Next Article in Journal
Investigation of the Surface Coating, Humidity Degradation, and Recovery of Perovskite Film Phase for Solar-Cell Applications
Previous Article in Journal
Controllable Phase Transformation and Enhanced Photocatalytic Performance of Nano-TiO2 by Using Oxalic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Ferroelectric Materials-Based Photoelectrochemical Reaction

1
Henan Engineering Center of New Energy Battery Materials, Henan D&A Engineering Center of Advanced Battery Materials, Shangqiu Normal University, Shangqiu 476000, China
2
Shandong Yuhuang New Energy Technology Co., Ltd., Heze 274000, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(17), 3026; https://0-doi-org.brum.beds.ac.uk/10.3390/nano12173026
Submission received: 31 July 2022 / Revised: 28 August 2022 / Accepted: 29 August 2022 / Published: 31 August 2022
(This article belongs to the Topic Catalysis for Sustainable Chemistry and Energy)

Abstract

:
Inorganic perovskite ferroelectric-based nanomaterials as sustainable new energy materials, due to their intrinsic ferroelectricity and environmental compatibility, are intended to play a crucial role in photoelectrochemical field as major functional materials. Because of versatile physical properties and excellent optoelectronic properties, ferroelectric-based nanomaterials attract much attention in the field of photocatalysis, photoelectrochemical water splitting and photovoltaic. The aim of this review is to cover the recent advances by stating the different kinds of ferroelectrics separately in the photoelectrochemical field as well as discussing how ferroelectric polarization will impact functioning of photo-induced carrier separation and transportation in the interface of the compounded semiconductors. In addition, the future prospects of ferroelectric-based nanomaterials are also discussed.

1. Introduction

Development and utilization of clean, renewable solar energy has proven to be an efficient way to cope with environmental concerns. Photocatalysis [1,2], photoelectrochemical (PEC) water splitting [3,4] and photovoltaic [5] are topics of effective use of solar energy with the central step of PEC. The photoactive species were categorized as inorganic semiconductors [6,7,8,9], organic semiconductors [10,11], organic-inorganic hybrid materials [12,13], polymer semiconductor materials [14], perovskites [15,16,17] and so on. Perovskite oxide ferroelectrics, generally referring to ABO3 and its derivatives, possess intrinsic ferroelectric polarization (P) with reversible external electric fields that have come to the forefront of nanoscience and nanotechnology ever since the pioneering discovery [18,19], followed by other ferroelectric ceramics, such as KNbO3, KTaO3, PbTiO3, etc. The ferroelectricity is due to the asymmetry of the crystal structure with the existence of electric dipoles inside because of the misalignment of the positive and negative charge centers [20]. Therefore, ferroelectrics’ splendid photoelectric properties include large nonlinear optical coefficients, stable chemical properties, as well as adjustable band gap. Furthermore, because of the excellent separation and migration efficiency of electron-hole pairs under depolarization field, photo-generated voltage above the energy gap can be obtained [21]. Despite the above-mentioned advantages, there are still some problems need to be investigated. For example, the low transport efficiency of the optical carrier has led to the knockdown quantum efficiency (10−5), which is far lower than conventional P-N junction devices (10−1). This phenomenon is mainly caused by two factors: (1) ferroelectrics can only absorb ultraviolet light because of wide bandgap values. (2) The low electrical conductivity properties seriously hinder the diversion of photo-generated electron holes, whereas the increment of specific conductance will cause electrical leakage inside the material so as that it cannot maintain strong electrical polarization [22]. Therefore, ion doping is adopted to reduce the bandgap. The combination of semiconductor and ferroelectrics is an efficient method to improve the separation efficiency of electron holes and will reduce the need for interface state control because interface is not a necessary channel of photo-carriers [23]. In all, as evidenced by gradual increasing academic papers, the advanced ferroelectric materials for photo-electric conversion are still currently in progress.
Given the pace of advances in this field, the ferroelectric materials for photovoltaic reactions and photocatalysis are the subject of the present review. Specifically, this work surveys the functional materials in a period of rapid development of ABO3 and 2D ferroelectrics. The recent progress, research technique, and future prospects of this orientation are also discussed and evaluated. We sincerely apologize to the researchers of ferroelectric-based photovoltaic and photocatalysis publications that were not deliberately overlooked and want to remind our readers that this article is not designed to cover comprehensive theoretical knowledge of ferroelectric-based photovoltaic reactions and photocatalysis but rather to present several typical ferroelectric materials. Furthermore, we do our utmost to emphasize some important publications which can accelerate further research in depth in this area. For learners seeking to acquire more knowledge on the fundamental theory about photovoltaic reactions and photocatalysis, we recommend they seek more information in other reviews [24,25,26,27,28].

2. Recent Advances of Different Species

The structure of ferroelectrics for photovoltaic reactions and photocatalysis comes in two kinds, with one category of perovskite-structured materials and another layered crystal structure. Generally, perovskite-structured materials are highly symmetrical in structure with a typical chemical formula ABO3, in which the cube’s corner points (A-site) was occupied by univalent or divalent metal ions, while the center of the cube (B-site) was taken up by metal ions and oxygen ions occupying the face-center position of the cube. Layered structure materials refer to 2D materials and often consist of either a single or a few atomic layers and several representative ferroelectric materials are listed below.

2.1. BaTiO3

BaTiO3, with the bandgap of 3.18 eV, is a widely used ferroelectric material. Steve Dunn proved that BaTiO3 with tetragonal crystal reveals a 3-fold increase in the Rhodamine B degradation compared with cubic material [29]. Although pure BaTiO3 owns a certain photocatalytic property, the wide band-gap precludes its photo-activity. Several strategies were applied for optimizing its optoelectronic properties: (1) loading noble metals onto the surface of ferroelectrics. Li et al. [30] assembled Ag loaded BaTiO3 nanotube arrays system and Ag act as the photogenerated electron traps, can inhibits the recombination of photoelectron and holes, leading to the increment of photo-degradation of MO efficiency. (2) doping ion, especially transition metal ions, can effectively increase the lifetime of carriers, reduce the recombination rate of electron hole pairs and increase the degree of separation of electron hole pairs. A certain number of studies reported that ion-doping, such as Li+ [31], Fe3+ [32], Mn2+ [33], Y3+ [34] and even NaNbO3 [35], KNbO3 [36], affect the crystal structure of intrinsic materials. (3) combining two semiconductors with apposite energy level and bandgap can change band bending, widen the light absorption range and accelerate carrier separation. As shown in Figure 1a,b, Wang et al. [37] reported that TiO2/BaTiO3 core/shell NWs obtained 67% photocurrent enhancement compared with pure TiO2 NWs, which can ascribe to adjustable ferroelectric polarization by external electric field poling, and this research proved that ferroelectric or piezoelectric potential-induced band structure engineering likely raise the capability of PEC electrode. Hu et al. [38] synthesized necklace-like BaTiO3 nanofibers by sol-gel assisted electrospinning and subsequently coating TiO2 on the nanofiber surface by wet-chemical. The TEM of necklace-like BaTiO3 nanofibers was shown in Figure 1c, which exhibited both high piezoelectric coefficient and ferroelectric. Due to the excellent piezo-photocatalytic property, the necklace-like BaTiO3@TiO2 core-shell demonstrated outstanding degradation of MO under both UV irradiation and ultrasound, as shown in Figure 1d.
Besides TiO2, other oxide semiconductors are also adopted to combine with BaTiO3, such as α-Fe2O3 [39,40], Ag2O [41,42], SnO2 [43,44], ZnO [45], MoO3 [46], WO3 [47], Cu2O [48,49]. Steve Dunn et al., proved that BaTiO3/α-Fe2O3 could form the island-like morphology with the minimum addition of α-Fe2O3, which allowed photons exciting both BaTiO3 and α-Fe2O3. More importantly, the discrete site of BaTiO3 and α-Fe2O3 could combine with dye molecule to form triple points and permit the dye molecule contact charge carriers directly, as shown in Figure 2a,b, and finally resulted in the observed photo-degradation [39]. The good news is that ferroelectric-based BaTiO3 has been applied in PEC bioanalysis. Yu et al. [47] reported that WO3 nanoflakes/BTO/Cu2O photoelectrode could realize the ultrasensitive detection of prostate specific antigen (PSA) with the limit of detection of 0.036 pg/mL. Due to the polar charge carriers-created electric field of BaTiO3, as illustrated in Figure 2d, the electrons and holes of WO3 and Cu2O can be induced to directional migration (Figure 2e), and finally realized the larger photocurrent density response (Figure 2f).
In addition to oxide semiconductors, metal sulfide semiconductor, such as CdS [50,51,52], Ag2S [53] etc., could enhance the photo-electric conversion by integrating with ferroelectric. Liu et al., reported that BaTiO3-CdS hybrid photocatalysts obtained H2 production rate of 483 µmol∙h−1∙g−1 with the 20 wt% CdS loading due to the spontaneous polarization electric field of BaTiO3 [52]. As illustrated in Figure 2c, with the aid of electric field driving force of BaTiO3, photo-induced electrons and holes in CdS would be separated effectively for followed water splitting for hydrogen preparation. Wang et al., introduced (Ag-Ag2S)/BaTiO3 hybrid ternary structure and reached a high MO degradation rate of 90% in 30 min on account of both synergistic exciton-plasmon interaction in Ag-Ag2S and piezoelectric polarization in Ag2S/BaTiO3 [53].
Figure 2. (a) Schematic island-like morphology of Fe2O3/BaTiO3 and the triple points between the dye solution, Fe2O3 and BaTiO3 provide active sites for redox reactions; (b) photodecolorization profiles of RhB with different photocatalysts under simulated sunlight. Adapted with permission from ref. [39], copyright: 2017, American Chemical Society. (c) Schematic of photoinduced hole and electron migration in BaTiO3–CdS composites and photocatalytic hydrogen process under visible light; Adapted with permission from ref. [52], copyright: 2018, Elsevier. (d) Schematic diagram of the promoting effect of the PCC electric field on the separation of electron-hole pairs; (e) DSCC Strategy of the WO3 Nanoflakes/BTO/Cu2O Photoelectrode; (f) photocurrent responses of WO3 nanoflakes (curve a), WO3/BTO (curve b), WO3/Cu2O (curve c), WO3/Cu2O/BTO (d), and WO3/BTO/Cu2O after poling (curve e). Adapted with permission from ref. [47], copyright: 2020, American Chemical Society.
Figure 2. (a) Schematic island-like morphology of Fe2O3/BaTiO3 and the triple points between the dye solution, Fe2O3 and BaTiO3 provide active sites for redox reactions; (b) photodecolorization profiles of RhB with different photocatalysts under simulated sunlight. Adapted with permission from ref. [39], copyright: 2017, American Chemical Society. (c) Schematic of photoinduced hole and electron migration in BaTiO3–CdS composites and photocatalytic hydrogen process under visible light; Adapted with permission from ref. [52], copyright: 2018, Elsevier. (d) Schematic diagram of the promoting effect of the PCC electric field on the separation of electron-hole pairs; (e) DSCC Strategy of the WO3 Nanoflakes/BTO/Cu2O Photoelectrode; (f) photocurrent responses of WO3 nanoflakes (curve a), WO3/BTO (curve b), WO3/Cu2O (curve c), WO3/Cu2O/BTO (d), and WO3/BTO/Cu2O after poling (curve e). Adapted with permission from ref. [47], copyright: 2020, American Chemical Society.
Nanomaterials 12 03026 g002

