Next Article in Journal
Cybersecurity Risk Analysis of Electric Vehicles Charging Stations
Previous Article in Journal
Development and Validation of an Observational Game Analysis Tool with Artificial Intelligence for Handball: Handball.ai
Previous Article in Special Issue
Starch-g-Acrylic Acid/Magnetic Nanochitin Self-Healing Ferrogels as Flexible Soft Strain Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanostructured Vanadium Dioxide Materials for Optical Sensing Applications

by
Jongwon Yoon
1,*,†,
Woong-Ki Hong
2,*,†,
Yonghun Kim
1 and
Seung-Young Park
2
1
Department of Energy & Electronic Materials, Surface & Nano Materials Division, Korea Institute of Materials Science, Changwon 51508, Republic of Korea
2
Center for Scientific Instrumentation, Korea Basic Science Institute, Daejeon 34133, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Sensors 2023, 23(15), 6715; https://doi.org/10.3390/s23156715
Submission received: 30 June 2023 / Revised: 21 July 2023 / Accepted: 23 July 2023 / Published: 27 July 2023
(This article belongs to the Special Issue Nanostructured Materials Systems for Optical Sensing)

Abstract

:
Vanadium dioxide (VO2) is one of the strongly correlated materials exhibiting a reversible insulator–metal phase transition accompanied by a structural transition from a low-temperature monoclinic phase to high-temperature rutile phase near room temperature. Due to the dramatic change in electrical resistance and optical transmittance of VO2, it has attracted considerable attention towards the electronic and optical device applications, such as switching devices, memory devices, memristors, smart windows, sensors, actuators, etc. The present review provides an overview of several methods for the synthesis of nanostructured VO2, such as solution-based chemical approaches (sol-gel process and hydrothermal synthesis) and gas or vapor phase synthesis techniques (pulsed laser deposition, sputtering method, and chemical vapor deposition). This review also presents stoichiometry, strain, and doping engineering as modulation strategies of physical properties for nanostructured VO2. In particular, this review describes ultraviolet-visible-near infrared photodetectors, optical switches, and color modulators as optical sensing applications associated with nanostructured VO2 materials. Finally, current research trends and perspectives are also discussed.

1. Introduction

The complex interplay between charge, spin, orbital, and lattice degrees of freedom results in the novel electronic and magnetic phenomena in strongly correlated materials (SCMs), as an interesting class of materials in condensed-matter physics [1]. Among SCMs, vanadium dioxide (VO2) has attracted considerable attention, due to the reversible and dramatic changes in conductance and transmittance during metal–insulator transition (MIT), which is a first-order phase transition accompanied by a crystal structure change from a low-temperature monoclinic phase to a high-temperature rutile phase at near-room-temperature (Tc ~ 340 K), as shown in Figure 1a [2,3]. VO2 is a tetragonal rutile (R) structure with space group P42/mnm and lattice constants a = b ≈ 4.55 Å and c ≈ 2.85 Å above Tc, whereas it is a monoclinic M1 structure with space group P21/c and lattice constants a ≈ 5.75 Å, b ≈ 4.53 Å, c ≈ 5.38 Å, b = 122.6° [4]. According to the band theory proposed by Goodenough, the vanadium (V) 3d orbitals are split into σ* (eg) symmetry and π* (t2g) symmetry states, and the t2g states are further split into two dπ orbitals and one d orbital [5]. In the R structure, the Fermi level falls between the π* band and the d band, whereas in the monoclinic structure, the d band is split into two energy bands (d and d*), and a forbidden band with the bandwidth of approximately 0.7 eV between the d band and the π* band is formed [5].
The driving mechanisms behind the MIT in VO2 have been a topic of controversy for decades whether the transition is driven by electron–electron correlations (Mott transition) or by a structure distortion (Peierls transition). Recently, a collaborative Mott-structural transition mechanism in the phase-transition process has also been proposed as an alternative to the two abovementioned mechanisms of the MIT, because both the structural and electron-correlation aspects are important for describing the MIT behavior in VO2 [6,7]. Park and co-workers studied a series of epitaxial VO2 films with different deposition temperatures to understand the cooperation effect between Peierls and Mott transitions in VO2 [6]. They proposed the diagram of band structures, which provides insights into the role of the strain and multivalent V states on the phase transition of VO2, as shown in Figure 1b [6]. In addition, they inferred electronic band structures corresponding to insulating M1 + M2 coexisting phases and metallic M1 and R phases, on the basis of experimental results through hydrogen incorporation in VO2, as shown in Figure 1c [8].
Figure 1. (a) Schematic of crystal and electronic band structures of VO2 in the high-temperature metallic rutile (R) phase and the low-temperature insulating monoclinic (M) phase. In VO2 (R), V4+ ions occupied the corner and center positions and each V4+ is surrounded by six O2−, where the closest V–V distance is approximately 2.85 Å in chains along the c-axis. In VO2 (M), the unit cell is a distorted rutile structure of VO2 (R) and two different V–V distances of 3.19 and 2.60 Å between the nearest V atoms form the zigzag atom chains. Reproduced with permission from [2], Copyright 2017, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Schematic diagram of band structures depicting Peierls, Mott, and collaborative Mott-Peierls transitions. Reproduced with permission from [6], Copyright 2020, American Chemical Society. (c) Band structures corresponding to M1, M2, and R phases. Reproduced with permission from [8], Copyright 2020, American Chemical Society.
Figure 1. (a) Schematic of crystal and electronic band structures of VO2 in the high-temperature metallic rutile (R) phase and the low-temperature insulating monoclinic (M) phase. In VO2 (R), V4+ ions occupied the corner and center positions and each V4+ is surrounded by six O2−, where the closest V–V distance is approximately 2.85 Å in chains along the c-axis. In VO2 (M), the unit cell is a distorted rutile structure of VO2 (R) and two different V–V distances of 3.19 and 2.60 Å between the nearest V atoms form the zigzag atom chains. Reproduced with permission from [2], Copyright 2017, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Schematic diagram of band structures depicting Peierls, Mott, and collaborative Mott-Peierls transitions. Reproduced with permission from [6], Copyright 2020, American Chemical Society. (c) Band structures corresponding to M1, M2, and R phases. Reproduced with permission from [8], Copyright 2020, American Chemical Society.
Sensors 23 06715 g001
Although the MIT mechanism is still unclear, the modification of the phase transition and the manipulation of physical properties in VO2 are possible by a variety of external stimuli, such as light, temperature, stress, stoichiometry, doping, pressure, electric field, and magnetic field [9]. The distinctive properties of MIT triggered by these stimuli have enabled the demonstration of a wide range of applications shown in Figure 2 [9], such as sensors, switches, smart windows, actuators, memory devices, camouflage, and memristors, including electromagnetic absorption materials [10,11].
In the present review, we focus on emerging optical sensing applications based on nanostructured VO2 materials. Firstly, we introduced briefly several synthesis methods of VO2 nanostructures and modification techniques of its physical properties. In addition, we describe the potential applications of VO2 nanostructures for optical sensing—e.g., photodetectors, optical switches, and color modulators. Finally, the current research trends and prospective research areas of VO2 in future applications are also briefly mentioned.

2. Synthesis of Nanostructured VO2 Materials and Modulation of Their Properties

2.1. Synthesis Methods of Nanostructured VO2

The morphology of VO2 depends on synthesis methods, which are primarily categorized solution- and gas-phase-based synthesis methods. For example, sol-gel process and hydrothermal synthesis are representative solution-based chemical approaches, while pulsed laser deposition (PLD), sputtering method, and chemical vapor deposition (CVD) are gas- or vapor-phase synthesis techniques. In previous reports [3,12,13,14,15], various techniques for the fabrication of nanostructured VO2 materials have been described in detail. The advantages and limitations for some of these synthesis methods are summarized in Table 1. Sol-gel or hydrothermal approaches have been used to synthesize nanostructured VO2, mainly for the application of thermochromic smart windows. Meanwhile, PLD, sputtering, and CVD have been used to fabricate high quality VO2 thin films or single-crystals for the application of MIT-related devices. The various nanostructures (e.g., nanowire, nanorod, nanobeam, nanosheet, nanoparticle, and nanoplate), as well as thin films, can be fabricated by using these synthesis methods. The optical sensing applications based on VO2 with different morphologies will be described in Section 3 and, in particular, nanostructured VO2-based photodetectors will be summarized in Table 2.

