Next Article in Journal
Investigation on Performances and Functions of Asphalt Mixtures Modified with Super Absorbent Polymer (SAP)
Next Article in Special Issue
Surface-Dependent Hydrogen Evolution Activity of Copper Foil
Previous Article in Journal
Theoretical Nanoarchitectonics of GaN Nanowires for Ultraviolet Irradiation-Dependent Electromechanical Properties
Previous Article in Special Issue
Crystal and Electronic Structures of New Two Dimensional 3-NH3-PyPbX4 Haloplumbate Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Photocatalysts for CO2 Conversion by Severe Plastic Deformation (SPD)

1
Department of Life Science and Applied Chemistry, Nagoya Institute of Technology, Tajimi 507-0071, Japan
2
WPI International Institute for Carbon-Neutral Energy Research (WPI-I2CNER), Kyushu University, Fukuoka 819-0395, Japan
3
Mitsui Chemicals, Inc.—Carbon Neutral Research Center (MCI-CNRC), Kyushu University, Fukuoka 819-0395, Japan
4
Department of Applied Chemistry, Faculty of Engineering, Kyushu University, Fukuoka 819-0395, Japan
5
Advanced Ceramics Research Center, Nagoya Institute of Technology, Tajimi 507-0071, Japan
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(3), 1081; https://doi.org/10.3390/ma16031081
Submission received: 31 December 2022 / Revised: 22 January 2023 / Accepted: 24 January 2023 / Published: 26 January 2023

Abstract

:
Excessive CO2 emission from fossil fuel usage has resulted in global warming and environmental crises. To solve this problem, the photocatalytic conversion of CO2 to CO or useful components is a new strategy that has received significant attention. The main challenge in this regard is exploring photocatalysts with high efficiency for CO2 photoreduction. Severe plastic deformation (SPD) through the high-pressure torsion (HPT) process has been effectively used in recent years to develop novel active catalysts for CO2 conversion. These active photocatalysts have been designed based on four main strategies: (i) oxygen vacancy and strain engineering, (ii) stabilization of high-pressure phases, (iii) synthesis of defective high-entropy oxides, and (iv) synthesis of low-bandgap high-entropy oxynitrides. These strategies can enhance the photocatalytic efficiency compared with conventional and benchmark photocatalysts by improving CO2 adsorption, increasing light absorbance, aligning the band structure, narrowing the bandgap, accelerating the charge carrier migration, suppressing the recombination rate of electrons and holes, and providing active sites for photocatalytic reactions. This article reviews recent progress in the application of SPD to develop functional ceramics for photocatalytic CO2 conversion.

1. Introduction

Nowadays, environmental crises, especially global warming caused by CO2 emission from burning fossil fuels and humankind activities, are considered one of the most significant challenges in the world. Reduction of CO2 to reactive CO gas or useful components and fuels, such as CH4 and CH3OH, using photocatalysts is one of the clean and new strategies, which is developing rapidly [1,2,3]. In photocatalytic CO2 conversion, excited electrons transfer from the valence band to the conduction band of the photocatalysts by solar irradiation and contribute to the reduction of CO2 to form desirable products, as shown in Figure 1a [3]. To perform these reduction reactions, some thermodynamic and kinetic conditions should be provided. From the viewpoint of thermodynamics, the standard potential of the reduction and oxidation reactions should be between the valence band and the conduction band of the photocatalyst [3,4]. On the other hand, from the kinetic viewpoint, the electrons should absorb the light, transfer to the conduction band, migrate to the surface of the photocatalyst, and take part in the reactions before combining with the holes [3,4]. To satisfy these kinetic and thermodynamic conditions, a photocatalyst should have some features, including high light absorbance, appropriate band structure, low recombination rate of electrons and holes, easy migration of charge carriers, and high surface affinity to adsorb CO2 with abundant active sites [3,4]. A combination of these thermodynamic and kinetic factors determines the speed of the reactions and the type of final products in photocatalysis.
Semiconductors, such as TiO2 [5,6,7], g-C3N4 [8,9], and BiVO4 [10,11,12], are typical photocatalysts that have been engineered by various strategies to enhance the catalytic efficiency for CO2 conversion. Doping with impurities, such as nitrogen, phosphorous, copper, and palladium [13,14,15]; defect engineering [16,17]; strain engineering [18,19]; mesoporous structure production [20]; and heterojunction introduction [21,22] are some of the most promising strategies that have been used so far to improve the optical properties and catalytic activity of various photocatalysts. Among these strategies, doping with impurities is the most investigated and feasible method, but impurities can increase the recombination rate of electrons and holes [13,14,15]. Therefore, finding new strategies to improve the photocatalytic activity and suppress the recombination rate of electrons and holes is a key issue. In this regard, severe plastic deformation (SPD) through the high-pressure torsion (HPT) method, which is mainly used for nanostructuring of metallic materials, has been used as a new tool to develop active photocatalysts for water splitting [23,24,25,26,27,28,29,30], dye degradation [31,32,33,34], and especially CO2 conversion [35,36,37,38]. This method not only does not increase the recombination rate of electrons and holes but also effectively suppresses it and improves some other optical properties. The SPD method has also been used effectively to synthesize new families of catalysts, such as high-pressure photocatalysts and high-entropy photocatalysts [23,27].
This article reviews recent publications on the impact of SPD through the HPT method on photocatalytic activity for CO2 conversion. The four main strategies used for this purpose are discussed in detail: (i) oxygen vacancy and strain engineering, (ii) stabilization of high-pressure phases, (iii) synthesis of defective high-entropy oxides, and (iv) synthesis of low-bandgap high-entropy oxynitrides.

2. Influence of HPT on Photocatalytic CO2 Conversion

HPT as an SPD method has been used since 1935 until now for grain refinement and the production of nanostructured materials. In addition to grain refinement, the introduction of various defects, such as vacancies and dislocations, is another feature of HPT, which resulta in the improvement of the functionality of materials proceeded by this method [39,40]. In the HPT method, both large shear strain and high pressure (in the range of several gigapascals) are simultaneously utilized to process or synthesize various ranges of materials [39,40]. Strain and pressure are applied to the material (disc or ring shape) using two anvils that rotate with respect to each other, as shown in Figure 1b [41]. Due to the high processing pressure in HPT, it is applicable to hard and less ductile materials, such as high-melting temperature metals (hafnium [42], molybdenum [43], and tungsten [44]), amorphous glasses [45,46], silicon-based semiconductors [47,48], and even ultrahard diamond [49,50]. Another advantage of HPT is its capacity to induce ultra-SPD (i.e., shear strains over 1000 for mechanical alloying) [51]. The inducing ultra-SPD [51] together with fast dynamic diffusion [52,53] introduces the HPT method as a unique path to mechanically synthesize new materials even from immiscible systems [54,55]. Due to these unique features of HPT, the method was even used for the process and synthesis of hard and brittle ceramics, but the number of publications on ceramics is quite limited despite the high potential of these materials for various applications [23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81]. Published studies regarding ceramics processed or synthesized by HPT are presented in Table 1, although there are other classic publications on HPT processing of ceramics mainly by physicists and geologists [40].
Figure 1. Schematic illustration of (a) photocatalytic CO2 conversion and (b) high-pressure torsion [41].
Figure 1. Schematic illustration of (a) photocatalytic CO2 conversion and (b) high-pressure torsion [41].
Materials 16 01081 g001
As given in Table 1, the recent usage of HPT to process and synthesize ceramics for photocatalysis, especially photocatalytic CO2 conversion, has shown a high potential of this method for the enhancement of photocatalytic activity [35,36,37,38]. The HPT method effectively leads to increased efficiency by narrowing the bandgap, increasing the light absorbance, aligning the band structure, introducing the interphases and active sites for chemical adsorption and reaction, and accelerating the charge carrier migration [35,36,37,38]. While the HPT method can control all these features simultaneously by simple mechanical treatment, chemical methods are not usually able to improve all these features at the same time. The main drawbacks of the HPT method are the small quantity of the sample and the low specific surface area of the catalyst due to the high pressure and strain utilized. However, upscaling the HPT method and increasing the specific surface area by a post-HPT treatment are issues that can be addressed in the future. The improvement of features of photocatalysts by HPT has been achieved using four main strategies, including simultaneous strain and oxygen vacancy engineering, the introduction of high-pressure phases, the formation of defective high-entropy phases, and the production of low-bandgap high-entropy oxynitride phases. The responsibility of each mentioned strategy to improve the photocatalytic CO2 conversion activity is discussed in detail as follows. It should be noted that all photocatalytic CO2 conversion experiments on HPT-processed catalysts were performed in an aqueous liquid phase inside a quartz photoreactor with a continuous flow of CO into the liquid phase and NaHCO3 as the sacrificial agent.

2.1. Simultaneous Strain and Oxygen Vacancy Engineering

Oxygen vacancy engineering is an effective method that has been used to improve photocatalytic CO2 conversion. Oxygen vacancies increase the photocatalytic efficiency by increasing the light absorbance, accelerating the charge carrier separation, and enhancing the surface reactions [35,36]. Oxygen vacancies on the surface of the photocatalysts act as active sites to trap the electrons for various ranges of reduction reactions. It was also observed that oxygen vacancies have a significant role in adsorbing and activating the CO2 molecules and increasing the local electronic density [35,36].
BiVO4 is one of the common photocatalysts utilized for photocatalytic CO2 conversion, but it suffers from a high recombination rate of electrons and holes and an inappropriate conduction band position [35]. Different strategies have been used to solve these problems, but in all of them, impurity atoms or a second phase are added to this material [35]. The HPT method was used to solve the problems of BiVO4 for photocatalytic CO2 conversion by simultaneous engineering of strain and oxygen vacancies without the addition of impurities. BiVO4 was processed by HPT for N = 0.25, 1, and 4 turns to investigate the impact of strain on photocatalytic properties and efficiency. Increasing the lattice strain and decreasing the crystallite size by increasing the HPT turns is shown in Figure 2a. The occurrence of lattice strain was also confirmed by Raman peak shift to lower wavenumbers, as shown in Figure 2b. It was also observed that the concentration of oxygen vacancies increases in BiVO4 by increasing the applied shear strain. Figure 2c illustrates the oxygen vacancy concentration, calculated by X-ray photoelectron spectroscopy (XPS), against the number of HPT turns, confirming that the concentration of vacancies increases by increasing the applied shear strain. Furthermore, strain and vacancy engineering led to an increase in light absorbance (Figure 2d) and a slight narrowing of the bandgap from 2.4 eV for the initial powder to 2.1 eV for the sample proceeded by HPT for N = 4 turns [35].
Simultaneous strain and oxygen vacancy engineering could significantly solve the problem of BiVO4 in terms of the high recombination rate of electrons and holes, as shown in Figure 2e. This figure demonstrates that the HPT method decreases the photoluminescence intensity, which is a piece of evidence for the suppression of recombination. Finally, this strategy was successful in improving the photocatalytic activity of BiVO4, as shown in Figure 2f. The CO production rate from CO2 photoreduction was effectively increased by increasing the number of HPT turns. This study was the first successful work that used simultaneous strain and oxygen vacancy engineering to improve the photocatalytic activity of BiVO4 without using impurities, suggesting SPD as a new path to improve the optical and electronic structure of photocatalysts for CO2 conversion [35].