2.2. PbTiO3

Tetragonal PbTiO3 is a well-known ferroelectric material with a band gap larger than 2.75 eV, and possess a certain photocatalytic effect under UV light irradiation with a similar crystal structure to BaTiO3. Through ion doping, such as Cu2+ [54], Fe3+ [55], Mo [56], the light absorption of the material can be enhanced and its photocatalytic effect can be improved. Additionally, due to local surface plasmon resonance phenomenon, noble metals, such as Ag [57] and Pt [58] elements can also be adhered to the surface of PbTiO3 to improve the photocatalytic efficiency. In another point, the formation of core-shell structure or heterojunction, such as TiO2@PbTiO3 [59,60], can significantly increase the photocurrent efficiency (up to 19%), which is almost 20 times higher than literature. By decorating CdS nanoparticle onto the pure Pb(Zr0.2Ti0.8)O3 film, 11 times photocurrent density can be obtained due to effectively separation of photo-induced carriers by spontaneous ferroelectric polarization and the broaden band gap by CdS nanoparticle [61]. Besides the above methods, the crystal morphology has a great influence on the performance of PbTiO3. Li et al., prepared the single-domain PbTiO3 nanoplates and confirmed the internal depolarization field play an vital role in advancing photogenerated charge separation rather than polarization-caused band bending on the surface [62]. PbTiO3 nanoplates [63] can obtained high H2 evolution by selectively depositing Pt particles on the positively specific orientation of (110) facets, while consolidating MnOx on the negatively (110) facets under the influence of ferroelectric field due to the directional diffusion of photo-generated electrons and holes.

2.3. BiFeO3

As a prototypical ferroelectric perovskite oxide, BiFeO3 (BFO) stands out as an excellent ferroelectric material for photo-electric conversion due to suitable band gap [64]. And it can be used as a photoelectrode alone, or as an auxiliary material to combine with classical photocatalytic materials such as C3N4 [65], CdS [66] or it can be modified with other materials (combined with Au [67], doping by Sr2+ at Bi-site [68], etc.). Cao et al. [69] prepared high-quality BiFeO3 polycrystalline films on ITO with thickness of the ITO and BiFeO3 100 nm and 300 nm, respectively. The barrier height between ITO and BiFeO3 is 1.24 eV, which is a typical Schottky junction. As shown in Figure 3a, the ferroelectric polarization P direction points to the electrolyte and its depolarization field –EP (ie Ebi) points to the bottom electrode, which leading to energy band upward on the left edge bends and the right edge bends downward after the BiFeO3 film was polarized by an external voltage of +8 V. However, as shown in Figure 3b, when the BiFeO3 film is polarized at −8 V, holes are difficult to transport from the electrolyte to BiFeO3, while electrons are difficult to transport from BiFeO3 to the electrolyte, and obtain the opposite result compared with +8 V polarization. Additionally, the current density also changes with the transition of the ferroelectric polarization states of BiFeO3 films. As shown in Figure 3c, on account of Schottky barrier at the BiFeO3/ITO interface, the photocurrents of polarized BiFeO3 were all negative, regardless of the BiFeO3 film was polarized by +8 V or −8 V. But when the working voltage of BiFeO3 electrode was 0 V, the photocurrent density of the BiFeO3 film after polarization at +8 V was −10 µA/cm2. In order to illustrate how the poling operation influence the charge transfer, the authors offered a sleek architectural design in which PEC data was gathered by immersing the polarized electrodes in the electrolyte with 50 µM Rhodamine B as the modifier. As shown in Figure 3d, the +8 V polarized BiFeO3 film presented a peak at 590 nm in the external quantum yield spectra, which is corresponding to the absorption spectrum of Rhodamine B. This result photocurrent signal derived from photo-excited hole injection of Rhodamine B rather than the BiFeO3 film and in turn had validated the correctness of mechanism of photo-excited charge transfer from BiFeO3 film to the KCl electrolyte shown in Figure 3a,b.
Huang et al. [70] achieved the enhancement of photocatalytic current density of 1.76 mA/cm2 by assemble BiFeO3 on Sn-doped TiO2 nanorods (Sn:TiO2@BiFeO3). The single as-synthesized Sn:TiO2@BiFeO3 nanorod with thin BiFeO3 film can be seen in Figure 4c. The PEC data revealed that coating BiFeO3 can significantly improve the photocatalytic performance of TiO2 nanorods. As shown in Figure 4a, when the working electrode potential is 1.23 V, the photocurrent density of Sn:TiO2@BiFeO3 increased to 1.51 mA/cm2. Furthermore, when the external voltage polarized BiFeO3 film, the ferroelectric polarization points to the electrolyte and its depolarization field points to Sn:TiO2, as shown in Figure 4d. Meanwhile, the energy band at the Sn:TiO2@BiFeO3 interface bends upwards, and its depletion layer widened. Thus, the built-in electric field of BiFeO3 increased the electron hole separation efficiency, and the photo-current density increases from 1.51 mA/cm2 to 1.76 mA/cm2 (Figure 4b). However, as shown in Figure 4e, when the applied polarization voltage points to the FTO, the depletion layer at the Sn:TiO2@BiFeO3 interface is narrowed, which reduces the electron hole separation efficiency, and the photocatalytic current density of the photoanode decreases from 1.51 mA/cm2 to 1.02 mA/cm2 (Figure 4b).
Except for normally semiconductor materials, BiFeO3 constantly be combined with other type ferroelectrics, such as LaNiO3 [71], XTiO3 (X = Sr, Zn, Pb) [72], BiVO3 [73] et al., which act as the buffer layer used to obtain high-quality BFO thin film to improve the related properties. Han et al. [71] synthesized BFO photocathode by introducing a thin LaNiO3 (LNO) film to improve the PEC H2O2 production. The current–voltage test proved that the photocurrent can raise from −0.3 to −0.9 mA cm−2 and the onset potential can positive shift from 1.27 eV to 1.38 eV due to the introduction of the LNO dense layer. At the same time, on account of the improvement of carrier dynamics, the incident photon conversion efficiency at 350 nm reached 12% for LNO/BFO photocathode, which was 2.7 times higher than bare BFO electrode. As a result, the researchers obtained excellent H2O2 production of 278 µmol/L, with doubled faradic efficiency, due to rich carrier collections and kinetics. Zhao Yu et al. [72] designed BiFeO3/XTiO3 (X = Sr, Zn, Pb) multilayer films on the LaNiO3 substrate and acquired ingenious photodetection of UV-visible light on account of the advantages of broaden UV absorption of XTiO3, enhanced electric field of the heterojunction and powerful built-in electric field of BiFeO3. The AEM images of BFO/ST film, BFO/ZT film, and BFO/PT film were present on Figure 5a–c. The data of current density-voltage (J–V) in Figure 5d–f showed these films presented favourable photovoltaic performance under monochromatic light. The current density increased with the optical power density up to 8 mW cm−2, while the dark current densities of all films kept low values due to the suitable energy band structure. The band structure under illumination of the BFO/ST film, BFO/ZT film, and BFO/PT film were present on Figure 5g–i, and the photoelectrons migrated from CB of XT to BFO, while the holes migrated from VB of BFO to XT. As a joint result of built-in electric field and heterojunction electric field, the photoelectron conversion of BFO/XT multilayer films stable output.
Due to the existence of a lone pair in Bi3+, BiFeO3 possesses the nature of ferroelectric polarization. Thus, last but not least, substitution of special elements at A-site can inhibit the volatility of Bi [74]. Additionally, the alternative B-site of a transition metal element can diminish the valence state fluctuations of Fe3+, and both of above substitutions can decrease the leakage current of BiFeO3 [75]. You et al. [76] reported the modify BiFeO3 nanofibers by A-site Pr ion and B-site Mn ion co-substitution, and as presented in Figure 6a, under no-poling condition, the photocurrent density was effectively enhanced from 8.2 μA⋅cm2, 17.2 μA/cm2, 31.9 μA/cm2 to 72.5 μA/cm2 for BiFeO3, BiPrFeO3, BiFeMnO3 and BiPrFeMnO3 photoelectrodes under one sun illumination, respectively. Furthermore, the PEC performance of all these photoelectrodes increased when the samples were negatively poled and the current density of BiPrFeMnO3 photoelectrode reached to 131.2 μA/cm2. Meanwhile, the linear sweep voltammetry characteristics in Figure 6b indicated the current density for negatively poled BiPrFeMnO3 photoelectrode up to 145 μA/cm2 and the onset potential left shift from −0.16 V to −0.18 V. As expected, the photocatalytic capability on RhB dye of the Pr and Mn co-doping BiFeO3 nanofibers boosted the degradation performance. As shown in Figure 6c,d, the degradation ratios of BiFeO3, BiPrFeO3, BiFeMnO3 of RhB dye were 17%, 25% and 29%, respectively. However, the obtained co-doped sample could degrade 49% for dye and the photocatalytic rate given rise to 0.0256 min−1. The excellent performance of co-doped BiFeO3 illustrates that proper ion-doping was an efficient way to improve the PEC property for ferroelectrics.