2.2. Modulation of Physical Properties of Nanostructured VO2

In recent years, considerable efforts have been devoted to manipulate physical properties (e.g., electrical and optical properties) of nanostructured VO2 materials for a variety of applications, such as optical switches, smart window coating, Mott transistors, memristors, sensors, and thermal actuators [12,15]. Most recently, Shi et al. [16] demonstrated the effective phase management of the metallic R phase and insulating phases of monoclinic (M1, M2) and triclinic (T) structures in single-crystalline VO2 microbeams through stoichiometry engineering, as shown in Figure 3 [16]. Figure 3a shows the synthesis process of VO2 microbeams in the nucleation/growth stage, driven by the reduction of high-valence vanadium precursors (V2O5 or V6O13) at T < 850 °C and the stoichiometry-modulation stage for the oxidation or deoxidation of VO2 under different partial pressures of oxygen ( P O 2 ) at T = 850 °C. Using these stoichiometry modulations by adding an appropriate amount of WO2, the single-crystalline W-doped VO2 actuator with a stoichiometry gradient and selective phase stability was proposed, as shown in Figure 3b. In Figure 3c, the VO2 microbeam actuators showed a clear laterally asymmetric configuration and evolution of domains and deflection with increasing temperature. The formation of a radially asymmetric M2-T-M1 domain pattern led to the initial bending at the beginning of the heating stage and with a further increase in temperature, the oxygen-deficient side was gradually occupied by R domains (the oxygen-rich side is occupied by M2 domains). At 60 °C, the entire VO2 beam was transformed into the pure R phase of the straight state. As mentioned in ref. [16], the stoichiometry engineering, which was used to selectively stabilize all the three insulating phases (M1, T, M2) in single-crystalline VO2 microbeams, may open opportunities for designing and controlling phase inhomogeneity and domains of VO2.
In addition to stoichiometry engineering, the ability to control domain structures and phase transitions of VO2 by strain or stress may lead to a deeper understanding of the correlated electron materials exhibiting the MIT, superconductivity, and magnetoresistance [15,17,18]. Cao et al. [19] demonstrated the manipulation of ordered arrays of metal (M) and insulator (I) domains along single-crystal VO2 microbeams by strain engineering, where uniaxial external stress was used to engineer M-I domains and to observe the Mott transition at room temperature, as shown in Figure 4a–c [19]. Figure 4a shows an array of triangle M-I domains which are nucleated and co-stabilized by tensile and compressive strain during heating in a mechanically bent VO2 microbeam. In the stress–temperature phase diagram (Figure 4b), when the M phase fraction η = 1 at high temperatures and high compressive stresses, the system was in pure M phase, while it was in pure I phase when η = 0 at low temperatures and high tensile stresses. The coexistence of M and I phases with the spatial arrangement and relative fraction was shown at intermediate temperatures and stresses. In Figure 4c, room-temperature I–V characteristics of a VO2 microbeam under different axial compressions show the significant reduction of threshold voltages and currents by the external compression upon MIT, implying the possibility of novel device applications using drastic reduction of the operation power through strain engineering of VO2.
The epitaxial VO2 nanostructures grown on single-crystal substrates can be strongly affected by the lattice mismatch with substrate or crystal orientations, resulting in determining the relationship between the stress and strain [17,20]. Figure 4d shows resistivity–temperature curves and the surface morphology of VO2 films grown on TiO2 and Al2O3 single crystals with various crystallographic orientations [21]. The results show that substrate-dependent strains in the VO2 films result in different MIT temperatures. This suggests an enhanced ability to manipulate the MIT properties of VO2 by using lattice strain control through the implementation of a microstructured buffer layer in heteroepitaxial oxide thin films. More recently, Shin et al. [22] demonstrated core-shell heterostructure-enabled stress engineering on MIT, providing accommodation of uniform axial stress and control of the phase-transition pathway and properties in VO2 nanobeams. In this previous study [22], core-shell VO2-Al2O3 (CS-VO2) nanobeams exhibited a simple and direct M1–R phase-transition pathway at a lower temperature without the appearance of metastable intermediate phases (M2 or T), compared to pristine VO2 nanobeams with an M1–M2–R transition pathway, as shown in Figure 4e. These results provide the unique insight that the formation of uniform stress states through core-shell architectures can be applied to the design of phase-transition paths and physical properties for VO2-based device applications using the MIT process.
Figure 4. (a) Optical images of coexisting triangular metallic (M, dark regions) and insulating (I, bright regions) domains nucleated by strain during heating. (b) Phase diagram showing the fraction of the M phase as a function of temperature (x-axis) and uniaxial stress (y-axis, left) or strain (y-axis, right). (c) Room-temperature current-voltage characteristic of a VO2 microbeam under different axial compressions, showing an MIT induced by Joule heating at different threshold voltages and currents. Reproduced with permission from [19], Copyright 2009, Springer Nature. (d) Resistivity-temperature curves for VO2 films grown on TiO2 and Al2O3 substrates. Illustration of the VO2/TiO2/M-Al2O3 heterostructure. Reproduced with permission from [21], Copyright 2017, American Chemical Society. (e) Cross-sectional transmission electron microscopy image and corresponding elemental maps of the rectangular core-shell VO2 (CS-VO2) nanobeam with a 20-nm-thick shell of Al2O3. Stress-temperature phase diagram for CS-VO2 nanobeams (purple-colored arrow) and pristine VO2 nanobeams (red-colored arrow). The arrows show the phase-transition traces on the phase diagram during heating. Temperature-dependent resistance during the heating process for CS-VO2 and pristine VO2 nanobeams. Reproduced with permission from [22], Copyright 2021, Elsevier.
Figure 4. (a) Optical images of coexisting triangular metallic (M, dark regions) and insulating (I, bright regions) domains nucleated by strain during heating. (b) Phase diagram showing the fraction of the M phase as a function of temperature (x-axis) and uniaxial stress (y-axis, left) or strain (y-axis, right). (c) Room-temperature current-voltage characteristic of a VO2 microbeam under different axial compressions, showing an MIT induced by Joule heating at different threshold voltages and currents. Reproduced with permission from [19], Copyright 2009, Springer Nature. (d) Resistivity-temperature curves for VO2 films grown on TiO2 and Al2O3 substrates. Illustration of the VO2/TiO2/M-Al2O3 heterostructure. Reproduced with permission from [21], Copyright 2017, American Chemical Society. (e) Cross-sectional transmission electron microscopy image and corresponding elemental maps of the rectangular core-shell VO2 (CS-VO2) nanobeam with a 20-nm-thick shell of Al2O3. Stress-temperature phase diagram for CS-VO2 nanobeams (purple-colored arrow) and pristine VO2 nanobeams (red-colored arrow). The arrows show the phase-transition traces on the phase diagram during heating. Temperature-dependent resistance during the heating process for CS-VO2 and pristine VO2 nanobeams. Reproduced with permission from [22], Copyright 2021, Elsevier.
Sensors 23 06715 g004
Meanwhile, doping in VO2 has attracted much attention as an effective way for its electrical and optical modulation for electronic and optical device applications [5,23,24,25]. Shao et al. [5] reviewed previous works by Yoon et al. [26] and Zou et al. [27]: (1) a two-step insulator (M-VO2)-to-metal (HxVO2)-to-insulator (HVO2) modulation as the hydrogen concentration increases in nano-sized Pt-island-decorated VO2 layers during annealing the samples at 120 °C, under forming gas containing 5% hydrogen gas (Figure 5a, upper panel); (2) a facile approach to hydrogenate monoclinic VO2 in an acidic solution under ambient conditions, by placing a small piece of low-work function metal (Al, Cu, Ag, Zn, or Fe) on the VO2 surface (Figure 5a, lower panel). Recently, Chet et al. [24] modulated the insertion/extraction of hydrogen into/from the VO2 lattice at room temperature through a solid electrolyte-assisted gating control, resulting in tristate phase transitions that enable the control of light transmittance, as shown in Figure 5b. Strelcov et al. [25] proposed a new high-yield method of doping VO2 nanostructures with aluminum, which could provide possible stabilization of the monoclinic M2 phase for realization of a purely electronic Mott transition field-effect transistor (Figure 5c). According to previous reports [28,29], uniaxial stress and doping can stabilize the M2 phase at ambient conditions. In the schematic diagram depicting phase transformations of VO2 phases by metal-ion dopants (Figure 5c), dopants of higher oxidation states (M = W6+, Nb5+, and Mo6+) lower the transition temperature, whereas dopants of lower oxidation states (M = Cr3+, Al3+, Fe3+, or Ga3+) stabilize the M2 and T phases of VO2 at room temperature [25,30,31]. This behavior shows the influences of reduction and oxidation of the V4+ ions, respectively, in which the oxidation effect is similar to the effect of application of uniaxial stress along the [110] direction of the R phase [25].

3. Optical Sensing Applications

In general, the detectable wavelength ranges for photodetectors are categorized as ultraviolet (UV, 10–400 nm), visible (400–760 nm), near-infrared (NIR, 760–1000 nm), short-wavelength infrared (SWIR, 1–3 μm), mid-wavelength infrared (MWIR, 3–5 μm), and long-wavelength infrared (LWIR, 8–12 μm) [32]. According to these detectable ranges, photodetectors can be used in a variety of potential applications, as shown in Figure 6. VO2 has an optical bandgap (~0.7 eV) and high-temperature coefficient of resistance (TCR), suggesting its potential for optical sensing applications over a wide wavelength range. In addition, the MIT of VO2 allows optical switching in the IR wavelength range and color modulation. In this section, we describe the representative photodetection mechanism and review up-to-date studies for VO2-based photodetectors. Other optical applications of transmission and color modulation are briefly reviewed.

3.1. Detection Mechanisms for VO2-Based Photodetectors

3.1.1. Light-Induced Phase Transition

VO2 typically has a phase-transition characteristic from the insulating M1 phase to metallic R phase above critical temperature (TC) of approximately 67 °C. Instead of thermal sources, optical excitation allows the phase transition of VO2 to occur on a picosecond time scale. However, the metallic state is not permanent and is transformed back into the insulating state [33,34]. Unlike a temporary phase transition, exposure of UV light on a VO2 can induce a permanent phase transition from an insulating to a metallic state [35]. Li et al. [36,37] also reported that a longer UV exposure duration led to a greater reduction in the resistance of the VO2 film, suggesting the possibility of UV detection. It is believed that photo-induced oxygen vacancies can induce an electronic phase transition leading to electrical resistance changes in the VO2 film during UV exposure.

3.1.2. Photoconductive Effect

By the incident photons with the energies greater than the energy bandgap (Eg), the electrical conductivity of the conduction channel can be modulated owing to the generation of electron-hole pairs, which change the carrier concentration in the conduction channel. The electron-hole pairs are separated by the applied electric field, generating a photocurrent (Iph). Iph depends on the applied potential, charge carrier mobility, and carrier lifetime. Because the reported Eg value of monoclinic VO2 was approximately 0.7 eV, the photoconductive effect can be a dominant detection mechanism for the photodetectors using a VO2 as a conductive channel under UV to NIR illumination

3.1.3. Photogating Effect

Because VO2 nanostructures have a large surface-to-volume ratio, oxygen molecules adsorbed on the surface can significantly affect their electrical conductivity by acting as electron acceptors. Under UV illumination, the adsorbed oxygen molecules are desorbed from the surface by recombination with photo-generated holes. Such absorption and desorption of oxygen molecules on the surface of VO2 nanostructures under illumination can modulate the carrier concentration. The incident light acts as a gate for carrier modulation (photogating effect).

3.1.4. Photovoltaic Effect

When the incident photon energy is higher than the Eg of the semiconductor materials, electrons are excited from the valence band to the conduction band, generating electron-hole pairs. Such photoexcited carriers are driven by a built-in electric field arising from semiconductor–semiconductor or semiconductor–metal junctions such as p-n junctions and Schottky junctions. Because a VO2 (M1) with an insulating phase is considered an n-type semiconductor, the formation of contacts with other semiconductor materials enables photodetection via the photovoltaic effect. Recently, heterojunctions between 2D materials and VO2 have been employed for photodetectors using the photovoltaic effect; the examples are explained in the following section.

3.1.5. Photobolometric Effect

The absorption of photon energy can induce resistance changes in materials by increasing temperature. The figure-of-merit of the photodetector based on the photobolometric effect is the TCR, where TCR = dR/R·dT. It is well known that VO2 has high TCR (−4%·K−1) and such a property enables commercial IR photodetectors (microbolometer) [38]. In VO2-based photodetectors, the photobolometric effect is believed to be dominant in the MWIR and LWIR regions.