2.2. Introducing High-Pressure Phases

The formation of high-pressure phases is one of the HPT effects that can occur for some ceramics, such as TiO2 [65], ZrO2 [58], ZnO [26], SiO2 [34], VO2 [78], Y2O3 [66], BaTiO3 [64], Al2O3 [27], and BN [68]. It was observed that these high-pressure phases contain defects, such as oxygen vacancies and dislocations, and have nanosized grains, which makes them attractive for photocatalytic applications. TiO2 with the anatase and rutile crystal structures is one of the most active photocatalysts for CO2 conversion. As shown in Figure 3a, in addition to anatase and rutile, TiO2 has a high-pressure TiO2-II (columbite) phase with the orthorhombic structure. Despite many studies on TiO2 photocatalysts, there was not any research work on photocatalytic CO2 conversion on the TiO2-II phase until 2021. Groups of current authors stabilized the TiO2-II phase by the HPT method and investigated it for photocatalytic CO2 conversion [36]. To decrease the fraction of oxygen vacancies in the bulk, which can act as recombination centers, an HPT-processed sample was further treated by annealing [36]. The formation of high-pressure TiO2-II was proved by X-ray diffraction (XRD), Raman spectroscopy, and transmission electron microscopy (TEM). Raman spectra along with the appearance of samples are shown in Figure 3b. New Raman peaks at wavenumbers 171, 283, 316, 340, 357, 428, 533, and 572 cm−1 correspond to the TiO2-II phase. The changes in the color of the sample from white to dark green after HPT processing and from dark green to white after annealing indicate that large fractions of oxygen vacancies are formed after HPT processing, while some of them are annihilated after annealing, a fact that was also proved by various characterization techniques [36].
The light absorbance of the TiO2-II phase produced by HPT processing was higher, and it had a narrower optical bandgap of 2.5 eV compared with anatase (3 eV), although the bandgap slightly increased to 2.7 eV after annealing [36]. Introducing the high-pressure TiO2-II phase using HPT suppressed the recombination rate of electrons and holes. It also had a positive impact on photocurrent generation, as shown in Figure 3c so that the annealed sample had the highest current density, suggesting the improvement of charge carrier separations by introducing the high-pressure TiO2-II phase. The potential of this new phase for CO2 adsorption was measured by attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy. It was observed that the annealed sample had the highest potential for CO2 adsorption, which can help with photocatalytic CO2 conversion. Finally, this new phase showed a higher potential for photocatalytic CO production compared with the anatase phase, as shown in Figure 3d. The introduction of the TiO2-II phase with an optimized fraction of oxygen vacancies significantly improved the activity so that the annealed sample had the highest efficiency for CO2-to-CO conversion. The formation of anatase–columbite interphases can also contribute to the high activity of the HPT-processed sample by increasing the electron–hole separation and migration. In conclusion, high-pressure phases show great potential to be used as photocatalysts, and SPD can be used to stabilize these high-pressure phases under ambient conditions [36].

2.3. Formation of Defective High-Entropy Phases

Introducing high-entropy ceramics as new materials with five or more principal elements opened a new path in the field of materials science to produce materials with high functionality for various applications [81,82]. High configurational entropy caused by a large number of elements in these materials leads to decreasing the Gibbs free energy and improving the phase stability. High-entropy ceramics have been utilized for various applications, and in many cases, they have shown better efficiencies than conventional materials [81,82]. Li-ion batteries [83], catalysts [84], dielectrics [85], magnetic components [86], thermal barrier coating [87], and so on are some of the applications of these materials. The high potential of high-entropy ceramics for various applications is attributed to their high stability, cocktail effect, lattice distortion, inherent defects, and valence electron distribution [81,82]. Despite the high functionality of these materials, their application for photocatalytic CO2 conversion was not investigated until a study was conducted by the current authors in 2022 [37].
The HPT method, followed by oxidation, was used to fabricate a defective high-entropy oxide (HEO) with the composition of TiZrNbHfTaO11 and dual crystal structure of monoclinic and orthorhombic [37]. The selection of elements for this high-entropy ceramic was conducted by considering the d0 electronic structure of cations that have shown high potential for photocatalysis. The oxidation states of anionic and cationic elements and their uniform distribution were proved by XPS and energy-dispersive X-ray spectroscopy (EDS), respectively. The microstructure of the oxide is shown in Figure 4a using scanning electron microscopy (SEM) and in Figure 4b using high-resolution TEM. In addition to a nanocrystalline dual-phase structure, the material exhibited the presence of various defects, such as vacancies and dislocations, as shown in Figure 4b. The formation of oxygen vacancies in this material was examined by electron paramagnetic resonance (EPR) spectroscopy. These oxygen vacancies can act as shallow traps between the valence band and the conduction band for easier charge carrier separation, as shown in Figure 4c [37].
TiZrNbHfTaO11 had a higher light absorbance and lower bandgap compared with the binary oxides, including TiO2, ZnO, Nb2O5, HfO3, and Ta2O5 [37]. This HEO could successfully generate photocurrent, which indicates its potential for easy separation of electrons and holes to improve photocatalytic activity. TiZrNbHfTaO11 showed higher activity for photocatalytic CO production compared with BiVO4 and TiO2 as two typical photocatalysts, as shown in Figure 4d. Additionally, it had the same photocatalytic activity as P25 TiO2 as a benchmark photocatalyst. The high activity of this HEO for photocatalytic CO2 conversion was attributed to the presence of defects, such as oxygen vacancies and dislocations; interphases; and high light absorbance. This study reported the first application of high-entropy ceramics for photocatalytic CO2 conversion and introduced a new way to design and synthesize highly efficient high-entropy photocatalysts by SPD processing [37].

2.4. Synthesis of Low-Bandgap High-Entropy Oxynitrides

Metal oxides are the most conventional photocatalysts for CO2 conversion but suffer from a large bandgap. On the other hand, metal nitrides have a low bandgap but suffer from low stability compared with metal oxides. Metal oxynitrides are rather new materials that can solve the problems of metal oxides and nitrides in terms of large bandgap and low stability, respectively [88]. Although oxynitrides have been used for photocatalytic water splitting in many research works, their application for photocatalytic CO2 conversion has been limitedly investigated mainly due to their limited chemical stability. The concept of high-entropy materials with high stability is one strategy used to produce high-entropy oxynitrides with low bandgap and high stability for CO2 photoreduction [38].
A high-entropy oxynitride (HEON) with the composition of TiZrNbHfTaO6N3 was fabricated by the HPT method, followed by oxidation and nitriding, and its photocatalytic performance was compared with a corresponding HEO TiZrNbHfTaO6 and P25 TiO2 benchmark photocatalyst [38]. This HEON had dual phases with face-centered cubic (FCC) and monoclinic structures with uniform distribution of elements. This HEON material had much higher light absorbance compared with P25 TiO2 and relevant HEO, as shown in Figure 5a. It showed a superior low bandgap of 1.6 eV as one of the lowest bandgaps reported in the literature for oxynitride photocatalysts. The improved electronic band structure of this HEON compared with P25 TiO2 and HEO is shown in Figure 5b. The recombination rate of electrons and holes in HEON was low so that its photoluminescence intensity was negligible compared with P25 TiO2 and HEO (Figure 5c). The shape of photocurrent spectra shown in Figure 5d also confirmed the low recombination rate of electrons and holes in this HEON compared with the HEO and P25 TiO2 catalysts. The potential of this HEON for CO2 adsorption was measured by diffuse reflectance infrared Fourier transform (DRIFT) spectrometry, which showed the higher physical adsorption and chemisorption (in the form of carbonate) of CO2 on this HEON compared with P25 TiO2 and HEO (Figure 5e).
This HEON successfully converted CO2 to CO with extremely high efficiency even compared with the P25 TiO2 benchmark photocatalyst, as shown in Figure 5f. Although HEON could adsorb the light in both visible and infrared regions of light, it could not convert CO2 in these regions within the detection limits of the gas chromatograph. The stability of HEON was examined by conducting a long-term photocatalytic test for 20 h after storage of the sample in the air for 6 months. The photocatalytic activity of the material was not degraded, and X-ray diffraction analysis confirmed that the crystal structure of HEON did not change after 6-month storage and the long-time photocatalytic reaction. In conclusion, the low-bandgap HEON catalysts synthesized by SPD can be considered a new family of highly efficient photocatalysts for CO2 conversion [38].

3. Discussion on Future Outlook

The application of SPD to synthesize new photocatalysts for CO2 conversion introduced significant findings from the viewpoints of photocatalysis and SPD. The significance of these issues and their impact on the future outlook of this research field are discussed here.
For all these photocatalysts developed by HPT, CO was the only product that was detected using a flame ionization detector. The nonproduction of other products, such as CH4, CH3OH, HCOOH, or CH2=CH2, can be explained by considering the thermodynamic and kinetic parameters. For instance, CH4 is a product that thermodynamically is more feasible to be produced than CO due to its lower standard potential. However, more electrons are required to produce this component compared with CO [89]. Therefore, from the viewpoint of kinetics, CO production is more feasible than CH4 formation. Another point that should be considered is that CO has no tendency to be adsorbed to the active sites of the photocatalysts after production, which leads to propelling the reaction to CO production [89]. The production of CO as the only product can also be explained by the pathway of the reaction. In photocatalytic CO2 conversion, the formation of a CO 2 intermediate product is the initial step. This intermediate product is formed by interchanging the electrons between CO2 and the surface of the catalyst. Adsorption modes of CO 2 to the surface of the photocatalyst specify the reaction pathway. The CO 2 intermediate product can be adsorbed to the surface of the photocatalyst by three modes, which include (i) oxygen coordination, (ii) carbon coordination, and (iii) combination of oxygen and carbon coordination [90]. Oxygen coordination occurs when the photocatalyst is formed from tin, lead, mercury, indium, and cadmium metals. In this case, OCHO and formic acid are produced as intermediate and final products, respectively. If the noble and transition metals are the elements forming the photocatalyst, carbon coordination occurs and CO and CO are the intermediate and final products, respectively [90]. The presence of copper atoms in the structure of photocatalysts leads to the formation of a combination of oxygen and carbon coordination to produce both OCHO and CO as intermediates and CO, CH4, and C2H5OH as final products. Since all photocatalysts investigated by HPT include the transition metals, CO is the final product, and the reaction pathway can be considered as follows [90].
CO 2 + e CO 2
CO 2 + 2 e + 2 H + CO +   H 2 O
Table 2 compares the photocatalytic activity of HEON synthesized by HPT with reported photocatalysts in the literature by normalizing the amount of CO production to catalyst mass and surface area [91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123]. Since the photocatalytic reaction occurs on the surface, comparing the results by normalizing them to the surface area is more reasonable. According to this table, the amount of CO production for HEON is 4.66 ± 0.3 µmolh−1m−1, which is higher than the best photocatalysts reported in the literature. This indicates that the contribution of SPD to introducing new families of photocatalysts will receive high appreciation in the future by considering the current demands in finding new strategies to deal with the CO2 emissions; however, the synthesis method and compositions are expected to be modified by the experts in the field of photocatalysis. For example, one main disadvantage of SPD for the process and synthesis of catalysts is the low surface area of the synthesized material, while large specific surface areas are desirable in catalysis [74]. Moreover, theoretical studies are required to clarify the mechanisms underlying the high activity of photocatalysts developed by SPD so that new catalysts can be designed.
The SPD field experienced significant progress in the past three decades, as discussed in several review papers [124,125,126,127,128,129], and more recently in a special issue in 2019 [130], which gathered overviews on both historical developments [131] and recent advancements [132]. A survey of these overviews indicates that despite significant progress on theoretical aspects [133,134], mechanisms [135,136], processing [137,138,139,140,141,142,143,144], microstructure [145,146,147,148,149], and mechanical properties [150,151,152,153,154,155] of metallic materials, there is a recent tendency to apply SPD to a wider range of materials (oxides [156], semiconductors [157], carbon polymorphs [158], glasses [159], and polymers [160]) to control phase transformations [161] and solid-state reaction [162,163,164] for achieving advanced functional properties [165,166,167,168,169,170,171,172]. CO2 conversion is perhaps the newest application of SPD to functional materials, which expanded the synthesis capability of SPD from metallic materials to ceramics [37,38]. Moreover, this application has led to the introduction of new benchmark photocatalysts, which can open new pathways and research directions in corresponding fields. Although the application of SPD for CO2 photoreduction is currently limited to the HPT method, which produces only small amounts of samples, the fundamentals developed by HPT should be used in the future to develop new methods with upscaled sample sizes and higher potential for industrial applications. This last issue is a general requirement of SPD for future commercialization in almost any application [173].