2.4. Bi2FeCrO6

Bi2FeCrO6 (BFO) has a special double perovskite structure, which has great advantages in visible light driven redox reactions due to its relatively small band gap. The Eg is controlled by mutual effect between Fe and Cr via O, while the ferroelectricity is actuated by Bi3+. In 2014, R. Nechache et al. [77] engineered the BFO thin-film by multilayer configuration for power conversion efficiency of 8.1% solar cells, which breached the efficiency previously reported. This article described in detail the depositional condition impact on the property of long-range ordering by characterizing the intensity ratio R and the ordered domain size D, as presented in Figure 7a. Due to the lower direct bandgap and low photogenerated carrier recombination rate, the photoelectric conversion efficiency reached to 3.3% for device with the active layer of low temperature of 580 °C and low-growth-rate of 2 Hz, while choosing SrRuO3 film as bottom electrodes and ITO arrays as transparent conducting electrodes and the schematic diagram was shown in Figure 7b,c. Moreover, after positively polarization of 25 V, the device with multi-active layer of laser repetition rate obtained excellent PEC efficiency of 8.1% as shown in Figure 7d,e. These excellent results demonstrated the enormous application potential in photovoltaic device of ferroelectric-based materials as active layers. Afterwards, in 2017, R. Nechache et al. [78] assembled special p-i-n structure with i-BFO thin-film as sandwich between p-type NiO and n-type Nb-doped SrTiO3(NSTO), as shown in Figure 7f, and yielded the PEC value in Figure 7g of 2.0%, which was four times enhancement compared with the double layer of i-BFO thin-film and n-type Nb-doped SrTiO3 on account of the efficient charge separation driven by the internal polarization as well as above-bandgap generated photovoltage. In addition, in 2019, as exhibited in Figure 7h,i, this team fabricated stable p-NiO/n-BFO heterojunction photoanodes [79] and achieved a high incident photon-to-current efficiency of 3.7% with an enhancement of photocurrent density of 0.4 mA cm−2 compared with the bare BFO devices. The photocatalytic degradation and PEC efficiency of ferroelectric materials has been shown in Table 1.

2.5. 2D Ferroelectrics

Since the discovery of graphene [83], the expanded family of two-dimensional materials has shown abundant physical properties. In terms of the exploration of 2D ferroelectric properties, due to the saturated interfacial chemical environment of layered materials and the weak interaction between layers, it is possible to prepare stable low-dimensional ultrathin ferroelectric films. Recently, a number of representative research reports on ferroelectricity in two-dimensional materials have emerged, and considering the lattice symmetry required for ferroelectrics, some experimental [84] and theoretical calculations [85] have confirmed the existence of ferroelectricity in 2D layered van der Waals materials. For readers seeking more information on the development of 2D layer ferroelectric materials, we recommend them to refer to other review for a broad scope in this area [86]. In 2015, A. Belianinov et al. [87] reported that the layered material CuInP2S6 exhibited ferroelectricity at room temperature. The researchers measured spontaneous ferroelectric domains with obvious hysteresis loops in samples thicker than 100 nm by piezoresponse force microscopy (PFM) and the ferroelectric polarization direction can be reversed under an applied electric field, while the ferroelectricity disappears with a thickness less than 50 nm due to the potential depolarization field. This work confirmed the possibility of ferroelectricity in two-dimensional layered materials at room-temperature and stimulated the exploration of thin-layer and even single-layer limit-thickness ferroelectrics. Then in 2016, also in the CuInP2S6 system, Liu et al. [84] re-studied its ferroelectric properties and reported that double-layer with the thickness of 2 nm CuInP2S6 undergoes out-of-plane spontaneous electrical polarization at room temperature (Figure 8a,b). The dominant peaks in Raman spectrum illustrated the crystal symmetry of CuInP2S6. In this work, they used PFM to probe the piezoelectric response, the presence of stable ferroelectric spontaneous poles can be determined from the amplitude map (Figure 8c), PFM phase (Figure 8d) change and the obvious butterfly-shaped hysteresis loop in Figure 8e. Furthermore, to verify the polarization-controlled erasing and writing of ferroelectrics, they demonstrated the construction of a back-shaped ferroelectric polarization domain could still be maintained, as shown in Figure 8f. In addition, through second harmonic wave (SHG) measurements, they demonstrated that the ferroelectric polarization inversion was accompanied by a structural phase transition, confirming that the transition temperature of the ferroelectric-paraelectric phase transition in 2D CuInP2S6 was above room temperature (~320 K). Additionally, Wang Dong et al. [88] broadened the family of metal phosphorous trichalcogenides M1M2P2X6 (M1 = Cu/Ag, M2 = In/Bi and X = S/Se) and proved that the M1-X bond in the Cu-based and S-based systems have the responsibility of high phase transition temperatures. Furthermore, the composite structure in CuInP2S6/Mn2P2S6 and CuInP2S6/Zn2P2Se6 could be detected via fast charge separation, which could provide the possibility in the area of photocatalytic water splitting. Kou et al. [89] also proved that photocatalytic activities and photoelectron conversion efficiency can be enhanced by monolayer AgBiP2Se6 system based on first principles calculations.
Although few-layer CuInP2S6 exhibits two-dimensional out-of-plane ferroelectricity, it’s not conducive to high-quality preparation and application promotion due to it is polycompound. In view of these shortcomings and also develop more practical two-dimensional out-of-plane ferroelectricity, in 2017, Ding et al. [85] predicted the room-temperature spontaneous ferroelectric polarization with in-plane and out-of-plane coupling in a binary monolayer α-In2Se3 and predicted that this characteristic could be generalized to other III2-VI3 2D van der Waals layered materials. At the same time, this study also demonstrated the controllability of ferroelectric α-In2Se3 by constructing a double-layer heterojunction composed of 2D ferroelectric α-In2Se3 and other 2D materials. Inspired by the above theoretical work, Zhu et al. [90] fabricated α-In2Se3 with a thickness of only 1.1 nm on SiO2/Si substrate by mechanical lift-off method, as shown in Figure 9a,b, and verified the room temperature ferroelectricity in α-phase In2Se3 experimentally, which is close to the reported monolayer c-In2Se3 thickness (~1 nm). The slightly red shift of Raman spectrum indicated the α-phase crystal structure of the In2Se3 thin layers. Due to the piezoelectric properties, ferroelectrics will undergo corresponding deformation in an external electric field. The authors observed spontaneous out-of-plane ferroelectric polarization on α-In2Se3 with a thickness of about 20 nm by PFM in the atmospheric environment, and a phase bias of 180° was be obtained between different ferroelectric domains (Figure 9c), indicating the existence of anti-parallel arrangement of electric dipoles in the out-of-plane direction. By applying an external electric field, the out-of-plane amplitudes (Figure 9e) have obvious butterfly-shaped and well phase nonlinear ferroelectric hysteresis (Figure 9f), indicating that the electric polarization direction was reversed under the external electric field. At the same time, the coercive field (Ec) of ferroelectric α-In2Se3 was about 200 kV/cm, much lower than the coercive field of the 2D layered material CuInP2S6 reported earlier (700 kV/cm) [84], which means that ferroelectric devices based on α-In2Se3 will consume less power. In order to verify the ferroelectric retention of α-In2Se3, as shown in Figure 9d, the authors controllably wrote the out-of-plane polarization by applying a positive bias voltage on a microscopic area of 1 μm × 1 μm, and then applied a reverse bias voltage on the central 0.5 μm × 0.5 μm area to write the electrode by PFM. In addition, the ferroelectric domain still remained after 24 h due to the ferroelectric polarization.