3.2. VO2-Based Photodetectors

3.2.1. Ultraviolet (UV) Photodetection

The structural change in VO2 from the M1 to the R phase at the phase-transition temperature can induce lattice expansion, leading to tensile stress in the interfaced film. Xin et al. [39] reported a UV photodetector based on a ZnO/VO2 heterostructure with a high responsivity in the UV range (Figure 7a). In this device architecture, the photodetectors exhibited high photoresponsivity and a fast photoresponse at the phase-transition temperature of the VO2 thin film. The structural phase transition of VO2 can induce tensile stress in ZnO thin films owing to lattice expansion. In addition, it is well known that ZnO has piezoelectric properties and that polarization can occur in the crystal. Therefore, the phase transition of VO2 could lead to an internal electric field in ZnO thin films. Such an electric field boosted the charge separation of the electron-hole pairs generated by UV illumination, resulting in a high photoresponsivity (R = 10.07 A·W−1) and a faster photoresponse (rise = 0.020 s, decay = 0.032 s), compared to the device below the phase-transition temperature (R = 0.32 A·W−1, rise = 2.49 s, decay = 2.06 s).
Basyooni et al. [40] reported a UV photodetector with a vertically stacked device configuration consisting of a VO2/MoS2/Si thin film and asymmetric metal contacts for energy-band alignment (Figure 7b). Here, the insertion of the VO2 thin film enabled higher conductivity and photocurrent owing to the high carrier mobility and enhanced photon absorption characteristics, compared to the MoS2/Si device. Consequently, an increase in the photoresponsivity and specific detectivity was observed in the VO2/MoS2/Si heterostructure. The maximum photoresponsivity of the device was 4.7 A·W−1.
Employing nanostructures into the photodetectors is promising due to their large surface-to-volume ratio, which allows for enhanced light absorption and response. Wu et al. [41] demonstrated a UV photodetector using a single VO2 microwire with a one-dimensional (1D) structure (Figure 7c). It is well known that a 1D structure has a relatively large surface-to-volume ratio, which significantly affects the sensing characteristics of devices. In particular, oxygen molecules can be easily adsorbed on the surface of the VO2 microwire, capturing conduction electrons. This can induce surface depletion of the VO2 microwire. When illuminated with UV light, electron-hole pairs can be generated, and the holes can migrate to the surface and recombine with the adsorbed oxygen molecules, reducing surface depletion. As a result, the device showed a significant enhancement in detection performance owing to the photogating effect. The reported photoresponsivity of the device was approximately 7069 A·W−1.
Figure 7. VO2-based UV photodetectors. (a) ZnO/VO2-heterojunction-based UV photodetectors using phase-transition-induced tensile stress in ZnO thin film. Reproduced with permission from [39], Copyright 2020, Royal Society of Chemistry. (b) VO2/MoS2/Si multilayer-based UV photodetectors using energy-band engineering. Reproduced with permission from [40], Copyright 2020, Springer Nature. (c) VO2 microwire-based UV photodetector using a photogating effect. Reproduced with permission from [41], Copyright 2014, American Chemical Society. (d) VO2 thin film-based UV photodetector via UV-induced phase transition. Reproduced with permission from [37], Copyright 2022, Springer Nature.
Figure 7. VO2-based UV photodetectors. (a) ZnO/VO2-heterojunction-based UV photodetectors using phase-transition-induced tensile stress in ZnO thin film. Reproduced with permission from [39], Copyright 2020, Royal Society of Chemistry. (b) VO2/MoS2/Si multilayer-based UV photodetectors using energy-band engineering. Reproduced with permission from [40], Copyright 2020, Springer Nature. (c) VO2 microwire-based UV photodetector using a photogating effect. Reproduced with permission from [41], Copyright 2014, American Chemical Society. (d) VO2 thin film-based UV photodetector via UV-induced phase transition. Reproduced with permission from [37], Copyright 2022, Springer Nature.
Sensors 23 06715 g007
UV exposure of VO2 can induce a phase transition as described previously. Li et al. [36,37] reported that UV illumination at an intensity of 64 mW·cm2 caused the nonvolatile and gradual conductance change in a VO2 by the phase transition, while the illumination of visible light induced volatile conductance change in a VO2 due to rapid recombination after turning off the light, as shown in Figure 6d. Based on these unique optoelectronic properties, the authors demonstrated the application of artificial synaptic devices.

3.2.2. Visible Photodetection

Regarding VO2-based photodetectors, interfacing with two-dimensional (2D) materials enables photodetection in the visible range because of the Eg of the 2D materials. Oliva et al. [42] demonstrated a photodetector based on a MoS2/VO2 heterojunction (Figure 8a). The heterojunction exhibited rectification behavior due to the energy band alignment, leading to a lower leakage current in the reverse bias region. In addition, it showed a relatively higher photoresponsivity in the visible range than multilayer MoS2 devices reported in other studies. The maximum photoresponsivity of the device was approximately 1.25 A·W−1. Luo et al. [43] introduced a WSe2/VO2 heterojunction to form a p-n junction, as shown in Figure 8b. The fabricated photodetector showed dual-mode operation depending on the VO2 phase transition temperature. At room temperature, VO2 had an insulating phase, and the photodetector exhibited photovoltaic properties due to the built-in potential formed by the p-n junction. However, when VO2 had a metallic phase at 90 °C, the photodetector was operated via photoconductive effect, forming a Schottky contact between WSe2 and VO2. The WSe2/VO2 photodetector indicated a relatively high photoresponsivity of 2.4 A·W−1 at room temperature and 6.6 A·W−1 at 90 °C.

3.2.3. Near-IR (NIR) Photodetection

Recently, a VO2-based NIR photodetector using the localized surface plasmon resonance (LSPR) effect was demonstrated to enhance the photodetector performance, as shown in Figure 9a [44]. The device showed significant enhancement in Iph under the illumination of light (λ = 808 nm, P = 8.59 W·cm2) due to the MIT property and LSPR effect leading to the hot electron injection. The fabricated NIR photodetector showed the maximum photoresponsivity of 502.1 mA·W−1.
Xie et al. [45] reported the highly ordered W-doped VO2 nanowire arrays for NIR detection (Figure 9b). One-dimensional nanowire arrays increased the effective area for photon absorption. In addition, it was reported that the Eg of W-doped VO2 was smaller than that of bare VO2 [46], and doped W could prevent the recombination of electron-hole pairs, thereby extending the exciton lifetime [47]. As a result, the W-doped VO2 nanowire arrays showed a much higher photocurrent and photoresponsivity than the bare VO2 nanowire arrays. The photoresponsivity of the W-doped VO2 nanowire array and bare VO2 nanowire array were 21.4 mA·W−1 and 0.29 mA·W−1, respectively.
A VO2/n-Si heterojuction for NIR photodetection, which showed low dark current and linear photoresponse characteristics, was also demonstrated as shown in Figure 9c [46]. In particular, enhanced photoresponsivity and a faster photoresponse were observed at relatively high electric fields and optical power densities. The enhanced optoelectronic performance could be resulted from the MIT of VO2 by the applied electric field and NIR illumination (λ = 940 nm), leading to the efficient collection of the photoexcited electron-hole pairs in n-Si, as shown in the energy band diagram in Figure 9c. The maximum photoresponsivity was 1.01 mA·W−1.
Figure 9. VO2-based NIR photodetectors. (a) Au nanoparticle-decorated VO2-based NIR photodetectors using an LSPR effect. Reproduced with permission from [44], Copyright 2021, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) NIR photodetectors based on W-doped VO2 nanowire arrays using photoconductive effect. Reproduced with permission from [45], Copyright 2018, Elsevier. (c) VO2/p-Si-heterojunction-based NIR photodetector. Phase transition of VO2 by applied external bias and NIR illumination allows the efficient hole collection as shown in the energy band diagram. Reproduced with permission from [48], Copyright 2023, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (d) NIR photodetector based on a CVD-grown VO2 thin film. The device was operated by photoconductive effect under NIR illumination (λ = 850 nm). Reproduced with permission from [49], Copyright 2021, Springer Nature.
Figure 9. VO2-based NIR photodetectors. (a) Au nanoparticle-decorated VO2-based NIR photodetectors using an LSPR effect. Reproduced with permission from [44], Copyright 2021, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) NIR photodetectors based on W-doped VO2 nanowire arrays using photoconductive effect. Reproduced with permission from [45], Copyright 2018, Elsevier. (c) VO2/p-Si-heterojunction-based NIR photodetector. Phase transition of VO2 by applied external bias and NIR illumination allows the efficient hole collection as shown in the energy band diagram. Reproduced with permission from [48], Copyright 2023, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (d) NIR photodetector based on a CVD-grown VO2 thin film. The device was operated by photoconductive effect under NIR illumination (λ = 850 nm). Reproduced with permission from [49], Copyright 2021, Springer Nature.
Sensors 23 06715 g009
Guo et al. [49] demonstrated an NIR photodetector based on a VO2 film synthesized via CVD in which a VO2 film was used as the photoconductor (Figure 9d). When the NIR light (λ = 850 nm) was illuminated on the VO2 channel, a photocurrent was generated by the photoexcited electron-hole pairs and an applied electric field (photoconductive effect). The photoresponsivity of the device was approximately 16 mA·W−1.

3.2.4. IR Photodetection

Infrared (IR) with wavelength longer than NIR is classified into SWIR (1–3 μm), MWIR (3–5 μm), and LWIR (8–12 μm). Because VO2 has a small Eg (~0.7 eV) and a high TCR, photoconductive and photobolometric effects are the dominant photodetection mechanisms. Rajeswaran et al. [50] demonstrated a VO2-based SWIR photodetector (Figure 10a) and observed its electrical properties under illumination (λ = 1550 nm). Because the absorbed photon energy (~0.8 eV) is higher than the Eg of VO2, the photoexcited electron-hole pairs can be generated and contribute to the photocurrent of the device. At high optical power density and applied bias region, the device showed a high photoresponse and then the maximum photoresponsivity was 7.13 × 10−2 mA·W−1.
Fu et al. [51] fabricated the photodetector based on vertically stacked 1D VO2 nanowire/carbon nanotube (CNT) composite film (Figure 10b). The composite-film-based photodetector exhibited both enhanced photoresponsivity and a faster photoresponse than the VO2-nanowire-based photodetector. The CNT film played a role as a medium for heat absorption and transfer between the CNT and VO2 films, leading to improved IR response characteristics. The photoresponsivity of the device was 17.83 mA·W−1. Ma et al. [52] used a VO2/silicon nitride (SN) composite film for IR photodetection. A flexible and freestanding thin film photodetector was fabricated using SN nanotubes. The device showed strong IR absorption and low heat capacity, leading to enhanced IR photodetection.