4. Conclusions

Global warming has become a significant concern in recent years, which seriously threatens the life of creatures. Conversion of CO2 molecules to other components, such as CO, is a way to stand this event. In this regard, photocatalytic CO2 conversion, which uses solar irradiation as a clean energy source, has been introduced as a new and promising strategy in recent years. Despite the introduction of various materials, which are modified by various strategies, the efficiency of CO2 photoreduction is still low compared with conventional methods for CO2 conversion. High-pressure torsion (HPT) as a severe plastic deformation (SPD) method has been used recently to produce some of the most active photocatalysts for CO2 conversion. The HPT method can increase the CO2 photoreduction efficiency by (i) oxygen vacancy and strain engineering, (ii) the stabilization of high-pressure phases, (iii) the formation of defective high-entropy oxides, and (iv) the synthesis of low-bandgap oxynitrides.

Author Contributions

Conceptualization, S.A., T.I., M.F. and K.E.; writing—review and editing, S.A., T.I., M.F. and K.E. All authors have read and agreed to the published version of the manuscript.

Funding

The author S.A. thanks Hosokawa Powder Technology Foundation, Japan, for a grant. The author K.E. was supported in part by the MEXT, Japan, through Grants-in-Aid for Scientific Research (JP19H05176, JP21H00150, and JP22K18737), and in part by Mitsui Chemicals, Inc.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Forkel, M.; Carvalhais, N.; Rodenbeck, C.; Keeling, R.; Heimann, M.; Thonicke, K.; Zaehle, S.; Reichstein, M. Enhanced seasonal CO2 exchange caused by amplified plant productivity in northern ecosystems. Science 2016, 351, 696–699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Morris, A.J.; Meyer, G.J.; Fujita, E. Molecular approaches to the photocatalytic reduction of carbon dioxide for solar fuels. Acc. Chem. Res. 2009, 42, 1983–1994. [Google Scholar] [CrossRef] [PubMed]
  3. Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-photocatalytic materials: Possibilities and challenges. Adv. Mater. 2012, 24, 229–251. [Google Scholar] [CrossRef]
  4. Li, X.; Yu, J.; Jiang, C. Chapter 1—Principle and surface science of photocatalysis. Interface Sci. Technol. 2020, 31, 1–38. [Google Scholar]
  5. Ola, O.; Maroto-Valer, M.M. Synthesis, characterization and visible light photocatalytic activity of metal based TiO2 monoliths for CO2 reduction. Chem. Eng. J. 2016, 283, 1244–1253. [Google Scholar] [CrossRef] [Green Version]
  6. Neatu, S.; Macia-Agullo, J.A.; Concepcion, P.; Garcia, H. Gold-copper nanoalloys supported on TiO2 as photocatalysts for CO2 reduction by water. J. Am. Chem. Soc. 2014, 136, 15969–15976. [Google Scholar] [CrossRef]
  7. Kocí, K.; Mateju, K.; Obalova, L.; Krejcikova, S.; Lacny, Z.; Placha, D.; Capek, L.; Hospodkova, A.; Solcova, O. Effect of silver doping on the TiO2 for photocatalytic reduction of CO2. Appl. Catal. B 2010, 96, 239–244. [Google Scholar] [CrossRef]
  8. Akhundi, A.; Habibi-Yangjeh, A.; Abitorabi, M.; Pouran, S.R. Review on photocatalytic conversion of carbon dioxide to value-added compounds and renewable fuels by graphitic carbon nitride-based photocatalysts. Catal. Rev. Sci. Eng. 2019, 61, 595–628. [Google Scholar] [CrossRef]
  9. Xia, P.; Antonietti, M.; Zhu, B.; Heil, T.; Yu, J.; Cao, S. Designing defective crystalline carbon nitride to enable selective CO2 photoreduction in the gas phase. Adv. Funct. Mater. 2019, 29, 1900093. [Google Scholar] [CrossRef]
  10. Wang, K.; Zhang, L.; Su, Y.; Sun, S.; Wang, Q.; Wang, H.; Wang, W. Boosted CO2 photoreduction to methane via Co doping in bismuth vanadate atomic layers. Catal. Sci. Technol. 2018, 8, 3115–3122. [Google Scholar] [CrossRef]
  11. Huang, L.; Duan, Z.; Song, Y.; Li, Q.; Chen, L. BiVO4 microplates with oxygen vacancies decorated with metallic Cu and Bi nanoparticles for CO2 photoreduction. ACS Appl. Nano Mater. 2021, 4, 3576–3585. [Google Scholar] [CrossRef]
  12. Zhu, Z.; Yang, C.X.; Hwang, Y.T.; Lin, Y.C.; Wu, R.J. Fuel generation through photoreduction of CO2 on novel Cu/BiVO4. Mater. Res. Bull. 2020, 130, 110955. [Google Scholar] [CrossRef]
  13. Lee, G.J.; Anandan, S.; Masten, S.J.; Wu, J.J. Photocatalytic hydrogen evolution from water splitting using Cu doped ZnS microspheres under visible light irradiation. Renew. Energy 2016, 89, 18–26. [Google Scholar] [CrossRef]
  14. Liu, B.; Ye, L.; Wang, R.; Yang, J.; Zhang, Y.; Guan, R.; Tian, L.; Chen, X. Phosphorus-doped graphitic carbon nitride nanotubes with amino-rich surface for efficient CO2 capture, enhanced photocatalytic activity, and product selectivity. ACS Appl. Mater. Interfaces 2018, 10, 4001–4009. [Google Scholar] [CrossRef]
  15. Kuvarega, A.T.; Krause, R.W.M.; Mamba, B.B. Nitrogen/palladium-codoped TiO2 for efficient visible light photocatalytic dye degradation. J. Phys. Chem. C 2011, 115, 22110–22120. [Google Scholar] [CrossRef]
  16. Bai, S.; Zhang, N.; Gao, C.; Xiong, Y. Defect engineering in photocatalytic materials. Nano Energy 2018, 53, 293–336. [Google Scholar] [CrossRef]
  17. Di, J.; Zhu, C.; Ji, M.; Duan, M.; Long, R.; Yan, C.; Gu, K.; Xiong, J.; She, Y.; Xia, J.; et al. Defect-rich Bi12O17Cl2 nanotubes self-accelerating charge separation for boosting photocatalytic CO2 reduction. Angew. Chem. 2018, 57, 14847–14851. [Google Scholar] [CrossRef]
  18. Cai, X.; Wang, F.; Wang, R.; Xi, Y.; Wang, A.; Wang, J.; Teng, B.; Baiy, S. Synergism of surface strain and interfacial polarization on Pd@Au core-shell cocatalysts for highly efficient photocatalytic CO2 reduction over TiO2. J. Mater. Chem. A 2020, 8, 7350–7359. [Google Scholar] [CrossRef]
  19. Liu, Z.; Menendez, C.; Shenoy, J.; Hart, J.N.; Sorrell, C.C.; Cazorl, C. Strain engineering of oxide thin films for photocatalytic applications. Nano Energy 2020, 72, 104732. [Google Scholar] [CrossRef] [Green Version]
  20. Li, Y.; Wang, W.N.; Zhan, Z.; Woo, M.H.; Wu, C.Y.; Biswas, P. Photocatalytic reduction of CO2 with H2O on mesoporous silica supported Cu/TiO2 catalysts. Appl. Catal. B 2010, 100, 386–392. [Google Scholar] [CrossRef]
  21. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D heterojunction of ultrathin MXene/ Bi2WO6 nanosheets for improved photocatalytic CO2 reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  22. Li, J.; Shao, W.; Geng, M.; Wan, S.; Ou, M.; Chen, Y. Combined Schottky junction and doping effect in CdxZn1-xS@Au/BiVO4 Z-Scheme photocatalyst with boosted carriers charge separation for CO2 reduction by H2O. J. Colloid Interface Sci. 2022, 606, 1469–1476. [Google Scholar] [CrossRef] [PubMed]
  23. Razavi-Khosroshahi, H.; Edalati, K.; Hirayama, M.; Emami, H.; Arita, M.; Yamauchi, M.; Hagiwara, H.; Ida, S.; Ishihara, T.; Akiba, E.; et al. Visible- light-driven photocatalytic hydrogen generation on nanosized TiO2-II stabilized by high-pressure torsion. ACS Catal. 2016, 6, 5103–5107. [Google Scholar] [CrossRef]
  24. Edalati, K.; Fujiwara, K.; Takechi, S.; Wang, Q.; Arita, M.; Watanabe, M.; Sauvage, X.; Ishihara, T.; Horita, Z. Improved photocatalytic hydrogen evolution on tantalate perovskites CsTaO3 and LiTaO3 by strain-induced vacancies. ACS Appl. Energy Mater. 2020, 3, 1710–1718. [Google Scholar] [CrossRef]
  25. Edalati, K.; Uehiro, R.; Takechi, S.; Wang, Q.; Arita, M.; Watanabe, M.; Ishihara, T.; Horita, Z. Enhanced photocatalytic hydrogen production on GaN-ZnO oxynitride by introduction of strain-induced nitrogen vacancy complexes. Acta Mater. 2020, 185, 149–156. [Google Scholar] [CrossRef]
  26. Wang, Q.; Edalati, K.; Koganemaru, Y.; Nakamura, S.; Watanabe, M.; Ishihara, T.; Horita, Z. Photocatalytic hydrogen generation on low-bandgap black zirconia (ZrO2) produced by high-pressure torsion. J. Mater. Chem. A 2020, 8, 3643–3650. [Google Scholar] [CrossRef]
  27. Edalati, P.; Wang, Q.; Razavi-Khosroshahi, H.; Fuji, M.; Ishihara, T.; Edalati, K. Photocatalytic hydrogen evolution on a high entropy oxide. J. Mater. Chem. A 2020, 8, 3814–3821. [Google Scholar] [CrossRef]
  28. Hidalgo-Jimeneza, J.; Wang, Q.; Edalatib, K.; Cubero-Sesína, J.M.; Razavi-Khosroshahid, H.; Ikomac, Y.; Gutiérrez-Fallase, D.; Dittel-Mezaa, F.A.; Rodríguez-Rufinoa, J.C.; Fujid, M.; et al. Phase transformations, vacancy formation and variations of optical and photocatalytic properties in TiO2-ZnO composites by high pressure torsion. Int. J. Plast. 2020, 124, 170–185. [Google Scholar] [CrossRef]
  29. Edalati, P.; Shen, X.F.; Watanabe, M.; Ishihara, T.; Arita, M.; Fuji, M.; Edalati, K. High-entropy oxynitride as a low-bandgap and stable photocatalyst for hydrogen production. J. Mater. Chem. A 2021, 9, 15076–15086. [Google Scholar] [CrossRef]
  30. Edalati, P.; Itagoe, Y.; Ishihara, H.; Ishihara, T.; Emami, H.; Arita, M.; Fuji, M.; Edalati, K. Visible-light photocatalytic oxygen production on a high-entropy oxide by multiple-heterojunction introduction. J. Photochem. Photobiol. A 2022, 433, 114167. [Google Scholar] [CrossRef]
  31. Razavi-Khosroshahi, H.; Edalati, K.; Wu, J.; Nakashima, Y.; Arita, M.; Ikoma, Y.; Sadakiyo, M.; Inagaki, Y.; Staykov, A.; Yamauchi, M.; et al. High-pressure zinc oxide phase as visible-light-active photocatalyst with narrow band gap. J. Mater. Chem. A 2017, 5, 20298–20303. [Google Scholar] [CrossRef]