3. Conclusions

Ferroelectric-based photocatalysis and photovoltaic serves a crucial approach for effective and serviceable determination of a target object, such as H2O, CO2, contaminants and so on. This review summarized the progress of ferroelectric materials by introducing the photovoltaic conversion properties of perovskite ferroelectric materials such as BaTiO3, BiFeO3, Bi2FeCrO6, CuInP2S6 and α-In2Se3. Even more to the point, single components exhibit low efficiency because of a broadened band gap and low electron conductivity. Strategies to improve photoelectric conversion performance could be depicted as follows: one proposed solution is ion doping, which can effectively increase the lifetime of carriers, reduce the recombination rate of electron hole pairs and increase the degree of separation of electron hole pairs inside the semiconductor. Another alternative is combining two semiconductors with apposite energy level and bandgap to change band bending, widening the light absorption range and accelerating carriers separation.
On account of the continued development in ferroelectrics as well as photocatalysis and photovoltaic reactions, research in this area is still growing at a high rate, meanwhile, more problems still need to be solved. We believe the future work will focus on improvement of charge separation efficiency, energy conversion efficiency, stability of structure and performance and environmentally friendly. In order to realize the expected above property, several aspects can be prioritized of electrodes in the future: (1) Design and preparation of ferroelectric photocatalytic materials. In consideration of the wider band gaps and lower carrier mobility of currently ferroelectric materials, it is necessary to prepare nanosheets or other nanostructures of ferroelectric with large specific surface area, and make the ferroelectric polarization face away from or face its surface and interface, so as to ensure that the depolarization field P can efficiently separate photogenerated electron hole pairs. (2) Systematic study of PEC processes. The performance determinants of ferroelectric material photoelectrodes should be carefully studied to elucidate the contribution of ferroelectric effects to separated electron hole pairs for further improvements for separation and transfer efficiency of photo-generated electron hole pairs. In more detail, the research on the interface of optoelectronics pole/electrolyte is very important. (3) Study on the stability of catalytic performance of ferroelectric. In practice, the requirements of ferroelectrics not only include the initial photocatalytic performance, but also the stability of catalysts. Up till now, most research has focused on improving the photocatalytic water splitting or pollutants of ferroelectric materials, but fewer articles have been done on stability. The advancement of these aspects would certainly lead to significant advantages compared to the current systems in terms of simplicity, high efficiency and stabilization. We believe that ferroelectric materials will play a pivotal role in the field of environmental applications in the near future.