3.2.5. Broadband Photodetection

Recently, VO2-based broadband photodetectors have been intensively studied due to their potential applications. Kabir et al. [53] demonstrated a broadband photodetector based on a VO2 thin film synthesized by DC sputtering and annealing in ambient air (Figure 11a). According to the literature, the broadband photodetection of a device is attributed to the photo-excitation, electrical excitation, and thermal excitation, simultaneously. In particular, the enhanced photoresponse after the phase transition from the insulating to the metallic phases of VO2 is attributed to free carriers. The highest photoresponsivity of the device was approximately 2 A·W−1 at the metallic phase of VO2 and in the visible range. Umar et al. reported similar device configurations but used different metal contacts (Ag). The authors suggested that the photocurrent was generated by a photon energy higher than the Eg of VO2 [54].
Hong et al. [55] introduced a glancing angle deposition method to form vertically aligned VO2 nanorods, as shown in Figure 11b. The porous VO2 nanorods provided a wide specific area for light absorption. In addition, Ag-nanoparticle-decorated VO2 was employed to induce an LSPR effect. The electric field at the interface between the Ag nanoparticles and VO2 nanorods enabled the broadband photodetector (visible to NIR) and enhanced the photoresponse of the device. Interestingly, the device showed high photoresponsivity from visible to NIR ranges and the maximum photoresponsivity was approximately 103 A·W−1 in the NIR range.
Hassan et al. [56] demonstrated a photodetector employing a vertical VO2/p-Si heterojunction (Figure 11c). The formed p-n junction enabled the self-powered operation of the photodetectors. The excited photocarriers could be separated using the built-in electric field of the device. The maximum photoresponsivity of the photodetector was approximately 0.02 mA·W−1. Jiang et al. [57] introduced a VO2/MoTe2 heterojunction into a photodetector to form a p-n junction (Figure 11d). The fabricated photodetector exhibited broadband photodetection properties in the visible-to-SWIR range. In addition, due to the high TCR of the VO2 film, the MWIR and LWIR were detectable. The photoresponsivity observed in the NIR range (λ = 830 nm) was approximately 0.22 A·W−1.
Figure 11. VO2-based broadband photodetectors. (a) Broadband photodetector that is detectable from UV to NIR. Such a photodetection property contributes to the photo-excitation, electrical excitation, and thermal excitation, simultaneously. Reproduced with permission from [53], Copyright 2020, Elsevier. (b) Vertically aligned and Ag-decorated VO2 nanorod-based broadband photodetector allowing detection from visible to NIR using LSPR effect by Ag nanoparticles. Reproduced with permission from [55], Copyright 2019, American Chemical Society. (c) VO2/p-Si heterojunction-based broadband photodetector (visible to NIR) using photovoltaic effect. The p-n junction forms the built-in potential leading to the efficient separation of photoexcited carriers. Reproduced with permission from [56], Copyright 2022, Elsevier. (d) VO2/MoTe2 heterojunction-based broadband photodetectors with dual-mode operation. The device operates by photovoltaic effect in visible-to-SWIR range, and by photobolometric effect in MWIR-to-LWIR range. Reproduced with permission from [57], Copyright 2020, Springer Nature.
Figure 11. VO2-based broadband photodetectors. (a) Broadband photodetector that is detectable from UV to NIR. Such a photodetection property contributes to the photo-excitation, electrical excitation, and thermal excitation, simultaneously. Reproduced with permission from [53], Copyright 2020, Elsevier. (b) Vertically aligned and Ag-decorated VO2 nanorod-based broadband photodetector allowing detection from visible to NIR using LSPR effect by Ag nanoparticles. Reproduced with permission from [55], Copyright 2019, American Chemical Society. (c) VO2/p-Si heterojunction-based broadband photodetector (visible to NIR) using photovoltaic effect. The p-n junction forms the built-in potential leading to the efficient separation of photoexcited carriers. Reproduced with permission from [56], Copyright 2022, Elsevier. (d) VO2/MoTe2 heterojunction-based broadband photodetectors with dual-mode operation. The device operates by photovoltaic effect in visible-to-SWIR range, and by photobolometric effect in MWIR-to-LWIR range. Reproduced with permission from [57], Copyright 2020, Springer Nature.
Sensors 23 06715 g011

3.3. Summary of VO2-Based Photodetectors

Table 2 shows a summary of the up-to-date VO2-based photodetectors. In this table, we summarize the materials, detectable wavelengths, deposition methods for VO2, photoresponsivity (R), and specific detectivity (D *) of the photodetectors.
Table 2. Summary of VO2-based photodetectors.
Table 2. Summary of VO2-based photodetectors.
Materials Typeλ (nm)MethodologyR (A/W)D * (Jones)Ref.
VO2375PLD--[36,37]
VO2/Nb-doped
TiO2
254/405DC sputter15.7 (@ 405 nm)
35.6 (@254 nm)
-[58]
MoWO3/VO2/
MoS2/Si
365RF sputter4.74.3 × 108[40]
VO2 (1D)360–400CVD70691.5 × 1014[41]
ZnO/VO2365PLD10.071.2 × 1010[39]
VO2/MoS2500–700DC Sputter1.25-[42]
VO2/WSe2532DC Sputter2.4 (@ RT)
6.6 (@ 90 °C)
1.9 × 1013 (@ RT)
1.8 × 1011 (@ 90 °C)
[43]
Au/VO2808DC Sputter0.261.14 × 1011[44]
W-doped VO2
(1D)
980Hydrothermal0.02-[45]
VO2/n-Si940ALD0.0011.0 × 1012[48]
H-doped VO2
nanoparticles
780Sol-gel3.6 × 1041.1 × 1013[59]
VO2/p-Si850RF sputter14.87.0 × 1012[60]
VO2850CVD0.02-[49]
VO21550CVD7.1 × 10−51.1 × 1011[50]
VO21064/1550DC sputter0.014
(@ 1064 nm)
1.7 × 1012 (@ 1064 nm)[61]
VO2 (1D)/CNTIRHydrothermal0.6 × 10−3-[51]
VO2/ZnO365/525/1064Hydrothermal0.5 × 10−3 (@ 365 nm)2.7 × 109 (@ 365 nm)[62]
VO2365–850DC sputter0.9 (@ 565 nm, M1)
2.1 (@ 850 nm, R)
9.4 × 109 (@ 565 nm, M1)
4.6 × 109 (@ 850 nm, M1)
[53]
VO2/Si650/980PLD0.35 (@ 650 nm)-[54]
VO2/MoTe2450–2000,
2800–10,000
DC sputter0.22 (@ 830 nm)3.0 × 1010 (@ 830 nm)[57]
VO2/p-Si456,515,950PLD2.0 × 10−5 (@ 950 nm)-[56]
Ag/VO2 (1D)400–1000E-beam evaporation4.1 × 1031.4 × 1014[55]

