concentration of NaOH on the Formation of Porous TiO2. J. Electrochem. Soc. 2005, 152 (12), 789; https://doi.org/10.1149/1.2073207.Search in Google Scholar

84. Morgan, D. L.; Zhu, H. Y.; Frost, R. L.; Waclawik, E. R. Determination of a Morphological Phase Diagram of Titania/titanate Nanostructures from Alkaline Hydrothermal Treatment of Degussa P25. Chem. Mater. 2008, 20 (12), 3800–3802; https://doi.org/10.1021/cm800077e.Search in Google Scholar

85. Peng, X.; Chen, A. Large Scale Synthesis and Characterization of TiO2 Based Nanostructures on Ti Substrates. Adv. Funct. Mater. 2006, 16 (10), 1355–1362; https://doi.org/10.1002/adfm.200500464.Search in Google Scholar

86. Yoon, Y.; Truong, P. L.; Lee, D.; Ko, S. H. Metal-oxide Nanomaterials Synthesis and Applications in Flexible and Wearable Sensors. ACS Nano 2021, 2 (2), 64–92; https://doi.org/10.1021/acsnanoscienceau.1c00029.Search in Google Scholar PubMed PubMed Central

87. Marschall, R.; Wang, L. Non-metal Doping of Transition Metal Oxides for Visible-Light Photocatalysis. Catal. Today 2014, 225, 111–135; https://doi.org/10.1016/j.cattod.2013.10.088.Search in Google Scholar

88. Anucha, C. B.; Altin, I.; Bacaksiz, E.; Stathopoulos, V. N. Titanium Dioxide (TiO2)-Based Photocatalyst Materials Activity Enhancement for Contaminants of Emerging Concern (CECs) Degradation: In the Light of Modification Strategies. Chem. Eng. J. Adv. 2022, 10, 100262; https://doi.org/10.1016/j.ceja.2022.100262.Search in Google Scholar

89. Asahi, R. Y. O. J. I.; Morikawa, T. A. K. E. S. H. I.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light Photocatalysis in Nitrogen-Doped Titanium Oxides. Sci 2001, 293 (5528), 269–271; https://doi.org/10.1126/science.1061051.Search in Google Scholar PubMed

90. Hou, X.; Wang, C. W.; Zhu, W. D.; Wang, X. Q.; Li, Y.; Wang, J.; Chen, J. B.; Gain, T.; Hu, Y. H.; Zhou, F. Preparation of Nitrogen-Doped Anatase TiO2 Nanoworm/nanotube Hierarchical Structures and its Photocatalytic Effect. Solid State Sci. 2014, 29, 27–33; https://doi.org/10.1016/j.solidstatesciences.2014.01.007.Search in Google Scholar

91. Favet, T.; Keller, V.; Cottineau, T.; El Khakani, M. A. Enhanced Visible-Light-Photoconversion Efficiency of TiO2 Nanotubes Decorated by Pulsed Laser Deposited CoNi Nanoparticles. Int. J. Hydrogen Energy 2019, 44 (54), 28656–28667; https://doi.org/10.1016/j.ijhydene.2019.08.179.Search in Google Scholar

92. Kim, D. H.; Han, H. S.; Cho, I. S.; Seong, W. M.; Park, I. J.; Park, J. H.; Sin, S.; Park, G. D.; Lee, S.; Hong, K. S. CdS-sensitized 1D Single-Crystalline Anatase TiO2 Nanowire Arrays for Photoelectrochemical Hydrogen Production. Int. J. Hydrogen Energy 2015, 40 (1), 863–869; https://doi.org/10.1016/j.ijhydene.2014.09.174.Search in Google Scholar

93. Ye, M.; Xin, X.; Lin, C.; Lin, Z. High Efficiency Dye-Sensitized Solar Cells Based on Hierarchically Structured Nanotubes. Nano Lett. 2011, 11 (8), 3214–3220; https://doi.org/10.1021/nl2014845.Search in Google Scholar PubMed