Author Contributions

Author Contributions: writing—original draft preparation, L.Y. and L.W.; review and editing, Y.Z., P.L. and J.L.; project administration, Y.D.; supervision, W.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Key R & D and Promotion Projects in Henan Province (Grant 212102210128), the Project of Scientific & Technological Innovation Team in Universities of Henan Province (Grant 20IRTSTHN007) and the Science & Technology Innovation Talents in Universities of Henan Province (No. 22HASTIT028) for support.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, L.; Huang, H. Ferroelectrics in Photocatalysis. Chem. Eur. J. 2022, 28, e202103975. [Google Scholar] [CrossRef] [PubMed]
  2. Xu, C.; Ravi Anusuyadevi, P.; Aymonier, C.; Luque, R.; Marre, S. Nanostructured Materials for Photocatalysis. Chem. Soc. Rev. 2019, 48, 3868–3902. [Google Scholar] [CrossRef]
  3. Joy, J.; Mathew, J.; George, S.C. Nanomaterials for Photoelectrochemical Water Splitting-Review. Int. J. Hydrogen Energy 2018, 43, 4804–4817. [Google Scholar] [CrossRef]
  4. Marwat, M.A.; Humayun, M.; Afridi, M.W.; Zhang, H.; Karim, M.R.A.; Ashtar, M.; Usman, M.; Waqar, S.; Ullah, H.; Wang, C.; et al. Advanced Catalysts for Photoelectrochemical Water Splitting. ACS Appl. Energy Mater. 2021, 4, 12007–12031. [Google Scholar] [CrossRef]
  5. A Decade of Perovskite Photovoltaics. Nat. Energy 2019, 4, 1. [CrossRef]
  6. Teng, F.; Hu, K.; Ouyang, W.; Fang, X. Photoelectric Detectors Based on Inorganic P-Type Semiconductor Materials. Adv. Mater. 2018, 30, 1706262. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, B.; Biesold, G.M.; Zhang, M.; Lin, Z. Amorphous Inorganic Semiconductors for the Development of Solar Cell, Photoelectrocatalytic and Photocatalytic Applications. Chem. Soc. Rev. 2021, 50, 6914–6949. [Google Scholar] [CrossRef] [PubMed]
  8. Naldoni, A.; Altomare, M.; Zoppellaro, G.; Liu, N.; Kment, Š.; Zbořil, R.; Schmuki, P. Photocatalysis with Reduced TiO2: From Black TiO2 to Cocatalyst-Free Hydrogen Production. ACS Catal. 2019, 9, 345–364. [Google Scholar] [CrossRef]
  9. Shariatinia, Z. Designing Novel Spiro Compounds as Favorable Hole Transport Materials for Quantum Dot Sensitized Photovoltaics. Sol. Energy 2022, 236, 548–560. [Google Scholar] [CrossRef]
  10. Zheng, B.; Huo, L. Recent Advances of Furan and Its Derivatives Based Semiconductor Materials for Organic Photovoltaics. Small Methods 2021, 5, 2100493. [Google Scholar] [CrossRef]
  11. Steier, L.; Bellani, S.; Rojas, H.C.; Pan, L.; Laitinen, M.; Sajavaara, T.; Di Fonzo, F.; Grätzel, M.; Antognazza, M.R.; Mayer, M.T. Stabilizing Organic Photocathodes by Low-Temperature Atomic Layer Deposition of TiO2. Sustain. Energy Fuels 2017, 1, 1915–1920. [Google Scholar] [CrossRef]
  12. Shao, D.; Cheng, Y.; He, J.; Feng, D.; Zheng, L.; Zheng, L.; Zhang, X.; Xu, J.; Wang, W.; Wang, W.; et al. A Spatially Separated Organic–Inorganic Hybrid Photoelectrochemical Cell for Unassisted Overall Water Splitting. ACS Catal. 2017, 7, 5308–5315. [Google Scholar] [CrossRef]
  13. Cui, W.; Bai, H.; Shang, J.; Wang, F.; Xu, D.; Ding, J.; Fan, W.; Shi, W. Organic-Inorganic Hybrid-Photoanode Built from Nife-Mof and TiO2 for Efficient Pec Water Splitting. Electrochim. Acta 2020, 349, 136383. [Google Scholar] [CrossRef]
  14. Xu, H.; Ding, Z.; Zhang, L.; Liu, J.; Hu, J.; Liu, Y. Cesium-Functionalized Pectin as a Cathode Interlayer for Polymer Solar Cells. J. Mater. Chem. C 2019, 7, 1592–1596. [Google Scholar] [CrossRef]
  15. Forgács, D.; Gil-Escrig, L.; Pérez-Del-Rey, D.; Momblona, C.; Werner, J.; Niesen, B.; Ballif, C.; Sessolo, M.; Bolink, H.J. Efficient Monolithic Perovskite/Perovskite Tandem Solar Cells. Adv. Energy Mater. 2017, 7, 1602121. [Google Scholar] [CrossRef]
  16. Clark, C.P.; Mann, J.E.; Bangsund, J.S.; Hsu, W.-J.; Aydil, E.S.; Holmes, R.J. Formation of Stable Metal Halide Perovskite/Perovskite Heterojunctions. ACS Energy Lett. 2020, 5, 3443–3451. [Google Scholar] [CrossRef]
  17. You, P.; Li, G.; Tang, G.; Cao, J.; Yan, F. Ultrafast Laser-Annealing of Perovskite Films for Efficient Perovskite Solar Cells. Energy Environ. Sci. 2020, 13, 1187–1196. [Google Scholar] [CrossRef]
  18. Busch, G. Early History of Ferroelectricity. Ferroelectrics 1987, 74, 267–284. [Google Scholar] [CrossRef]
  19. Kanzig, W. History of Ferroelectricity 1938–1955. Ferroelectrics 1987, 74, 285–291. [Google Scholar] [CrossRef]
  20. Gao, H.; Yang, Y.; Wang, Y.; Chen, L.; Wang, J.; Yuan, G.; Liu, J.-M. Transparent, Flexible, Fatigue-Free, Optical-Read, and Nonvolatile Ferroelectric Memories. ACS Appl. Mater. Inter. 2019, 11, 35169–35176. [Google Scholar] [CrossRef]
  21. Seidel, J.; Fu, D.; Yang, S.-Y.; Alarcón-Lladó, E.; Wu, J.; Ramesh, R.; Ager, J.W. Efficient Photovoltaic Current Generation at Ferroelectric Domain Walls. Phys. Rev. Lett. 2011, 107, 126805. [Google Scholar] [CrossRef]
  22. Huang, H. Ferroelectric Photovoltaics. Nat. Photonics 2010, 4, 134–135. [Google Scholar] [CrossRef]
  23. Singh, S.; Khare, N. Coupling of Piezoelectric, Semiconducting and Photoexcitation Properties in NaNbO3 Nanostructures for Controlling Electrical Transport: Realizing an Efficient Piezo-Photoanode and Piezo-Photocatalyst. Nano Energy 2017, 38, 335–341. [Google Scholar] [CrossRef]
  24. Wenderich, K.; Mul, G. Methods, Mechanism, and Applications of Photodeposition in Photocatalysis: A Review. Chem. Rev. 2016, 116, 14587–14619. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, X.; Wang, D. Photocatalysis: From Fundamental Principles to Materials and Applications. ACS Appl. Energy Mater. 2018, 1, 6657–6693. [Google Scholar] [CrossRef]
  26. Richards, B.S.; Hudry, D.; Busko, D.; Turshatov, A.; Howard, I.A. Photon Upconversion for Photovoltaics and Photocatalysis: A critical review. Chem. Rev. 2021, 121, 9165–9195. [Google Scholar] [CrossRef]
  27. Bai, Y.; Mora-Seró, I.; De Angelis, F.; Bisquert, J.; Wang, P. Titanium Dioxide Nanomaterials for Photovoltaic Applications. Chem. Rev. 2014, 114, 10095–10130. [Google Scholar] [CrossRef] [PubMed]
  28. Jena, A.K.; Kulkarni, A.; Miyasaka, T. Halide Perovskite Photovoltaics: Background, Status, and Future Prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef] [PubMed]
  29. Cui, Y.; Briscoe, J.; Dunn, S. Effect of Ferroelectricity on Solar-Light-Driven Photocatalytic Activity of BaTiO3—Influence on the Carrier Separation and Stern Layer Formation. Chem. Mater. 2013, 25, 4215–4223. [Google Scholar] [CrossRef]
  30. Liu, J.; Sun, Y.; Li, Z. Ag Loaded Flower-Like BaTiO3 Nanotube Arrays: Fabrication and Enhanced Photocatalytic Property. CrystEngComm 2012, 14, 1473–1478. [Google Scholar] [CrossRef]
  31. Lou, Q.; Shi, X.; Ruan, X.; Zeng, J.; Man, Z.; Zheng, L.; Park, C.H.; Li, G. Ferroelectric Properties of Li-Doped BaTiO3 Ceramics. J. Am. Ceram. Soc. 2018, 101, 3597–3604. [Google Scholar] [CrossRef]
  32. Cortés-Vega, F.D.; Montero-Tavera, C.; Yañez-Limón, J.M. Influence of Diluted Fe3+ Doping on the Physical Properties of BaTiO3. J. Alloys Compd. 2020, 847, 156513. [Google Scholar] [CrossRef]
  33. Maldonado-Orozco, M.C.; Ochoa-Lara, M.T.; Sosa-Márquez, J.E.; Talamantes-Soto, R.P.; Hurtado-Macías, A.; López Antón, R.; González, J.A.; Holguín-Momaca, J.T.; Olive-Méndez, S.F.; Espinosa-Magaña, F. Absence of Ferromagnetism in Ferroelectric Mn-Doped BaTiO3 Nanofibers. J. Am. Ceram. Soc. 2019, 102, 2800–2809. [Google Scholar] [CrossRef]
  34. Ren, P.; Wang, Q.; Wang, X.; Wang, L.; Wang, J.; Fan, H.; Zhao, G. Effects of Doping Sites on Electrical Properties of Yttrium Doped BaTiO3. Mater. Lett. 2016, 174, 197–200. [Google Scholar] [CrossRef]
  35. Cao, W.Q.; Xu, L.F.; Ismail, M.M.; Huang, L.L. Colossal Dielectric Constant of Nanbo Doped Batio Ceramics. Mater. Sci.-Poland 2016, 34, 322–329. [Google Scholar] [CrossRef]
  36. Yang, S.; Dong, G.; Guan, R.; Wu, D. Effect of Knn Doping on the Dielectric Properties of BaTiO3 Lead-Free Ceramics. J. Mater. Sci. Mater. Electron. 2022, 33, 810–816. [Google Scholar] [CrossRef]