3.4. Optical Switching and Color Modulator Applications

VO2 has been extensively investigated as a tunable material for optical modulation systems due to the drastic change in the IR transmittance and the refractive index across the MIT [63,64,65]. Long and co-workers demonstrated static visible light tunability and dynamic NIR modulation of two-dimensional SiO2-VO2 core-shell photonic crystal films, as shown in Figure 12a [63]. The SiO2-VO2 core-shell photonic crystal-based thermochromic smart window can show the tunable functionality via selectively reflecting and blocking light (indicated by red, blue, orange, and green arrows) and simultaneously maintaining (attenuating) IR transmission at low temperature (at high temperature) (Figure 12(ai)). The finite difference time domain simulations for transmission spectra of the photonic crystal structure predict that the transmittance can be tuned across the visible spectrum, while maintaining good solar regulation efficiency (ΔTsol = 11.0%) and high solar transmittance (Tlum = 49.6) (Figure 12(aii,iii)).
Liang et al. [64] presented dual-band modulation of visible and NIR transmittance through voltage and temperature in a hybrid micro–nano composite film, which contains the microsized liquid crystals domains with a negative dielectric constant and tungsten-doped vanadium dioxide (W-VO2) nanocrystals (Figure 12b). The light modulation performance of the films with 2.5 and 5.0 wt.% W-VO2 nanocrystals showed transparency in the visible region and a drastic change in NIR transmittance at different temperatures (Figure 12(bi)). This result indicates that NIR light transmittance of the hybrid composite film can be passively modulated according to the temperature variations. In addition, the visible light transmittance of the hybrid composite film can be independently and dynamically regulated by the external voltages (Figure 12(bii)). Specifically, Vis/NIR spectra of the films with 2.5 wt.% of W-VO2/PVP nanocrystals showed that the visible light transmittance of the film gradually decreased due to a spatial variation of IR between the micro-liquid crystal domains and the polymer during the increase in the applied voltages from 0 to 35 V, resulting from the parallel alignment to the direction of the electric field of liquid crystals (Figure 12(bii)).
Wan et al. [65] demonstrated a VO2-based limiting optical diode as a nonlinear device that features asymmetric transmission of light, which was bidirectionally transparent at low power but opaque during illumination of a sufficiently intense light incident from a particular direction. The proof-of-concept of a VO2-based limiting optical diode comprising a transparent sapphire substrate, a thin VO2 layer, and a semitransparent gold film shows the asymmetric absorption of a VO2 thin film to selectively trigger the MIT, enabling asymmetric transmission (Figure 12(ci)). For the case of forward incidence, a significant amount of power is reflected before it reaches the VO2, whereas there is substantially more absorption of the light by the VO2 in the case of backward incidence. This could be supported by the temperature-dependent infrared refractive indices of the VO2 film. Due to a large change in n and κ for a relatively small change in temperature, a VO2-based limiting optical diode can be designed to operate over a broad wavelength range (1–3.5 μm) (Figure 12(cii)) and, by using VO2 films with narrower transitions, maximal asymmetry can be reached in the simple thin-film geometry (Figure 12(ciii)).
In addition to light transmission modulation of VO2, the device applications using dynamic color modulation based on the phase transition of VO2, in combination with nanostructured metals, were recently reported [66,67,68]. Shu et al. [66] demonstrated color generation for display and imaging applications through the integration of plasmonic nanostructures with periodic silver-nanodisk arrays on VO2 film (Figure 13a). As illustrated in Figure 13a, the reflection images of samples can readily be tuned by adjusting the geometric parameters (diameter and periodicity) of the nanodisks, and the color of the sample changes with the increase in nanodisk diameter at both 20 and 80 °C. In Figure 13a, the scanning electron microscopy (SEM) images for four different patterns of VO2 and periodic silver-nanodisk arrays showed distinctively different reflection color images of the patterns at 20 and 80 °C, indicating the realization of abundant color variation due to the MIT of VO2 and surface plasmon effect of metal nanostructures. Liu and co-workers demonstrated reconfigurable multistate optical systems enabled by phase transitions in VO2, which could be modulated by thermal tuning, hydrogen (H)-doping, and electron (e)-doping, as shown in Figure 13b [67]. Specifically, they presented a quadruple-state dynamic plasmonic display based on stacked structures of aluminum (Al)/Al2O3 nanodisks on a VO2/Au mirror substrate with Pd dots, in response to a combination of temperature and H-doping (Figure 13b, bottom panels). Meanwhile, In et al. [68] proposed composites of self-organized gold network (SGN) and VO2 as promising templates for photonic applications, combining advantages of both MIT hysteresis and strong light–matter interactions. They demonstrated thermoactive cyan–magenta–yellow color filters based on SGN–VO2 hybrid films, which were fabricated on 2 in. sapphire wafers with various VO2 thickness, as shown in Figure 13c.
Figure 12. (a) (i) Working mechanism of photonic VO2 smart window; (ii) illustration of designed structures for simulation; and (iii) simulated transmittance spectrum, where transmittance peaks and troughs are indicated by the solid and dashed arrows, respectively. The colorful background in (iii) denotes the visible spectrum from 370 to 770 nm. Reproduced with permission from [63], Copyright 2016, American Chemical Society. (b) Schematic illustration of the optical behaviors of the as-prepared film upon the heating or cooling (i, upper panel) and by applying or removing the voltage at low environmental temperature (ii, upper panel). Vis/NIR transmittance spectra from 400 nm to 2500 nm for the films with W-VO2 nanocrystals at 20 and 55 °C (i, lower panel), and under the application of various voltages at 20 °C (ii, lower panel). Reproduced with permission from [64], Copyright 2017, American Chemical Society. (c) (i) Schematic of a planar-limiting optical diode made from sputtered VO2 and its real and imaginary refractive indices in its insulating (25 °C) and metallic (90 °C) phases. (ii) Transmission spectra of the limiting optical diode in the unswitched (blue, insulating) and switched (red, metallic) states at room temperature and 90 °C, respectively. (iii) Simulation results, including VO2 films with a narrower phase transition [69]. Reproduced with permission from [65], Copyright 2018, American Chemical Society.
Figure 12. (a) (i) Working mechanism of photonic VO2 smart window; (ii) illustration of designed structures for simulation; and (iii) simulated transmittance spectrum, where transmittance peaks and troughs are indicated by the solid and dashed arrows, respectively. The colorful background in (iii) denotes the visible spectrum from 370 to 770 nm. Reproduced with permission from [63], Copyright 2016, American Chemical Society. (b) Schematic illustration of the optical behaviors of the as-prepared film upon the heating or cooling (i, upper panel) and by applying or removing the voltage at low environmental temperature (ii, upper panel). Vis/NIR transmittance spectra from 400 nm to 2500 nm for the films with W-VO2 nanocrystals at 20 and 55 °C (i, lower panel), and under the application of various voltages at 20 °C (ii, lower panel). Reproduced with permission from [64], Copyright 2017, American Chemical Society. (c) (i) Schematic of a planar-limiting optical diode made from sputtered VO2 and its real and imaginary refractive indices in its insulating (25 °C) and metallic (90 °C) phases. (ii) Transmission spectra of the limiting optical diode in the unswitched (blue, insulating) and switched (red, metallic) states at room temperature and 90 °C, respectively. (iii) Simulation results, including VO2 films with a narrower phase transition [69]. Reproduced with permission from [65], Copyright 2018, American Chemical Society.
Sensors 23 06715 g012
Figure 13. (a) Optical properties of VO2 film comprising periodic silver-nanodisk array with different diameters. Reflection images of samples with nanodisk diameters ranging from 100 to 190 nm in steps of 10 nm at 20 and 80 °C. Magnified SEM images of the four different pattern-containing regions, where all scale bars are 400 nm in length. SEM image of the patterns (I, II, III, and IV); the sale bar is 40 µm in length. Reflection images of the patterns with different SiO2 thickness at 20 °C and at 80 °C; the scale bars are 40 µm in length. Reproduced with permission from [66], Copyright 2018, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Illustration depicting reconfigurable multistate optical systems based on VO2 phase transitions modulated by temperature, e-doping, and H-doping. Schematic of stacked Al2O3/Al nanodisks on a VO2/Au mirror substrate with nanosized Pd dots, where the Pd dots were utilized to facilitate its hydrogenation and dehydrogenation. SEM images of a palette square. Scale bar: 200 nm. Optical micrographs of a color palette with stepwise tuning of diameter (D) and gap (g) at the four different states, monoclinic (m)-VO2, rutile (r)-VO2, H-VO2 (M), and H-VO2(I). Reproduced with permission from [67], Copyright 2020, American Chemical Society. (c) Self-organized gold network (SGN)–VO2 hybrid films for color filter application. Photographs of bare VO2 (80 nm) film (1) and SGN–VO2 hybrid films (2–4) for various Au thicknesses (VO2 thickness fixed at 80 nm). Photographs of SGN–VO2 hybrid films (5–8) for various VO2 thicknesses (Au thickness fixed at 50 nm). Reproduced with permission from [68], Copyright 2020, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 13. (a) Optical properties of VO2 film comprising periodic silver-nanodisk array with different diameters. Reflection images of samples with nanodisk diameters ranging from 100 to 190 nm in steps of 10 nm at 20 and 80 °C. Magnified SEM images of the four different pattern-containing regions, where all scale bars are 400 nm in length. SEM image of the patterns (I, II, III, and IV); the sale bar is 40 µm in length. Reflection images of the patterns with different SiO2 thickness at 20 °C and at 80 °C; the scale bars are 40 µm in length. Reproduced with permission from [66], Copyright 2018, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Illustration depicting reconfigurable multistate optical systems based on VO2 phase transitions modulated by temperature, e-doping, and H-doping. Schematic of stacked Al2O3/Al nanodisks on a VO2/Au mirror substrate with nanosized Pd dots, where the Pd dots were utilized to facilitate its hydrogenation and dehydrogenation. SEM images of a palette square. Scale bar: 200 nm. Optical micrographs of a color palette with stepwise tuning of diameter (D) and gap (g) at the four different states, monoclinic (m)-VO2, rutile (r)-VO2, H-VO2 (M), and H-VO2(I). Reproduced with permission from [67], Copyright 2020, American Chemical Society. (c) Self-organized gold network (SGN)–VO2 hybrid films for color filter application. Photographs of bare VO2 (80 nm) film (1) and SGN–VO2 hybrid films (2–4) for various Au thicknesses (VO2 thickness fixed at 80 nm). Photographs of SGN–VO2 hybrid films (5–8) for various VO2 thicknesses (Au thickness fixed at 50 nm). Reproduced with permission from [68], Copyright 2020, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Sensors 23 06715 g013

4. Summary and Outlook

In this review, we first introduced several solution-based and gas-phase-based synthesis methods of nanostructured VO2 and modulation approaches of its properties, such as stoichiometry, strain (or stress), and doping. Among the potential applications of VO2 nanostructures and optical sensing devices, including photodetectors, optical switches, and color modulators, were discussed. Specifically, we reviewed and summarized photodetection mechanisms and VO2-based photodetectors for UV, visible, NIR, and IR lights, including optical transmission and dynamic color modulations. Many researchers have devoted to modulate the MIT properties and have demonstrated the design of various device applications.
In addition to the applications mentioned in this review, very recently, VO2 has been considered as one of the promising materials for energy-efficient neuromorphic computing applications, owing to the rise of artificial intelligence related to the fourth industrial revolution [23,70,71,72,73,74]. Moreover, VO2 has attracted much attention as a promising material for adaptive radiative cooling due to a thermochromic property, offering a potential way to reduce energy consumption in buildings [75,76,77]. However, there are still some challenges, such as scalable and reliable fabrication of nanostructured VO2 materials and precise control of their phase-transition properties (transition temperature, resistivity ratio, hysteresis, transition pathway, phase coexistence, etc.). Thus, it is still necessary to achieve a comprehensive understanding of the MIT in VO2, substantial progress for practical uses, and enhanced regulation of the physical and chemical processes associated with future research fields, including the applications discussed above. Finally, further advancements in VO2 will pave the way for new possibilities and opportunities, enabling the expansion of VO2-based devices into a wider range of innovative functional applications.