  32. Edalati, K.; Fujita, I.; Takechi, S.; Nakashima, Y.; Kumano, K.; Razavi-Khosroshahi, H.; Arita, M.; Watanabe, M.; Sauvage, X.; Akbay, T.; et al. Photocatalytic activity of aluminum oxide by oxygen vacancy generation using high-pressure torsion straining. Scr. Mater. 2019, 173, 120–124. [Google Scholar] [CrossRef]
  33. Fujita, I.; Edalati, K.; Wang, Q.; Arita, M.; Watanabe, M.; Munetoh, S.; Ishihara, T.; Horita, Z. High-pressure torsion to induce oxygen vacancies in nanocrystals of magnesium oxide: Enhanced light absorbance, photocatalysis and significance in geology. Materialia 2020, 11, 100670. [Google Scholar] [CrossRef]
  34. Wang, Q.; Edalati, K.; Fujita, I.; Watanabe, M.; Ishihara, T.; Horita, Z. High-pressure torsion of SiO2 quartz sand: Phase transformation, optical properties, and significance in geology. J. Am. Ceram. Soc. 2020, 103, 6594–6602. [Google Scholar] [CrossRef]
  35. Akrami, S.; Murakami, Y.; Watanabe, M.; Ishihara, T.; Arita, M.; Guo, Q.; Fuji, M.; Edalati, K. Enhanced CO2 conversion on highly-strained and oxygen-deficient BiVO4 photocatalyst. Chem. Eng. J. 2022, 442, 136209. [Google Scholar] [CrossRef]
  36. Akrami, S.; Watanabe, M.; Ling, T.H.; Ishihara, T.; Arita, M.; Fuji, M.; Edalati, K. High pressure TiO2-II polymorph as an active photocatalyst for CO2 to CO conversion. Appl. Catal. B 2021, 298, 120566. [Google Scholar] [CrossRef]
  37. Akrami, S.; Murakami, Y.; Watanabe, M.; Ishihara, T.; Arita, M.; Fuji, M.; Edalati, K. Defective high-entropy oxide photocatalyst with high activity for CO2 conversion. Appl. Catal. B 2022, 303, 120896. [Google Scholar] [CrossRef]
  38. Akrami, S.; Edalati, P.; Shundo, Y.; Watanabe, M.; Ishihara, T.; Fuji, M.; Edalati, K. Significant CO2 photoreduction on a high-entropy oxynitride. Chem. Eng. J. 2022, 449, 137800. [Google Scholar]
  39. Edalati, K.; Bachmaier, A.; Beloshenko, V.A.; Beygelzimer, Y.; Blank, V.D.; Botta, W.J.; Bryła, K.; Čížek, J.; Divinski, S.; Enikeev, N.A.; et al. Nanomaterials by severe plastic deformation: Review of historical developments and recent advances. Mater. Res. Lett. 2022, 10, 163–256. [Google Scholar] [CrossRef]
  40. Edalati, K.; Horita, Z. A review on high-pressure torsion (HPT) from 1935 to 1988. Mater. Sci. Eng. A 2016, 652, 325–352. [Google Scholar] [CrossRef]
  41. Edalati, K.; Horita, Z. Scaling-up of high pressure torsion using ring shape. Mater. Trans. 2009, 50, 92–95. [Google Scholar] [CrossRef] [Green Version]
  42. Edalati, K.; Horita, Z.; Mine, Y. High-pressure torsion of hafnium. Mater. Sci. Eng. A 2010, 527, 2136–2141. [Google Scholar] [CrossRef]
  43. Lee, S.; Edalati, K.; Horita, Z. Microstructures and mechanical properties of pure V and Mo processed by high-pressure torsion. Mater. Trans. 2010, 51, 1072–1079. [Google Scholar] [CrossRef] [Green Version]
  44. Chen, L.; Ping, L.; Ye, T.; Lingfeng, L.; Kemin, X.; Meng, Z. Observations on the ductility and thermostability of tungsten processed from micropowder by improved high-pressure torsion. Rare Met. Mater. Eng. 2016, 45, 3089–3094. [Google Scholar] [CrossRef] [Green Version]
  45. Edalati, K.; Yokoyama, Y.; Horita, Z. High-pressure torsion of machining chips and bulk discs of amorphous Zr50Cu30Al10Ni10. Mater. Trans. 2010, 51, 23–26. [Google Scholar] [CrossRef] [Green Version]
  46. Wang, Y.B.; Qu, D.D.; Wang, X.H.; Cao, Y.; Liao, X.Z.; Kawasaki, M.; Ringer, S.P.; Shan, Z.W.; Langdon, T.G.; Shen, J. Introducing a strain-hardening capability to improve the ductility of bulk metallic glasses via severe plastic deformation. Acta Mater. 2012, 60, 253–260. [Google Scholar] [CrossRef]
  47. Edalati, K.; Horita, Z. Correlations between hardness and atomic bond parameters of pure metals and semi-metals after processing by high-pressure torsion. Scr. Mater. 2011, 64, 161–164. [Google Scholar] [CrossRef]
  48. Ikoma, Y.; Hayano, K.; Edalati, K.; Saito, K.; Guo, Q.; Horita, Z. Phase transformation and nanograin refinement of silicon by processing through high-pressure torsion. Appl. Phys. Lett. 2012, 101, 121908. [Google Scholar] [CrossRef]
  49. Blank, V.D.; Churkin, V.D.; Kulnitskiy, B.A.; Perezhogin, I.A.; Kirichenko, A.N.; Erohin, S.V.; Sorokin, P.B.; Popov, M.Y. Pressure-induced transformation of graphite and diamond to onions. Crystals 2018, 8, 68. [Google Scholar] [CrossRef] [Green Version]
  50. Gao, Y.; Ma, Y.; An, Q.; Levitas, V.I.; Zhang, Y.; Feng, B.; Chaudhuri, J.; Goddard, W.A. III: Shear driven formation of nano-diamonds at sub-gigapascals and 300 K. Carbon 2019, 146, 364–368. [Google Scholar] [CrossRef] [Green Version]
  51. Edalati, K.; Uehiro, R.; Fujiwara, K.; Ikeda, Y.; Li, H.W.; Sauvage, X.; Valiev, R.Z.; Akiba, E.; Tanaka, I.; Horita, Z. Ultra-severe plastic deformation: Evolution of microstructure, phase transformation and hardness in immiscible magnesium-based systems. Mater. Sci. Eng. A 2017, 701, 158–166. [Google Scholar] [CrossRef]
  52. Oberdorfer, B.; Lorenzoni, B.; Unger, K.; Sprengel, W.; Zehetbauer, M.; Pippan, R.; Wurschum, R. Absolute concentration of free volume-type defects in ultrafine-grained Fe prepared by high-pressure torsion. Scr. Mater. 2010, 63, 452–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Straumal, B.B.; Mazilkin, A.A.; Baretzky, B.; Schütz, G.; Rabkin, E.; Valiev, R.Z. Accelerated diffusion and phase transformations in Co-Cu alloys driven by the severe plastic deformation. Mater. Trans. 2012, 53, 63–71. [Google Scholar] [CrossRef] [Green Version]
  54. Edalati, K.; Emami, H.; Staykov, A.; Smith, D.J.; Akiba, E.; Horita, Z. Formation of metastable phases in magnesium-titanium system by high-pressure torsion and their hydrogen storage performance. Acta Mater. 2015, 50, 150–156. [Google Scholar] [CrossRef]
  55. Kormout, K.S.; Pippan, R.; Bachmaier, A. Deformation-induced supersaturation in immiscible material systems during high-pressure torsion. Adv. Eng. Mater. 2017, 19, 1600675. [Google Scholar] [CrossRef]
  56. Bridgman, P.W. Effects of high shearing stress combined with high hydrostatic pressure. Phys. Rev. 1935, 48, 825–847. [Google Scholar] [CrossRef]
  57. Edalati, K.; Horita, Z. Application of high-pressure torsion for consolidation of ceramic Powders. Scr. Mater. 2010, 63, 174–177. [Google Scholar] [CrossRef]
  58. Edalati, K.; Toh, S.; Ikomaa, Y.; Horita, Z. Plastic deformation and allotropic phase transformations in zirconia ceramics during high-pressure torsion. Scr. Mater. 2011, 65, 974–977. [Google Scholar] [CrossRef]
  59. Makhnev, A.A.; Nomerovannaya, L.V.; Gizhevskii, B.A.; Naumov, S.V.; Kostromitin, N.V. Effect of redistribution of the optical spectral weight in CuO nanostructured ceramics. Solid State Phenom. 2011, 168–169, 285–288. [Google Scholar] [CrossRef]
  60. Gizhevskii, B.A.; Sukhorukov, Y.P.; Nomerovannaya, L.V.; Makhnev, A.A.; Ponosov, Y.S.; Telegin, A.V.; Mostovshchikov, E.V. Features of optical properties and the electronic structure of nanostructured oxides of 3D-Metals. Solid State Phenom. 2011, 168–169, 317–320. [Google Scholar] [CrossRef]
  61. Delogu, F. Are processing conditions similar in ball milling and high-pressure torsion? The case of the tetragonal-to-monoclinic phase transition in ZrO2 powders. Scr. Mater. 2012, 67, 340–343. [Google Scholar] [CrossRef]
  62. Mostovshchikova, E.V.; Gizhevskii, B.A.; Loshkareva, N.N.; Galakhov, V.R.; Naumov, S.V.; Ponosov, Y.S.; Ovechkina, N.A.; Kostromitina, N.V.; Buling, A.; Neumann, M. Infrared and X-ray absorption spectra of Cu2O and CuO nanoceramics. Solid State Phenom. 2012, 190, 683–686. [Google Scholar] [CrossRef]
  63. Telegin, A.V.; Gizhevskii, B.A.; Nomerovannaya, L.V.; Makhnev, A.A. The optical and magneto-optical properties of nanostructured oxides of 3d-metals. J. Supercond. Nov. Magn. 2012, 25, 2683–2686. [Google Scholar] [CrossRef]
  64. Edalati, K.; Arimura, M.; Ikoma, Y.; Daio, T.; Miyata, M.; Smith, D.J.; Horita, Z. Plastic deformation of BaTiO3 ceramics by high-pressure torsion and changes in phase transformations, optical and dielectric properties. Mater. Res. Lett. 2015, 3, 216–221. [Google Scholar] [CrossRef]
  65. Razavi-Khosroshahi, H.; Edalati, K.; Arita, M.; Horita, Z.; Fuji, M. Plastic strain and grain size effect on high-pressure phase transformations in nanostructured TiO2 ceramics. Scr. Mater. 2016, 124, 59–62. [Google Scholar] [CrossRef]
  66. Razavi-Khosroshahi, H.; Edalati, K.; Emami, H.; Akiba, E.; Horita, Z.; Fuji, M. Optical properties of nanocrystalline monoclinic Y2O3 stabilized by grain size and plastic strain effects via high-pressure torsion. Inorg. Chem. 2017, 56, 2576–2580. [Google Scholar] [CrossRef] [PubMed]
  67. Kuznetsova, E.I.; Degtyarev, M.V.; Zyuzeva, N.A.; Bobylev, I.B.; Pilyugin, V.P. Microstructure of YBa2Cu3Oy subjected to severe plastic deformation by high pressure torsion. Phys. Met. Metallogr. 2017, 118, 815–823. [Google Scholar] [CrossRef]
  68. Feng, B.; Levitas, V.I. Coupled elastoplasticity and plastic strain-induced phase transformation under high pressure and large strains: Formulation and application to BN sample compressed in a diamond anvil cell. Int. J. Plast. 2017, 96, 156–181. [Google Scholar] [CrossRef]