  37. Yang, W.; Yu, Y.; Starr, M.B.; Yin, X.; Li, Z.; Kvit, A.; Wang, S.; Zhao, P.; Wang, X. Ferroelectric Polarization-Enhanced Photoelectrochemical Water Splitting in TiO2–BaTiO3 Core–Shell Nanowire Photoanodes. Nano Lett. 2015, 15, 7574–7580. [Google Scholar] [CrossRef]
  38. Fu, B.; Li, J.; Jiang, H.; He, X.; Ma, Y.; Wang, J.; Hu, C. Modulation of Electric Dipoles inside Electrospun BaTiO3@TiO2 Core-Shell Nanofibers for Enhanced Piezo-Photocatalytic Degradation of Organic Pollutants. Nano Energy 2022, 93, 106841. [Google Scholar] [CrossRef]
  39. Cui, Y.; Briscoe, J.; Wang, Y.; Tarakina, N.V.; Dunn, S. Enhanced Photocatalytic Activity of Heterostructured Ferroelectric BaTiO3/A-Fe2O3 and the Significance of Interface Morphology Control. ACS Appl. Mater. Inter. 2017, 9, 24518–24526. [Google Scholar] [CrossRef]
  40. Buscaglia, M.T.; Buscaglia, V.; Curecheriu, L.; Postolache, P.; Mitoseriu, L.; Ianculescu, A.C.; Vasile, B.S.; Zhe, Z.; Nanni, P. Fe2O3@BaTiO3 Core-Shell Particles as Reactive Precursors for the Preparation of Multifunctional Composites Containing Different Magnetic Phases. Chem. Mater. 2010, 22, 4740–4748. [Google Scholar] [CrossRef]
  41. Zhao, W.; Zhang, Q.; Wang, H.; Rong, J.; Lei, E.; Dai, Y. Enhanced Catalytic Performance of Ag2O/BaTiO3 Heterostructure Microspheres by the Piezo/Pyro-Phototronic Synergistic Effect. Nano Energy 2020, 73, 104783. [Google Scholar] [CrossRef]
  42. Liu, Z.; Wang, L.; Yu, X.; Zhang, J.; Yang, R.; Zhang, X.; Ji, Y.; Wu, M.; Deng, L.; Li, L.; et al. Piezoelectric-Effect-Enhanced Full-Spectrum Photoelectrocatalysis in P-N Heterojunction. Adv. Funct. Mater. 2019, 29, 1807279. [Google Scholar] [CrossRef]
  43. Li, R.; Zhang, G.; Wang, Y.; Lin, Z.; He, C.; Li, Y.; Ren, X.; Zhang, P.; Mi, H. Fast Ion Diffusion Kinetics Based on Ferroelectric and Piezoelectric Effect of SnO2/BaTiO3 Heterostructures for High-Rate Sodium Storage. Nano Energy 2021, 90, 106591. [Google Scholar] [CrossRef]
  44. Nageri, M.; Shalet, A.B.; Kumar, V. SnO2-Loaded BaTiO3 Nanotube Arrays: Fabrication and Visible-Light Photocatalytic Application. J. Mater. Sci. Mater. Electron. 2017, 28, 9770–9776. [Google Scholar] [CrossRef]
  45. Wang, L.; Haugen, N.O.; Wu, Z.; Shu, X.; Jia, Y.; Ma, J.; Yu, S.; Li, H.; Chai, Q. Ferroelectric BaTiO3@ZnO Heterostructure Nanofibers with Enhanced Pyroelectrically-Driven-Catalysis. Ceram. Int. 2019, 45, 90–95. [Google Scholar] [CrossRef]
  46. Alex, K.V.; Prabhakaran, A.; Jayakrishnan, A.R.; Kamakshi, K.; Silva, J.P.B.; Sekhar, K.C. Charge Coupling Enhanced Photocatalytic Activity of BaTiO3/MoO3 Heterostructures. ACS Appl. Mater. Inter. 2019, 11, 40114–40124. [Google Scholar] [CrossRef]
  47. Gao, C.; Yu, H.; Zhang, L.; Zhao, Y.; Xie, J.; Li, C.; Cui, K.; Yu, J. Ultrasensitive Paper-Based Photoelectrochemical Sensing Platform Enabled by the Polar Charge Carriers-Created Electric Field. Anal. Chem. 2020, 92, 2902–2906. [Google Scholar] [CrossRef]
  48. Sharma, D.; Upadhyay, S.; Satsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; Dass, S. Nanostructured BaTiO3/Cu2O Heterojunction with Improved Photoelectrochemical Activity for H2 Evolution: Experimental and First-Principles Analysis. Appl. Catal. B Environ. 2016, 189, 75–85. [Google Scholar] [CrossRef]
  49. Li, C.; Fang, T.; Hu, H.; Wang, Y.; Liu, X.; Zhou, S.; Fu, J.; Wang, W. Synthesis and Enhanced Bias-Free Photoelectrochemical Water-Splitting Activity of Ferroelectric BaTiO3/Cu2O Heterostructures under Solar Light Irradiation. Ceram. Int. 2021, 47, 11379–11386. [Google Scholar] [CrossRef]
  50. Meng, K.; Surolia, P.K.; Thampi, K.R. BaTiO3 Photoelectrodes for CdS Quantum Dot Sensitized Solar Cells. J. Mater. Chem. A 2014, 2, 10231–10238. [Google Scholar] [CrossRef]
  51. Fang, T.; Hu, H.; Liu, J.; Jiang, M.; Zhou, S.; Fu, J.; Wang, W.; Yang, Y. Type-II Band Alignment Enhances Unassisted Photoelectrochemical Water-Splitting Performance of the BaTiO3/CdS Ferroelectric Heterostructure Photoanode under Solar Light Irradiation. J. Phys. Chem. C 2021, 125, 18734–18742. [Google Scholar] [CrossRef]
  52. Huang, X.; Wang, K.; Wang, Y.; Wang, B.; Zhang, L.; Gao, F.; Zhao, Y.; Feng, W.; Zhang, S.; Liu, P. Enhanced Charge Carrier Separation to Improve Hydrogen Production Efficiency by Ferroelectric Spontaneous Polarization Electric Field. Appl. Catal. B Environ. 2018, 227, 322–329. [Google Scholar] [CrossRef]
  53. Lei, Y.; Xu, S.; Ding, M.; Li, L.; Sun, Q.; Wang, Z.L. Enhanced Photocatalysis by Synergistic Piezotronic Effect and Exciton-Plasmon Interaction Based on (Ag-Ag2S)/BaTiO3 Heterostructures. Adv. Func. Mater. 2020, 30, 2005716. [Google Scholar] [CrossRef]
  54. Reddy, K.H.; Parida, K. Fabrication, Characterization, and Photoelectrochemical Properties of Cu-Doped PbTiO3 and Its Hydrogen Production Activity. ChemCatChem 2013, 5, 3812–3820. [Google Scholar] [CrossRef]
  55. Yuxiang, H.; Wen, D.; Fengang, Z.; Liang, F.; Mingrong, S. Fe(Iii) Doped and Grafted PbTiO3 Film Photocathode with Enhanced Photoactivity for Hydrogen Production. Appl. Phys. Lett. 2014, 105, 082903. [Google Scholar]
  56. Prajapati, P.; Singh, A.K. Band Gap Tuning of Ferroelectric PbTiO3 by Mo Doping. J. Mater. Sci. Mater. Electron. 2022, 33, 2550–2565. [Google Scholar] [CrossRef]
  57. Abirami, R.; Senthil, T.S.; Kalpana, S.; Kungumadevi, L.; Kang, M. Hydrothermal Synthesis of Pure PbTiO3 and Silver Doped PbTiO3 Perovskite Nanoparticles for Enhanced Photocatalytic Activity. Mater. Lett. 2020, 279, 128507. [Google Scholar] [CrossRef]
  58. Yin, S.; Zhu, Y.; Ren, Z.; Chao, C.; Li, X.; Wei, X.; Shen, G.; Han, Y.; Han, G. Facile Synthesis of PbTiO3 Truncated Octahedra Via Solid-State Reaction and Their Application in Low-Temperature Co Oxidation by Loading Pt Nanoparticles. J. Mater. Chem. A 2014, 2, 9035–9039. [Google Scholar] [CrossRef]
  59. Tabari, T.; Ebadi, M.; Singh, D.; Caglar, B.; Yagci, M.B. Efficient Synthesis of Perovskite-Type Oxide Photocathode by Nonhydrolytic Sol-Gel Method with an Enhanced Photoelectrochemical Activity. J. Alloys Compd. 2018, 750, 248–257. [Google Scholar] [CrossRef]
  60. Li, W.; Wang, F.; Li, M.; Chen, X.; Ren, Z.; Tian, H.; Li, X.; Lu, Y.; Han, G. Polarization-Dependent Epitaxial Growth and Photocatalytic Performance of Ferroelectric Oxide Heterostructures. Nano Energy 2018, 45, 304–310. [Google Scholar] [CrossRef]
  61. He, Y.; Shen, P.; Liu, Y.; Chen, M.; Cao, D.; Yan, X. Integrated Heterostructure of PZT/CdS Containing the Synergistic Effect between Heterojunction Structure and Ferroelectric Polarization for Photoelectrochemical Applications. Mater. Sci. Semicond. Process. 2021, 121, 105351. [Google Scholar] [CrossRef]
  62. Liu, Y.; Ye, S.; Xie, H.; Zhu, J.; Shi, Q.; Ta, N.; Chen, R.; Gao, Y.; An, H.; Nie, W.; et al. Internal-Field-Enhanced Charge Separation in a Single-Domain Ferroelectric PbTiO3 Photocatalyst. Adv. Mater. 2020, 32, 1906513. [Google Scholar] [CrossRef] [PubMed]
  63. Zhen, C.; Yu, J.C.; Liu, G.; Cheng, H.-M. Selective Deposition of Redox Co-Catalyst(S) to Improve the Photocatalytic Activity of Single-Domain Ferroelectric PbTiO3 Nanoplates. Chem. Commun. 2014, 50, 10416–10419. [Google Scholar] [CrossRef] [PubMed]
  64. Manal, B.; Sébastien, S.; Nitul, S.R.; Matthieu, C.; EI Mimoun, M.; Mustapha, J. Experimental and Theoretical Investigations of Low-Dimensional BiFeO3 System for Photocatalytic Applications. Catalysts 2022, 12, 215. [Google Scholar]
  65. Deng, X.-Z.; Song, C.; Tong, Y.-L.; Yuan, G.; Gao, F.; Liu, D.-Q.; Zhang, S.-T. Enhanced Photocatalytic Efficiency of C3N4/BiFeO3 Heterojunctions: The Synergistic Effects of Band Alignment and Ferroelectricity. Phys. Chem. Chem. Phys. 2018, 20, 3648–3657. [Google Scholar] [CrossRef]
  66. Long, J.; Ren, T.; Han, J.; Li, N.; Chen, D.; Xu, Q.; Li, H.; Lu, J. Heterostructured BiFeO3@CdS Nanofibers with Enhanced Piezoelectric Response for Efficient Piezocatalytic Degradation of Organic Pollutants. Sep. Purif. Technol. 2022, 290, 120861. [Google Scholar] [CrossRef]
  67. Huang, Y.-L.; Chang, W.S.; Van, C.N.; Liu, H.-J.; Tsai, K.-A.; Chen, J.-W.; Kuo, H.-H.; Tzeng, W.-Y.; Chen, Y.-C.; Wu, C.-L.; et al. Tunable Photoelectrochemical Performance of Au/BiFeO3 Heterostructure. Nanoscale 2016, 8, 15795–15801. [Google Scholar] [CrossRef]