Author Contributions

Conceptualization, W.-K.H. and J.Y.; Validation, W.-K.H. and J.Y.; Resources, W.-K.H. and J.Y.; Data Curation, W.-K.H. and J.Y.; writing—original draft preparation, W.-K.H. and J.Y.; writing—review and editing, W.-K.H., J.Y., Y.K. and S.-Y.P.; visualization, W.-K.H. and J.Y.; project administration, W.-K.H. and J.Y.; funding acquisition, W.-K.H., J.Y., Y.K. and S.-Y.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Korea Basic Science Institute (KBSI) (C330210), the Ministry of Science and ICT (Project Number: 2023-22030002-10) and Commercialization Promotion Agency for R&D Outcomes (COMPA), and the National Research Council of Science and Technology (NST) grant by the Korean Government (MSIT) (CPS22191-100).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brahlek, M.; Zhang, L.; Lapano, J.; Zhang, H.-T.; Engel-Herbert, R.; Shukla, N.; Datta, S.; Paik, H.; Schlom, D.G. Opportunities in vanadium-based strongly correlated electron systems. MRS Commun. 2017, 7, 27–52. [Google Scholar] [CrossRef] [Green Version]
  2. Li, M.; Magdassi, S.; Gao, Y.; Long, Y. Hydrothermal synthesis of VO2 polymorphs: Advantages, challenges and prospects for the application of energy efficient smart windows. Small 2017, 13, 1701147. [Google Scholar] [CrossRef]
  3. Devthade, V.; Lee, S. Synthesis of vanadium dioxide thin films and nanostructures. J. Appl. Phys. 2020, 128, 231101. [Google Scholar] [CrossRef]
  4. Liu, K.; Lee, S.; Yang, S.; Delaire, O.; Wu, J. Recent progresses on physics and applications of vanadium dioxide. Mater. Today 2018, 21, 875–896. [Google Scholar] [CrossRef] [Green Version]
  5. Shao, Z.; Cao, X.; Luo, H.; Jin, P. Recent progress in the phase-transition mechanism and modulation of vanadium dioxide materials. NPG Asia Mater. 2018, 10, 581–605. [Google Scholar] [CrossRef] [Green Version]
  6. Lee, D.; Yang, D.; Kim, H.; Kim, J.; Song, S.; Choi, K.S.; Bae, J.-S.; Lee, J.; Lee, J.; Lee, Y.; et al. Deposition-temperature-mediated selective phase transition mechanism of VO2 films. J. Phys. Chem. C 2020, 124, 17282–17289. [Google Scholar] [CrossRef]
  7. Zhou, Y.; Ramanathan, S. Mott memory and neuromorphic devices. Proc. IEEE 2015, 103, 1289–1310. [Google Scholar] [CrossRef]
  8. Lee, D.; Min, T.; Lee, G.; Kim, J.; Song, S.; Lee, J.; Bae, J.-S.; Kang, H.; Lee, J.; Park, S. Understanding the phase transition evolution mechanism of partially M2 phased VO2 film by hydrogen incorporation. J. Phys. Chem. Lett. 2020, 11, 9680–9688. [Google Scholar] [CrossRef]
  9. Ke, Y.; Wang, S.; Liu, G.; Li, M.; White, T.J.; Long, Y. Vanadium dioxide: The multistimuli responsive material and its applications. Small 2018, 14, 1802025. [Google Scholar]
  10. Mao, F.; Fan, X.; Long, L.; Li, Y.; Chen, H.; Zhou, W. Constructing 3D hierarchical CNTs/VO2 composite microspheres with superior electromagnetic absorption performance. Ceram. Int. 2023, 49 Pt A, 16924–16931. [Google Scholar] [CrossRef]
  11. Mao, F.; Long, L.; Zeng, G.; Chen, H.; Li, Y.; Zhou, W. Achieving excellent electromagnetic wave absorption property by constructing VO2 coated biomass carbon heterostructures. Diam. Relat. Mater. 2022, 130, 109422. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Xiong, W.; Chen, W.; Zheng, Y. Recent progress on vanadium dioxide nanostructures and devices: Fabrication, properties, applications and perspectives. Nanomaterials 2021, 11, 338. [Google Scholar] [CrossRef]
  13. Wang, S.; Owusu, K.A.; Mai, L.; Ke, Y.; Zhou, Y.; Hu, P.; Magdassi, S.; Long, Y. Vanadium dioxide for energy conservation and energy storage applications: Synthesis and performance improvement. Appl. Energy 2018, 211, 200–217. [Google Scholar]
  14. Zeng, W.; Chen, N.; Xie, W. Research progress on the preparation methods for VO2 nanoparticles and their application in smart windows. CrystEngComm 2020, 22, 851–869. [Google Scholar] [CrossRef]
  15. Liu, C.; Wang, Y.; Tian, Z.; Mei, Y. Low-dimensional vanadium dioxide nanomaterials: Fabrication, properties and applications. J. Phys. Mater. 2020, 3, 032007. [Google Scholar]
  16. Shi, R.; Chen, Y.; Cai, X.; Lian, Q.; Zhang, Z.; Shen, N.; Amini, A.; Wang, N.; Cheng, C. Phase management in single-crystalline vanadium dioxide beams. Nat. Commun. 2021, 12, 4214. [Google Scholar] [CrossRef]
  17. Hong, B.; Yang, Y.; Hu, K.; Dong, Y.; Zhou, J.; Zhang, Y.; Zhao, W.; Luo, Z.; Gao, C. Strain engineering on the metal-insulator transition of VO2/TiO2 epitaxial films dependent on the strain state of vanadium dimers. Appl. Phys. Lett. 2019, 115, 251605. [Google Scholar] [CrossRef]
  18. Ma, H.; Xiao, X.; Wang, Y.; Sun, Y.; Wang, B.; Gao, X.; Wang, E.; Jiang, K.; Liu, K.; Zhang, X. Wafer-scale freestanding vanadium dioxide film. Sci. Adv. 2021, 7, eabk3438. [Google Scholar] [CrossRef] [PubMed]
  19. Cao, J.; Ertekin, E.; Srinivasan, V.; Fan, W.; Huang, S.; Zheng, H.; Yim, J.W.L.; Khanal, D.R.; Ogletree, D.F.; Grossman, J.C.; et al. Strain engineering and one-dimensional organization of metal–insulator domains in single-crystal vanadium dioxide beams. Nat. Nanotechnol. 2009, 4, 732–737. [Google Scholar] [CrossRef] [PubMed]
  20. White, S.T.; Thompson, E.A.; Brown, P.F.; Haglund, R.F. Substrate chemistry and lattice effects in vapor transport growth of vanadium dioxide microcrystals. Cryst. Growth Des. 2021, 21, 3770–3778. [Google Scholar] [CrossRef]
  21. Breckenfeld, E.; Kim, H.; Burgess, K.; Charipar, N.; Cheng, S.-F.; Stroud, R.; Piqué, A. Strain effects in epitaxial VO2 thin films on columnar buffer-layer TiO2/Al2O3 virtual substrates. ACS Appl. Mater. Interfaces 2017, 9, 1577–1584. [Google Scholar] [CrossRef]
  22. Shin, K.H.; Bae, J.Y.; Lee, S.Y.; Ahn, D.; Cho, J.; Yoon, J.; Hong, W.-K.; Sohn, J.I. Core-shell heterostructure-enabled stress engineering in vanadium dioxide nanobeams. Appl. Mater. Today 2021, 25, 101244. [Google Scholar] [CrossRef]
  23. Deng, S.; Yu, H.; Park, T.J.; Islam, A.N.M.N.; Manna, S.; Pofelski, A.; Wang, Q.; Zhu, Y.; Sankaranarayanan, S.K.R.S.; Sengupta, A.; et al. Selective area doping for Mott neuromorphic electronics. Sci. Adv. 2023, 9, eade4838. [Google Scholar] [CrossRef]
  24. Chen, S.; Wang, Z.; Ren, H.; Chen, Y.; Yan, W.; Wang, C.; Li, B.; Jiang, J.; Zou, C. Gate-controlled VO2 phase transition for high-performance smart windows. Sci. Adv. 2019, 5, eaav6815. [Google Scholar] [CrossRef] [Green Version]
  25. Strelcov, E.; Tselev, A.; Ivanov, I.; Budai, J.D.; Zhang, J.; Tischler, J.Z.; Kravchenko, I.; Kalinin, S.V.; Kolmakov, A. Doping-based stabilization of the M2 phase in free-standing VO2 nanostructures at room temperature. Nano Lett. 2012, 12, 6198–6205. [Google Scholar] [CrossRef] [PubMed]
  26. Yoon, H.; Choi, M.; Lim, T.-W.; Kwon, H.; Ihm, K.; Kim, J.K.; Choi, S.-Y.; Son, J. Reversible phase modulation and hydrogen storage in multivalent VO2 epitaxial thin films. Nat. Mater. 2016, 15, 1113–1119. [Google Scholar] [CrossRef]
  27. Chen, Y.; Wang, Z.; Chen, S.; Ren, H.; Wang, L.; Zhang, G.; Lu, Y.; Jiang, J.; Zou, C.; Luo, Y. Non-catalytic hydrogenation of VO2 in acid solution. Nat. Commun. 2018, 9, 818. [Google Scholar] [CrossRef] [Green Version]
  28. Guo, H.; Chen, K.; Oh, Y.; Wang, K.; Dejoie, C.; Asif, S.A.S.; Warren, O.L.; Shan, Z.W.; Wu, J.; Minor, A.M. Mechanics and dynamics of the strain-induced M1–M2 structural phase transition in individual VO2 nanowires. Nano Lett. 2011, 11, 3207–3213. [Google Scholar] [CrossRef]
  29. Zhang, Y.-Q.; Chen, K.; Shen, H.; Wang, Y.-C.; Hedhill, M.N.; Zhang, X.; Li, J.; Shan, Z.-W. Achieving room-temperature M2-phase VO2 nanowires for superior thermal actuation. Nano Res. 2021, 14, 4146–4153. [Google Scholar] [CrossRef]
  30. Marezio, M.; McWhan, D.B.; Remeika, J.P.; Dernier, P.D. Structural aspects of the metal-insulator transitions in Cr-doped VO2. Phys. Rev. B 1972, 5, 2541–2551. [Google Scholar] [CrossRef]
  31. Brückner, W.; Gerlach, U.; Thuss, B. Phase diagram of V1-xAlxO2. Phys. Status Solidi A 1977, 40, K131–K134. [Google Scholar] [CrossRef]
  32. Zha, J.; Luo, M.; Ye, M.; Ahmed, T.; Yu, X.; Lien, D.-H.; He, Q.; Lei, D.; Ho, J.C.; Bullock, J.; et al. Infrared photodetectors based on 2D materials and nanophotonics. Adv. Funct. Mater. 2022, 32, 2111970. [Google Scholar]
  33. Cavalleri, A.; Tóth, C.; Siders, C.W.; Squier, J.A. Femtosecond structural dynamics in VO2 during an ultrafast solid-solid phase transition. Phys. Rev. Lett. 2001, 87, 237401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Baum, P.; Yang, D.-S.; Zewail, A.H. 4D visualization of transitional structures in phase transformations by electron diffraction. Science 2007, 318, 788–792. [Google Scholar] [CrossRef]
  35. Zhang, H.-T.; Guo, L.; Stone, G.; Zhang, L.; Zheng, Y.-X.; Freeman, E.; Keefer, D.W.; Chaudhuri, S.; Paik, H.; Moyer, J.A.; et al. Imprinting of local metallic states into VO2 with ultraviolet light. Adv. Funct. Mater. 2016, 26, 6612–6618. [Google Scholar] [CrossRef]
  36. Li, G.; Xie, D.; Zhang, Z.; Zhou, Q.; Zhang, H.; Ni, H.; Wang, J.; Guo, E.-J.; He, M.; Wang, C.; et al. Flexible VO2 films for in-sensor computing with ultraviolet light. Adv. Funct. Mater. 2022, 32, 2203074. [Google Scholar] [CrossRef]
  37. Li, G.; Xie, D.; Zhong, H.; Zhang, Z.; Fu, X.; Zhou, Q.; Li, Q.; Ni, H.; Wang, J.; Guo, E.-J.; et al. Photo-induced non-volatile VO2 phase transition for neuromorphic ultraviolet sensors. Nat. Commun. 2022, 13, 1729. [Google Scholar] [CrossRef]