94. Wang, M.; Sun, L.; Lin, Z.; Cai, J.; Xie, K.; Lin, C. p -n Heterojunction Photoelectrodes Composed of Cu2O Loaded TiO2 Nanotube Arrays with Enhanced Photoelectrochemical and Photoelectrocatalytic Activities. Energy Environ. Sci. 2013, 6 (4), 1211–1220; https://doi.org/10.1039/c3ee24162a.Search in Google Scholar

95. Xu, P.; Ding, C.; Li, Z.; Yu, R.; Cui, H.; Gao, S. Photocatalytic Degradation of Air Pollutant by Modified Nano Titanium Oxide (TiO2) in a Fluidized Bed Photoreactor: Optimizing and Kinetic Modeling. Chemosphere 2023, 319, 137995; https://doi.org/10.1016/j.chemosphere.2023.137995.Search in Google Scholar PubMed

96. Natarajan, T. S.; Lee, J. Y.; Bajaj, H. C.; Jo, W. K.; Tayade, R. J. Synthesis of Multiwall Carbon nanotubes/TiO2 Nanotube Composites with Enhanced Photocatalytic Decomposition Efficiency. Catal. Today 2017, 282, 13–23; https://doi.org/10.1016/j.cattod.2016.03.018.Search in Google Scholar

97. Yuan, B.; Wang, Y.; Bian, H.; Shen, T.; Wu, Y.; Chen, Z. Nitrogen Doped TiO2 Nanotube Arrays with High Photoelectrochemical Activity for Photocatalytic Applications. Appl. Surf. Sci. 2013, 280, 523–529; https://doi.org/10.1016/j.apsusc.2013.05.021.Search in Google Scholar

98. Al-Hajji, L. A.; Ismail, A. A.; Al-Hazza, A.; Ahmed, S. A.; Alsaidi, M.; Almutawa, F.; Bumajdad, A. Impact of Calcination of Hydrothermally Synthesized TiO2 Nanowires on Their Photocatalytic Efficiency. J. Mol. Struct. 2020, 1200, 127153; https://doi.org/10.1016/j.molstruc.2019.127153.Search in Google Scholar

99. Yao, K.; Basnet, P.; Sessions, H.; Larsen, G. K.; Murph, S. E. H.; Zhao, Y. Fe2O3-TiO2 Core–Shell Nanorod Arrays for Visible Light Photocatalytic Applications. Catal. Today 2016, 270, 51–58; https://doi.org/10.1016/j.cattod.2015.10.026.Search in Google Scholar

100. Ding, M. Y.; Meng, D. S.; Tang, Y. H.; Liu, C. B.; Luo, S. L. One-dimensional Porous Ag/AgBr/TiO2 Nanofibres with Enhanced Visible Light Photocatalytic Activity. Chem. Papers 2015, 69, 1411–1420; https://doi.org/10.1515/chempap-2015-0151.Search in Google Scholar

101. Guo, Z.; Li, P.; Che, H.; Wang, G.; Wu, C.; Zhang, X.; Mu, J. One-dimensional Spindle-like BiVO4/TiO2 Nanofibers Heterojunction Nanocomposites with Enhanced Visible Light Photocatalytic Activity. Ceram. Int. 2016, 42 (3), 4517–4525; https://doi.org/10.1016/j.ceramint.2015.11.142.Search in Google Scholar

102. Liu, E.; Gong, P.; Wang, L.; Bi, X.; Yin, X.; Yu, S.; Wang, B.; Yang, X. Preparation of 2H Phase MoS2 on One-Dimensional Titanium Dioxide Nanotube Array to Construct High Performance Z-type Heterojunction Photocatalytic Decomposition of Rhodamine B. Appl. Phys. A 2022, 128 (9), 781; https://doi.org/10.1007/s00339-022-05932-z.Search in Google Scholar

103. Shao, J.; Sheng, W.; Wang, M.; Li, S.; Chen, J.; Zhang, Y.; Cao, S. In Situ Synthesis of Carbon-Doped TiO2 Single-Crystal Nanorods with a Remarkably Photocatalytic Efficiency. Appl. Catal. B Environ. 2017, 209, 311–319; https://doi.org/10.1016/j.apcatb.2017.03.008.Search in Google Scholar