  68. Tripathi, H.S.; Dutta, A.; Sinha, T.P. Tailoring Structural and Electrochemical Properties in Sr2+ Incorporated Nanostructured BiFeO3 for Enhanced Asymmetric Solidstate Supercapacitor. Electrochim. Acta 2022, 421, 140505. [Google Scholar] [CrossRef]
  69. Cao, D.; Wang, Z.; Nasori; Wen, L.; Mi, Y.; Lei, Y. Switchable Charge-Transfer in the Photoelectrochemical Energy-Conversion Process of Ferroelectric BiFeO3 Photoelectrodes. Angew. Chem. 2014, 53, 11027–11031. [Google Scholar] [CrossRef]
  70. Huang, J.; Wang, Y.; Liu, X.; Li, Y.; Hu, X.; He, B.; Shu, Z.; Li, Z.; Zhao, Y. Synergistically Enhanced Charge Separation in BiFeO3/Sn:TiO2 Nanorod Photoanode Via Bulk and Surface Dual Modifications. Nano Energy 2019, 59, 33–40. [Google Scholar] [CrossRef]
  71. Zhang, Z.; Tan, B.; Ma, W.; Liu, B.; Sun, M.; Cooper, J.K.; Han, W. BiFeO3 Photocathode for Efficient H2O2 Production via Charge Carrier Dynamics Engineering. Mater. Horiz. 2022, 9, 1999–2006. [Google Scholar] [CrossRef]
  72. Li, Z.; Zhao, Y.; Li, W.; Peng, Y.; Zhao, W.; Wang, Z.; Shi, L.; Fei, W. A Self-Powered Flexible UV-Visible Photodetector with High Photosensitivity Based on BiFeO3/XTiO3 (Sr, Zn, Pb) Multilayer Films. J. Mater. Chem. A 2022, 10, 8772–8783. [Google Scholar] [CrossRef]
  73. Ng, Y.H.; Iwase, A.; Kudo, A.; Amal, R. Reducing Graphene Oxide on a Visible-Light BiVO4 Photocatalyst for an Enhanced Photoelectrochemical Water Splitting. J. Phys. Chem. Lett. 2010, 1, 2607–2612. [Google Scholar] [CrossRef]
  74. Tie, Y.; Ma, S.Y.; Pei, S.T.; Zhang, Q.X.; Zhu, K.M.; Zhang, R.; Xu, X.H.; Han, T.; Liu, W.W. Pr Doped BiFeO3 Hollow Nanofibers Via Electrospinning Method as a Formaldehyde Sensor. Sens. Actuators B Chem. 2020, 308, 127689. [Google Scholar] [CrossRef]
  75. Liang, X.-L.; Dai, J.-Q. Prominent Ferroelectric Properties in Mn-Doped BiFeO3 Spin-Coated Thin Films. J. Alloys Compd. 2021, 886, 161168. [Google Scholar] [CrossRef]
  76. You, D.; Liu, L.; Yang, Z.; Xing, X.; Li, K.; Mai, W.; Guo, T.; Xiao, G.; Xu, C. Polarization-Induced Internal Electric Field to Manipulate Piezo-Photocatalytic and Ferro-Photoelectrochemical Performance in Bismuth Ferrite Nanofibers. Nano Energy 2022, 93, 106852. [Google Scholar] [CrossRef]
  77. Nechache, R.; Harnagea, C.; Li, S.; Cardenas, L.; Huang, W.; Chakrabartty, J.; Rosei, F. Bandgap Tuning of Multiferroic Oxide Solar Cells. Nat. Photonics 2015, 9, 61–67. [Google Scholar] [CrossRef]
  78. Huang, W.; Harnagea, C.; Benetti, D.; Chaker, M.; Rosei, F.; Nechache, R. Multiferroic Bi2FeCrO6 Based p-i-n Heterojunction Photovoltaic Devices. J. Mater. Chem. A 2017, 5, 10355–10364. [Google Scholar] [CrossRef]
  79. Huang, W.; Harnagea, C.; Tong, X.; Benetti, D.; Sun, S.; Chaker, M.; Rosei, F.; Nechache, R. Epitaxial Bi2FeCrO6 Multiferroic Thin-Film Photoanodes with Ultrathin P-Type NiO Layers for Improved Solar Water Oxidation. ACS Appl. Mater. Inter. 2019, 11, 13185–13193. [Google Scholar] [CrossRef]
  80. Singh, S.; Sangle, A.L.; Wu, T.; Khare, N.; MacManus-Driscoll, J.L. Growth of Doped SrTiO3 Ferroelectric Nanoporous Thin Films and Tuning of Photoelectrochemical Properties with Switchable Ferroelectric Polarization. ACS Appl. Mater. Inter. 2019, 11, 45683–45691. [Google Scholar] [CrossRef]
  81. Li, S.; Zhang, J.; Zhang, B.-P.; Huang, W.; Harnagea, C.; Nechache, R.; Zhu, L.; Zhang, S.; Lin, Y.H.; Ni, L.; et al. Manipulation of Charge Transfer in Vertically Aligned Epitaxial Ferroelectric KNbO3 Nanowire Array Photoelectrodes. Nano Energy 2017, 35, 92–100. [Google Scholar] [CrossRef]
  82. Qing, L.; Yang, Z.; Lu, Y.; Junling, W.; Mingrong, S.; Liang, F. Enhanced Ferroelectric Photoelectrochemical Properties of Polycrystalline BiFeO3 Film by Decorating with Ag Nanoparticles. Appl. Phys. Lett. 2016, 108, 022902. [Google Scholar]
  83. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Liu, F.; You, L.; Seyler, K.L.; Li, X.; Yu, P.; Lin, J.; Wang, X.; Zhou, J.; Wang, H.; He, H.; et al. Room-Temperature Ferroelectricity in CuInP2S6 Ultrathin Flakes. Nat. Commun. 2016, 7, 12357. [Google Scholar] [CrossRef] [PubMed]
  85. Ding, W.; Zhu, J.; Wang, Z.; Gao, Y.; Xiao, D.; Gu, Y.; Zhang, Z.; Zhu, W. Prediction of Intrinsic Two-Dimensional Ferroelectrics in In2Se3 and Other III2-VI3 Van Der Waals Materials. Nat. Commun. 2017, 8, 14956. [Google Scholar] [CrossRef] [PubMed]
  86. Qi, L.; Ruan, S.; Zeng, Y.-J. Review on Recent Developments in 2d Ferroelectrics: Theories and Applications. Adv. Mater. 2021, 33, 2005098. [Google Scholar] [CrossRef] [PubMed]
  87. Belianinov, A.; He, Q.; Dziaugys, A.; Maksymovych, P.; Eliseev, E.; Borisevich, A.; Morozovska, A.; Banys, J.; Vysochanskii, Y.; Kalinin, S.V. CuInP2S6 Room Temperature Layered Ferroelectric. Nano Lett. 2015, 15, 3808–3814. [Google Scholar] [CrossRef]
  88. Huang, S.; Shuai, Z.; Wang, D. Ferroelectricity in 2D Metal Phosphorus Trichalcogenides and Van Der Waals Heterostructures for Photocatalytic Water Splitting. J. Mater. Chem. A 2021, 9, 2734–2741. [Google Scholar] [CrossRef]
  89. Ju, L.; Shang, J.; Tang, X.; Kou, L. Tunable Photocatalytic Water Splitting by the Ferroelectric Switch in a 2d AgBiP2Se6 Monolayer. J. Am. Chem. Soc. 2020, 142, 1492–1500. [Google Scholar] [CrossRef]
  90. Wan, S.; Li, Y.; Li, W.; Mao, X.; Zhu, W.; Zeng, H. Room-Temperature Ferroelectricity and a Switchable Diode Effect in Two-Dimensional A-In2Se3 Thin Layers. Nanoscale 2018, 10, 14885–14892. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) The scheme of ferroelectric polarization-endowed band engineering of TiO2/BTO core/shell nanowires. (b) Photocurrent density-potential curves of TiO2 NWs and TiO2/BTO NWs under AM1.5G illumination. Adapted with permission from ref. [37]. Copyright: 2015, American Chemical Society. (c) TEM and SAED of necklace-like BaTiO3 nanofibers; (d) the comparison of the corresponding K values for degradation of MO by BT6H@TiO2 core-shell nanofibers under various catalytic conditions. Adapted with permission from ref. [38]. Copyright: 2022, Elsevier.
Figure 1. (a) The scheme of ferroelectric polarization-endowed band engineering of TiO2/BTO core/shell nanowires. (b) Photocurrent density-potential curves of TiO2 NWs and TiO2/BTO NWs under AM1.5G illumination. Adapted with permission from ref. [37]. Copyright: 2015, American Chemical Society. (c) TEM and SAED of necklace-like BaTiO3 nanofibers; (d) the comparison of the corresponding K values for degradation of MO by BT6H@TiO2 core-shell nanofibers under various catalytic conditions. Adapted with permission from ref. [38]. Copyright: 2022, Elsevier.
Nanomaterials 12 03026 g001
Figure 3. Schematic representations of the mechanisms in photo-excited charge transfer from BFO films to the electrolyte ① and from excited surface modifiers to the BFO films ② after the BFO films were (a) positively and (b) negatively poled of 8 V. (c) Photocurrent–potential characteristics of the photoelectrodes with different polarization states. (d) External quantum yield spectra of the BFO electrodes measured with 50 mm Rhodamine B. Inset: the absorption spectrum of Rhodamine B in water. Adapted with permission from ref. [69], copyright: 2014, Wiley.
Figure 3. Schematic representations of the mechanisms in photo-excited charge transfer from BFO films to the electrolyte ① and from excited surface modifiers to the BFO films ② after the BFO films were (a) positively and (b) negatively poled of 8 V. (c) Photocurrent–potential characteristics of the photoelectrodes with different polarization states. (d) External quantum yield spectra of the BFO electrodes measured with 50 mm Rhodamine B. Inset: the absorption spectrum of Rhodamine B in water. Adapted with permission from ref. [69], copyright: 2014, Wiley.
Nanomaterials 12 03026 g003
Figure 4. (a) LSV performance of TiO2, Sn–doped TiO2 and BFO/Sn:TiO2 NRs; (b) J−V curves of BFO/Sn:TiO2 NRs with no poling, positive poling and negative poling; (c) TEM image of a single BFO/Sn:TiO2 NRs; Schematic electronic band diagram of BFO/Sn:TiO2 NRs with (d) positive poling, and (e) negative poling. Adapted with permission from ref. [70], copyright: 2019, Elsevier.