  69. Qian, C.; He, Z.; Liang, C.; Ji, W. New in situ synthesis method for Fe3O4/flake graphite nanosheet composite structure and its application in anode materials of lithium-ion batteries. J. Nanomater. 2018, 2018, 2417949. [Google Scholar] [CrossRef] [Green Version]
  70. Qi, Y.; Kosinova, A.; Kilmametov, A.R.; Straumal, B.B.; Rabkin, E. Plastic flow and microstructural instabilities during high-pressure torsion of Cu/ZnO composites. Mater. Charact. 2018, 145, 389–401. [Google Scholar] [CrossRef]
  71. Shabashov, V.A.; Makarov, A.V.; Kozlov, K.A.; Sagaradze, V.V.; Zamatovskii, A.E.; Volkova, E.G.; Luchko, S.N. Deformation-induced dissolution and precipitation of nitrides in austenite and ferrite of a high-nitrogen stainless steel. Phys. Met. Metallogr. 2018, 119, 193–204. [Google Scholar] [CrossRef]
  72. Feng, B.; Levitas, V.I.; Li, W. FEM modeling of plastic flow and strain-induced phase transformation in BN under high pressure and large shear in a rotational diamond anvil cell. Int. J. Plast. 2019, 113, 236–254. [Google Scholar] [CrossRef] [Green Version]
  73. Edalati, K.; Wang, Q.; Eguchi, H.; Razavi-Khosroshahi, H.; Emami, H.; Yamauchi, M.; Fuji, M.; Horita, Z. Impact of TiO2-II phase stabilized in anatase matrix by high-pressure torsion on electrocatalytic hydrogen production. Mater. Res. Lett. 2019, 7, 334–339. [Google Scholar] [CrossRef] [Green Version]
  74. Edalati, K. Review on recent advancements in severe plastic deformation of oxides by high-pressure torsion (HPT). Adv. Eng. Mater. 2019, 21, 1800272. [Google Scholar] [CrossRef]
  75. Permyakova, I.; Glezer, A. Amorphous-nanocrystalline composites prepared by high-pressure torsion. Metals 2020, 10, 511. [Google Scholar] [CrossRef] [Green Version]
  76. Fujita, I.; Edalati, P.; Wang, Q.; Watanabe, M.; Arita, M.; Munetoh, S.; Ishihara, T.; Edalati, K. Novel black bismuth oxide (Bi2O3) with enhanced photocurrent generation, produced by high-pressure torsion straining. Scr. Mater. 2020, 187, 366–370. [Google Scholar] [CrossRef]
  77. Wang, Q.; Watanabe, M.; Edalati, K. Visible-light photocurrent in nanostructured high-pressure TiO2-II (Columbite) phase. J. Phys. Chem. C 2020, 124, 13930–13935. [Google Scholar] [CrossRef]
  78. Edalati, K.; Fujita, I.; Sauvage, X.; Arita, M.; Horita, Z. Microstructure and phase transformations of silica glass and vanadium oxide by severe plastic deformation via high-pressure torsion straining. J. Alloys Compd. 2019, 779, 394–398. [Google Scholar] [CrossRef]
  79. Edalati, K.; Wang, Q.; Razavi-Khosroshahi, H.; Emami, H.; Fuji, M.; Horita, Z. Low-temperature anatase-to-rutile phase transformation and unusual grain coarsening in titanium oxide nanopowders by high-pressure torsion straining. Scr. Mater. 2019, 162, 341–344. [Google Scholar] [CrossRef]
  80. Qi, Y.; Kauffmann, Y.; Kosinova, A.; Kilmametov, A.R.; Straumal, B.B.; Rabkin, E. Gradient bandgap narrowing in severely deformed ZnO nanoparticles. Matter. Res. Lett. 2021, 9, 58–64. [Google Scholar] [CrossRef]
  81. Akrami, S.; Edalati, P.; Edalati, K.; Fuji, M. High-entropy ceramics: A review of principles, production and applications. Mater. Sci. Eng. R 2021, 146, 100644. [Google Scholar] [CrossRef]
  82. Oses, C.; Toher, C.; Curtarolo, S. High-entropy ceramics. Nat. Rev. Mater. 2020, 5, 295–309. [Google Scholar] [CrossRef]
  83. Sarkar, A.; Velasco, L.; Wang, D.; Wang, Q.; Talasila, G.; de Biasi, L.; Kübel, C.; Brezesinski, T.; Bhattacharya, S.S.; Hahn, H.; et al. High entropy oxides for reversible energy storage. Nat. Commun. 2018, 9, 3400. [Google Scholar] [CrossRef] [Green Version]
  84. Chen, H.; Lin, W.; Zhang, Z.; Jie, K.; Mullins, D.R.; Sang, X.; Yang, S.Z.; Jafta, C.J.; Bridges, C.A.; Hu, X. Mechanochemical synthesis of high entropy oxide materials under ambient conditions: Dispersion of catalysts via entropy maximization. ACS, Mater. Lett. 2019, 1, 83–88. [Google Scholar] [CrossRef]
  85. Radon, A.; Hawełek, D.; Łukowiec, J.; Kubacki, P.; Włodarczyk, P. Dielectric and electromagnetic interference shielding properties of high entropy (Zn, Fe, Ni, Mg, Cd)Fe2O4 ferrite. Sci. Rep. 2019, 9, 20078. [Google Scholar] [CrossRef] [Green Version]
  86. Witte, R.; Sarkar, A.; Kruk, R.; Eggert, B.; Brand, R.A.; Wende, H.; Hahn, H. High entropy oxides: An emerging prospect for magnetic rare-earth transition metal perovskites. Phys. Rev. Mater. 2019, 3, 034406. [Google Scholar] [CrossRef] [Green Version]
  87. Wright, A.J.; Huang, C.; Walock, M.J.; Ghoshal, A.; Murugan, M.; Luo, J. Sand corrosion, thermal expansion, and ablation of medium-and high-entropy compositionally complex fluorite oxides. J. Am. Ceram. Soc. 2021, 104, 448–462. [Google Scholar] [CrossRef]
  88. Takata, T.; Pan, C.; Domen, K. Recent progress in oxynitride photocatalysts for visible-light-driven water splitting. Sci. Technol. Adv. Mater. 2015, 16, 033506. [Google Scholar] [CrossRef]
  89. Wang, K.; Lu, J.; Lu, y.; Lau, C.H.; Zheng, Y.; Fan, X. Unravelling the C-C coupling in CO2 photocatalytic reduction with H2O on Au/TiO2-x: Combination of plasmonic excitation and oxygen vacancy. Appl. Catal. B 2021, 292, 120147. [Google Scholar] [CrossRef]
  90. Lu, H.; Tournet, J.; Dastafkan, K.; Liu, Y.; Ng, Y.H.; Karuturi, S.K.; Zhao, C.; Yin, Z. No-blemetal- free multicomponent nanointegration for sustainable energy conversion. Chem. Rev. 2021, 121, 10271–10366. [Google Scholar] [CrossRef]
  91. Wang, Y.; Chen, Y.; Zuo, Y.; Wang, F.; Yao, J.; Li, B.; Kang, S.; Li, X.; Cui, L. Hierarchically mesostructured TiO2/graphitic carbon composite as a new efficient photocatalyst for the reduction of CO2 under simulated solar irradiation. Catal. Sci. Technol. 2013, 3, 3286–3291. [Google Scholar] [CrossRef]
  92. Liu, L.J.; Pitts, D.T.; Zhao, H.L.; Zhao, C.Y.; Li, Y. Silver-incorporated bicrystalline (anatase/brookite) TiO2 microspheres for CO2 photoreduction with water in the presence of methanol. Appl. Catal. A 2013, 467, 474–482. [Google Scholar] [CrossRef]
  93. Sekizawa, K.; Maeda, K.; Domen, K.; Koike, K.; Ishitani, O. Artificial Z-scheme constructed with a supramolecular metal complex and semiconductor for the photocatalytic reduction of CO2. J. Am. Chem. Soc. 2013, 135, 4596–4599. [Google Scholar] [CrossRef] [PubMed]
  94. Tahir, M.; Amin, N.S. Photocatalytic reduction of carbon dioxide with water vapors over montmorillonite modified TiO2 nanocomposites. Appl. Catal. B 2013, 142–143, 512–522. [Google Scholar] [CrossRef]
  95. Zhang, Z.Y.; Wang, Z.; Cao, S.W.; Xue, C. Au/Pt nanoparticle-decorated TiO2 nanofibers with plasmon-enhanced photocatalytic activities for solar-to-fuel conversion. J. Phys. Chem. C 2013, 117, 25939–25947. [Google Scholar] [CrossRef]
  96. He, Z.; Wen, L.; Wang, D.; Xue, Y.; Lu, Q.; Wu, C.; Chen, J.; Song, S. Photocatalytic reduction of CO2 in aqueous solution on surface-fuorinated anatase TiO2 nanosheets with exposed {001} fcets. Energy Fuels 2014, 28, 3982–3993. [Google Scholar] [CrossRef]
  97. Fang, B.Z.; Bonakdarpour, A.; Reilly, K.; Xing, Y.L.; Taghipour, F.; Wilkinson, D.P. Large-scale synthesis of TiO2 microspheres with hierarchical nanostructure for highly efficient photodriven reduction of CO2 to CH4. ACS Appl. Mater. Interfaces 2014, 6, 15488–15498. [Google Scholar] [CrossRef]
  98. Li, K.; Lin, L.; Peng, T.; Guo, Y.; Li, R.; Zhang, J. Asymmetric zinc porphyrin-sensitized nanosized TiO2 for efficient visible-light-driven CO2 photoreduction to CO/CH4. J. Chem. Commun. 2015, 51, 12443–12446. [Google Scholar] [CrossRef]
  99. Fang, B.Z.; Xing, Y.L.; Bonakdarpour, A.; Zhang, S.C.; Wilkinson, D.P. Hierarchical CuO-TiO2 hollow microspheres for highly efficient photodriven reduction of CO2 to CH4. ACS Sustain. Chem. Eng. 2015, 3, 2381–2388. [Google Scholar] [CrossRef]
  100. Tahir, M.; Amin, N.S. Indium-doped TiO2 nanoparticles for photocatalytic CO2 reduction with H2O vapors to CH4. Appl. Catal. B 2015, 162, 98–109. [Google Scholar] [CrossRef]
  101. Xiong, Z.; Wang, H.B.; Xu, N.Y.; Li, H.L.; Fang, B.Z.; Zhao, Y.C.; Zhang, J.Y.; Zheng, C.G. Photocatalytic reduction of CO2 on Pt2+–Pt0/TiO2 nanoparticles under UV/Vis light irradiation: A combination of Pt2+ doping and Pt nanoparticles deposition. Int. J. Hydrog. Energy 2015, 40, 10049–10062. [Google Scholar] [CrossRef]
  102. Ye, L.; Wang, H.; Jin, X.; Su, Y.; Wang, D.; Xie, H.; Liu, X.; Liu, X. Synthesis of olive- green few-layered BiOI for efficient photoreduction of CO2 into solar fuels under visible/near-infrared light. Sol. Energy Mater. Sol. Cells 2016, 144, 732–739. [Google Scholar] [CrossRef]
  103. Nakada, A.; Nakashima, T.; Sekizawa, K.; Maeda, K.; Ishitani, O. Visible-light-driven CO2 reduction on a hybrid photocatalyst consisting of a Ru(II) binuclear complex and a Ag-loaded TaON in aqueous solutions. Chem. Sci. 2016, 7, 4364–4371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Ye, T.; Huang, W.; Zeng, L.; Li, M.; Shi, J. CeO2-x platelet from monometallic cerium layered double hydroxides and its photocatalytic reduction of CO2. Appl. Catal. B 2017, 210, 141–148. [Google Scholar] [CrossRef]
  105. Pastor, E.; Pesci, F.; Reynal, A.; Handoko, A.; Guo, M.; An, X.; Cowan, A.; Klug, D.; Durrant, J.; Tang, J. Interfacial charge separation in Cu2O/RuOx as a visible light driven CO2 reduction catalyst. Phys. Chem. Chem. Phys. 2014, 16, 5922–5926. [Google Scholar] [CrossRef] [Green Version]