104. Wang, H.; Wu, X.; Zhao, H.; Quan, X. Enhanced Photocatalytic Degradation of Tetracycline Hydrochloride by Molecular Imprinted Film Modified TiO2 Nanotubes. Sci. Bull. 2012, 57, 601–605; https://doi.org/10.1007/s11434-011-4897-x.Search in Google Scholar

105. Ridha, N. J.; Alosfur, F. K. M.; Kadhim, H. B. A.; Ahmed, L. M. Synthesis of Ag Decorated TiO2 Nanoneedles for Photocatalytic Degradation of Methylene Blue Dye. Mater. Res. Express 2021, 8 (12), 125013; https://doi.org/10.1088/2053-1591/ac4408.Search in Google Scholar

106. Mohamadpour, F.; Amani, A. M. Photocatalytic Systems: Reactions, Mechanism, and Applications. RSC Adv. 2024, 14 (29), 20609–20645; https://doi.org/10.1039/d4ra03259d.Search in Google Scholar PubMed PubMed Central

107. Yang, D.; Sun, Y.; Tong, Z.; Tian, Y.; Li, Y.; Jiang, Z. Synthesis of Ag/TiO2 Nanotube Heterojunction with Improved Visible-Light Photocatalytic Performance Inspired by Bioadhesion. J. Phys. Chem. C 2015, 119 (11), 5827–5835; https://doi.org/10.1021/jp511948p.Search in Google Scholar

108. Sopyan, I. Kinetic Analysis on Photocatalytic Degradation of Gaseous Acetaldehyde, Ammonia and Hydrogen Sulfide on Nanosized Porous TiO2 Films. Sci. Techn. Advan. Mater. 2007, 8 (1-2), 33–39; https://doi.org/10.1016/j.stam.2006.10.004.Search in Google Scholar

109. Khalilzadeh, A.; Fatemi, S. Modification of Nano-TiO2 by Doping with Nitrogen and Fluorine and Study Acetaldehyde Removal under Visible Light Irradiation. Clean Technol. Environ. Policy 2014, 16, 629–636; https://doi.org/10.1007/s10098-013-0666-7.Search in Google Scholar

110. Montalvo, C.; Gines, R. S.; Cantu, D.; Ruiz, A.; Aguilar, C. A.; Perez, I.; Ceron, R. M. Fluidized Bed Photoreactor for the Removal of Acetaminophen and Pyridine Using Al-Doped TiO2 Supported on Alumina. Iran. Cataly. 2022, 12 (3), 295–313.Search in Google Scholar

111. Qi, B.; Xu, P.; Wu, C. Analysis of the Infiltration and Water Storage Performance of Recycled Brick Mix Aggregates in Sponge City Construction. Water 2023, 15 (2), 363; https://doi.org/10.3390/w15020363.Search in Google Scholar

112. Wang, S. D.; Ma, Q.; Liu, H.; Wang, K.; Ling, L. Z.; Zhang, K. Q. Robust Electrospinning Cellulose acetate@TiO2 Ultrafine Fibers for Dyeing Water Treatment by Photocatalytic Reactions. RSC Adv. 2015, 5 (51), 40521–40530; https://doi.org/10.1039/c5ra03797b.Search in Google Scholar

113. Zhang, J.; Wang, Y.; Yu, C.; Shu, X.; Jiang, L.; Cui, J.; Chen, Z.; Xie, T.; Wu, Y. Enhanced Visible-Light Photoelectrochemical Behaviour of Heterojunction Composite with Cu2O Nanoparticles-Decorated TiO2 Nanotube Arrays. New J. Chem. 2014, 38 (10), 4975–4984; https://doi.org/10.1039/c4nj00787e.Search in Google Scholar

114. Singh, N.; Kapoor, A.; Mehra, R. M. Quantum Dot Sensitized Solar Cells. J. Renew. Energy. 2013, 3 (3), 133–177.Search in Google Scholar