Figure 4. (a) LSV performance of TiO2, Sn–doped TiO2 and BFO/Sn:TiO2 NRs; (b) J−V curves of BFO/Sn:TiO2 NRs with no poling, positive poling and negative poling; (c) TEM image of a single BFO/Sn:TiO2 NRs; Schematic electronic band diagram of BFO/Sn:TiO2 NRs with (d) positive poling, and (e) negative poling. Adapted with permission from ref. [70], copyright: 2019, Elsevier.
Nanomaterials 12 03026 g004
Figure 5. (ac) AFM images of the BFO/XT film; (df) J–V curves under the illumination of 365 nm with different power density of BFO/XT film; (gi) The band structure of the BFO/XT multilayer films under illumination. Adapted with permission from ref. [72], copyright: 2022, Royal Society of Chemistry.
Figure 5. (ac) AFM images of the BFO/XT film; (df) J–V curves under the illumination of 365 nm with different power density of BFO/XT film; (gi) The band structure of the BFO/XT multilayer films under illumination. Adapted with permission from ref. [72], copyright: 2022, Royal Society of Chemistry.
Nanomaterials 12 03026 g005
Figure 6. (a) Transient photocurrent responses of BiFeO3, BiPrFeO3, BiFeMnO3 and BiPrFeMnO3 photoelectrode with different polarization conditions; (b) linear sweep voltammetry curves of BiPrFeMnO3 photo-anodes with different polarization condition; (c) photocatalysis of BiFeO3, BiPrFeO3, BiFeMnO3, BiPrFeMnO3 nanofibers and control sample (blank test without catalyst) in the reaction of aqueous RhB within 25 min, respectively; (d) the corresponding reaction rate constant of photocatalytic activity. Adapted with permission from ref. [76], copyright: 2022, Elsevier.
Figure 6. (a) Transient photocurrent responses of BiFeO3, BiPrFeO3, BiFeMnO3 and BiPrFeMnO3 photoelectrode with different polarization conditions; (b) linear sweep voltammetry curves of BiPrFeMnO3 photo-anodes with different polarization condition; (c) photocatalysis of BiFeO3, BiPrFeO3, BiFeMnO3, BiPrFeMnO3 nanofibers and control sample (blank test without catalyst) in the reaction of aqueous RhB within 25 min, respectively; (d) the corresponding reaction rate constant of photocatalytic activity. Adapted with permission from ref. [76], copyright: 2022, Elsevier.
Nanomaterials 12 03026 g006
Figure 7. (a) R–D relationship in BFCO films grown under different PLD conditions; (b) device layout of the tested BFCO single-layer-based structure; (c) J–V characteristics of BFCO single-layer devices; (d) Device geometry of the tested BFCO multilayer structure; (e) J–V characteristics of BFCO multilayer devices. Adapted with permission from ref. [77], copyright: 2015, Nature Phonos; (f) Layout of the devices for p–i–n; (g) J–V characteristics for p–i–n devices; Adapted with permission from ref. [78], copyright: 2017, Royal Society of Chemistry. (h) J–V vs. RHE for the photoanodes coated with 10 and 20 nm NiO layer; The inset shows the corresponding Voc vs. time; (i) IPCE spectra for BFCO photoanodes with/without NiO layer at 1.23 V (vs. RHE). Adapted with permission from ref. [79], copyright: 2019, American Chemical Society.
Figure 7. (a) R–D relationship in BFCO films grown under different PLD conditions; (b) device layout of the tested BFCO single-layer-based structure; (c) J–V characteristics of BFCO single-layer devices; (d) Device geometry of the tested BFCO multilayer structure; (e) J–V characteristics of BFCO multilayer devices. Adapted with permission from ref. [77], copyright: 2015, Nature Phonos; (f) Layout of the devices for p–i–n; (g) J–V characteristics for p–i–n devices; Adapted with permission from ref. [78], copyright: 2017, Royal Society of Chemistry. (h) J–V vs. RHE for the photoanodes coated with 10 and 20 nm NiO layer; The inset shows the corresponding Voc vs. time; (i) IPCE spectra for BFCO photoanodes with/without NiO layer at 1.23 V (vs. RHE). Adapted with permission from ref. [79], copyright: 2019, American Chemical Society.
Nanomaterials 12 03026 g007
Figure 8. AFM topography (a) PFM amplitude (c) and phase (d) of 2–4 layer thick CIPS on Au coated SiO2/Si substrate. Scale bar in a, 500 nm. (b) the height (black) and PFM amplitude (blue) profile along the lines shown in a and c, respectively. L, layers. (e) The PFM amplitude (black) and phase (blue) hysteresis loops during the switching process for 4 nm thick CIPS flakes. (f) The PFM phase images for 4 nm thick CIPS flakes with written box-in-box patterns with reverse DC bias. Scale bar, 1 μm. Adapted with permission from ref. [84], copyright: 2017, Nature.
Figure 8. AFM topography (a) PFM amplitude (c) and phase (d) of 2–4 layer thick CIPS on Au coated SiO2/Si substrate. Scale bar in a, 500 nm. (b) the height (black) and PFM amplitude (blue) profile along the lines shown in a and c, respectively. L, layers. (e) The PFM amplitude (black) and phase (blue) hysteresis loops during the switching process for 4 nm thick CIPS flakes. (f) The PFM phase images for 4 nm thick CIPS flakes with written box-in-box patterns with reverse DC bias. Scale bar, 1 μm. Adapted with permission from ref. [84], copyright: 2017, Nature.
Nanomaterials 12 03026 g008
Figure 9. (a) AFM topography (10 × 10 µm2) of exfoliated α–In2Se3 thin layers on SiO2/Si substrate; (b) corresponding height profiles taken along the red dash line in (a); (c) the phase profile of different ferroelectric domains; (d) PFM amplitude image of domain engineering in α–In2Se3 with film thickness 12 nm; (e) PFM phase hysteresis loop measured and (f) PFM amplitude from α–In2Se3 thin layers. Adapted with permission from ref. [90], copyright: 2018, Royal Society of Chemistry.
Figure 9. (a) AFM topography (10 × 10 µm2) of exfoliated α–In2Se3 thin layers on SiO2/Si substrate; (b) corresponding height profiles taken along the red dash line in (a); (c) the phase profile of different ferroelectric domains; (d) PFM amplitude image of domain engineering in α–In2Se3 with film thickness 12 nm; (e) PFM phase hysteresis loop measured and (f) PFM amplitude from α–In2Se3 thin layers. Adapted with permission from ref. [90], copyright: 2018, Royal Society of Chemistry.
Nanomaterials 12 03026 g009
Table 1. Photocatalytic degradation and PEC performance using a variety of catalytic methods of ferroelectric material under one sun illumination [38,39,42,61,64,65,67,70,76,80,81,82].
Table 1. Photocatalytic degradation and PEC performance using a variety of catalytic methods of ferroelectric material under one sun illumination [38,39,42,61,64,65,67,70,76,80,81,82].
Material and StructureBandgap/eVLight SourcePCE/%Photocurrent Density/mA·cm−2Catalytic DegradantsCatalytic ActivityReference
BiFeO3 nanoparticle2.20visible light0.4[64]
Au/BiFeO32.62visible light100.55[67]
g–C3N4/BiFeO32.16visible lightRhodamine B90%/60 min[65]
Ag/BiFeO32.46–2.32 full spectrum0.35[82]
KNbO3 nanowire3.28 full spectrum0.115[81]
BaTiO3/α–Fe2O33.14 full spectrumRhodamine B0.0153 min−1[39]
BaTiO3@TiO2 core-shellfull spectrumMethyl orange0.06 min−1[38]
PbTiO3/CdS3.2full spectrum7.40.106[61]
Ag/Nb:SrTiO3full spectrum1.30[80]
BiPrFeMnO32.08full spectrum1.31Rhodamine B0.1352 min−1[76]
Sn:TiO2@BiFeO32.84full spectrumIPCE:82%1.76[70]
TiO2/BaTiO3/Ag2Ofull spectrum291.8[42]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, L.; Wang, L.; Dou, Y.; Zhang, Y.; Li, P.; Li, J.; Wei, W. Recent Advances in Ferroelectric Materials-Based Photoelectrochemical Reaction. Nanomaterials 2022, 12, 3026. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12173026

AMA Style

Yu L, Wang L, Dou Y, Zhang Y, Li P, Li J, Wei W. Recent Advances in Ferroelectric Materials-Based Photoelectrochemical Reaction. Nanomaterials. 2022; 12(17):3026. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12173026

Chicago/Turabian Style

Yu, Limin, Lijing Wang, Yanmeng Dou, Yongya Zhang, Pan Li, Jieqiong Li, and Wei Wei. 2022. "Recent Advances in Ferroelectric Materials-Based Photoelectrochemical Reaction" Nanomaterials 12, no. 17: 3026. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12173026

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