  106. Jiao, J.; Wei, Y.; Zhao, Y.; Zhao, Z.; Duan, A.; Liu, J.; Pang, Y.; Li, J.; Jiang, G.; Wang, Y. AuPd/3DOM-TiO2 catalysts for photocatalytic reduction of CO2: High efficient separation of photogenerated charge carriers. Appl. Catal. B 2017, 209, 228–239. [Google Scholar] [CrossRef]
  107. Kim, C.; Cho, K.M.; Al-Saggaf, A.; Gereige, I.; Jung, H.T. Z-scheme photocatalytic CO2 conversion on three-dimensional BiVO4/carbon-coated Cu2O nanowire arrays under visible light. ACS Catal. 2018, 8, 4170–4177. [Google Scholar] [CrossRef]
  108. Wang, J.; Qin, C.; Wang, H.; Chu, M.; Zada, A.; Zhang, X.; Li, J.; Raziq, F.; Qu, Y.; Jing, L. Exceptional photocatalytic activities for CO2 conversion on AlO bridged g-C3N4/α-Fe2O3 z-scheme nanocomposites and mechanism insight with isotopes Z. Appl. Catal. B 2018, 224, 459–466. [Google Scholar] [CrossRef]
  109. Jiang, H.; Katsumata, K.; Hong, J.; Yamaguchi, A.; Nakata, K.; Terashima, C.; Matsushita, N.; Miyauchi, M.; Fujishima, A. Photocatalytic reduction of CO2 on Cu2O-loaded Zn-Cr layered double hydroxides. Appl. Catal. B 2018, 224, 783–790. [Google Scholar] [CrossRef]
  110. Crake, A.; Christoforidis, K.C.; Godin, R.; Moss, B.; Kafizas, A.; Zafeiratos, S.; Durrant, J.R.; Petit, C. Titanium dioxide/carbon nitride nanosheet nanocomposites for gas phase CO2 photoreduction under UV-visible irradiation. Appl. Catal. B 2019, 242, 369–378. [Google Scholar] [CrossRef]
  111. Li, Y.; Wang, C.; Song, M.; Li, D.; Zhang, X.; Liu, Y. TiO2-x/CoOx Photocatalyst sparkles in photothermocatalytic reduction of CO2 with H2O steam. Appl. Catal. B 2019, 243, 760–770. [Google Scholar] [CrossRef]
  112. Bai, Y.; Yang, P.; Wang, L.; Yang, B.; Xie, H.; Zhou, Y.; Ye, L. Ultrathin Bi4O5Br2 nanosheets for selective photocatalytic CO2 conversion into CO. Chem. Eng. J. 2019, 360, 473–482. [Google Scholar] [CrossRef]
  113. Maiti, D.; Meier, A.J.; Cairns, J.; Ramani, S.; Martinet, K.; Kuhn, J.N.; Bhethanabotla, V.R. Intrinsically strained noble metal-free oxynitrides for solar photoreduction of CO2. Dalton Trans. 2019, 48, 12738–12748. [Google Scholar] [CrossRef] [PubMed]
  114. Kozlova, E.A.; Lyulyukin, M.N.; Markovskaya, D.V.; Selishchev, D.S.; Cherepanova, S.V.; Kozlov, D.V. Synthesis of Cd1-xZnxS photocatalysts for gas-phase CO2 reduction under visible light. Photochem. Photobiol. Sci. 2019, 18, 871–877. [Google Scholar] [CrossRef] [PubMed]
  115. Jin, X.; Lv, C.; Zhou, X.; Ye, L.; Xie, H.; Liu, Y.; Su, H.; Zhang, B.; Chen, G. Oxygen vacancies engineering Bi24O31Cl10 photocatalyst for boosted CO2 conversion. ChemSusChem 2019, 12, 2740–2747. [Google Scholar] [CrossRef]
  116. Guo, S.; Di, J.; Chen, C.; Zhu, C.; Duan, M.; Lian, C.; Ji, M.; Zhou, W.; Xu, M.; Song, P.; et al. Oxygen vacancy mediated bismuth stannate ultra-small nanoparticle towards photocatalytic CO2-to-CO conversion. Appl. Catal. B 2020, 276, 119156. [Google Scholar] [CrossRef]
  117. Duan, Z.; Zhao, X.; Wei, C.; Chen, L. Ag-Bi/BiVO4 chain-like hollow microstructures with enhanced photocatalytic activity for CO2 conversion. Appl. Catal. A 2020, 594, 117459. [Google Scholar] [CrossRef]
  118. Chen, Y.; Wang, F.; Cao, Y.; Zhang, F.; Zou, Y.; Huang, Z.; Ye, L.; Zhou, Y. Interfacial oxygen vacancy engineered two-dimensional g-C3N4/BiOCl heterostructures with boosted photocatalytic conversion of CO2. ACS Appl. Energy Mater. 2020, 3, 4610–4618. [Google Scholar] [CrossRef]
  119. Dao, X.Y.; Xie, X.F.; Guo, J.H.; Zhang, X.Y.; Kang, Y.S.; Sun, W.Y. Boosting photocatalytic CO2 reduction efficiency by heterostructues of NH2-MIL-101 (Fe)/g-C3N4. ACS Appl. Energy Mater. 2020, 3, 3946–3954. [Google Scholar] [CrossRef]
  120. Zhang, X.; Ren, G.; Zhang, C.; Li, R.; Zhao, Q.; Fan, C. Photocatalytic reduction of CO2 to CO over 3D Bi2MoO6 microspheres: Simple synthesis, high efficiency and selectivity, reaction mechanism. Catal. Lett. 2020, 150, 2510–2516. [Google Scholar] [CrossRef]
  121. Xie, Y.; Zhuo, Y.; Liu, S.; Lin, Y.; Zuo, D.; Wu, X.; Li, C.; Wong, P.K. Ternary g-C3N4/ZnNCN@ZIF-8 hybrid photocatalysts with robust interfacial interactions and enhanced CO2 reduction. Solar RRL 2020, 4, 1900440. [Google Scholar] [CrossRef]
  122. Lin, N.; Lin, Y.; Qian, X.; Wang, X.; Su, W. Construction of a 2D/2D WO3/LaTiO2N direct Z-scheme photocatalyst for enhanced CO2 reduction performance under visible light. ACS Sustain. Chem. Eng. 2021, 9, 13686–13694. [Google Scholar] [CrossRef]
  123. Song, J.; Lu, Y.; Lin, Y.; Liu, Q.; Wang, X.; Su, W. A direct Z-scheme α-Fe2O3/LaTiO2N visible-light photocatalyst for enhanced CO2 reduction activity. Appl. Catal. B 2021, 292, 120185. [Google Scholar] [CrossRef]
  124. Valiev, R.Z.; Islamgaliev, R.K.; Alexandrov, I.V. Bulk nanostructured materials from severe plastic deformation. Prog. Mater. Sci. 2000, 45, 103–189. [Google Scholar] [CrossRef]
  125. Valiev, R.Z.; Estrin, Y.; Horita, Z.; Langdon, T.G.; Zehetbauer, M.J.; Zhu, Y.T. Producing bulk ultrafine-grained materials by severe plastic deformation. JOM 2006, 58, 33–39. [Google Scholar] [CrossRef] [Green Version]
  126. Azushima, A.; Kopp, R.; Korhonen, A.; Yang, D.Y.; Micari, F.; Lahoti, G.D.; Groche, P.; Yanagimoto, J.; Tsuji, N.; Rosochowski, A.; et al. Severe plastic deformation (SPD) processes for metals. CRIP Ann. Mauuf. Technol. 2008, 57, 716–735. [Google Scholar] [CrossRef]
  127. Segal, V. Review: Modes and processes of severe plastic deformation (SPD). Materials 2018, 11, 1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Zehetbauer, M.; Grossinger, R.; Krenn, H.; Krystian, M.; Pippan, R.; Rogl, P.; Waitz, T.; Wurschum, R. Bulk nanostructured functional ‘materials by severe plastic deformation. Adv. Eng. Mater. 2010, 12, 692–700. [Google Scholar] [CrossRef]
  129. Estrin, Y.; Vinogradov, A. Extreme grain refinement by severe plastic deformation: A wealth of challenging science. Acta Mater. 2013, 61, 782–817. [Google Scholar] [CrossRef]
  130. Edalati, K.; Horita, Z. Special issue on severe plastic deformation for nanomaterials with advanced functionality. Mater. Trans. 2019, 60, 1103. [Google Scholar] [CrossRef]
  131. Bryła, K.; Edalati, K. Historical studies by polish scientist on ultrafine-grained materials by severe plastic deformation. Mater. Trans. 2019, 60, 1553–1560. [Google Scholar] [CrossRef] [Green Version]
  132. Horita, Z.; Edalati, K. Severe plastic deformation for nanostructure controls. Mater. Trans. 2020, 61, 2241–2247. [Google Scholar] [CrossRef]
  133. Pereira, P.H.R.; Figueiredo, R.B. Finite element modelling of high-pressure torsion: An overview. Mater. Trans. 2019, 60, 1139–1150. [Google Scholar] [CrossRef] [Green Version]
  134. Levitas, V.I. High-pressure phase transformations under severe plastic deformation by torsion in rotational anvils. Mater. Trans. 2019, 60, 1294–1301. [Google Scholar] [CrossRef] [Green Version]
  135. Tsuji, N.; Gholizadeh, R.; Ueji, R.; Kamikawa, N.; Zhao, L.; Tian, Y.; Bai, Y.; Shibata, A. Formation mechanism of ultrafine grained microstructures: Various possibilities for fabricating bulk nanostructured metals and alloys. Mater. Trans. 2019, 60, 1518–1532. [Google Scholar] [CrossRef] [Green Version]
  136. Renk, O.; Pippan, R. Saturation of grain refinement during severe plastic deformation of single phase materials: Reconsiderations, current status and open questions. Mater. Trans. 2019, 60, 1270–1282. [Google Scholar] [CrossRef] [Green Version]
  137. Popov, V.V.; Popova, E.N. Behavior of Nb and Cu¬Nb composites under severe plastic deformation and annealing. Mater. Trans. 2019, 60, 1209–1220. [Google Scholar] [CrossRef] [Green Version]
  138. Skrotzki, W. Deformation heterogeneities in equal channel angular pressing. Mater. Trans. 2019, 60, 1331–1343. [Google Scholar] [CrossRef] [Green Version]
  139. Miura, H.; Iwama, Y.; Kobayashi, M. Comparisons of microstructures and mechanical properties of heterogeneous nano-structure induced by heavy cold rolling and ultrafine-grained structure by multi-directional forging of Cu¬Al alloy. Mater. Trans. 2019, 60, 1111–1115. [Google Scholar] [CrossRef] [Green Version]
  140. Faraji, G.; Torabzadeh, H. An overview on the continuous severe plastic deformation methods. Mater. Trans. 2019, 60, 1316–1330. [Google Scholar] [CrossRef] [Green Version]
  141. Toth, L.S.; Chen, C.; Pougis, A.; Arzaghi, M.; Fundenberger, J.J.; Massion, R.; Suwas, S. High pressure tube twisting for producing ultra fine grained materials: A review. Mater. Trans. 2019, 60, 1177–1191. [Google Scholar] [CrossRef] [Green Version]
  142. Masuda, T.; Horita, Z. Grain refinement of AZ31 and AZ61 Mg alloys through room temperature processing by up-scaled high-pressure torsion. Mater. Trans. 2019, 60, 1104–1110. [Google Scholar] [CrossRef] [Green Version]
  143. Yang, X.; Pan, H.; Zhang, J.; Gao, H.; Shu, B.; Gong, Y.; Zhu, X. Progress in mechanical properties of gradient structured metallic materials induced by surface mechanical attrition treatment. Mater. Trans. 2019, 60, 1543–1552. [Google Scholar] [CrossRef] [Green Version]
  144. Grosdidier, T.; Novelli, M. Recent developments in the application of surface mechanical attrition treatments for improved gradient structures: Processing parameters and surface reactivity. Mater. Trans. 2019, 60, 1344–1355. [Google Scholar] [CrossRef] [Green Version]