115. Shaker, L. M.; Al-Amiery, A. A.; Hanoon, M. M.; Al-Azzawi, W. K.; Kadhum, A. A. H. Examining the Influence of Thermal Effects on Solar Cells: a Comprehensive Review. Sustain. Energy Res. 2024, 11 (1), 1–6; https://doi.org/10.1186/s40807-024-00100-8.Search in Google Scholar

116. Wang, X.; Liu, Y.; Zhou, X.; Li, B.; Wang, H.; Zhao, W.; Huang, H.; Liang, C.; Yu, X.; Liu, Z.; Shen, H. Synthesis of Long TiO2 Nanowire Arrays with High Surface Areas via Synergistic Assembly Route for Highly Efficient Dye-Sensitized Solar Cells. J. Mater. Chem. 2012, 22 (34), 17531–17538; https://doi.org/10.1039/c2jm32883f.Search in Google Scholar

117. Zhou, M.; Liu, Y.; Wu, B.; Zhang, X. Different Crystalline Phases of Aligned TiO2 Nanowires and Their Ethanol Gas Sensing Properties. Physica E Low Dimens. Syst. Nanostruct. 2019, 114, 113601; https://doi.org/10.1016/j.physe.2019.113601.Search in Google Scholar

118. Mahmoud, M. S.; Akhtar, M. S.; Mohamed, I. M.; Hamdan, R.; Dakka, Y. A.; Barakat, N. A. Demonstrated Photons to Electron Activity of S-Doped TiO2 Nanofibers as Photoanode in the DSSC. Mater. Lett. 2018, 225, 77–81; https://doi.org/10.1016/j.matlet.2018.04.108.Search in Google Scholar

119. Ng, S.; Kuberský, P.; Krbal, M.; Prikryl, J.; Gärtnerová, V.; Moravcová, D.; Sopha, H.; Zazpe, R.; Kwong Yam, F.; Jager, A.; Hromádko, L.; Beneš, L.; Hamáček, A.; Macak, J. M. ZnO Coated Anodic 1D TiO2 Nanotube Layers: Efficient Photo electrochemical and Gas Sensing Heterojunction. Adv. Eng. Mater. 2018, 20 (2), 1700589; https://doi.org/10.1002/adem.201700589.Search in Google Scholar

120. Gao, J.; Qiu, G.; Li, H.; Li, M.; Li, C.; Qian, L.; Yang, B. Boron-Doped graphene/TiO2 Nanotube-Based Aqueous Lithium-Ion Capacitors with High Energy Density. Electrochim. Acta 2020, 329, 135175; https://doi.org/10.1016/j.electacta.2019.135175.Search in Google Scholar

121. Ma, Y.; Li, Y.; Li, D.; Liu, Y.; Zhang, J. Uniformly Distributed TiO2 Nanorods on Reduced Graphene Oxide Composites as Anode Material for High-Rate Lithium-Ion Batteries. J. Alloys Compd. 2019, 771, 885–891; https://doi.org/10.1016/j.jallcom.2018.08.251.Search in Google Scholar

122. Li, Y.; Hang, X.; Liang, J.; Lu, C.; Ye, K.; Hou, C.; Yu, K. TiO2 Nanoflakes as Anode Material for Lithium-Ion Batteries. SYN React. Inorg. Metaog. Nanometal Chem. 2016, 46 (10), 1480–1484; https://doi.org/10.1080/15533174.2015.1068808.Search in Google Scholar

123. Li, Y.; Han, X.; Liang, J.; Leng, X.; Ye, K.; Hou, C.; Yu, K. Preparation of TiO2 Nanoflakes and Their Influence on Lithium-Ion Battery Storage Performance. Chem. Res. Chin. Univ. 2015, 31, 332–336; https://doi.org/10.1007/s40242-015-4421-y.Search in Google Scholar

124. Feng, Z.; Zhang, L.; Chen, W.; Peng, Z.; Li, Y. A Strategy for Supportless Sensors: Fluorine Doped TiO2 Nanosheets Directly Grown onto Ti Foam Enabling Highly Sensitive Detection toward Acetone. Sens. Actuators B: Chem. 2020, 322, 128633; https://doi.org/10.1016/j.snb.2020.128633.Search in Google Scholar