  38. Richards, P.L. Bolometers for infrared and millimeter waves. J. Appl. Phys. 1994, 76, 1–24. [Google Scholar] [CrossRef]
  39. Xin, Y.; Hu, L.; Ruan, S.; Ye, Z.; Zeng, Y.-J. Phase-transition-induced superior ultraviolet photodetection of a ZnO/VO2 bilayer. J. Mater. Chem. C 2020, 8, 11399–11406. [Google Scholar] [CrossRef]
  40. Basyooni, M.A.; Zaki, S.E.; Shaban, M.; Eker, Y.R.; Yilmaz, M. Efficient MoWO3/VO2/MoS2/Si UV Schottky photodetectors; MoS2 optimization and monoclinic VO2 surface modifications. Sci. Rep. 2020, 10, 15926. [Google Scholar] [CrossRef]
  41. Wu, J.M.; Chang, W.E. Ultrahigh responsivity and external quantum efficiency of an ultraviolet-light photodetector based on a single VO2 microwire. ACS Appl. Mater. Interfaces 2014, 6, 14286–14292. [Google Scholar] [CrossRef]
  42. Oliva, N.; Casu, E.A.; Yan, C.; Krammer, A.; Rosca, T.; Magrez, A.; Stolichnov, I.; Schueler, A.; Martin, O.J.F.; Ionescu, A.M. Van der Waals MoS2/VO2 heterostructure junction with tunable rectifier behavior and efficient photoresponse. Sci. Rep. 2017, 7, 14250. [Google Scholar] [CrossRef] [Green Version]
  43. Luo, H.; Wang, B.; Wang, E.; Wang, X.; Sun, Y.; Li, Q.; Fan, S.; Cheng, C.; Liu, K. Phase-transition modulated, high-performance dual-mode photodetectors based on WSe2/VO2 heterojunctions. Appl. Phys. Rev. 2019, 6, 041407. [Google Scholar] [CrossRef]
  44. Zhou, X.; Zhao, L.; Zhen, W.; Lin, Y.; Wang, C.; Pan, T.; Li, L.; Du, G.; Lu, L.; Cao, X.; et al. Phase-transition-induced VO2 thin film IR photodetector and threshold switching selector for optical neural network applications. Adv. Electron. Mater. 2021, 7, 2001254. [Google Scholar] [CrossRef]
  45. Xie, B.H.; Fu, W.B.; Fei, G.T.; Xu, S.H.; Gao, X.D.; Zhang, L.D. Preparation and enhanced infrared response properties of ordered W-doped VO2 nanowire array. Appl. Surf. Sci. 2018, 436, 1061–1066. [Google Scholar] [CrossRef]
  46. Xiao, Y.; Zhai, Z.H.; Shi, Q.W.; Zhu, L.G.; Li, J.; Huang, W.X.; Yue, F.; Hu, Y.Y.; Peng, Q.X.; Li, Z.R. Ultrafast terahertz modulation characteristic of tungsten doped vanadium dioxide nanogranular film revealed by time-resolved terahertz spectroscopy. Appl. Phys. Lett. 2015, 107, 031906. [Google Scholar] [CrossRef]
  47. Lu, J.; Liu, H.; Deng, S.; Zheng, M.; Wang, Y.; van Kan, J.-A.; Tang, S.H.; Zhang, X.; Sow, C.H.; Mhaisalkar, S.G. Highly sensitive and multispectral responsive phototransistor using tungsten-doped VO2 nanowires. Nanoscale 2014, 6, 7619–7627. [Google Scholar] [CrossRef] [PubMed]
  48. Kumar, M.; Lim, S.; Kim, J.; Seo, H. Picoampere dark current and electro-opto-coupled sub-to-super linear response from Mott-transition enabled infrared photodetector for near-sensor vision processing. Adv. Mater. 2023, 35, 221907. [Google Scholar] [CrossRef]
  49. Guo, X.; Tan, Y.; Hu, Y.; Zafar, Z.; Liu, J.; Zou, J. High quality VO2 thin films synthesized from V2O5 powder for sensitive near-infrared detection. Sci. Rep. 2021, 11, 21749. [Google Scholar] [CrossRef] [PubMed]
  50. Rajeswaran, B.; Tadeo, I.J.; Urnarji, A.M. IR photoresponsvie VO2 thin films and electrically assisted transition prepared by single-step chemical vapor deposition. J. Mater. Chem. C 2020, 8, 12543–12550. [Google Scholar] [CrossRef]
  51. Fu, W.B.; Ma, H.; Wei, Y.; Jiang, K.; Fei, G.T.; Zhang, L.D. Preparation and infrared response properties of vanadium dioxide nanowire/carbon nanotube composite film. J. Mater. Sci. 2017, 52, 7224–7231. [Google Scholar] [CrossRef]
  52. Ma, H.; Wang, Y.; Fu, Y.; Zhang, X. A bottom-up strategy toward a flexible vanadium dioxide/silicon nitride composite film with infrared sensing performance. Nanoscale 2020, 12, 11865–11867. [Google Scholar]
  53. Kabir, S.; Nirantar, S.; Zhu, L.; Ton-That, C.; Jain, S.K.; Kayani, A.B.A.; Murdoch, B.J.; Sriram, S.; Walia, S.; Bhaskaran, M. Phase change vanadium dioxide light sensors. Appl. Mater. Today 2020, 21, 100833. [Google Scholar] [CrossRef]
  54. Umar, Z.A.; Ahmed, R.; Asghar, H.; Liaqat, U.; Fayyaz, A.; Baig, M.A. VO2 thin film based highly responsive and fast VIS/IR photodetector. Mater. Chem. Phys. 2022, 290, 126655. [Google Scholar]
  55. Hong, K.T.; Moon, C.W.; Suh, J.M.; Lee, T.H.; Kim, S.-I.; Lee, S.; Jang, H.W. Daylight-induced metal-insulator transition in Ag decorated vanadium dioxide nanorod arrays. ACS Appl. Mater. Interfaces 2019, 11, 11568–11578. [Google Scholar] [CrossRef]
  56. Hassan, Z.U.; Siddique, R.; Sajjad, S.A.; Umer, Z.A.; Bukhari, S.; Anwar-UI-Haq, M.; Rafique, M.; Raza, S.R.A. Self-powered and temperature-tunable infrared-visible photodetector based on a VO2/Si heterojunction. J. Phys. D Appl. Phys. 2021, 54, 165109. [Google Scholar] [CrossRef]
  57. Jiang, W.; Zheng, T.; Wu, B.; Jiao, H.; Wang, X.; Chen, Y.; Zhang, X.; Peng, M.; Wang, H.; Lin, T.; et al. A versatile photodetector assisted by photovoltaic and bolometric effect. Light Sci. Appl. 2020, 9, 160. [Google Scholar] [PubMed]
  58. Creeden, J.A.; Madaras, S.E.; Beringer, D.B.; Novikova, I.; Lukaszew, R.A. Growth and characterization of vanadium dioxide/niobium doped titanium dioxide heterostructures for ultraviolet detection. Adv. Opt. Mater. 2019, 7, 1901143. [Google Scholar] [CrossRef]
  59. Kim, M.-W.; Jo, Y.-R.; Lee, C.; Moon, W.-J.; Shim, J.H.; Kim, B.-J. Ultrafast infrared photoresponse from heavily hydrogen-doped VO2 Single Crystalline Nanoparticles. Nano Lett. 2020, 20, 2733–2740. [Google Scholar] [CrossRef]
  60. Selman, A.M.; Kadhim, M.J. Fabrication of high sensitivity and fast response IR photodetector based on VO2 nanocrystalline thin films prepared on the silicon substrate. Opt. Mater. 2022, 131, 112664. [Google Scholar] [CrossRef]
  61. Tadeo, I.J.; Krupanidhi, S.B.; Umarji, A.M. Enhanced phase transition and infrared photoresponse characteristics in VO2(M1) thin films synthesized by DC reactive sputtering on different substrates. Mater. Adv. 2021, 2, 3726–3735. [Google Scholar] [CrossRef]
  62. Xu, D.; Zhu, X.; An, J.; Chen, G.; Bao, J.; Xu, X. UV-vis-IR broad spectral photodetectors based on VO2-ZnO nanocrystal films. ACS Omega 2022, 7, 37078–37084. [Google Scholar] [CrossRef] [PubMed]
  63. Ke, Y.; Balin, I.; Wang, N.; Lu, Q.; Tok, A.I.Y.; White, T.J.; Magdassi, S.; Abdulhalim, I.; Long, Y. Two-dimensional SiO2/VO2 photonic crystals with statically visible and dynamically infrared modulated for smart window deployment. ACS Appl. Mater. Interfaces 2016, 8, 33112–33120. [Google Scholar] [CrossRef] [PubMed]
  64. Liang, X.; Chen, M.; Guo, S.; Zhang, L.; Li, F.; Yang, H. Dual-band modulation of visible and near-infrared light transmittance in an all-solution-processed hybrid micro-nano composite film. ACS Appl. Mater. Interfaces 2017, 9, 40810–40819. [Google Scholar] [CrossRef]
  65. Wan, C.; Horak, E.H.; King, J.; Salman, J.; Zhang, Z.; Zhou, Y.; Roney, P.; Gundlach, B.; Ramanathan, S.; Goldsmith, R.H.; et al. Limiting optical diodes enabled by the phase transition of vanadium dioxide. ACS Photonics 2018, 5, 2688–2692. [Google Scholar] [CrossRef] [Green Version]
  66. Shu, F.-Z.; Yu, F.-F.; Peng, R.-W.; Zhu, Y.-Y.; Xiong, B.; Fan, R.-H.; Wang, Z.-H.; Liu, Y.; Wang, M. Dynamic plasmonic color generation based on phase transition of vanadium dioxide. Adv. Opt. Mater. 2018, 6, 1700939. [Google Scholar] [CrossRef]
  67. Duan, X.; White, S.T.; Cui, Y.; Neubrech, F.; Gao, Y.; Haglund, R.F.; Liu, N. Reconfigurable multistate optical systems enabled by VO2 phase transitions. ACS Photonics 2020, 7, 2958–2965. [Google Scholar] [CrossRef]
  68. In, S.; Cho, J.; Park, J.; Kim, S.Y.; Kim, H.-T.; Noh, T.W.; Park, N. Self-organized gold network-vanadium dioxide hybrid film for dynamic modulation of visible-to-near-infrared light. Adv. Photonics Res. 2020, 1, 2000050. [Google Scholar] [CrossRef]
  69. Zhang, H.-T.; Zhang, L.; Mukherjee, D.; Zheng, Y.-X.; Haislmaier, R.C.; Alem, N.; Engel-Herbert, R. Wafer-scale growth of VO2 thin films using a combinatorial approach. Nat. Commun. 2015, 6, 8475. [Google Scholar] [CrossRef] [Green Version]
  70. Goi, E.; Zhang, Q.; Chen, X.; Luan, H.; Gu, M. Perspective on photonic memristive neuromorphic computing. PhotoniX 2020, 1, 3. [Google Scholar] [CrossRef] [Green Version]
  71. Yuan, R.; Duan, Q.; Tiw, P.J.; Li, G.; Xiao, Z.; Jing, Z.; Yang, K.; Liu, C.; Ge, C.; Huang, R.; et al. A calibratable sensory neuron based on epitaxial VO2 for spike-based neuromorphic multisensory system. Nat. Commun. 2022, 13, 3973. [Google Scholar] [CrossRef]
  72. Gea, C.; Lia, G.; Zhou, Q.-l.; Du, J.-y.; Guo, E.-j.; He, M.; Wang, C.; Yang, G.-z.; Jin, K.-j. Gating-induced reversible HxVO2 phase transformations for neuromorphic computing. Nano Energy 2020, 67, 104268. [Google Scholar] [CrossRef]
  73. Oh, S.; Shi, Y.; del Valle, J.; Salev, P.; Lu, Y.; Huang, Z.; Kalcheim, Y.; Schuller, I.K.; Kuzum, D. Energy-efficient Mott activation neuron for full-hardware implementation of neural networks. Nat. Nanotechnol. 2021, 16, 680–687. [Google Scholar] [CrossRef]
  74. Nikoo, M.S.; Soleimanzadeh, R.; Krammer, A.; Marega, G.M.; Park, Y.; Son, J.; Schueler, A.; Kis, A.; Moll, P.J.W.; Matioli, E. Electrical control of glass-like dynamics in vanadium dioxide for data storage and processing. Nat. Electron. 2022, 5, 596–603. [Google Scholar] [CrossRef]