  145. Suwas, S.; Mondal, S. Texture evolution in severe plastic deformation processes. Mater. Trans. 2019, 60, 1457–1471. [Google Scholar] [CrossRef] [Green Version]
  146. Sauvage, X.; Duchaussoy, A.; Zaher, G. Strain induced segregations in severely deformed materials. Mater. Trans. 2019, 60, 1151–1158. [Google Scholar] [CrossRef] [Green Version]
  147. Wilde, G.; Divinski, S. Grain boundaries and diffusion phenomena in severely deformed materials. Mater. Trans. 2019, 60, 1302–1315. [Google Scholar] [CrossRef] [Green Version]
  148. Gubicza, J. Lattice defects and their influence on the mechanical properties of bulk materials processed by severe plastic deformation. Mater. Trans. 2019, 60, 1230–1242. [Google Scholar] [CrossRef] [Green Version]
  149. Čížek, J.; Janeček, M.; Vlasák, T.; Smola, B.; Melikhova, O.; Islamgaliev, R.K.; Dobatkin, S.V. The development of vacancies during severe plastic deformation. Mater. Trans. 2019, 60, 1533–1542. [Google Scholar] [CrossRef] [Green Version]
  150. Kunimine, T.; Watanabe, M. A comparative study of hardness in nanostructured Cu-Zn, Cu-Si and Cu-Ni solid-solution alloys processed by severe plastic deformation. Mater. Trans. 2019, 60, 1484–1488. [Google Scholar] [CrossRef] [Green Version]
  151. Kuramoto, S.; Furuta, T. Severe plastic deformation to achieve high strength and high ductility in Fe¬Ni based alloys with lattice softening. Mater. Trans. 2019, 60, 1116–1122. [Google Scholar] [CrossRef] [Green Version]
  152. Kawasaki, M.; Langdon, T.G. The contribution of severe plastic deformation to research on superplasticity. Mater. Trans. 2019, 60, 1123–1130. [Google Scholar] [CrossRef] [Green Version]
  153. Demirtas, M.; Purcek, G. Room temperature superplaticity in fine/ultrafine grained materials subjected to severe plastic deformation. Mater. Trans. 2019, 60, 1159–1167. [Google Scholar] [CrossRef] [Green Version]
  154. Moreno-Valle, E.C.; Pachla, W.; Kulczyk, M.; Sabirov, I.; Hohenwarter, A. Anisotropy of tensile and fracture behavior of pure titanium after hydrostatic extrusion. Mater. Trans. 2019, 60, 2160–2167. [Google Scholar] [CrossRef] [Green Version]
  155. Kral, P.; Dvorak, J.; Sklenicka, V.; Langdon, T.G. The characteristics of creep in metallic materials processed by severe plastic deformation. Mater. Trans. 2019, 60, 1506–1517. [Google Scholar] [CrossRef] [Green Version]
  156. Razavi-Khosroshahi, H.; Fuji, M. Development of metal oxide high-pressure phases for photocatalytic properties by severe plastic deformation. Mater. Trans. 2019, 60, 1203–1208. [Google Scholar] [CrossRef] [Green Version]
  157. Ikoma, Y. Severe plastic deformation of semiconductor materials using high-pressure torsion. Mater. Trans. 2019, 60, 1168–1176. [Google Scholar] [CrossRef] [Green Version]
  158. Blank, V.D.; Popov, M.Y.; Kulnitskiy, B.A. The effect of severe plastic deformations on phase transitions and structure of solids. Mater. Trans. 2019, 60, 1500–1505. [Google Scholar] [CrossRef] [Green Version]
  159. Révész, Á.; Kovács, Z. Severe plastic deformation of amorphous alloys. Mater. Trans. 2019, 60, 1283–1293. [Google Scholar] [CrossRef] [Green Version]
  160. Beloshenko, V.; Vozniak, I.; Beygelzimer, Y.; Estrin, Y.; Kulagin, R. Severe plastic deformation of polymers. Mater. Trans. 2019, 60, 1192–1202. [Google Scholar] [CrossRef] [Green Version]
  161. Mazilkin, A.; Straumal, B.; Kilmametov, A.; Straumal, P.; Baretzky, B. Phase transformations induced by severe plastic deformation. Mater. Trans. 2019, 60, 1489–1499. [Google Scholar] [CrossRef] [Green Version]
  162. Bachmaier, A.; Pippan, R. High-pressure torsion deformation induced phase transformations and formations: New material combinations and advanced properties. Mater. Trans. 2019, 60, 1256–1269. [Google Scholar] [CrossRef] [Green Version]
  163. Han, J.K.; Jang, J.I.; Langdon, T.G.; Kawasaki, M. Bulk-state reactions and improving the mechanical properties of metals through high-pressure torsion. Mater. Trans. 2019, 60, 1131–1138. [Google Scholar] [CrossRef] [Green Version]
  164. Edalati, K. Metallurgical alchemy by ultra-severe plastic deformation via high-pressure torsion process. Mater. Trans. 2019, 60, 1221–1229. [Google Scholar] [CrossRef] [Green Version]
  165. Mito, M.; Shigeoka, S.; Kondo, H.; Noumi, N.; Kitamura, Y.; Irie, K.; Nakamura, K.; Takagi, S.; Deguchi, H.; Tajiri, T.; et al. Hydrostatic compression effects on fifth-group element superconductors V, Nb, and Ta subjected to high-pressure torsion. Mater. Trans. 2019, 60, 1472–1483. [Google Scholar] [CrossRef] [Green Version]
  166. Nishizaki, T.; Edalati, K.; Lee, S.; Horita, Z.; Akune, T.; Nojima, T.; Iguchi, S.; Sasaki, T. Critical temperature in bulk ultrafine-grained superconductors of Nb, V, and Ta processed by high-pressure torsion. Mater. Trans. 2019, 60, 1367–1376. [Google Scholar] [CrossRef] [Green Version]
  167. Rogl, G.; Zehetbauer, M.J.; Rogl, P.F. The effect of severe plastic deformation on thermoelectric performance of skutterudites, half-Heuslers and Bi-tellurides. Mater. Trans. 2019, 60, 2071–2085. [Google Scholar] [CrossRef] [Green Version]
  168. Enikeev, N.A.; Shamardin, V.K.; Radiguet, B. Radiation tolerance of ultrafine-grained materials fabricated by severe plastic deformation. Mater. Trans. 2019, 60, 1723–1731. [Google Scholar] [CrossRef] [Green Version]
  169. Leiva, D.R.; Jorge, A.M.; Ishikawa, T.T., Jr.; Botta, W.J. Hydrogen storage in Mg and Mg-based alloys and composites processed by severe plastic deformation. Mater. Trans. 2019, 60, 1561–1570. [Google Scholar] [CrossRef] [Green Version]
  170. Huot, J.; Tousignant, M. Effect of cold rolling on metal hydrides. Mater. Trans. 2019, 60, 1571–1576. [Google Scholar] [CrossRef] [Green Version]
  171. Miyamoto, H.; Yuasa, M.; Rifai, M.; Fujiwara, H. Corrosion behavior of severely deformed pure and single-phase materials. Mater. Trans. 2019, 60, 1243–1255. [Google Scholar] [CrossRef] [Green Version]
  172. Valiev, R.Z.; Parfenov, E.V.; Parfenova, L.V. Developing nanostructured metals for manufacturing of medical implants with improved design and biofunctionality. Mater. Trans. 2019, 60, 1356–1366. [Google Scholar] [CrossRef] [Green Version]
  173. Lowe, T.C.; Valiev, R.Z.; Li, X.; Ewing, B.R. Commercialization of bulk nanostructured metals and alloys. MRS Bull. 2021, 46, 265–272. [Google Scholar] [CrossRef]
Figure 2. Improvement of light absorbance, suppression of recombination, and enhancement of photocatalytic CO2 conversion for BiVO4 by simultaneous strain and oxygen vacancy engineering using high-pressure torsion (HPT). (a) Crystallite size and volumetric strain versus the number of HPT turns (N), (b) Raman spectroscopy of initial and HPT-processed samples (inset: the appearance of samples), (c) oxygen vacancy concentration versus the number of HPT turns calculated by X-ray photoelectron spectroscopy, (d) UV–VIS spectroscopy, (e) photoluminescence spectra, and (f) photocatalytic CO production rate versus time for initial powder and sample proceeded by HPT for N = 0.25, 1, and 4 turns [35].
Figure 2. Improvement of light absorbance, suppression of recombination, and enhancement of photocatalytic CO2 conversion for BiVO4 by simultaneous strain and oxygen vacancy engineering using high-pressure torsion (HPT). (a) Crystallite size and volumetric strain versus the number of HPT turns (N), (b) Raman spectroscopy of initial and HPT-processed samples (inset: the appearance of samples), (c) oxygen vacancy concentration versus the number of HPT turns calculated by X-ray photoelectron spectroscopy, (d) UV–VIS spectroscopy, (e) photoluminescence spectra, and (f) photocatalytic CO production rate versus time for initial powder and sample proceeded by HPT for N = 0.25, 1, and 4 turns [35].
Materials 16 01081 g002
Figure 3. Improved charge carrier migration and photocatalytic CO2 conversion by introducing the high-pressure TiO2-II phase. (a) Pressure–temperature phase diagram of TiO2. (b) Raman spectra, (c) photocurrent spectra, and (d) photocatalytic CO production rate versus time for TiO2 before and after high-pressure torsion processing and after annealing [36].
Figure 3. Improved charge carrier migration and photocatalytic CO2 conversion by introducing the high-pressure TiO2-II phase. (a) Pressure–temperature phase diagram of TiO2. (b) Raman spectra, (c) photocurrent spectra, and (d) photocatalytic CO production rate versus time for TiO2 before and after high-pressure torsion processing and after annealing [36].
Materials 16 01081 g003
Figure 4. High photocatalytic CO2 conversion on defective high-entropy oxide TiZrNbHfTaO11 synthesized by high-pressure torsion. Microstructure of TiZrNbHfTaO11 by (a) scanning electron microscopy and (b) high-resolution transmission electron microscopy. (c) Electronic band structure of TiZrNbHfTaO11. (d) Photocatalytic CO production rate on TiZrNbHfTaO11 versus time compared with P25 TiO2, BiVO4, and anatase TiO2 [37].
Figure 4. High photocatalytic CO2 conversion on defective high-entropy oxide TiZrNbHfTaO11 synthesized by high-pressure torsion. Microstructure of TiZrNbHfTaO11 by (a) scanning electron microscopy and (b) high-resolution transmission electron microscopy. (c) Electronic band structure of TiZrNbHfTaO11. (d) Photocatalytic CO production rate on TiZrNbHfTaO11 versus time compared with P25 TiO2, BiVO4, and anatase TiO2 [37].