125. Chen, Q.; Wang, H.; Wang, C.; Guan, R.; Duan, R.; Fang, Y.; Hu, X. Activation of Molecular Oxygen in Selectively Photocatalytic Organic Conversion upon Defective TiO2 Nanosheets with Boosted Separation of Charge Carriers. Appl. Catal. B Environ. 2020, 262, 118258; https://doi.org/10.1016/j.apcatb.2019.118258.Search in Google Scholar

126. Kim, W. R.; Park, H.; Choi, W. Y. TiO2 Micro-flowers Composed of Nanotubes and Their Application to Dye-Sensitized Solar Cells. Nanoscale Res. Lett. 2014, 9, 1–10; https://doi.org/10.1186/1556-276x-9-93.Search in Google Scholar

127. Liao, J. Y.; Lei, B. X.; Kuang, D. B.; Su, C. Y. Tri-functional Hierarchical TiO2 Spheres Consisting of Anatase Nanorods and Nanoparticles for High Efficiency Dye-Sensitized Solar Cells. EES or Energ. Environ. Sci. 2011, 4 (10), 4079–4085; https://doi.org/10.1039/c1ee01574e.Search in Google Scholar

128. Liu, L.; Li, X.; He, G.; Zhang, G.; Su, G.; Fang, C. SiO@C/TiO2 Nanospheres with Dual Stabilized Architecture as Anode Material for High-Performance Li-Ion Battery. J. Alloys Compd. 2020, 836, 155407; https://doi.org/10.1016/j.jallcom.2020.155407.Search in Google Scholar

129. Cao, J.; Song, X. Z.; Kang, X.; Dai, Z.; Tan, Z. One-pot Synthesis of Oleic Acid Modified Monodispersed Mesoporous TiO2 Nanospheres with Enhanced Visible Light Photocatalytic Performance. Adv. Powder Technol. 2018, 29 (8), 1925–1932; https://doi.org/10.1016/j.apt.2018.05.005.Search in Google Scholar

130. Li, M.; Jiang, Y.; Ding, R.; Song, D.; Yu, H.; Chen, Z. Hydrothermal Synthesis of Anatase TiO2 Nanoflowers on a Nanobelt Framework for Photocatalytic Applications. J. Electron. Mater. 2013, 42, 1290–1296; https://doi.org/10.1007/s11664-013-2593-0.Search in Google Scholar

131. Wu, D.; Jiang, Y.; Li, S.; Li, F.; Li, J.; Lan, X.; Zhang, Y.; Wu, C.; Luo, L.; Jie, J. Construction of High-Quality CdS: Ga Nanoribbon/silicon Heterojunctions and Their Nano-Optoelectronic Applications. Nanotechnology 2011 (40), 405–201; https://doi.org/10.1088/0957-4484/22/40/405201.Search in Google Scholar PubMed

132. Britt, J.; Ferekides, C. Thin-film CdS/CdTe Solar Cell with 15.8% Efficiency. Appl. Phys. Lett. 1993, 62, 2851–2852; https://doi.org/10.1063/1.109629.Search in Google Scholar

133. Zhou, J.; Gu, Y.; Hu, Y.; Mai, W.; Yeh, P. H.; Bao, G.; Sood, A. K.; Polla, D. L.; Wang, Z. L. Gigantic Enhancement in Response and Reset Time of ZnO UV Nanosensor by Utilizing Schottky Contact and Surface Functionalization. Appl. Phys. Lett. 2009, 94, 191103; https://doi.org/10.1063/1.3133358.Search in Google Scholar PubMed PubMed Central

134. Zhang, X.; Wu, D.; Geng, H. Heterojunctions Based on II-VI Compound Semiconductor One-Dimensional Nanostructures and Their Optoelectronic Applications. Crystals 2017 (10), 307; https://doi.org/10.3390/cryst7100307.Search in Google Scholar