  75. Tang, K.; Dong, K.; Li, J.; Gordon, M.P.; Reichertz, F.G.; Kim, H.; Rho, Y.; Wang, Q.; Lin, C.-Y.; Grigoropoulos, C.P.; et al. Temperature-adaptive radiative coating for all-season household thermal regulation. Science 2021, 374, 1504–1509. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, S.; Jiang, T.; Yang, Y.M.; Tan, G.; Long, Y. Scalable thermochromic smart windows with passive radiative cooling regulation. Science 2021, 374, 1501–1504. [Google Scholar] [CrossRef]
  77. Liu, M.; Li, X.; Li, L.; Li, L.; Zhao, S.; Lu, K.; Chen, K.; Zhu, J.; Zhou, T.; Hu, C.; et al. Continuous photothermal and radiative cooling energy harvesting by VO2 smart coatings with switchable broadband infrared emission. ACS Nano 2023, 17, 9501–9509. [Google Scholar] [CrossRef]
Figure 2. VO2-based device applications related to the category of input stimuli and output response. Reproduced with permission from [9], Copyright 2018, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 2. VO2-based device applications related to the category of input stimuli and output response. Reproduced with permission from [9], Copyright 2018, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Sensors 23 06715 g002
Figure 3. (a) Schematics of synthesis process of VO2 microbeams in stage 1 (T < 850 °C) and stage 2 (T = 850 °C) under different oxygen partial pressures ( P O 2 ). (b) Working mechanism for a single-crystalline W-doped VO2 actuator. The array of yellow gradient bars indicates the traces of actuation process upon heating. (c) Optical images of a single-crystalline W-doped VO2 actuator at 30, 54, and 60 °C (scale bars are 40 μm). Temperature-dependent optical images of another single-crystalline VO2 actuator showing a clear domain evolution process (scale bars are 5 μm). Black dashed lines indicate the positions of domain walls between M2 and M1/T domains. Reproduced with permission from [16], Copyright 2021, Springer Nature.
Figure 3. (a) Schematics of synthesis process of VO2 microbeams in stage 1 (T < 850 °C) and stage 2 (T = 850 °C) under different oxygen partial pressures ( P O 2 ). (b) Working mechanism for a single-crystalline W-doped VO2 actuator. The array of yellow gradient bars indicates the traces of actuation process upon heating. (c) Optical images of a single-crystalline W-doped VO2 actuator at 30, 54, and 60 °C (scale bars are 40 μm). Temperature-dependent optical images of another single-crystalline VO2 actuator showing a clear domain evolution process (scale bars are 5 μm). Black dashed lines indicate the positions of domain walls between M2 and M1/T domains. Reproduced with permission from [16], Copyright 2021, Springer Nature.
Sensors 23 06715 g003
Figure 5. (a) Scanning transmission electron microscopy images of the (100) fully hydrogenated HVO2 film grown on TiO2 (100) substrates and the temperature-dependent sheet resistance in HxVO2 films with increasing hydrogen content (x) from plot I to plot IV, indicating a transition from insulator (M-VO2)-to-metal (HxVO2)-to-insulator (HVO2) (upper panel). A schematic illustration of the contiguous electron–proton co-doping mechanism with the metal-acid treatment of a semiconductor (lower panel). Reproduced with permission from [5], Copyright 2018, Springer Nature. (b) The gating diagram and hydrogen ion movement under gating control for the VO2 device with source, drain, and gate electrodes. The reversible insulator-metal-insulator tristate phase transitions of VO2 by tuning hydrogenating level with positive or negative gating voltages. Reproduced with permission from [24], Copyright 2019, American Association for the Advancement of Science. (c) Temperature-composition phase diagram showing influence of doping and uniaxial stress along the [110]R crystallographic direction of the V1−xMxO2+y. Reproduced with permission from [25], Copyright 2012, American Chemical Society.
Figure 5. (a) Scanning transmission electron microscopy images of the (100) fully hydrogenated HVO2 film grown on TiO2 (100) substrates and the temperature-dependent sheet resistance in HxVO2 films with increasing hydrogen content (x) from plot I to plot IV, indicating a transition from insulator (M-VO2)-to-metal (HxVO2)-to-insulator (HVO2) (upper panel). A schematic illustration of the contiguous electron–proton co-doping mechanism with the metal-acid treatment of a semiconductor (lower panel). Reproduced with permission from [5], Copyright 2018, Springer Nature. (b) The gating diagram and hydrogen ion movement under gating control for the VO2 device with source, drain, and gate electrodes. The reversible insulator-metal-insulator tristate phase transitions of VO2 by tuning hydrogenating level with positive or negative gating voltages. Reproduced with permission from [24], Copyright 2019, American Association for the Advancement of Science. (c) Temperature-composition phase diagram showing influence of doping and uniaxial stress along the [110]R crystallographic direction of the V1−xMxO2+y. Reproduced with permission from [25], Copyright 2012, American Chemical Society.
Sensors 23 06715 g005
Figure 6. Spectral range for the typical photodetectors, and their potential applications corresponding to each wavelength range (UV, Visible, and IR). IR is divided into four different ranges; NIR (0.76–1 μm), SWIR (1–3 μm), MWIR (3–5 μm), and LWIR (8–12 μm).
Figure 6. Spectral range for the typical photodetectors, and their potential applications corresponding to each wavelength range (UV, Visible, and IR). IR is divided into four different ranges; NIR (0.76–1 μm), SWIR (1–3 μm), MWIR (3–5 μm), and LWIR (8–12 μm).
Sensors 23 06715 g006
Figure 8. VO2-based visible photodetectors. (a) MoS2/VO2-heterojunction-based visible photodetector by using photovoltaic effect. Reproduced with permission from [42], Copyright 2017, Springer Nature. (b) WSe2/VO2 heterojunction-based visible photodetector by using both photovoltaic and photoconductive effects. Reproduced with permission from [43], Copyright 2019, American Institute of Physics.
Figure 8. VO2-based visible photodetectors. (a) MoS2/VO2-heterojunction-based visible photodetector by using photovoltaic effect. Reproduced with permission from [42], Copyright 2017, Springer Nature. (b) WSe2/VO2 heterojunction-based visible photodetector by using both photovoltaic and photoconductive effects. Reproduced with permission from [43], Copyright 2019, American Institute of Physics.
Sensors 23 06715 g008
Figure 10. VO2-based IR photodetectors. (a) VO2-based SWIR photodetector using photoconductive effect. Reproduced with permission from [50], Copyright 2020, The Royal Society of Chemistry. (b) VO2 nanowire/carbon nanotube (CNT) composite-based IR photodetector via photobolometric effect. Reproduced with permission from [51], Copyright 2017, Springer.
Figure 10. VO2-based IR photodetectors. (a) VO2-based SWIR photodetector using photoconductive effect. Reproduced with permission from [50], Copyright 2020, The Royal Society of Chemistry. (b) VO2 nanowire/carbon nanotube (CNT) composite-based IR photodetector via photobolometric effect. Reproduced with permission from [51], Copyright 2017, Springer.
Sensors 23 06715 g010
Table 1. Synthesis methods of nanostructured VO2 [3,12,13,14,15].
Table 1. Synthesis methods of nanostructured VO2 [3,12,13,14,15].
Synthesis MethodAdvantagesLimitations
Sol-gel
  • Simple, cheap, and precise control of variables
  • Creation of amorphous materials
  • High chemical reactivity of precursors
  • Post-deposition heat treatment
  • Long preparation and curing times
  • Poor adhesion
Hydrothermal
synthesis
  • Easy regulation of size, shape, and composition
  • Higher chemical purity
  • Expensive autoclaves
  • Longer reaction times
  • Poor adhesion
PLD 1
  • Growth of compatible and consistent films
  • No restrictions on the types of PLD targets
  • Slow process and high cost preparation
  • Not suitable for vast area deposition
  • Material loss due to evaporation
Sputtering
  • Deposition of stable and uniform films
  • Easy deposition of hybrid materials
  • Good film adhesion
  • Expensive and sophisticated equipment
  • Low purity
CVD 2
  • Growth of nanostructures with various shapes
  • High-quality crystalline
  • Small-scale crystals at low cost and in a short time
  • Rapid and inexpensive method
  • Numerous growth parameters: growth time, temperature, precursors, ambient conditions, and substrate, etc.
  • Influences of growth substrates: crystal size, shape, and orientation, etc.
1 PLD: pulsed laser deposition. 2 CVD: chemical vapor deposition.
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

Yoon, J.; Hong, W.-K.; Kim, Y.; Park, S.-Y. Nanostructured Vanadium Dioxide Materials for Optical Sensing Applications. Sensors 2023, 23, 6715. https://doi.org/10.3390/s23156715

AMA Style

Yoon J, Hong W-K, Kim Y, Park S-Y. Nanostructured Vanadium Dioxide Materials for Optical Sensing Applications. Sensors. 2023; 23(15):6715. https://doi.org/10.3390/s23156715

Chicago/Turabian Style

Yoon, Jongwon, Woong-Ki Hong, Yonghun Kim, and Seung-Young Park. 2023. "Nanostructured Vanadium Dioxide Materials for Optical Sensing Applications" Sensors 23, no. 15: 6715. https://doi.org/10.3390/s23156715

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