Materials 16 01081 g004
Figure 5. High light absorbance, appropriate band structure, suppressed recombination, significant CO2 adsorption, and high photocatalytic CO2 conversion for high-entropy oxynitride (HEON) TiZrNbHfTaO6N 3. (a) UV–VIS light absorbance spectra, (b) electronic band structure together with the appearance of samples, (c) photoluminescence spectra, (d) photocurrent density versus time, (e) diffuse reflectance infrared Fourier transform spectra (peaks at 665 and 2350 cm−1 represent chemisorption and physisorption of CO2), and (f) photocatalytic CO production rate versus time for TiZrNbHfTaO6N3 compared with P25 TiO2 and high-entropy oxide (HEO) TiZrNbHfTaO11 [38].
Figure 5. High light absorbance, appropriate band structure, suppressed recombination, significant CO2 adsorption, and high photocatalytic CO2 conversion for high-entropy oxynitride (HEON) TiZrNbHfTaO6N 3. (a) UV–VIS light absorbance spectra, (b) electronic band structure together with the appearance of samples, (c) photoluminescence spectra, (d) photocurrent density versus time, (e) diffuse reflectance infrared Fourier transform spectra (peaks at 665 and 2350 cm−1 represent chemisorption and physisorption of CO2), and (f) photocatalytic CO production rate versus time for TiZrNbHfTaO6N3 compared with P25 TiO2 and high-entropy oxide (HEO) TiZrNbHfTaO11 [38].
Materials 16 01081 g005
Table 1. Summary of major publications about ceramics treated by high-pressure torsion and their major properties and applications.
Table 1. Summary of major publications about ceramics treated by high-pressure torsion and their major properties and applications.
MaterialsInvestigated Properties and ApplicationsReference
Various MaterialsImpact of pressure and strain on allotropyBridgman (1935) [56]
α-Al2O3Microstructure and mechanical propertiesEdalati et al. (2010) [57]
ZrO2Allotropic phase transformationsEdalati et al. (2011) [58]
CuODielectric propertiesMakhnev et al. (2011) [59]
CuO, Y3Fe5O12, FeBO3Optical properties and electronic structureGizhevskii et al. (2011) [60]
ZrO2Phase transformationDelogu et al. (2012) [61]
Cu2O, CuOMiddle infrared absorption and X-ray absorptionMostovshchikova et al. (2012) [62]
CuO, Y3Fe5O12, FeBO3Optical propertiesTelegin et al. (2012) [63]
BaTiO3Optical and dielectric propertiesEdalati et al. (2015) [64]
TiO2-IIPhotocatalytic activity for hydrogen productionRazavi-Khosroshahi et al. (2016) [23]
Various MaterialsReview on HPTEdalati et al. (2016) [40]
TiO2Plastic strain and phase transformationRazavi-Khosroshahi et al. (2016) [65]
Y2O3Optical propertiesRazavi-Khosroshahi et al. (2016) [66]
YBa2Cu3OyMicrostructural investigationKuznetsova et al. (2017) [67]
BNCoupled elastoplasticity and plastic strain-induced phase transformationFeng et al. (2017) [68]
ZnOPhotocatalytic activity for dye degradationRazavi-Khosroshahi et al. (2017) [26]
Fe3O4Lithium-ion batteriesQian et al. (2018) [69]
ZnOPlastic flow and microstructural instabilitiesQi et al. (2018) [70]
Fe71.2Cr22.7Mn1.3N4.8Microstructural featuresShabashov et al. (2018) [71]
BNModeling of plastic flow and strain-induced phase transformationFeng et al. (2019) [72]
TiO2-IIElectrocatalysis for hydrogen generationEdalati et al. (2019) [73]
γ-Al2O3Photocatalytic activity for dye degradationEdalati et al. (2019) [27]
Various OxidesReview on HPT of oxidesEdalati et al. (2019) [74]
MgOPhotocatalytic activity for dye degradationFujita et al. (2020) [28]
ZrO2Photocatalytic activity for hydrogen productionWang et al. (2020) [26]
SiO2Photocatalytic activity for dye degradationWang et al. (2020) [34]
CsTaO3, LiTaO3Photocatalytic activity for hydrogen productionEdalati et al. (2020) [24]
GaN-ZnOPhotocatalytic activity for hydrogen productionEdalati et al. (2020) [25]
Fe53.3Ni26.5B20.2, Co28.2Fe38.9Cr15.4Si0.3B17.2Microstructure and mechanical propertiesPermyakova et al. (2020) [75]
TiHfZrNbTaO11Photocatalytic activity for hydrogen productionEdalati et al. (2020) [27]
TiO2-ZnOPhotocatalytic activity for hydrogen productionHidalgo-Jimeneza et al. (2020) [28]
Bi2O3Enhanced photocurrent generationFujita et al. (2020) [76]
TiO2-IIVisible-light photocurrent generationWang et al. (2020) [77]
TiO2-IIPhotocatalytic activity for CO2 conversionAkrami et al. (2021) [30]
TiZrHfNbTaO6N3Photocatalytic activity for hydrogen productionEdalati et al. (2021) [29]
SiO2, VO2Phase transformationEdalati et al. (2021) [78]
TiO2Grain coarsening and phase transformationEdalati et al. (2021) [79]
ZnOBandgap narrowingQi et al. (2021) [80]
BiVO4Photocatalytic activity for CO2 conversionAkrami et al. (2022) [29]
TiHfZrNbTaO11Photocatalytic activity for CO2 conversionAkrami et al. (2022) [31]
TiZrNbTaWO12Photocatalytic activity for oxygen productionEdalati et al. (2022) [30]
TiZrHfNbTaO6N3Photocatalytic activity for CO2 conversionAkrami et al. (2022) [32]
Table 2. Photocatalytic CO production rate on high-entropy oxynitride TiZrNbHfTaO6N3 synthesized by high-pressure torsion compared with photocatalysts reported in the literature.
Table 2. Photocatalytic CO production rate on high-entropy oxynitride TiZrNbHfTaO6N3 synthesized by high-pressure torsion compared with photocatalysts reported in the literature.
PhotocatalystCatalyst Concentration Light SourceCO Production Rate (µmolh−1g−1)CO Production Rate (µmolh−1m−1)Ref.
TiO2/Graphitic Carbon100 mg (Gas System)300 W Xenon10.160.04Wang et al. (2013) [91]
Bicrystalline Anatase/Brookite TiO2 Microspheres30 mg (Gas System)150 W Solar Simulator1450.95Liu et al. (2013) [92]
Ag/TaON/RuBLRu′2 gL−1 (Liquid System)500 W High-Pressure Mercury0.056----Sekizawa et al. (2013) [93]
10 wt % Montmorillonite-Loaded TiO250 mg (Gas System)500 W Mercury1031.25Tahir et al. (2013) [94]
Anatase TiO2 Nanofibers50 gL−1 (Liquid System)500 W Mercury Flash40-----Zhang et al. (2013) [95]
TiO2 Nanosheets Exposed {001} Facet 1 gL−1 (Liquid System)Two 18 W Low-Pressure Mercury0.120.00095He et al. (2014) [96]
Anatase TiO2 Hierarchical Microspheres200 mg (Gas System)40 W Mercury UV 18.50.37Fang et al. (2014) [97]
TiO2 and Zn(II) Porphyrin Mixed Phases60 mg (Gas System)300 W Xenon 80.062Li et al. (2015) [98]
Anatase TiO2 Hollow Sphere100 mg (Gas System)40 W Mercury UV 140.16Fang et al. (2015) [99]
10 wt % In-Doped Anatase TiO2250 mg (Gas System)500 W Mercury Flash811.33Tahir et al. (2015) [100]
Pt2+–Pt0/TiO2100 mg (Gas System)300 W Xenon~12.140.7Xiong et al. (2015) [101]
BiOI150 mg (Gas System)300 W High-Pressure Xenon 4.10.03Ye et al. (2016) [102]
RuRu/Ag/TaON1 gL−1 (Liquid System)High-Pressure Mercury5----Nakada et al. (2016) [103]
RuRu/TaON1 gL−1 (Liquid System)High-Pressure Mercury3.33 ----Nakada et al. (2016) [103]
CeO2-x50 mg (Gas System)300 W Xenon1.650.08Ye et al. (2017) [104]
Cu2O/RuOx500 mg (Gas System)150 W Xenon0.88---Pastor et al. (2017) [105]
TiO2 3D Ordered Microporous/Pd100 mg (Gas System)300 W Xenon3.90.066Jiao et al. (2017) [106]
BiVO4/C/Cu2O---300 W Xenon3.01----Kim et al. (2018) [107]
g-C3N4/α-Fe2O3200 mg (Gas System)300 W Xenon5.7-----Wang et al. (2018) [108]
xCu2O/Zn2-2xCr4 gL−1 (Liquid System)200 W Mercury-Xenon 2.50.018Jiang et al. (2018) [109]
TiO2/Carbon Nitride Nanosheet25 mg (Gas System)150 W Xenon2.04----Crake et al. (2019) [110]
TiO2/CoOx Hydrogenated50 mg (Gas System)150 W UV 1.240.0045Li et al. (2019) [111]
Bi4O5Br220 mg (Gas System)300 W High-Pressure Xenon63.130.58Bai et al. (2019) [112]
ZnGaON---1600 W Xenon1.05---Maiti et al. (2019) [113]
C3N4 by Thermal Condensation100 mg (Gas System) 350 W Mercury 4.83------Xia et al. (2019) [9]
Cd1-xZnxS45 mg (Gas System)UV-LED Irradiation 2.90.015Kozlova et al. (2019) [114]
Bi24O31Cl1050 mg (Gas System)300 W High-Pressure Xenon0.9---Jin et al. (2019) [115]
Bi2Sn2O70.4 gL−1 (Liquid System)300 W Xenon14.880.24Guo et al. (2020) [116]
Ag/Bi/BiVO410 mg (Gas System)300 W Xenon Illuminator5.190.42Duan et al. (2020) [117]
g-C3N4/BiOCl20 mg (Gas System)300 W High-Pressure Xenon4.73---Chen et al. (2020) [118]
Fe/g-C3N41 gL−1 (Liquid System)300 W Xenon~22.50.06Dao et al. (2020) [119]
Bi2MoO60.7 gL−1 (Liquid System)300 W Xenon41.51.26Zhang et al. (2020) [120]
g-C3N4/Zinc Carbodiimide/Zeolitic Imidazolate Framework100 mg (Gas System)300 W Xenon~0.450.014Xie et al. (2020) [121]
WO3/LaTiO2N10 mg (Gas System)300 W Xenon2.210.4Lin et al. (2021) [122]
α-Fe2O3/LaTiO2N20 mg (Gas System)300 W Xenon9.70.65Song et al. (2021) [123]
TiZrHfNbTaO6N30.2 gL−1 (Liquid System)400 W High-Pressure Mercury10.72 ± 1.774.66 ± 0.3Akrami et al. (2022) [32]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Akrami, S.; Ishihara, T.; Fuji, M.; Edalati, K. Advanced Photocatalysts for CO2 Conversion by Severe Plastic Deformation (SPD). Materials 2023, 16, 1081. https://doi.org/10.3390/ma16031081

AMA Style

Akrami S, Ishihara T, Fuji M, Edalati K. Advanced Photocatalysts for CO2 Conversion by Severe Plastic Deformation (SPD). Materials. 2023; 16(3):1081. https://doi.org/10.3390/ma16031081

Chicago/Turabian Style

Akrami, Saeid, Tatsumi Ishihara, Masayoshi Fuji, and Kaveh Edalati. 2023. "Advanced Photocatalysts for CO2 Conversion by Severe Plastic Deformation (SPD)" Materials 16, no. 3: 1081. https://doi.org/10.3390/ma16031081

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

Article Metrics

Back to TopTop