135. Zhang, X.; Mao, J.; Shao, Z.; Diao, S.; Hu, D.; Tang, Z.; Wu, H.; Jie, J. Efficient Photovoltaic Devices Based on P-ZnSe/n-CdS Core-Shell Heterojunctions with High Open-Circuit Voltage. J. Mater. Chem. C. 2017 (8), 2107–2113; https://doi.org/10.1039/c6tc04960e.Search in Google Scholar

136. Ohsaki, Y.; Masaki, N.; Kitamura, T.; Wada, Y.; Okamoto, T.; Sekino, T.; Niihara, K.; Yanagida, S. Dye-sensitized TiO2 Nanotube Solar Cells: Fabrication and Electronic Characterization. Phys. Chem. Chem. Phys. 2005, 7 (24), 4157–4163; https://doi.org/10.1039/b511016e.Search in Google Scholar PubMed

137. Hu, S.; Chen, W.; Zhou, J.; Yin, F.; Uchaker, E.; Zhang, Q.; Cao, G. Preparation of Carbon Coated MoS2 Flower-like Nanostructure with Self-Assembled Nanosheets as High-Performance Lithium-Ion Battery Anodes. J. Mater. Chem. A 2014, 2 (21), 7862–7872; https://doi.org/10.1039/c4ta01247j.Search in Google Scholar

138. Zhang, L.; Wang, Y.; Peng, B.; Yu, W.; Wang, H.; Wang, T.; Deng, W.; Chai, L.; Zhang, K.; Wang, J. Preparation of a Macroscopic, Robust Carbon-Fiber Monolith from Filamentous Fungi and its Application in Li-S Batteries. Green Chem. 2014, 16 (8), 3926–3934; https://doi.org/10.1039/c4gc00761a.Search in Google Scholar

139. Li, J.; Tang, Z.; Zhang, Z. Layered Hydrogen Titanate Nanowires with Novel Lithium Intercalation Properties. Chem. of Maters. 2005, 17 (23), 5848–5855; https://doi.org/10.1021/cm0516199.Search in Google Scholar

140. Wang, C.; Wu, L.; Wang, H.; Zuo, W.; Li, Y.; Liu, J. Fabrication and Shell Optimization of Synergistic TiO2 MoO3 Core-Shell Nanowire Array Anode for High Energy and Power Density Lithium‐ion Batteries. Adv. Funct. Mater. 2015, 25 (23), 3524–3533; https://doi.org/10.1002/adfm.201500634.Search in Google Scholar

141. Tammawat, P.; Meethong, N. Synthesis and Characterization of Stable and Binder free Electrodes of TiO2 Nanofibers for Li‐ion Batteries. J. Nanomater 2013 (1), 413692; https://doi.org/10.1155/2013/413692.Search in Google Scholar

142. Wang, H.; Guan, C.; Wang, X.; Fan, H. J. A High Energy and Power Li‐ion Capacitor Based on a TiO2 Nanobelt Array Anode and a Graphene Hydrogel Cathode. Small 2015, 11 (12), 1470–1477; https://doi.org/10.1002/smll.201402620.Search in Google Scholar PubMed

143. Zhu, K.; Liu, X.; Du, J.; Tian, J.; Wang, Y.; Liu, S.; Shan, Z. In Situ Synthesis of Mesoporous Single-Grain Layer Anatase TiO2 Nanosheets without Additives via a Mild and Simple Process for a Long-Term Li-Ion Battery. J. Mater. Chem. A 2015, 3 (12), 6455–6463; https://doi.org/10.1039/c5ta00236b.Search in Google Scholar

144. Wu, F.; Wang, Z.; Li, X.; Guo, H. Simple Preparation of Petal-like TiO2 Nanosheets as Anode Materials for Lithium-Ion Batteries. Ceram. Int. 2014, 40 (10), 16805–16810; https://doi.org/10.1016/j.ceramint.2014.07.060.Search in Google Scholar

Received: 2024-11-21
Accepted: 2025-04-04
Published Online: 2025-04-22

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 10.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/revic-2024-0130/html
Scroll to top button